Sample records for vapor-induced phase separation

  1. Vapor-liquid phase separator studies

    NASA Technical Reports Server (NTRS)

    Yuan, S. W. K.; Lee, J. M.; Kim, Y. I.; Hepler, W. A.; Frederking, T. H. K.

    1983-01-01

    Porous plugs serve as both entropy rejection devices and phase separation components separating the vapor phase on the downstream side from liquid Helium 2 upstream. The liquid upstream is the cryo-reservoir fluid needed for equipment cooling by means of Helium 2, i.e Helium-4 below its lambda temperature in near-saturated states. The topics outlined are characteristic lengths, transport equations and plug results.

  2. Vapor-liquid phase separator permeability results

    NASA Technical Reports Server (NTRS)

    Yuan, S. W. K.; Frederking, T. H. K.

    1981-01-01

    Continued studies are described in the area of vapor-liquid phase separator work with emphasis on permeabilities of porous sintered plugs (stainless steel, nominal pore size 2 micrometer). The temperature dependence of the permeability has been evaluated in classical fluid using He-4 gas at atmospheric pressure and in He-2 on the basis of a modified, thermosmotic permeability of the normal fluid.

  3. Superfluid helium 2 liquid-vapor phase separation: Technology assessment

    NASA Technical Reports Server (NTRS)

    Lee, J. M.

    1984-01-01

    A literature survey of helium 2 liquid vapor phase separation is presented. Currently, two types of He 2 phase separators are being investigated: porous, sintered metal plugs and the active phase separator. The permeability K(P) shows consistency in porous plug geometric characterization. Both the heat and mass fluxes increase with K(P). Downstream pressure regulation to adjust for varying heat loads and both temperatures is possible. For large dynamic heat loads, the active phase separator shows a maximum heat rejection rate of up to 2 W and bath temperature stability of 0.1 mK. Porous plug phase separation performance should be investigated for application to SIRTF and, in particular, that plugs of from 10 to the minus ninth square centimeters to 10 to the minus eighth square centimeters in conjunction with downstream pressure regulation be studied.

  4. Space cryogenics components based on the thermomechanical effect - Vapor-liquid phase separation

    NASA Technical Reports Server (NTRS)

    Yuan, S. W. K.; Frederking, T. H. K.

    1989-01-01

    Applications of the thermomechanical effect has been qualified including incorporation in large-scale space systems in the area of vapor-liquid phase separation (VLPS). The theory of the porous-plug phase separator is developed for the limit of a high thermal impedance of the solid-state grains. Extensions of the theory of nonlinear turbulent flow are presented based on experimental results.

  5. Vapors-liquid phase separator. [infrared telescope heat sink

    NASA Technical Reports Server (NTRS)

    Frederking, T. H. K.; Brown, G. S.; Chuang, C.; Kamioka, Y.; Kim, Y. I.; Lee, J. M.; Yuan, S. W. K.

    1980-01-01

    The use of porous plugs, mostly with in the form of passive devices with constant area were considered as vapor-liquid phase separators for helium 2 storage vessels under reduced gravity. The incorporation of components with variable cross sectional area as a method of flow rate modification was also investigated. A particular device which uses a shutter-type system for area variation was designed and constructed. This system successfully permitted flor rate changes of up to plus or minus 60% from its mean value.

  6. Simplified thermodynamic functions for vapor-liquid phase separation and fountain effect pumps

    NASA Technical Reports Server (NTRS)

    Yuan, S. W. K.; Hepler, W. A.; Frederking, T. H. K.

    1984-01-01

    He-4 fluid handling devices near 2 K require novel components for non-Newtonian fluid transport in He II. Related sizing of devices has to be based on appropriate thermophysical property functions. The present paper presents simplified equilibrium state functions for porous media components which serve as vapor-liquid phase separators and fountain effect pumps.

  7. Motility-Induced Phase Separation

    NASA Astrophysics Data System (ADS)

    Cates, Michael E.; Tailleur, Julien

    2015-03-01

    Self-propelled particles include both self-phoretic synthetic colloids and various microorganisms. By continually consuming energy, they bypass the laws of equilibrium thermodynamics. These laws enforce the Boltzmann distribution in thermal equilibrium: The steady state is then independent of kinetic parameters. In contrast, self-propelled particles tend to accumulate where they move more slowly. They may also slow down at high density for either biochemical or steric reasons. This creates positive feedback, which can lead to motility-induced phase separation (MIPS) between dense and dilute fluid phases. At leading order in gradients, a mapping relates variable-speed, self-propelled particles to passive particles with attractions. This deep link to equilibrium phase separation is confirmed by simulations but generally breaks down at higher order in gradients: New effects, with no equilibrium counterpart, then emerge. We give a selective overview of the fast-developing field of MIPS, focusing on theory and simulation but including a brief speculative survey of its experimental implications.

  8. Laser-induced phase separation of silicon carbide

    PubMed Central

    Choi, Insung; Jeong, Hu Young; Shin, Hyeyoung; Kang, Gyeongwon; Byun, Myunghwan; Kim, Hyungjun; Chitu, Adrian M.; Im, James S.; Ruoff, Rodney S.; Choi, Sung-Yool; Lee, Keon Jae

    2016-01-01

    Understanding the phase separation mechanism of solid-state binary compounds induced by laser–material interaction is a challenge because of the complexity of the compound materials and short processing times. Here we present xenon chloride excimer laser-induced melt-mediated phase separation and surface reconstruction of single-crystal silicon carbide and study this process by high-resolution transmission electron microscopy and a time-resolved reflectance method. A single-pulse laser irradiation triggers melting of the silicon carbide surface, resulting in a phase separation into a disordered carbon layer with partially graphitic domains (∼2.5 nm) and polycrystalline silicon (∼5 nm). Additional pulse irradiations cause sublimation of only the separated silicon element and subsequent transformation of the disordered carbon layer into multilayer graphene. The results demonstrate viability of synthesizing ultra-thin nanomaterials by the decomposition of a binary system. PMID:27901015

  9. Laser vaporization/ionization interface for coupling microscale separation techniques with mass spectrometry

    DOEpatents

    Yeung, Edward S.; Chang, Yu-chen

    1999-06-29

    The present invention provides a laser-induced vaporization and ionization interface for directly coupling microscale separation processes to a mass spectrometer. Vaporization and ionization of the separated analytes are facilitated by the addition of a light-absorbing component to the separation buffer or solvent.

  10. Laser vaporization/ionization interface for coupling microscale separation techniques with mass spectrometry

    DOEpatents

    Yeung, E.S.; Chang, Y.C.

    1999-06-29

    The present invention provides a laser-induced vaporization and ionization interface for directly coupling microscale separation processes to a mass spectrometer. Vaporization and ionization of the separated analytes are facilitated by the addition of a light-absorbing component to the separation buffer or solvent. 8 figs.

  11. An Experimental Visualization and Image Analysis of Electrohydrodynamically Induced Vapor-Phase Silicon Oil Flow under DC Corona Discharge

    NASA Astrophysics Data System (ADS)

    Ohyama, Ryu-Ichiro; Fukumoto, Masaru

    A DC corona discharge induced electrohydrodynamic (EHD) flow phenomenon for a multi-phase fluid containing a vapor-phase dielectric liquid in the fresh air was investigated. The experimental electrode system was a simple arrangement of needle-plate electrodes for the corona discharges and high-resistivity silicon oil was used as the vapor-phase liquid enclosure. The qualitative observation of EHD flow patterns was conducted by an optical processing on computer tomography and the time-series of discharge current pulse generations at corona discharge electrode were measured simultaneously. These experimental results were analyzed in relationship between the EHD flow motions and the current pulse generations in synchronization. The current pulses and the EHD flow motions from the corona discharge electrode presented a continuous mode similar to the ionic wind in the fresh air and an intermittent mode. In the intermittent mode, the observed EHD flow motion was synchronized with the separated discharge pulse generations. From these experimental results, it was expected that the existence of silicon oil vapor trapped charges gave an occasion to the intermittent generations of the discharge pulses and the secondary EHD flow.

  12. Electron irradiation induced phase separation in a sodium borosilicate glass

    NASA Astrophysics Data System (ADS)

    Sun, K.; Wang, L. M.; Ewing, R. C.; Weber, W. J.

    2004-06-01

    Electron irradiation induced phase separation in a sodium borosilicate glass was studied in situ by analytical electron microscopy. Distinctly separate phases that are rich in boron and silicon formed at electron doses higher than 4.0 × 10 11 Gy during irradiation. The separated phases are still in amorphous states even at a much high dose (2.1 × 10 12 Gy). It indicates that most silicon atoms remain tetrahedrally coordinated in the glass during the entire irradiation period, except some possible reduction to amorphous silicon. The particulate B-rich phase that formed at high dose was identified as amorphous boron that may contain some oxygen. Both ballistic and ionization processes may contribute to the phase separation.

  13. Re-entrant phase behavior for systems with competition between phase separation and self-assembly

    NASA Astrophysics Data System (ADS)

    Reinhardt, Aleks; Williamson, Alexander J.; Doye, Jonathan P. K.; Carrete, Jesús; Varela, Luis M.; Louis, Ard A.

    2011-03-01

    In patchy particle systems where there is a competition between the self-assembly of finite clusters and liquid-vapor phase separation, re-entrant phase behavior can be observed, with the system passing from a monomeric vapor phase to a region of liquid-vapor phase coexistence and then to a vapor phase of clusters as the temperature is decreased at constant density. Here, we present a classical statistical mechanical approach to the determination of the complete phase diagram of such a system. We model the system as a van der Waals fluid, but one where the monomers can assemble into monodisperse clusters that have no attractive interactions with any of the other species. The resulting phase diagrams show a clear region of re-entrance. However, for the most physically reasonable parameter values of the model, this behavior is restricted to a certain range of density, with phase separation still persisting at high densities.

  14. Polymer-induced phase separation and crystallization in immunoglobulin G solutions.

    PubMed

    Li, Jianguo; Rajagopalan, Raj; Jiang, Jianwen

    2008-05-28

    We study the effects of the size of polymer additives and ionic strength on the phase behavior of a nonglobular protein-immunoglobulin G (IgG)-by using a simple four-site model to mimic the shape of IgG. The interaction potential between the protein molecules consists of a Derjaguin-Landau-Verwey-Overbeek-type colloidal potential and an Asakura-Oosawa depletion potential arising from the addition of polymer. Liquid-liquid equilibria and fluid-solid equilibria are calculated by using the Gibbs ensemble Monte Carlo technique and the Gibbs-Duhem integration (GDI) method, respectively. Absolute Helmholtz energy is also calculated to get an initial coexisting point as required by GDI. The results reveal a nonmonotonic dependence of the critical polymer concentration rho(PEG) (*) (i.e., the minimum polymer concentration needed to induce liquid-liquid phase separation) on the polymer-to-protein size ratio q (equivalently, the range of the polymer-induced depletion interaction potential). We have developed a simple equation for estimating the minimum amount of polymer needed to induce the liquid-liquid phase separation and show that rho(PEG) (*) approximately [q(1+q)(3)]. The results also show that the liquid-liquid phase separation is metastable for low-molecular weight polymers (q=0.2) but stable at large molecular weights (q=1.0), thereby indicating that small sizes of polymer are required for protein crystallization. The simulation results provide practical guidelines for the selection of polymer size and ionic strength for protein phase separation and crystallization.

  15. Cross-stacked carbon nanotubes assisted self-separation of free-standing GaN substrates by hydride vapor phase epitaxy

    NASA Astrophysics Data System (ADS)

    Wei, Tongbo; Yang, Jiankun; Wei, Yang; Huo, Ziqiang; Ji, Xiaoli; Zhang, Yun; Wang, Junxi; Li, Jinmin; Fan, Shoushan

    2016-06-01

    We report a novel method to fabricate high quality 2-inch freestanding GaN substrate grown on cross-stacked carbon nanotubes (CSCNTs) coated sapphire by hydride vapor phase epitaxy (HVPE). As nanoscale masks, these CSCNTs can help weaken the interface connection and release the compressive stress by forming voids during fast coalescence and also block the propagation of threading dislocations (TDs). During the cool-down process, thermal stress-induced cracks are initiated at the CSCNTs interface with the help of air voids and propagated all over the films which leads to full self-separation of FS-GaN substrate. Raman and photoluminescence spectra further reveal the stress relief and crystalline improvement of GaN with CSCNTs. It is expected that the efficient, low cost and mass-producible technique may enable new applications for CNTs in nitride optoelectronic fields.

  16. Cross-stacked carbon nanotubes assisted self-separation of free-standing GaN substrates by hydride vapor phase epitaxy.

    PubMed

    Wei, Tongbo; Yang, Jiankun; Wei, Yang; Huo, Ziqiang; Ji, Xiaoli; Zhang, Yun; Wang, Junxi; Li, Jinmin; Fan, Shoushan

    2016-06-24

    We report a novel method to fabricate high quality 2-inch freestanding GaN substrate grown on cross-stacked carbon nanotubes (CSCNTs) coated sapphire by hydride vapor phase epitaxy (HVPE). As nanoscale masks, these CSCNTs can help weaken the interface connection and release the compressive stress by forming voids during fast coalescence and also block the propagation of threading dislocations (TDs). During the cool-down process, thermal stress-induced cracks are initiated at the CSCNTs interface with the help of air voids and propagated all over the films which leads to full self-separation of FS-GaN substrate. Raman and photoluminescence spectra further reveal the stress relief and crystalline improvement of GaN with CSCNTs. It is expected that the efficient, low cost and mass-producible technique may enable new applications for CNTs in nitride optoelectronic fields.

  17. Cross-stacked carbon nanotubes assisted self-separation of free-standing GaN substrates by hydride vapor phase epitaxy

    PubMed Central

    Wei, Tongbo; Yang, Jiankun; Wei, Yang; Huo, Ziqiang; Ji, Xiaoli; Zhang, Yun; Wang, Junxi; Li, Jinmin; Fan, Shoushan

    2016-01-01

    We report a novel method to fabricate high quality 2-inch freestanding GaN substrate grown on cross-stacked carbon nanotubes (CSCNTs) coated sapphire by hydride vapor phase epitaxy (HVPE). As nanoscale masks, these CSCNTs can help weaken the interface connection and release the compressive stress by forming voids during fast coalescence and also block the propagation of threading dislocations (TDs). During the cool-down process, thermal stress-induced cracks are initiated at the CSCNTs interface with the help of air voids and propagated all over the films which leads to full self-separation of FS-GaN substrate. Raman and photoluminescence spectra further reveal the stress relief and crystalline improvement of GaN with CSCNTs. It is expected that the efficient, low cost and mass-producible technique may enable new applications for CNTs in nitride optoelectronic fields. PMID:27340030

  18. Melt-Vapor Phase Diagram of the Te-S System

    NASA Astrophysics Data System (ADS)

    Volodin, V. N.; Trebukhov, S. A.; Kenzhaliyev, B. K.; Nitsenko, A. V.; Burabaeva, N. M.

    2018-03-01

    The values of partial pressure of saturated vapor of the constituents of the Te-S system are determined from boiling points. The boundaries of the melt-vapor phase transition at atmospheric pressure and in vacuum of 2000 and 100 Pa are calculated on the basis of partial pressures. A phase diagram that includes vapor-liquid equilibrium fields whose boundaries allow us to assess the behavior of elements upon distillation fractioning is plotted. It is established that the separation of elements is possible at the first evaporation-condensation cycle. Complications can be caused by crystallization of a sulfur solid solution in tellurium.

  19. Hydrogen isotope systematics of phase separation in submarine hydrothermal systems: Experimental calibration and theoretical models

    USGS Publications Warehouse

    Berndt, M.E.; Seal, R.R.; Shanks, Wayne C.; Seyfried, W.E.

    1996-01-01

    Hydrogen isotope fractionation factors were measured for coexisting brines and vapors formed by phase separation of NaCl/H2O fluids at temperatures ranging from 399-450??C and pressures from 277-397 bars. It was found that brines are depleted in D compared to coexisting vapors at all conditions studied. The magnitude of hydrogen isotope fractionation is dependent on the relative amounts of Cl in the two phases and can be empirically correlated to pressure using the following relationship: 1000 ln ??(vap-brine) = 2.54(??0.83) + 2.87(??0.69) x log (??P), where ??(vap-brine) is the fractionation factor and ??P is a pressure term representing distance from the critical curve in the NaCl/H2O system. The effect of phase separation on hydrogen isotope distribution in subseafloor hydrothermal systems depends on a number of factors, including whether phase separation is induced by heating at depth or by decompression of hydrothermal fluids ascending to the seafloor. Phase separation in most subseafloor systems appears to be a simple process driven by heating of seawater to conditions within the two-phase region, followed by segregation and entrainment of brine or vapor into a seawater dominated system. Resulting vent fluids exhibit large ranges in Cl concentration with no measurable effect on ??D. Possible exceptions to this include hydrothermal fluids venting at Axial and 9??N on the East Pacific Rise. High ??D values of low Cl fluids venting at Axial are consistent with phase separation taking place at relatively shallow levels in the oceanic crust while negative ??D values in some low Cl fluids venting at 9??N suggest involvement of a magmatic fluid component or phase separation of D-depleted brines derived during previous hydrothermal activity.

  20. Texas A&M vortex type phase separator

    NASA Astrophysics Data System (ADS)

    Best, Frederick

    2000-01-01

    Phase separation is required for regenerative biological and chemical process systems as well as thermal transport and rejection systems. Liquid and gas management requirements for future spacecraft will demand small, passive systems able to operate over wide ranges of inlet qualities. Conservation and recycling of air and water is a necessary part of the construction and operation of the International Space Station as well as future long duration space missions. Space systems are sensitive to volume, mass, and power. Therefore, it is necessary to develop a method to recycle wastewater with minimal power consumption. Regenerative life support systems currently being investigated require phase separation to separate the liquid from the gas produced. The microgravity phase separator designed and fabricated at Texas A&M University relies on centripetal driven buoyancy forces to form a gas-liquid vortex within a fixed, right-circular cylinder. Two-phase flow is injected tangentially along the inner wall of this cylinder producing a radial acceleration gradient. The gradient produced from the intrinsic momentum of the injected mixture results in a rotating flow that drives the buoyancy process by the production of a hydrostatic pressure gradient. Texas A&M has flown several KC-135 flights with separator. These flights have included scaling studies, stability and transient investigations, and tests for inventory instrumentation. Among the hardware tested have been passive devices for separating mixed vapor/liquid streams into single-phase streams of vapor only and liquid only. .

  1. Atomic vapor laser isotope separation process

    DOEpatents

    Wyeth, R.W.; Paisner, J.A.; Story, T.

    1990-08-21

    A laser spectroscopy system is utilized in an atomic vapor laser isotope separation process. The system determines spectral components of an atomic vapor utilizing a laser heterodyne technique. 23 figs.

  2. Laser-induced separation of hydrogen isotopes in the liquid phase

    DOEpatents

    Freund, Samuel M.; Maier, II, William B.; Beattie, Willard H.; Holland, Redus F.

    1980-01-01

    Hydrogen isotope separation is achieved by either (a) dissolving a hydrogen-bearing feedstock compound in a liquid solvent, or (b) liquefying a hydrogen-bearing feedstock compound, the liquid phase thus resulting being kept at a temperature at which spectral features of the feedstock relating to a particular hydrogen isotope are resolved, i.e., a clear-cut isotope shift is delineated, irradiating the liquid phase with monochromatic radiation of a wavelength which at least preferentially excites those molecules of the feedstock containing a first hydrogen isotope, inducing photochemical reaction in the excited molecules, and separating the reaction product containing the first isotope from the liquid phase.

  3. Polymerization- and Solvent-Induced Phase Separation in Hydrophilic-rich Dentin Adhesive Mimic

    PubMed Central

    Abedin, Farhana; Ye, Qiang; Good, Holly J; Parthasarathy, Ranganathan; Spencer, Paulette

    2014-01-01

    Current dental resin undergoes phase separation into hydrophobic-rich and hydrophilic-rich phases during infiltration of the over-wet demineralized collagen matrix. Such phase separation undermines the integrity and durability of the bond at the composite/tooth interface. This study marks the first time that the polymerization kinetics of model hydrophilic-rich phase of dental adhesive has been determined. Samples were prepared by adding varying water content to neat resins made from 95 and 99wt% hydroxyethylmethacrylate (HEMA) and 5 and 1wt% (2,2-bis[4-(2-hydroxy-3-methacryloxypropoxy)phenyl1]-propane (BisGMA) prior to light curing. Viscosity of the formulations decreased with increased water content. The photo-polymerization kinetics study was carried out by time-resolved FTIR spectrum collector. All of the samples exhibited two-stage polymerization behavior which has not been reported previously for dental resin formulation. The lowest secondary rate maxima were observed for water content of 10-30%wt. Differential scanning calorimetry (DSC) showed two glass transition temperatures for the hydrophilic-rich phase of dental adhesive. The DSC results indicate that the heterogeneity within the final polymer structure decreased with increased water content. The results suggest a reaction mechanism involving both polymerization-induced phase separation (PIPs) and solvent-induced phase separation (SIPs) for the model hydrophilic-rich phase of dental resin. PMID:24631658

  4. Synthesis of TiO2 Nanoparticles from Ilmenite Through the Mechanism of Vapor-Phase Reaction Process by Thermal Plasma Technology

    NASA Astrophysics Data System (ADS)

    Samal, Sneha

    2017-11-01

    Synthesis of nanoparticles of TiO2 was carried out by non-transferred arc thermal plasma reactor using ilmenite as the precursor material. The powder ilmenite was vaporized at high temperature in plasma flame and converted to a gaseous state of ions in the metastable phase. On cooling, chamber condensation process takes place on recombination of ions for the formation of nanoparticles. The top-to-bottom approach induces the disintegration of complex ilmenite phases into simpler compounds of iron oxide and titanium dioxide phases. The vapor-phase reaction mechanism was carried out in thermal plasma zone for the synthesis of nanoparticles from ilmenite compound in a plasma reactor. The easy separation of iron particles from TiO2 was taken place in the plasma chamber with deposition of light TiO2 particles at the top of the cooling chamber and iron particles at the bottom. The dissociation and combination process of mechanism and synthesis are studied briefly in this article. The product TiO2 nanoparticle shows the purity with a major phase of rutile content. TiO2 nanoparticles produced in vapor-phase reaction process shows more photo-induced capacity.

  5. Kinetics of motility-induced phase separation and swim pressure

    NASA Astrophysics Data System (ADS)

    Patch, Adam; Yllanes, David; Marchetti, M. Cristina

    Active Brownian particles (ABPs) represent a minimal model of active matter consisting of self-propelled spheres with purely repulsive interactions and rotational noise. We correlate the time evolution of the mean pressure towards its steady state value with the kinetics of motility-induced phase separation. For parameter values corresponding to phase separated steady states, we identify two dynamical regimes. The pressure grows monotonically in time during the initial regime of rapid cluster formation, overshooting its steady state value and then quickly relaxing to it, and remains constant during the subsequent slower period of cluster coalescence and coarsening. The overshoot is a distinctive feature of active systems. NSF-DMR-1305184, NSF-DGE-1068780, ACI-1341006, FIS2015-65078-C02, BIFI-ZCAM.

  6. Dynamics of polymerization induced phase separation in reactive polymer blends

    NASA Astrophysics Data System (ADS)

    Lee, Jaehyung

    Mechanisms and dynamics of phase decomposition following polymerization induced phase separation (PIPS) of reactive polymer blends have been investigated experimentally and theoretically. The phenomenon of PIPS is a non-equilibrium and non-linear dynamic process. The mechanism of PIPS has been thought to be a nucleation and growth (NG) type originally, however, newer results indicate spinodal decomposition (SD). In PIPS, the coexistence curve generally passes through the reaction temperature at off-critical compositions, thus phase separation has to be initiated first in the metastable region where nucleation occurs. When the system farther drifts from the metastable to unstable region, the NG structure transforms to the SD bicontinuous morphology. The crossover behavior of PIPS may be called nucleation initiated spinodal decomposition (NISD). The formation of newer domains between the existing ones is responsible for the early stage of PIPS. Since PIPS is non- equilibrium kinetic process, it would not be surprising to discern either or both structures. The phase separation dynamics of DGEBA/CTBN mixtures having various kinds of curing agents from low reactivity to high reactivity and various amount of curing agents were examined at various reaction temperatures. The phase separation behavior was monitored by a quantity of scattered light intensity experimentally and by a quantity of collective structure factor numerically. Prior to the study of phase separation dynamics, a preliminary investigation on the isothermal cure behavior of the mixtures were executed in order to determine reaction kinetics parameters. The cure behavior followed the overall second order reaction kinetics. Next, based on the knowledge obtained from the phase separation dynamics study of DGEBA/CTBN mixtures, the phase separation dynamics of various composition of DGEBA/R45EPI mixtures having MDA as a curing agent were investigated. The phase separation behavior was quite dependent upon the

  7. Phase-separation induced extraordinary toughening of magnetic hydrogels

    NASA Astrophysics Data System (ADS)

    Tang, Jingda; Li, Chenghai; Li, Haomin; Lv, Zengyao; Sheng, Hao; Lu, Tongqing; Wang, T. J.

    2018-05-01

    Phase separation markedly influences the physical properties of hydrogels. Here, we find that poly (N, N-dimethylacrylamide) (PDMA) hydrogels suffer from phase separation in aqueous sodium hydroxide solutions when the concentration is higher than 2 M. The polymer volume fraction and mechanical properties show an abrupt change around the transition point. We utilize this phase separation mechanism to synthesize tough magnetic PDMA hydrogels with the in-situ precipitation method. For comparison, we also prepared magnetic poly (2-acrylamido-2-methyl-propane sulfonic acid sodium) (PNaAMPS) magnetic hydrogels, where no phase separation occurs. The phase-separated magnetic PDMA hydrogels exhibit an extraordinarily high toughness of ˜1000 J m-2; while non-phase-separated magnetic PNaAMPS hydrogels only show a toughness of ˜1 J m-2, three orders of magnitude lower than that of PDMA hydrogels. This phase separation mechanism may become a new approach to prepare tough magnetic hydrogels and inspire more applications.

  8. Moisture-Induced Amorphous Phase Separation of Amorphous Solid Dispersions: Molecular Mechanism, Microstructure, and Its Impact on Dissolution Performance.

    PubMed

    Chen, Huijun; Pui, Yipshu; Liu, Chengyu; Chen, Zhen; Su, Ching-Chiang; Hageman, Michael; Hussain, Munir; Haskell, Roy; Stefanski, Kevin; Foster, Kimberly; Gudmundsson, Olafur; Qian, Feng

    2018-01-01

    Amorphous phase separation (APS) is commonly observed in amorphous solid dispersions (ASD) when exposed to moisture. The objective of this study was to investigate: (1) the phase behavior of amorphous solid dispersions composed of a poorly water-soluble drug with extremely low crystallization propensity, BMS-817399, and PVP, following exposure to different relative humidity (RH), and (2) the impact of phase separation on the intrinsic dissolution rate of amorphous solid dispersion. Drug-polymer interaction was confirmed in ASDs at different drug loading using infrared (IR) spectroscopy and water vapor sorption analysis. It was found that the drug-polymer interaction could persist at low RH (≤75% RH) but was disrupted after exposure to high RH, with the advent of phase separation. Surface morphology and composition of 40/60 ASD at micro-/nano-scale before and after exposure to 95% RH were also compared. It was found that hydrophobic drug enriched on the surface of ASD after APS. However, for the 40/60 ASD system, the intrinsic dissolution rate of amorphous drug was hardly affected by the phase behavior of ASD, which may be partially attributed to the low crystallization tendency of amorphous BMS-817399 and enriched drug amount on the surface of ASD. Intrinsic dissolution rate of PVP decreased resulting from APS, leading to a lower concentration in the dissolution medium, but supersaturation maintenance was not anticipated to be altered after phase separation due to the limited ability of PVP to inhibit drug precipitation and prolong the supersaturation of drug in solution. This study indicated that for compounds with low crystallization propensity and high hydrophobicity, the risk of moisture-induced APS is high but such phase separation may not have profound impact on the drug dissolution performance of ASDs. Therefore, application of ASD technology on slow crystallizers could incur low risks not only in physical stability but also in dissolution performance

  9. Continuous Determination of High-Vapor Phase Concentrations of Tetrachloroethylene Using On-Line Mass Spectrometry

    EPA Science Inventory

    A method was developed to determine the vapor concentration of tetrachloroethylene (PCE) at and below its equilibrium vapor phase concentration, 168,000 μg/L (25°C). Vapor samples were drawn by vacuum into a six-port sampling valve and injected through a jet separator into an io...

  10. Dynamics of crowding-induced mixing in phase separated lipid bilayers

    DOE PAGES

    Zeno, Wade F.; Johnson, Kaitlin E.; Sasaki, Darryl Y.; ...

    2016-10-10

    We use fluorescence microscopy to examine the dynamics of the crowding-induced mixing transition of liquid ordered (L o)–liquid disordered (L d) phase separated lipid bilayers when the following particles of increasing size bind to either the L o or L d phase: Ubiquitin, green fluorescent protein (GFP), and nanolipoprotein particles (NLPs) of two diameters. These proteinaceous particles contained histidine-tags, which were phase targeted by binding to iminodiacetic acid (IDA) head groups, via a Cu 2+ chelating mechanism, of lipids that specifically partition into either the Lo phase or Ld phase. The degree of steric pressure was controlled by varying themore » size of the bound particle (10–240 kDa) and the amount of binding sites present (i.e., DPIDA concentrations of 9 and 12 mol%) in the supported lipid multibilayer platform used here. We develop a mass transfer-based diffusional model to analyze the observed L o phase domain dissolution that, along with visual observations and activation energy calculations, provides insight into the sequence of events in crowding-induced mixing. Furthermore, our results suggest that the degree of steric pressure and target phase influence not only the efficacy of steric-pressure induced mixing, but the rate and controlling mechanism for which it occurs.« less

  11. Vapor phase pyrolysis

    NASA Technical Reports Server (NTRS)

    Steurer, Wolfgang

    1992-01-01

    The vapor phase pyrolysis process is designed exclusively for the lunar production of oxygen. In this concept, granulated raw material (soil) that consists almost entirely of metal oxides is vaporized and the vapor is raised to a temperature where it dissociates into suboxides and free oxygen. Rapid cooling of the dissociated vapor to a discrete temperature causes condensation of the suboxides, while the oxygen remains essentially intact and can be collected downstream. The gas flow path and flow rate are maintained at an optimum level by control of the pressure differential between the vaporization region and the oxygen collection system with the aid of the environmental vacuum.

  12. Fabrication of PVDF-based blend membrane with a thin hydrophilic deposition layer and a network structure supporting layer via the thermally induced phase separation followed by non-solvent induced phase separation process

    NASA Astrophysics Data System (ADS)

    Wu, Zhiguo; Cui, Zhenyu; Li, Tianyu; Qin, Shuhao; He, Benqiao; Han, Na; Li, Jianxin

    2017-10-01

    A simple strategy of thermally induced phase separation followed by non-solvent induced phase separation (TIPS-NIPS) is reported to fabricate poly (vinylidene fluoride) (PVDF)-based blend membrane. The dissolved poly (styrene-co-maleic anhydride) (SMA) in diluent prevents the crystallization of PVDF during the cooling process and deposites on the established PVDF matrix in the later extraction. Compared with traditional coating technique, this one-step TIPS-NIPS method can not only fabricate a supporting layer with an interconnected network structure even via solid-liquid phase separation of TIPS, but also form a uniform SMA skin layer approximately as thin as 200 nm via surface deposition of NIPS. Besides the better hydrophilicity, what's interesting is that the BSA rejection ratio increases from 48% to 94% with the increase of SMA, which indicates that the separation performance has improved. This strategy can be conveniently extended to the creation of firmly thin layer, surface functionalization and structure controllability of the membrane.

  13. Solvent annealing induced phase separation and dewetting in PMMA∕SAN blend film: film thickness and solvent dependence.

    PubMed

    You, Jichun; Zhang, Shuangshuang; Huang, Gang; Shi, Tongfei; Li, Yongjin

    2013-06-28

    The competition between "dewetting" and "phase separation" behaviors in polymer blend films attracts significant attention in the last decade. The simultaneous phase separation and dewetting in PMMA∕SAN [poly(methyl methacrylate) and poly(styrene-ran-acrylonitrile)] blend ultrathin films upon solvent annealing have been observed for the first time in our previous work. In this work, film thickness and annealing solvent dependence of phase behaviors in this system has been investigated using atomic force microscopy and grazing incidence small-angle X-ray scattering (GISAXS). On one hand, both vertical phase separation and dewetting take place upon selective solvent vapor annealing, leading to the formation of droplet∕mimic-film structures with various sizes (depending on original film thickness). On the other hand, the whole blend film dewets the substrate and produces dispersed droplets on the silicon oxide upon common solvent annealing. GISAXS results demonstrate the phase separation in the big dewetted droplets resulted from the thicker film (39.8 nm). In contrast, no period structure is detected in small droplets from the thinner film (5.1 nm and 9.7 nm). This investigation indicates that dewetting and phase separation in PMMA∕SAN blend film upon solvent annealing depend crucially on the film thickness and the atmosphere during annealing.

  14. Kinetics of motility-induced phase separation and swim pressure

    NASA Astrophysics Data System (ADS)

    Patch, Adam; Yllanes, David; Marchetti, M. Cristina

    2017-01-01

    Active Brownian particles (ABPs) represent a minimal model of active matter consisting of self-propelled spheres with purely repulsive interactions and rotational noise. Here we examine the pressure of ABPs in two dimensions in both closed boxes and systems with periodic boundary conditions and show that its nonmonotonic behavior with density is a general property of ABPs and is not the result of finite-size effects. We correlate the time evolution of the mean pressure towards its steady-state value with the kinetics of motility-induced phase separation. For parameter values corresponding to phase-separated steady states, we identify two dynamical regimes. The pressure grows monotonically in time during the initial regime of rapid cluster formation, overshooting its steady-state value and then quickly relaxing to it, and remains constant during the subsequent slower period of cluster coalescence and coarsening. The overshoot is a distinctive feature of active systems.

  15. Substrate-induced phase of a [1]benzothieno[3,2-b]benzothiophene derivative and phase evolution by aging and solvent vapor annealing.

    PubMed

    Jones, Andrew O F; Geerts, Yves H; Karpinska, Jolanta; Kennedy, Alan R; Resel, Roland; Röthel, Christian; Ruzié, Christian; Werzer, Oliver; Sferrazza, Michele

    2015-01-28

    Substrate-induced phases (SIPs) are polymorphic phases that are found in thin films of a material and are different from the single crystal or "bulk" structure of a material. In this work, we investigate the presence of a SIP in the family of [1]benzothieno[3,2-b]benzothiophene (BTBT) organic semiconductors and the effect of aging and solvent vapor annealing on the film structure. Through extensive X-ray structural investigations of spin coated films, we find a SIP with a significantly different structure to that found in single crystals of the same material forms; the SIP has a herringbone motif while single crystals display layered π-π stacking. Over time, the structure of the film is found to slowly convert to the single crystal structure. Solvent vapor annealing initiates the same structural evolution process but at a greatly increased rate, and near complete conversion can be achieved in a short period of time. As properties such as charge transport capability are determined by the molecular structure, this work highlights the importance of understanding and controlling the structure of organic semiconductor films and presents a simple method to control the film structure by solvent vapor annealing.

  16. Effects of temperature and solvent condition on phase separation induced molecular fractionation of gum arabic/hyaluronan aqueous mixtures.

    PubMed

    Hu, Bing; Han, Lingyu; Gao, Zhiming; Zhang, Ke; Al-Assaf, Saphwan; Nishinari, Katsuyoshi; Phillips, Glyn O; Yang, Jixin; Fang, Yapeng

    2018-05-14

    Effects of temperature and solvent condition on phase separation-induced molecular fractionation of gum arabic/hyaluronan (GA/HA) mixed solutions were investigated. Two gum arabic samples (EM10 and STD) with different molecular weights and polydispersity indices were used. Phase diagrams, including cloud and binodal curves, were established by visual observation and GPC-RI methods. The molecular parameters of control and fractionated GA, from upper and bottom phases, were measured by GPC-MALLS. Fractionation of GA increased the content of arabinogalactan-protein complex (AGP) from ca. 11% to 18% in STD/HA system and 28% to 55% in EM10/HA system. The phase separation-induced molecular fractionation was further studied as a function of temperature and solvent condition (varying ionic strength and ethanol content). Increasing salt concentration (from 0.5 to 5 mol/L) greatly reduced the extent of phase separation-induced fractionation. This effect may be ascribed to changes in the degree of ionization and shielding of the acid groups. Increasing temperature (from 4 °C to 80 °C) also exerted a significant influence on phase separation-induced fractionation. The best temperature for GA/HA mixture system was 40 °C while higher temperature negatively affected the fractionation due to denaturation and possibly degradation in mixed solutions. Increasing the ethanol content up to 30% showed almost no effect on the phase separation induced fractionation. Copyright © 2018 Elsevier B.V. All rights reserved.

  17. Conducting polymer networks synthesized by photopolymerization-induced phase separation

    NASA Astrophysics Data System (ADS)

    Yamashita, Yuki; Komori, Kana; Murata, Tasuku; Nakanishi, Hideyuki; Norisuye, Tomohisa; Yamao, Takeshi; Tran-Cong-Miyata, Qui

    2018-03-01

    Polymer mixtures composed of double networks of a polystyrene derivative (PSAF) and poly(methyl methacrylate) (PMMA) were alternatively synthesized by using ultraviolet (UV) and visible (Vis) light. The PSAF networks were generated by UV irradiation to photodimerize the anthracene (A) moieties labeled on the PSAF chains, whereas PMMA networks were produced by photopolymerization of methyl methacrylate (MMA) monomer and the cross-link reaction using ethylene glycol dimethacrylate (EGDMA) under Vis light irradiation. It was found that phase separation process of these networks can be independently induced and promptly controlled by using UV and Vis light. The characteristic length scale distribution of the resulting co-continuous morphology can be well regulated by the UV and Vis light intensity. In order to confirm and utilize the connectivity of the bicontinuous morphology observed by confocal microscopy, a very small amount, 0.1 wt%, of multi-walled carbon nanotubes (MWCNTs) was introduced into the mixture and the current-voltage (I-V) relationship was subsequently examined. Preliminary data show that MWCNTs are preferentially dispersed in the PSAF-rich continuous domains and the whole mixture became electrically conducting, confirming the connectivity of the observed bi-continuous morphology. The experimental data obtained in this study reveal a promising method to design various scaffolds for conducting soft matter taking advantages of photopolymerization-induced phase separation.

  18. Amide-induced phase separation of hexafluoroisopropanol-water mixtures depending on the hydrophobicity of amides.

    PubMed

    Takamuku, Toshiyuki; Wada, Hiroshi; Kawatoko, Chiemi; Shimomura, Takuya; Kanzaki, Ryo; Takeuchi, Munetaka

    2012-06-21

    Amide-induced phase separation of hexafluoro-2-propanol (HFIP)-water mixtures has been investigated to elucidate solvation properties of the mixtures by means of small-angle neutron scattering (SANS), (1)H and (13)C NMR, and molecular dynamics (MD) simulation. The amides included N-methylformamide (NMF), N-methylacetamide (NMA), and N-methylpropionamide (NMP). The phase diagrams of amide-HFIP-water ternary systems at 298 K showed that phase separation occurs in a closed-loop area of compositions as well as an N,N-dimethylformamide (DMF) system previously reported. The phase separation area becomes wider as the hydrophobicity of amides increases in the order of NMF < NMA < DMF < NMP. Thus, the evolution of HFIP clusters around amides due to the hydrophobic interaction gives rise to phase separation of the mixtures. In contrast, the disruption of HFIP clusters causes the recovery of the homogeneity of the ternary systems. The present results showed that HFIP clusters are evolved with increasing amide content to the lower phase separation concentration in the same mechanism among the four amide systems. However, the disruption of HFIP clusters in the NMP and DMF systems with further increasing amide content to the upper phase separation concentration occurs in a different way from those in the NMF and NMA systems.

  19. Modification of linear prepolymers to tailor heterogeneous network formation through photo-initiated Polymerization-Induced Phase Separation

    PubMed Central

    Szczepanski, Caroline R.; Stansbury, Jeffrey W.

    2015-01-01

    Polymerization-induced phase separation (PIPS) was studied in ambient photopolymerizations of triethylene glycol dimethacrylate (TEGDMA) modified by poly(methyl methacrylate) (PMMA). The molecular weight of PMMA and the rate of network formation (through incident UV-irradiation) were varied to influence both the promotion of phase separation through increases in overall free energy, as well as the extent to which phase development occurs during polymerization through diffusion prior to network gelation. The overall free energy of the polymerizing system increases with PMMA molecular weight, such that PIPS is promoted thermodynamically at low loading levels (5 wt%) of a higher molecular weight PMMA (120 kDa), while a higher loading level (20 wt%) is needed to induce PIPS with lower PMMA molecular weight (11 kDa), and phase separation was not promoted at any loading level tested of the lowest molecular weight PMMA (1 kDa). Due to these differences in overall free energy, systems modified by PMMA (11 kDa) underwent phase separation via Nucleation and Growth, and systems modified by PMMA (120 kDa), followed the Spinodal Decomposition mechanism. Despite differences in phase structure, all materials form a continuous phase rich in TEGDMA homopolymer. At high irradiation intensity (Io=20mW/cm2), the rate of network formation prohibited significant phase separation, even when thermodynamically preferred. A staged curing approach, which utilizes low intensity irradiation (Io=300µW/cm2) for the first ~50% of reaction to allow phase separation via diffusion, followed by a high intensity flood-cure to achieve a high degree of conversion, was employed to form phase-separated networks with reduced polymerization stress yet equivalent final conversion and modulus. PMID:26190865

  20. Spatial patterning of P granules by RNA-induced phase separation of the intrinsically-disordered protein MEG-3

    PubMed Central

    Smith, Jarrett; Calidas, Deepika; Schmidt, Helen; Lu, Tu; Rasoloson, Dominique; Seydoux, Geraldine

    2016-01-01

    RNA granules are non-membrane bound cellular compartments that contain RNA and RNA binding proteins. The molecular mechanisms that regulate the spatial distribution of RNA granules in cells are poorly understood. During polarization of the C. elegans zygote, germline RNA granules, called P granules, assemble preferentially in the posterior cytoplasm. We present evidence that P granule asymmetry depends on RNA-induced phase separation of the granule scaffold MEG-3. MEG-3 is an intrinsically disordered protein that binds and phase separates with RNA in vitro. In vivo, MEG-3 forms a posterior-rich concentration gradient that is anti-correlated with a gradient in the RNA-binding protein MEX-5. MEX-5 is necessary and sufficient to suppress MEG-3 granule formation in vivo, and suppresses RNA-induced MEG-3 phase separation in vitro. Our findings suggest that MEX-5 interferes with MEG-3’s access to RNA, thus locally suppressing MEG-3 phase separation to drive P granule asymmetry. Regulated access to RNA, combined with RNA-induced phase separation of key scaffolding proteins, may be a general mechanism for controlling the formation of RNA granules in space and time. DOI: http://dx.doi.org/10.7554/eLife.21337.001 PMID:27914198

  1. Thermocapillary-Induced Phase Separation with Coalescence

    NASA Technical Reports Server (NTRS)

    Davis, Robert H.

    2003-01-01

    Research has been undertaken on interactions of two or more deformable drops (or bubbles) in a viscous fluid and subject to a temperature, gravitational, or flow field. An asymptotic theory for nearly spherical drops shows that small deformations reduce the coalescence and phase separation rates. Boundary-integral simulations for large deformations show that bubbles experience alignment and enhanced coalescence, whereas more viscous drops may break as a result of hydrodynamic interactions. Experiments for buoyancy motion confirm these observations. Simulations of the sedimentation of many drops show clustering phenomena due to deformations, which lead to enhanced phase separation rates, and simulations of sheared emulsions show that deformations cause a reduction in the effective viscosity.

  2. Initiated Chemical Vapor Deposition (iCVD) of Highly Cross-Linked Polymer Films for Advanced Lithium-Ion Battery Separators.

    PubMed

    Yoo, Youngmin; Kim, Byung Gon; Pak, Kwanyong; Han, Sung Jae; Song, Heon-Sik; Choi, Jang Wook; Im, Sung Gap

    2015-08-26

    We report an initiated chemical vapor deposition (iCVD) process to coat polyethylene (PE) separators in Li-ion batteries with a highly cross-linked, mechanically strong polymer, namely, polyhexavinyldisiloxane (pHVDS). The highly cross-linked but ultrathin pHVDS films can only be obtained by a vapor-phase process, because the pHVDS is insoluble in most solvents and thus infeasible with conventional solution-based methods. Moreover, even after the pHVDS coating, the initial porous structure of the separator is well preserved owing to the conformal vapor-phase deposition. The coating thickness is delicately controlled by deposition time to the level that the pore size decreases to below 7% compared to the original dimension. The pHVDS-coated PE shows substantially improved thermal stability and electrolyte wettability. After incubation at 140 °C for 30 min, the pHVDS-coated PE causes only a 12% areal shrinkage (versus 90% of the pristine separator). The superior wettability results in increased electrolyte uptake and ionic conductivity, leading to significantly improved rate performance. The current approach is applicable to a wide range of porous polymeric separators that suffer from thermal shrinkage and poor electrolyte wetting.

  3. Rapid variations in fluid chemistry constrain hydrothermal phase separation at the Main Endeavour Field

    NASA Astrophysics Data System (ADS)

    Love, Brooke; Lilley, Marvin; Butterfield, David; Olson, Eric; Larson, Benjamin

    2017-02-01

    Previous work at the Main Endeavour Field (MEF) has shown that chloride concentration in high-temperature vent fluids has not exceeded 510 mmol/kg (94% of seawater), which is consistent with brine condensation and loss at depth, followed by upward flow of a vapor phase toward the seafloor. Magmatic and seismic events have been shown to affect fluid temperature and composition and these effects help narrow the possibilities for sub-surface processes. However, chloride-temperature data alone are insufficient to determine details of phase separation in the upflow zone. Here we use variation in chloride and gas content in a set of fluid samples collected over several days from one sulfide chimney structure in the MEF to constrain processes of mixing and phase separation. The combination of gas (primarily magmatic CO2 and seawater-derived Ar) and chloride data, indicate that neither variation in the amount of brine lost, nor mixing of the vapor phase produced at depth with variable quantities of (i) brine or (ii) altered gas rich seawater that has not undergone phase separation, can explain the co-variation of gas and chloride content. The gas-chloride data require additional phase separation of the ascending vapor-like fluid. Mixing and gas partitioning calculations show that near-critical temperature and pressure conditions can produce the fluid compositions observed at Sully vent as a vapor-liquid conjugate pair or as vapor-liquid pair with some remixing, and that the gas partition coefficients implied agree with theoretically predicted values.Plain Language SummaryWhen the chemistry of fluids from deep sea hot springs changes over a short time span, it allows us to narrow down the conditions and processes that created those fluids. This gives us a better idea what is happening under the seafloor where the water is interacting with hot rocks and minerals, boiling, and taking on the character it will have when it emerges at</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012TRACE...2...59I','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012TRACE...2...59I"><span>Characteristics of Evaporator with a Lipuid-<span class="hlt">Vapor</span> <span class="hlt">Separator</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ikeguchi, Masaki; Tanaka, Naoki; Yumikura, Tsuneo</p> <p></p> <p>Flow pattern of refrigerant in a heat exchanger tube changes depending on <span class="hlt">vapor</span> quality, tube diameter, refrigerant flow rate and refrigerant properties. High flow rate causes mist flow where the quality is from 0.8 to 1.0. 1n this flow pattern, the liquid film detaches from the tube wall so that the heat flow is intervened. The heat transfer coefficient generally increases with the flow rate. But the pressure drop of refrigerant flow simultaneously increases and the region of the mist flow enlarges. In order to reduce the pressure drop and suppress the mist flow, we have developped a small liquid-<span class="hlt">vapor</span> <span class="hlt">separator</span> that removes the <span class="hlt">vapor</span> from the evaporating refrigerant flow. This <span class="hlt">separator</span> is equipped in the middle of the evaporator where the flow pattern is annular. The experiments to evaluate the effect of this <span class="hlt">separator</span> were carried out and the following conclutions were obtained. (1) Average heat transfer coefficient increases by 30-60 %. (2) Pressure drop reduces by 20-30 %. (3) Cooling Capacity increases by 2-9 %.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=191247&Lab=NRMRL&keyword=fermentation&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50','EPA-EIMS'); return false;" href="https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=191247&Lab=NRMRL&keyword=fermentation&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50"><span>Membrane-assisted <span class="hlt">vapor</span> stripping: energy efficient hybrid distillation-<span class="hlt">vapor</span> permeation process for alcohol-water <span class="hlt">separation</span></span></a></p> <p><a target="_blank" href="http://oaspub.epa.gov/eims/query.page">EPA Science Inventory</a></p> <p></p> <p></p> <p>BACKGROUND: Energy efficient alternatives to distillation for alcohol recovery from dilute solution are needed to improve biofuel sustainability. A process integrating steam stripping with a <span class="hlt">vapor</span> compression step and a <span class="hlt">vapor</span> permeation membrane <span class="hlt">separation</span> step is proposed. The...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26906600','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26906600"><span>Simultaneous imaging of fuel <span class="hlt">vapor</span> mass fraction and gas-<span class="hlt">phase</span> temperature inside gasoline sprays using two-line excitation tracer planar laser-<span class="hlt">induced</span> fluorescence.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zigan, Lars; Trost, Johannes; Leipertz, Alfred</p> <p>2016-02-20</p> <p>This paper reports for the first time, to the best of our knowledge, on the simultaneous imaging of the gas-<span class="hlt">phase</span> temperature and fuel <span class="hlt">vapor</span> mass fraction distribution in a direct-injection spark-ignition (DISI) spray under engine-relevant conditions using tracer planar laser-<span class="hlt">induced</span> fluorescence (TPLIF). For measurements in the spray, the fluorescence tracer 3-pentanone is added to the nonfluorescent surrogate fuel iso-octane, which is excited quasi-simultaneously by two different excimer lasers for two-line excitation LIF. The gas-<span class="hlt">phase</span> temperature of the mixture of fuel <span class="hlt">vapor</span> and surrounding gas and the fuel <span class="hlt">vapor</span> mass fraction can be calculated from the two LIF signals. The measurements are conducted in a high-temperature, high-pressure injection chamber. The fluorescence calibration of the tracer was executed in a flow cell and extended significantly compared to the existing database. A detailed error analysis for both calibration and measurement is provided. Simultaneous single-shot gas-<span class="hlt">phase</span> temperature and fuel <span class="hlt">vapor</span> mass fraction fields are processed for the assessment of cyclic spray fluctuations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JChPh.148o4902W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JChPh.148o4902W"><span><span class="hlt">Phase</span> <span class="hlt">separation</span> and large deviations of lattice active matter</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Whitelam, Stephen; Klymko, Katherine; Mandal, Dibyendu</p> <p>2018-04-01</p> <p>Off-lattice active Brownian particles form clusters and undergo <span class="hlt">phase</span> <span class="hlt">separation</span> even in the absence of attractions or velocity-alignment mechanisms. Arguments that explain this phenomenon appeal only to the ability of particles to move persistently in a direction that fluctuates, but existing lattice models of hard particles that account for this behavior do not exhibit <span class="hlt">phase</span> <span class="hlt">separation</span>. Here we present a lattice model of active matter that exhibits motility-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> in the absence of velocity alignment. Using direct and rare-event sampling of dynamical trajectories, we show that clustering and <span class="hlt">phase</span> <span class="hlt">separation</span> are accompanied by pronounced fluctuations of static and dynamic order parameters. This model provides a complement to off-lattice models for the study of motility-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26243640','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26243640"><span>Enhanced water <span class="hlt">vapor</span> <span class="hlt">separation</span> by temperature-controlled aligned-multiwalled carbon nanotube membranes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jeon, Wonjae; Yun, Jongju; Khan, Fakhre Alam; Baik, Seunghyun</p> <p>2015-09-14</p> <p>Here we present a new strategy of selectively rejecting water <span class="hlt">vapor</span> while allowing fast transport of dry gases using temperature-controlled aligned-multiwalled carbon nanotubes (aligned-MWNTs). The mechanism is based on the water <span class="hlt">vapor</span> condensation at the entry region of nanotubes followed by removing aggregated water droplets at the tip of the superhydrophobic aligned-MWNTs. The first condensation step could be dramatically enhanced by decreasing the nanotube temperature. The permeate-side relative humidity was as low as ∼17% and the helium-water <span class="hlt">vapor</span> <span class="hlt">separation</span> factor was as high as 4.62 when a helium-water <span class="hlt">vapor</span> mixture with a relative humidity of 100% was supplied to the aligned-MWNTs. The flow through the interstitial space of the aligned-MWNTs allowed the permeability of single dry gases an order of magnitude higher than the Knudsen prediction regardless of membrane temperature. The water <span class="hlt">vapor</span> <span class="hlt">separation</span> performance of hydrophobic polytetrafluoroethylene membranes could also be significantly enhanced at low temperatures. This work combines the membrane-based <span class="hlt">separation</span> technology with temperature control to enhance water <span class="hlt">vapor</span> <span class="hlt">separation</span> performance.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JPS...384..408L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JPS...384..408L"><span>Facile fabrication of multilayer <span class="hlt">separators</span> for lithium-ion battery via multilayer coextrusion and thermal <span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Yajie; Pu, Hongting</p> <p>2018-04-01</p> <p>Polypropylene (PP)/polyethylene (PE) multilayer <span class="hlt">separators</span> with cellular-like submicron pore structure for lithium-ion battery are efficiently fabricated by the combination of multilayer coextrusion (MC) and thermal <span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> (TIPS). The as-prepared <span class="hlt">separators</span>, referred to as MC-TIPS PP/PE, not only show efficacious thermal shutdown function and wider shutdown temperature window, but also exhibit higher thermal stability than the commercial <span class="hlt">separator</span> with trilayer construction of PP and PE (Celgard® 2325). The dimensional shrinkage of MC-TIPS PP/PE can be negligible until 160 °C. In addition, compared to the commercial <span class="hlt">separator</span>, MC-TIPS PP/PE exhibits higher porosity and electrolyte uptake, leading to higher ionic conductivity and better battery performances. The above-mentioned fascinating characteristics with the convenient preparation process make MC-TIPS PP/PE a promising candidate for the application as high performance lithium-ion battery <span class="hlt">separators</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22489686','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22489686"><span>Significant vertical <span class="hlt">phase</span> <span class="hlt">separation</span> in solvent-<span class="hlt">vapor</span>-annealed poly(3,4-ethylenedioxythiophene):poly(styrene sulfonate) composite films leading to better conductivity and work function for high-performance indium tin oxide-free optoelectronics.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yeo, Jun-Seok; Yun, Jin-Mun; Kim, Dong-Yu; Park, Sungjun; Kim, Seok-Soon; Yoon, Myung-Han; Kim, Tae-Wook; Na, Seok-In</p> <p>2012-05-01</p> <p>In the present study, a novel polar-solvent <span class="hlt">vapor</span> annealing (PSVA) was used to <span class="hlt">induce</span> a significant structural rearrangement in poly(3,4-ethylenedioxythiophene):poly(styrene sulfonate) (PEDOT:PSS) films in order to improve their electrical conductivity and work function. The effects of polar-solvent <span class="hlt">vapor</span> annealing on PEDOT:PSS were systematically compared with those of a conventional solvent additive method (SAM) and investigated in detail by analyzing the changes in conductivity, morphology, top and bottom surface composition, conformational PEDOT chains, and work function. The results confirmed that PSVA <span class="hlt">induces</span> significant <span class="hlt">phase</span> <span class="hlt">separation</span> between excess PSS and PEDOT chains and a spontaneous formation of a highly enriched PSS layer on the top surface of the PEDOT:PSS polymer blend, which in turn leads to better 3-dimensional connections between the conducting PEDOT chains and higher work function. The resultant PSVA-treated PEDOT:PSS anode films exhibited a significantly enhanced conductivity of up to 1057 S cm(-1) and a tunable high work function of up to 5.35 eV. The PSVA-treated PEDOT:PSS films were employed as transparent anodes in polymer light-emitting diodes (PLEDs) and polymer solar cells (PSCs). The cell performances of organic optoelectronic devices with the PSVA-treated PEDOT:PSS anodes were further improved due to the significant vertical <span class="hlt">phase</span> <span class="hlt">separation</span> and the self-organized PSS top surface in PSVA-treated PEDOT:PSS films, which can increase the anode conductivity and work function and allow the direct formation of a functional buffer layer between the active layer and the polymeric electrode. The results of the present study will allow better use and understanding of polymeric-blend materials and will further advance the realization of high-performance indium tin oxide (ITO)-free organic electronics.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/866165','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/866165"><span>Combination downflow-upflow <span class="hlt">vapor</span>-liquid <span class="hlt">separator</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Kidwell, John H.; Prueter, William P.; Eaton, Andrew M.</p> <p>1987-03-10</p> <p>An improved <span class="hlt">vapor</span>-liquid <span class="hlt">separator</span> having a vertically disposed conduit for flow of a mixture. A first, second and third plurality of curved arms penetrate and extend within the conduit. A cylindrical member is radially spaced from the conduit forming an annulus therewith and having perforations and a retaining lip at its upper end.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016GeCoA.183..125S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016GeCoA.183..125S"><span>Fractionation of Cl/Br during fluid <span class="hlt">phase</span> <span class="hlt">separation</span> in magmatic-hydrothermal fluids</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Seo, Jung Hun; Zajacz, Zoltán</p> <p>2016-06-01</p> <p>Brine and <span class="hlt">vapor</span> inclusions were synthesized to study Cl/Br fractionation during magmatic-hydrothermal fluid <span class="hlt">phase</span> <span class="hlt">separation</span> at 900 °C and pressures of 90, 120, and 150 MPa in Li/Na/K halide salt-H2O systems. Laser ablation ICP-MS microanalysis of high-density brine inclusions show an elevated Cl/Br ratio compared to the coexisting low-density <span class="hlt">vapor</span> inclusions. The degree of Cl/Br fractionation between <span class="hlt">vapor</span> and brine is significantly dependent on the identity of the alkali metal in the system: stronger <span class="hlt">vapor</span> partitioning of Br occurs in the Li halide-H2O system compared to the systems of K and Na halide-H2O. The effect of the identity of alkali-metals in the system is stronger compared to the effect of <span class="hlt">vapor</span>-brine density contrast. We infer that competition between alkali-halide and alkali-OH complexes in high-temperature fluids might cause the Cl/Br fractionation, consistent with the observed molar imbalances of alkali metals compared to halides in the analyzed brine inclusions. Our experiments show that the identity of alkali metals controls the degrees of Cl/Br fractionation between the <span class="hlt">separating</span> aqueous fluid <span class="hlt">phases</span> at 900 °C, and suggest that a significant variability in the Cl/Br ratios of magmatic fluids can arise in Li-rich systems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=215069&keyword=huang&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50','EPA-EIMS'); return false;" href="https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=215069&keyword=huang&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50"><span>Experimental Validation of Hybrid Distillation-<span class="hlt">Vapor</span> Permeation Process for Energy Efficient Ethanol-Water <span class="hlt">Separation</span></span></a></p> <p><a target="_blank" href="http://oaspub.epa.gov/eims/query.page">EPA Science Inventory</a></p> <p></p> <p></p> <p>The energy demand of distillation-based systems for ethanol recovery and dehydration can be significant, particularly for dilute solutions. An alternative <span class="hlt">separation</span> process integrating <span class="hlt">vapor</span> stripping with a <span class="hlt">vapor</span> compression step and a <span class="hlt">vapor</span> permeation membrane <span class="hlt">separation</span> step...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=215907&keyword=biotechnology&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50','EPA-EIMS'); return false;" href="https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=215907&keyword=biotechnology&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50"><span>Experimental Validation of Hybrid Distillation-<span class="hlt">Vapor</span> Permeation Process for Energy Efficient Ethanol-Water <span class="hlt">Separation</span></span></a></p> <p><a target="_blank" href="http://oaspub.epa.gov/eims/query.page">EPA Science Inventory</a></p> <p></p> <p></p> <p>The energy demand of distillation-based systems for ethanol recovery and dehydration can be significant, particularly for dilute solutions. An alternative <span class="hlt">separation</span> process integrating <span class="hlt">vapor</span> stripping with a <span class="hlt">vapor</span> compression step and a <span class="hlt">vapor</span> permeation membrane <span class="hlt">separation</span> step,...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22036489-identification-vapor-phase-chemical-warfare-agent-simulants-rocket-fuels-using-laser-induced-breakdown-spectroscopy','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22036489-identification-vapor-phase-chemical-warfare-agent-simulants-rocket-fuels-using-laser-induced-breakdown-spectroscopy"><span>Identification of <span class="hlt">vapor-phase</span> chemical warfare agent simulants and rocket fuels using laser-<span class="hlt">induced</span> breakdown spectroscopy</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Stearns, Jaime A.; McElman, Sarah E.; Dodd, James A.</p> <p>2010-05-01</p> <p>Application of laser-<span class="hlt">induced</span> breakdown spectroscopy (LIBS) to the identification of security threats is a growing area of research. This work presents LIBS spectra of <span class="hlt">vapor-phase</span> chemical warfare agent simulants and typical rocket fuels. A large dataset of spectra was acquired using a variety of gas mixtures and background pressures and processed using partial least squares analysis. The five compounds studied were identified with a 99% success rate by the best method. The temporal behavior of the emission lines as a function of chamber pressure and gas mixture was also investigated, revealing some interesting trends that merit further study.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28915000','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28915000"><span>Ductile Glass of Polyrotaxane Toughened by Stretch-<span class="hlt">Induced</span> Intramolecular <span class="hlt">Phase</span> <span class="hlt">Separation</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kato, Kazuaki; Nemoto, Kaito; Mayumi, Koichi; Yokoyama, Hideaki; Ito, Kohzo</p> <p>2017-09-27</p> <p>A new class of ductile glasses is created from a thermoplastic polyrotaxane. The hard glass, which has a Young's modulus of 1 GPa, shows crazing, necking, and strain hardening with a total elongation of 330%. Stress concentration is prevented through a unique stretch-<span class="hlt">induced</span> intramolecular <span class="hlt">phase</span> <span class="hlt">separation</span> of the cyclic components and the exposed backbone. In situ synchrotron X-ray scattering studies indicate that the backbone polymer chains slip through the cyclic components in the regions where the stress is concentrated.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017CPL...685..263S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017CPL...685..263S"><span>Reaction-mediated entropic effect on <span class="hlt">phase</span> <span class="hlt">separation</span> in a binary polymer system</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sun, Shujun; Guo, Miaocai; Yi, Xiaosu; Zhang, Zuoguang</p> <p>2017-10-01</p> <p>We present a computer simulation to study the <span class="hlt">phase</span> <span class="hlt">separation</span> behavior <span class="hlt">induced</span> by polymerization in a binary system comprising polymer chains and reactive monomers. We examined the influence of interaction parameter between components and monomer concentration on the reaction-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span>. The simulation results demonstrate that increasing interaction parameter (enthalpic effect) would accelerate <span class="hlt">phase</span> <span class="hlt">separation</span>, while entropic effect plays a key role in the process of <span class="hlt">phase</span> <span class="hlt">separation</span>. Furthermore, scanning electron microscopy observations illustrate identical morphologies as found in theoretical simulation. This study may enrich our comprehension of <span class="hlt">phase</span> <span class="hlt">separation</span> in polymer mixture.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014JTePh..59.1101B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014JTePh..59.1101B"><span><span class="hlt">Phase</span> transformations during the growth of paracetamol crystals from the <span class="hlt">vapor</span> <span class="hlt">phase</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Belyaev, A. P.; Rubets, V. P.; Antipov, V. V.; Bordei, N. S.</p> <p>2014-07-01</p> <p><span class="hlt">Phase</span> transformations during the growth of paracetamol crystals from the <span class="hlt">vapor</span> <span class="hlt">phase</span> are studied by differential scanning calorimetry. It is found that the <span class="hlt">vapor</span>-crystal <span class="hlt">phase</span> transition is actually a superposition of two <span class="hlt">phase</span> transitions: a first-order <span class="hlt">phase</span> transition with variable density and a second-order <span class="hlt">phase</span> transition with variable ordering. The latter, being a diffuse <span class="hlt">phase</span> transition, results in the formation of a new, "pretransition," <span class="hlt">phase</span> irreversibly spent in the course of the transition, which ends in the appearance of orthorhombic crystals. X-ray diffraction data and micrograph are presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29089498','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29089498"><span>Multifunctional nanocomposite hollow fiber membranes by solvent transfer <span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Haase, Martin F; Jeon, Harim; Hough, Noah; Kim, Jong Hak; Stebe, Kathleen J; Lee, Daeyeon</p> <p>2017-11-01</p> <p>The decoration of porous membranes with a dense layer of nanoparticles imparts useful functionality and can enhance membrane <span class="hlt">separation</span> and anti-fouling properties. However, manufacturing of nanoparticle-coated membranes requires multiple steps and tedious processing. Here, we introduce a facile single-step method in which bicontinuous interfacially jammed emulsions are used to form nanoparticle-functionalized hollow fiber membranes. The resulting nanocomposite membranes prepared via solvent transfer-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> and photopolymerization have exceptionally high nanoparticle loadings (up to 50 wt% silica nanoparticles) and feature densely packed nanoparticles uniformly distributed over the entire membrane surfaces. These structurally well-defined, asymmetric membranes facilitate control over membrane flux and selectivity, enable the formation of stimuli responsive hydrogel nanocomposite membranes, and can be easily modified to introduce antifouling features. This approach forms a foundation for the formation of advanced nanocomposite membranes comprising diverse building blocks with potential applications in water treatment, industrial <span class="hlt">separations</span> and as catalytic membrane reactors.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010JTST...19..502V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010JTST...19..502V"><span><span class="hlt">Vapor</span> <span class="hlt">Phase</span> Deposition Using Plasma Spray-PVD™</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>von Niessen, K.; Gindrat, M.; Refke, A.</p> <p>2010-01-01</p> <p>Plasma spray—physical <span class="hlt">vapor</span> deposition (PS-PVD) is a low pressure plasma spray technology to deposit coatings out of the <span class="hlt">vapor</span> <span class="hlt">phase</span>. PS-PVD is a part of the family of new hybrid processes recently developed by Sulzer Metco AG (Switzerland) on the basis of the well-established low pressure plasma spraying (LPPS) technology. Included in this new process family are plasma spray—chemical <span class="hlt">vapor</span> deposition (PS-CVD) and plasma spray—thin film (PS-TF) processes. In comparison to conventional vacuum plasma spraying and LPPS, these new processes use a high energy plasma gun operated at a work pressure below 2 mbar. This leads to unconventional plasma jet characteristics which can be used to obtain specific and unique coatings. An important new feature of PS-PVD is the possibility to deposit a coating not only by melting the feed stock material which builds up a layer from liquid splats, but also by <span class="hlt">vaporizing</span> the injected material. Therefore, the PS-PVD process fills the gap between the conventional PVD technologies and standard thermal spray processes. The possibility to <span class="hlt">vaporize</span> feedstock material and to produce layers out of the <span class="hlt">vapor</span> <span class="hlt">phase</span> results in new and unique coating microstructures. The properties of such coatings are superior to those of thermal spray and EB-PVD coatings. This paper reports on the progress made at Sulzer Metco to develop functional coatings build up from <span class="hlt">vapor</span> <span class="hlt">phase</span> of oxide ceramics and metals.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_1");'>1</a></li> <li><a href="#" onclick='return showDiv("page_2");'>2</a></li> <li class="active"><span>3</span></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_3 --> <div id="page_4" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_2");'>2</a></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li class="active"><span>4</span></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="61"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvM...2c3402W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvM...2c3402W"><span><span class="hlt">Phase</span>-field model of <span class="hlt">vapor</span>-liquid-solid nanowire growth</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, Nan; Upmanyu, Moneesh; Karma, Alain</p> <p>2018-03-01</p> <p>We present a multiphase-field model to describe quantitatively nanowire growth by the <span class="hlt">vapor</span>-liquid-solid (VLS) process. The free-energy functional of this model depends on three nonconserved order parameters that distinguish the <span class="hlt">vapor</span>, liquid, and solid <span class="hlt">phases</span> and describe the energetic properties of various interfaces, including arbitrary forms of anisotropic γ plots for the solid-<span class="hlt">vapor</span> and solid-liquid interfaces. The evolution equations for those order parameters describe basic kinetic processes including the rapid (quasi-instantaneous) equilibration of the liquid catalyst to a droplet shape with constant mean curvature, the slow incorporation of growth atoms at the droplet surface, and crystallization within the droplet. The standard constraint that the sum of the <span class="hlt">phase</span> fields equals unity and the conservation of the number of catalyst atoms, which relates the catalyst volume to the concentration of growth atoms inside the droplet, are handled via <span class="hlt">separate</span> Lagrange multipliers. An analysis of the model is presented that rigorously maps the <span class="hlt">phase</span>-field equations to a desired set of sharp-interface equations for the evolution of the <span class="hlt">phase</span> boundaries under the constraint of force balance at three-<span class="hlt">phase</span> junctions (triple points) given by the Young-Herring relation that includes torque term related to the anisotropy of the solid-liquid and solid-<span class="hlt">vapor</span> interface excess free energies. Numerical examples of growth in two dimensions are presented for the simplest case of vanishing crystalline anisotropy and the more realistic case of a solid-liquid γ plot with cusped minima corresponding to two sets of (10 ) and (11 ) facets. The simulations reproduce many of the salient features of nanowire growth observed experimentally, including growth normal to the substrate with tapering of the side walls, transitions between different growth orientations, and crawling growth along the substrate. They also reproduce different observed relationships between the nanowire growth</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004APS..MARS22007R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004APS..MARS22007R"><span>Wetting phenomenon in the liquid-<span class="hlt">vapor</span> <span class="hlt">phase</span> coexistence of a partially miscible Lennard-Jones binary mixture</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ramírez-Santiago, Guillermo; Díaz-Herrera, Enrique; Moreno Razo, José A.</p> <p>2004-03-01</p> <p>We have carried out extensive equilibrium MD simulations to study wetting phenomena in the liquid-<span class="hlt">vapor</span> <span class="hlt">phase</span> coexistence of a partially miscible binary LJ mixture. We find that in the temperature range 0.60 ≤ T^* < 0.80, the system <span class="hlt">separates</span> forming a liquid A-liquid B interface in coexistence with the <span class="hlt">vapor</span> <span class="hlt">phase</span>. At higher temperatures, 0.80 ≤ T^* < 1.25 the liquid <span class="hlt">phases</span> are wet by the <span class="hlt">vapor</span> <span class="hlt">phase</span>. By studying the behavior of the surface tension as a function of temperature we estimate the wetting transition temperature (WTT) to be T^*_w≃ 0.80. The adsorption of molecules at the liquid-liquid interface shows a discontinuity at about T^*≃ 0.79 suggesting that the wetting transition is a first order <span class="hlt">phase</span> transition. These results are in agreement with some experiments carried out in fluid binary mixtures. In addition, we estimated the consolute temperature to be T^* _cons≃ 1.25. The calculated <span class="hlt">phase</span> diagram of the mixture suggest the existence of a tricritical point.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27869711','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27869711"><span>Preparation and Characterization of Hydrophilically Modified PVDF Membranes by a Novel Nonsolvent Thermally <span class="hlt">Induced</span> <span class="hlt">Phase</span> <span class="hlt">Separation</span> Method.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hu, Ningen; Xiao, Tonghu; Cai, Xinhai; Ding, Lining; Fu, Yuhua; Yang, Xing</p> <p>2016-11-18</p> <p>In this study, a nonsolvent thermally-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> (NTIPS) method was first proposed to fabricate hydrophilically-modified poly(vinylidene fluoride) (PVDF) membranes to overcome the drawbacks of conventional thermally-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> (TIPS) and nonsolvent-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> (NIPS) methods. Hydrophilically-modified PVDF membranes were successfully prepared by blending in hydrophilic polymer polyvinyl alcohol (PVA) at 140 °C. A series of PVDF/PVA blend membranes was prepared at different total polymer concentrations and blend ratios. The morphological analysis via SEM indicated that the formation mechanism of these hydrophilically-modified membranes was a combined NIPS and TIPS process. As the total polymer concentration increased, the tensile strength of the membranes increased; meanwhile, the membrane pore size, porosity and water flux decreased. With the PVDF/PVA blend ratio increased from 10:0 to 8:2, the membrane pore size and water flux increased. The dynamic water contact angle of these membranes showed that the hydrophilic properties of PVDF/PVA blend membranes were prominently improved. The higher hydrophilicity of the membranes resulted in reduced membrane resistance and, hence, higher permeability. The total resistance R t of the modified PVDF membranes decreased significantly as the hydrophilicity increased. The irreversible fouling related to pore blocking and adsorption fouling onto the membrane surface was minimal, indicating good antifouling properties.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=238590&Lab=NRMRL&keyword=fermentation&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50','EPA-EIMS'); return false;" href="https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=238590&Lab=NRMRL&keyword=fermentation&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50"><span>Pervaporation & <span class="hlt">Vapor</span> Permeation Membrane Processes for the Selective <span class="hlt">Separation</span> of Liquid and <span class="hlt">Vapor</span> Mixtures</span></a></p> <p><a target="_blank" href="http://oaspub.epa.gov/eims/query.page">EPA Science Inventory</a></p> <p></p> <p></p> <p>Pervaporation and <span class="hlt">vapor</span> permeation are membrane-based processes which have been proposed as alternatives to conventional <span class="hlt">separation</span> technologies. Applications range from organic solvent removal from water, ethanol or butanol recovery from dilute fermentation broths, solvent/biofu...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24138255','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24138255"><span>Continuum theory of <span class="hlt">phase</span> <span class="hlt">separation</span> kinetics for active Brownian particles.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Stenhammar, Joakim; Tiribocchi, Adriano; Allen, Rosalind J; Marenduzzo, Davide; Cates, Michael E</p> <p>2013-10-04</p> <p>Active Brownian particles (ABPs), when subject to purely repulsive interactions, are known to undergo activity-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> broadly resembling an equilibrium (attraction-<span class="hlt">induced</span>) gas-liquid coexistence. Here we present an accurate continuum theory for the dynamics of <span class="hlt">phase-separating</span> ABPs, derived by direct coarse graining, capturing leading-order density gradient terms alongside an effective bulk free energy. Such gradient terms do not obey detailed balance; yet we find coarsening dynamics closely resembling that of equilibrium <span class="hlt">phase</span> <span class="hlt">separation</span>. Our continuum theory is numerically compared to large-scale direct simulations of ABPs and accurately accounts for domain growth kinetics, domain topologies, and coexistence densities.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19580250','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19580250"><span>Freezing-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> and spatial microheterogeneity in protein solutions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Dong, Jinping; Hubel, Allison; Bischof, John C; Aksan, Alptekin</p> <p>2009-07-30</p> <p>Amid decades of research, the basic mechanisms of lyo-/cryostabilization of proteins and more complex organisms have not yet been fully established. One major bottleneck is the inability to probe into and control the molecular level interactions. The molecular interactions are responsible for the significant differences in the outcome of the preservation processes. (1) In this communication, we have utilized confocal Raman microspectroscopy to quantify the freezing-<span class="hlt">induced</span> microheterogeneity and <span class="hlt">phase</span> <span class="hlt">separation</span> (solid and liquid) in a frozen solution composed of a model protein (lysozyme) and a lyo-/cryoprotectant (trehalose), which experienced different degrees of supercooling. Detailed quantitative spectral analysis was performed across the ice, the freeze-concentrated liquid (FCL) <span class="hlt">phases</span>, and the interface region between them. It was established that the characteristics of the microstructures observed after freezing depended not only on the concentration of trehalose in the solution but also on the degree of supercooling. It was shown that, when samples were frozen after high supercooling, small amounts of lysozyme and trehalose were occluded in the ice <span class="hlt">phase</span>. Lysozyme preserved its native-like secondary structure in the FCL region but was denatured in the ice <span class="hlt">phase</span>. Also, it was observed that induction of freezing after a high degree of supercooling of high trehalose concentrations resulted in aggregation of the sugar and the protein.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20040062056&hterms=ballistic&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dballistic','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20040062056&hterms=ballistic&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dballistic"><span>Non-Ballistic <span class="hlt">Vapor</span>-Driven Ejecta</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Wrobel, K. E.; Schultz, P. H.; Heineck, J. T.</p> <p>2004-01-01</p> <p>Impact-<span class="hlt">induced</span> <span class="hlt">vaporization</span> is a key component of early-time cratering mechanics. Previous experimental [1,2] and computational [e.g., 3] studies focused on the generation and expansion of <span class="hlt">vapor</span> clouds in an attempt to better understand <span class="hlt">vaporization</span> in hypervelocity impacts. Presented here is a new experimental approach to the study of impact-<span class="hlt">induced</span> <span class="hlt">vaporization</span>. The three-dimensional particle image velocimetry (3D PIV) system captures interactions between expanding <span class="hlt">vapor</span> <span class="hlt">phases</span> and fine particulates. Particles ejected early in the cratering process may be entrained in expanding gas <span class="hlt">phases</span> generated at impact, altering their otherwise ballistic path of flight. 3D PIV allows identifying the presence of such non-ballistic ejecta from very early times in the cratering process.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5192403','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5192403"><span>Preparation and Characterization of Hydrophilically Modified PVDF Membranes by a Novel Nonsolvent Thermally <span class="hlt">Induced</span> <span class="hlt">Phase</span> <span class="hlt">Separation</span> Method</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Hu, Ningen; Xiao, Tonghu; Cai, Xinhai; Ding, Lining; Fu, Yuhua; Yang, Xing</p> <p>2016-01-01</p> <p>In this study, a nonsolvent thermally-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> (NTIPS) method was first proposed to fabricate hydrophilically-modified poly(vinylidene fluoride) (PVDF) membranes to overcome the drawbacks of conventional thermally-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> (TIPS) and nonsolvent-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> (NIPS) methods. Hydrophilically-modified PVDF membranes were successfully prepared by blending in hydrophilic polymer polyvinyl alcohol (PVA) at 140 °C. A series of PVDF/PVA blend membranes was prepared at different total polymer concentrations and blend ratios. The morphological analysis via SEM indicated that the formation mechanism of these hydrophilically-modified membranes was a combined NIPS and TIPS process. As the total polymer concentration increased, the tensile strength of the membranes increased; meanwhile, the membrane pore size, porosity and water flux decreased. With the PVDF/PVA blend ratio increased from 10:0 to 8:2, the membrane pore size and water flux increased. The dynamic water contact angle of these membranes showed that the hydrophilic properties of PVDF/PVA blend membranes were prominently improved. The higher hydrophilicity of the membranes resulted in reduced membrane resistance and, hence, higher permeability. The total resistance Rt of the modified PVDF membranes decreased significantly as the hydrophilicity increased. The irreversible fouling related to pore blocking and adsorption fouling onto the membrane surface was minimal, indicating good antifouling properties. PMID:27869711</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28705623','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28705623"><span>Water-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> of miconazole-poly (vinylpyrrolidone-co-vinyl acetate) amorphous solid dispersions: Insights with confocal fluorescence microscopy.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Saboo, Sugandha; Taylor, Lynne S</p> <p>2017-08-30</p> <p>The aim of this study was to evaluate the utility of confocal fluorescence microscopy (CFM) to study the water-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> of miconazole-poly (vinylpyrrolidone-co-vinyl acetate) (mico-PVPVA) amorphous solid dispersions (ASDs), <span class="hlt">induced</span> during preparation, upon storage at high relative humidity (RH) and during dissolution. Different fluorescent dyes were added to drug-polymer films and the location of the dyes was evaluated using CFM. Orthogonal techniques, in particular atomic force microscopy (AFM) coupled with nanoscale infrared spectroscopy (AFM-nanoIR), were used to provide additional analysis of the drug-polymer blends. The initial miscibility of mico-PVPVA ASDs prepared under low humidity conditions was confirmed by AFM-nanoIR. CFM enabled rapid identification of drug-rich and polymer-rich <span class="hlt">phases</span> in <span class="hlt">phase</span> <span class="hlt">separated</span> films prepared under high humidity conditions. The identity of drug- and polymer-rich domains was confirmed using AFM-nanoIR imaging and localized IR spectroscopy, together with Lorentz contact resonance (LCR) measurements. The CFM technique was then utilized successfully to further investigate <span class="hlt">phase</span> <span class="hlt">separation</span> in mico-PVPVA films exposed to high RH storage and to visualize <span class="hlt">phase</span> <span class="hlt">separation</span> dynamics following film immersion in buffer. CFM is thus a promising new approach to study the <span class="hlt">phase</span> behavior of ASDs, utilizing drug and polymer specific dyes to visualize the evolution of heterogeneity in films exposed to water. Copyright © 2017 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19800005954','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19800005954"><span>Methods for calculation of engineering parameters for gas <span class="hlt">separation</span>. [<span class="hlt">vapor</span> pressure and solubility of gases in organic liquids</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Lawson, D. D.</p> <p>1979-01-01</p> <p>A group additivity method is generated which allows estimation, from the structural formulas alone, of the energy of <span class="hlt">vaporization</span> and the molar volume at 25 C of many nonpolar organic liquids. Using these two parameters and appropriate thermodynamic relations, the <span class="hlt">vapor</span> pressure of the liquid <span class="hlt">phase</span> and the solubility of various gases in nonpolar organic liquids are predicted. It is also possible to use the data to evaluate organic and some inorganic liquids for use in gas <span class="hlt">separation</span> stages or liquids as heat exchange fluids in prospective thermochemical cycles for hydrogen production.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28779144','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28779144"><span>Rationalizing the light-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> of mixed halide organic-inorganic perovskites.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Draguta, Sergiu; Sharia, Onise; Yoon, Seog Joon; Brennan, Michael C; Morozov, Yurii V; Manser, Joseph S; Kamat, Prashant V; Schneider, William F; Kuno, Masaru</p> <p>2017-08-04</p> <p>Mixed halide hybrid perovskites, CH 3 NH 3 Pb(I 1-x Br x ) 3 , represent good candidates for low-cost, high efficiency photovoltaic, and light-emitting devices. Their band gaps can be tuned from 1.6 to 2.3 eV, by changing the halide anion identity. Unfortunately, mixed halide perovskites undergo <span class="hlt">phase</span> <span class="hlt">separation</span> under illumination. This leads to iodide- and bromide-rich domains along with corresponding changes to the material's optical/electrical response. Here, using combined spectroscopic measurements and theoretical modeling, we quantitatively rationalize all microscopic processes that occur during <span class="hlt">phase</span> <span class="hlt">separation</span>. Our model suggests that the driving force behind <span class="hlt">phase</span> <span class="hlt">separation</span> is the bandgap reduction of iodide-rich <span class="hlt">phases</span>. It additionally explains observed non-linear intensity dependencies, as well as self-limited growth of iodide-rich domains. Most importantly, our model reveals that mixed halide perovskites can be stabilized against <span class="hlt">phase</span> <span class="hlt">separation</span> by deliberately engineering carrier diffusion lengths and injected carrier densities.Mixed halide hybrid perovskites possess tunable band gaps, however, under illumination they undergo <span class="hlt">phase</span> <span class="hlt">separation</span>. Using spectroscopic measurements and theoretical modelling, Draguta and Sharia et al. quantitatively rationalize the microscopic processes that occur during <span class="hlt">phase</span> <span class="hlt">separation</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/7104294','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/7104294"><span>Role of lipid <span class="hlt">phase</span> <span class="hlt">separations</span> and membrane hydration in phospholipid vesicle fusion.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hoekstra, D</p> <p>1982-06-08</p> <p>The relationship between lipid <span class="hlt">phase</span> <span class="hlt">separation</span> and fusion of small unilamellar phosphatidylserine-containing vesicles was investigated. The kinetics of <span class="hlt">phase</span> <span class="hlt">separation</span> were monitored by following the increase of self-quenching of the fluorescent phospholipid analogue N-(7-nitro-2,1,3-benzoxadiazol-4-yl)phosphatidylethanolamine, which occurs when the local concentration of the probe increases upon Ca2+-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> in phosphatidylserine (PS) bilayers [Hoekstra, D. (1982) Biochemistry 21, 1055-1061]. Fusion was determined by using the resonance energy transfer fusion assay [Struck, D. K., Hoekstra, D., & Pagano, R. E. (1981) Biochemistry 20, 4093-4099], which monitors the mixing of fluorescent lipid donor and acceptor molecules, resulting in an increase in energy transfer efficiency. The results show that in the presence of Ca2+, fusion proceeds much more rapidly (t 1/2 less than 5 s) than the process of <span class="hlt">phase</span> <span class="hlt">separation</span> (T 1/2 congruent to 1 min). Mg2+ also <span class="hlt">induced</span> fusion, albeit at higher concentrations than Ca2+. Mg2+-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> were not detected, however. Subthreshold concentrations of Ca2+ (0.5 mM) or Mg2+ (2 mM) <span class="hlt">induced</span> extensive fusion of PS-containing vesicles in poly(ethylene glycol) containing media. This effect did not appear to be a poly(ethylene glycol)-facilitated enhancement of cation binding to the bilayer, and consequently Ca2+-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> was not observed. The results suggest that macroscopic <span class="hlt">phase</span> <span class="hlt">separation</span> may facilitate but does not <span class="hlt">induced</span> the fusion process and is therefore, not directly involved in the actual fusion mechanism. The fusion experiments performed in the presence of poly(ethylene glycol) suggest that the degree of bilayer dehydration and the creation of "point defects" in the bilayer without rigorous structural rearrangements in the membrane are dominant factors in the initial fusion events.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3481168','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3481168"><span>A new approach to network heterogeneity: Polymerization <span class="hlt">Induced</span> <span class="hlt">Phase</span> <span class="hlt">Separation</span> in photo-initiated, free-radical methacrylic systems</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Szczepanski, Caroline R.; Pfeifer, Carmem S.; Stansbury, Jeffrey W.</p> <p>2012-01-01</p> <p>Non-reactive, thermoplastic prepolymers (poly- methyl, ethyl and butyl methacrylate) were added to a model homopolymer matrix composed of triethylene glycol dimethacrylate (TEGDMA) to form heterogeneous networks via polymerization <span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> (PIPS). PIPS creates networks with distinct <span class="hlt">phase</span> structure that can partially compensate for volumetric shrinkage during polymerization through localized internal volume expansion. This investigation utilizes purely photo-initiated, free-radical systems, broadening the scope of applications for PIPS since these processing conditions have not been studied previously. The introduction of prepolymer into TEGDMA monomer resulted in stable, homogeneous monomer formulations, most of which underwent PIPS upon photo-irradiation, creating heterogeneous networks. During polymerization the presence of prepolymer enhanced autoacceleration, allowing for a more extensive ambient cure of the material. <span class="hlt">Phase</span> <span class="hlt">separation</span>, as characterized by dynamic changes in sample turbidity, was monitored simultaneously with monomer conversion and either preceded or was coincident with network gelation. Dynamic mechanical analysis shows a broadening of the tan delta peak and secondary peak formation, characteristic of <span class="hlt">phase-separated</span> materials, indicating one <span class="hlt">phase</span> rich in prepolymer and another depleted form upon <span class="hlt">phase</span> <span class="hlt">separation</span>. In certain cases, PIPS leads to an enhanced physical reduction of volumetric shrinkage, which is attractive for many applications including dental composite materials. PMID:23109733</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1999PhDT.......306B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1999PhDT.......306B"><span>Formation of anisotropic hollow-fiber membranes via thermally <span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Batarseh, Melanie Turkett</p> <p></p> <p>The goal of this research project was to study the formation of anisotropic hollow fiber membranes via thermally <span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> (TIPS). This objective included developing a fundamental knowledge of the factors that contribute to anisotropy and studying how anisotropy can be controlled via operational parameters in hollow fiber spinning. The objective was met by creating a model to simulate the mass and heat transfer in the fiber wall during spinning and by experimentally varying spinning parameters and observing the affect on the membrane microstructure. The TIPS membrane formation process consists of forming a homogeneous solution of polymer and diluent and extruding the solution through a spinneret to form a hollow fiber. The fiber is cooled in an air gap followed by a quench bath, which results in <span class="hlt">phase</span> <span class="hlt">separation</span> of the solution into a diluent-rich <span class="hlt">phase</span> dispersed in a continuous polymer-rich liquid <span class="hlt">phase</span>. The diluent-rich domains grow in size until the polymer-rich <span class="hlt">phase</span> crystallizes. Then the diluent is removed, and the spaces left behind become the pores of the microporous membrane. Therefore, the size of the diluent-rich domains when the polymer solidifies is related to the size of the pores in the finished membrane. Increasing the polymer concentration of the homogeneous solution or increasing the cooling rate of the <span class="hlt">phase</span> <span class="hlt">separated</span> solution decreases the domain size, and thus decreases pore size. An anisotropic membrane, which has a gradation of pore size from small pores at the feed-side to large pores at the permeate-side, can be formed by creating a concentration gradient or a cooling rate gradient across the membrane. In hollow fiber spinning, a concentration gradient can be created by allowing diluent to evaporate from the outside wall of the fiber in the air gap, and a cooling rate gradient can be created by quenching the fiber in a liquid bath. The spinning model calculates concentration and temperature profiles across the hollow fiber</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1024115','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/1024115"><span>Liquid-<span class="hlt">phase</span> and <span class="hlt">vapor-phase</span> dehydration of organic/water solutions</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Huang, Yu [Palo Alto, CA; Ly, Jennifer [San Jose, CA; Aldajani, Tiem [San Jose, CA; Baker, Richard W [Palo Alto, CA</p> <p>2011-08-23</p> <p>Processes for dehydrating an organic/water solution by pervaporation or <span class="hlt">vapor</span> <span class="hlt">separation</span> using fluorinated membranes. The processes are particularly useful for treating mixtures containing light organic components, such as ethanol, isopropanol or acetic acid.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017NanoL..17.1028B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017NanoL..17.1028B"><span>Origin of Reversible Photoinduced <span class="hlt">Phase</span> <span class="hlt">Separation</span> in Hybrid Perovskites</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bischak, Connor G.; Hetherington, Craig L.; Wu, Hao; Aloni, Shaul; Ogletree, D. Frank; Limmer, David T.; Ginsberg, Naomi S.</p> <p>2017-02-01</p> <p>Nonequilibrium processes occurring in functional materials can significantly impact device efficiencies and are often difficult to characterize due to the broad range of length and time scales involved. In particular, mixed halide hybrid perovskites are promising for optoelectronics, yet the halides reversibly <span class="hlt">phase</span> <span class="hlt">separate</span> when photo-excited, significantly altering device performance. By combining nanoscale imaging and multiscale modeling, we elucidate the mechanism underlying this phenomenon, demonstrating that local strain <span class="hlt">induced</span> by photo-generated polarons promotes halide <span class="hlt">phase</span> <span class="hlt">separation</span> and leads to nucleation of light-stabilized iodide-rich clusters. This effect relies on the unique electromechanical properties of hybrid materials, characteristic of neither their organic nor inorganic constituents alone. Exploiting photo-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> and other nonequilibrium phenomena in hybrid materials, generally, could enable new opportunities for expanding the functional applications in sensing, photoswitching, optical memory, and energy storage.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017NatMa..16.1022T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017NatMa..16.1022T"><span>Formation of porous crystals via viscoelastic <span class="hlt">phase</span> <span class="hlt">separation</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tsurusawa, Hideyo; Russo, John; Leocmach, Mathieu; Tanaka, Hajime</p> <p>2017-10-01</p> <p>Viscoelastic <span class="hlt">phase</span> <span class="hlt">separation</span> of colloidal suspensions can be interrupted to form gels either by glass transition or by crystallization. With a new confocal microscopy protocol, we follow the entire kinetics of <span class="hlt">phase</span> <span class="hlt">separation</span>, from homogeneous <span class="hlt">phase</span> to different arrested states. For the first time in experiments, our results unveil a novel crystallization pathway to sponge-like porous crystal structures. In the early stages, we show that nucleation requires a structural reorganization of the liquid <span class="hlt">phase</span>, called stress-driven ageing. Once nucleation starts, we observe that crystallization follows three different routes: direct crystallization of the liquid <span class="hlt">phase</span>, the Bergeron process, and Ostwald ripening. Nucleation starts inside the reorganized network, but crystals grow past it by direct condensation of the gas <span class="hlt">phase</span> on their surface, driving liquid evaporation, and producing a network structure different from the original <span class="hlt">phase</span> <span class="hlt">separation</span> pattern. We argue that similar crystal-gel states can be formed in monatomic and molecular systems if the liquid <span class="hlt">phase</span> is slow enough to <span class="hlt">induce</span> viscoelastic <span class="hlt">phase</span> <span class="hlt">separation</span>, but fast enough to prevent immediate vitrification. This provides a novel pathway to form nanoporous crystals of metals and semiconductors without dealloying, which may be important for catalytic, optical, sensing, and filtration applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70025943','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70025943"><span>Extraordinary <span class="hlt">phase</span> <span class="hlt">separation</span> and segregation in vent fluids from the southern East Pacific Rise</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Von Damm, Karen L.; Lilley, M.D.; Shanks, Wayne C.; Brockington, M.; Bray, A.M.; O'Grady, K. M.; Olson, E.; Graham, A.; Proskurowski, G.</p> <p>2003-01-01</p> <p>The discovery of Brandon vent on the southern East Pacific Rise is providing new insights into the controls on midocean ridge hydrothermal vent fluid chemistry. The physical conditions at the time ofsampling (287 bar and 405??C) place the Brandon fluids very close to the critical point of seawater (298 bar and 407??C). This permits in situ study of the effects of near criticalphenomena, which are interpreted to be the primary cause of enhanced transition metal transport in these fluids. Of the five orifices on Brandon sampled, three were venting fluids with less than seawater chlorinity, and two were venting fluids with greater than seawater chlorinity. The liquid <span class="hlt">phase</span> orifices contain 1.6-1.9 times the chloride content of the <span class="hlt">vapors</span>. Most other elements, excluding the gases, have this same ratio demonstrating the conservative nature of <span class="hlt">phase</span> <span class="hlt">separation</span> and the lack of subsequent water-rock interaction. The <span class="hlt">vapor</span> and liquid <span class="hlt">phases</span> vent at the same time from orifices within meters of each other on the Brandon structure. Variations in fluid compositions occur on a time scale of minutes. Our interpretation is that <span class="hlt">phase</span> <span class="hlt">separation</span> and segregation must be occurring 'real time' within the sulfide structure itself. Fluids from Brandon therefore provide an unique opportunity to understand in situ <span class="hlt">phase</span> <span class="hlt">separation</span> without the overprinting of continued water-rock interaction with the oceanic crust, as well as critical phenomena. ?? 2002 Elsevier Science B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014Nanos...6.1467G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014Nanos...6.1467G"><span>APTS and rGO co-functionalized pyrenated fluorescent nanonets for representative <span class="hlt">vapor</span> <span class="hlt">phase</span> nitroaromatic explosive detection</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Guo, Linjuan; Zu, Baiyi; Yang, Zheng; Cao, Hongyu; Zheng, Xuefang; Dou, Xincun</p> <p>2014-01-01</p> <p>For the first time, flexible PVP/pyrene/APTS/rGO fluorescent nanonets were designed and synthesized via a one-step electrospinning method to detect representative subsaturated nitroaromatic explosive <span class="hlt">vapor</span>. The functional fluorescent nanonets, which were highly stable in air, showed an 81% quenching efficiency towards TNT <span class="hlt">vapor</span> (~10 ppb) with an exposure time of 540 s at room temperature. The nice performance of the nanonets was ascribed to the synergistic effects <span class="hlt">induced</span> by the specific adsorption properties of APTS, the fast charge transfer properties and the effective π-π interaction with pyrene and TNT of rGO. Compared to the analogues of TNT, the PVP/pyrene/APTS/rGO nanonets showed notable selectivity towards TNT and DNT <span class="hlt">vapors</span>. The explored functionalization method opens up brand new insight into sensitive and selective detection of <span class="hlt">vapor</span> <span class="hlt">phase</span> nitroaromatic explosives.For the first time, flexible PVP/pyrene/APTS/rGO fluorescent nanonets were designed and synthesized via a one-step electrospinning method to detect representative subsaturated nitroaromatic explosive <span class="hlt">vapor</span>. The functional fluorescent nanonets, which were highly stable in air, showed an 81% quenching efficiency towards TNT <span class="hlt">vapor</span> (~10 ppb) with an exposure time of 540 s at room temperature. The nice performance of the nanonets was ascribed to the synergistic effects <span class="hlt">induced</span> by the specific adsorption properties of APTS, the fast charge transfer properties and the effective π-π interaction with pyrene and TNT of rGO. Compared to the analogues of TNT, the PVP/pyrene/APTS/rGO nanonets showed notable selectivity towards TNT and DNT <span class="hlt">vapors</span>. The explored functionalization method opens up brand new insight into sensitive and selective detection of <span class="hlt">vapor</span> <span class="hlt">phase</span> nitroaromatic explosives. Electronic supplementary information (ESI) available: <span class="hlt">Vapor</span> pressure of TNT and its analogues, fluorescence quenching kinetics, fluorescence quenching efficiencies and additional SEM images. See DOI: 10.1039/c3nr04960d</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18558771','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18558771"><span><span class="hlt">Vapor-phase</span> infrared laser spectroscopy: from gas sensing to forensic urinalysis.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bartlome, Richard; Rey, Julien M; Sigrist, Markus W</p> <p>2008-07-15</p> <p>Numerous gas-sensing devices are based on infrared laser spectroscopy. In this paper, the technique is further developed and, for the first time, applied to forensic urinalysis. For this purpose, a difference frequency generation laser was coupled to an in-house-built, high-temperature multipass cell (HTMC). The continuous tuning range of the laser was extended to 329 cm(-1) in the fingerprint C-H stretching region between 3 and 4 microm. The HTMC is a long-path absorption cell designed to withstand organic samples in the <span class="hlt">vapor</span> <span class="hlt">phase</span> (Bartlome, R.; Baer, M.; Sigrist, M. W. Rev. Sci. Instrum. 2007, 78, 013110). Quantitative measurements were taken on pure ephedrine and pseudoephedrine <span class="hlt">vapors</span>. Despite featuring similarities, the <span class="hlt">vapor-phase</span> infrared spectra of these diastereoisomers are clearly distinguishable with respect to a vibrational band centered at 2970.5 and 2980.1 cm(-1), respectively. Ephedrine-positive and pseudoephedrine-positive urine samples were prepared by means of liquid-liquid extraction and directly evaporated in the HTMC without any preliminary chromatographic <span class="hlt">separation</span>. When 10 or 20 mL of ephedrine-positive human urine is prepared, the detection limit of ephedrine, prohibited in sports as of 10 microg/mL, is 50 or 25 microg/mL, respectively. The laser spectrometer has room for much improvement; its potential is discussed with respect to doping agents detection.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_2");'>2</a></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li class="active"><span>4</span></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_4 --> <div id="page_5" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li class="active"><span>5</span></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="81"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=241491&Lab=NRMRL&keyword=fermentation&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50','EPA-EIMS'); return false;" href="https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=241491&Lab=NRMRL&keyword=fermentation&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50"><span>Pervaporation and <span class="hlt">Vapor</span> Permeation Tutorial: Membrane Processes for the Selective <span class="hlt">Separation</span> of Liquid and <span class="hlt">Vapor</span> Mixtures</span></a></p> <p><a target="_blank" href="http://oaspub.epa.gov/eims/query.page">EPA Science Inventory</a></p> <p></p> <p></p> <p>Pervaporation and <span class="hlt">vapor</span> permeation are membrane-based processes proposed as alternatives to conventional <span class="hlt">separation</span> technologies. Applications range from organic solvent removal from water, ethanol or butanol recovery from fermentation broths, solvent/biofuel dehydration to meet ...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1998APS..MAR.K1505K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1998APS..MAR.K1505K"><span>Visualization of Two-<span class="hlt">Phase</span> Fluid Distribution Using Laser <span class="hlt">Induced</span> Exciplex Fluorescence</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kim, J. U.; Darrow, J.; Schock, H.; Golding, B.; Nocera, D.; Keller, P.</p> <p>1998-03-01</p> <p>Laser-<span class="hlt">induced</span> exciplex (excited state complex) fluorescence has been used to generate two-dimensional images of dispersed liquid and <span class="hlt">vapor</span> <span class="hlt">phases</span> with spectrally resolved two-color emissions. In this method, the <span class="hlt">vapor</span> <span class="hlt">phase</span> is tagged by the monomer fluorescence while the liquid <span class="hlt">phase</span> is tracked by the exciplex fluorescence. A new exciplex visualization system consisting of DMA and 1,4,6-TMN in an isooctane solvent was developed.(J.U. Kim et al., Chem. Phys. Lett. 267, 323-328 (1997)) The direct ca</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JaJAP..56dCJ04S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JaJAP..56dCJ04S"><span>Thermodynamic considerations of the <span class="hlt">vapor</span> <span class="hlt">phase</span> reactions in III-nitride metal organic <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sekiguchi, Kazuki; Shirakawa, Hiroki; Chokawa, Kenta; Araidai, Masaaki; Kangawa, Yoshihiro; Kakimoto, Koichi; Shiraishi, Kenji</p> <p>2017-04-01</p> <p>We analyzed the metal organic <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxial growth mechanism of the III-nitride semiconductors GaN, AlN, and InN by first-principles calculations and thermodynamic analyses. In these analyses, we investigated the decomposition processes of the group III source gases X(CH3)3 (X = Ga, Al, In) at finite temperatures and determined whether the (CH3)2GaNH2 adduct can be formed or not. The results of our calculations show that the (CH3)2GaNH2 adduct cannot be formed in the gas <span class="hlt">phase</span> in GaN metal organic <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy (MOVPE), whereas, in AlN MOVPE, the formation of the (CH3)2AlNH2 adduct in the gas <span class="hlt">phase</span> is exclusive. In the case of GaN MOVPE, trimethylgallium (TMG, [Ga(CH3)3]) decomposition into Ga gas on the growth surface with the assistance of H2 carrier gas, instead of the formation of the (CH3)2GaNH2 adduct, occurs almost exclusively. Moreover, in the case of InN MOVPE, the formation of the (CH3)2InNH2 adduct does not occur and it is relatively easy to produce In gas even without H2 in the carrier gas.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1425742','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1425742"><span>Rationalizing the light-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> of mixed halide organic–inorganic perovskites</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Draguta, Sergiu; Sharia, Onise; Yoon, Seog Joon</p> <p></p> <p>Mixed halide hybrid perovskites, CH 3NH 3Pb(I 1-xBrx) 3' represent good candidates for lowcost, high efficiency photovoltaic, and light-emitting devices. Their band gaps can be tuned from 1.6 to 2.3 eV, by changing the halide anion identity. Unfortunately, mixed halide perovskites undergo <span class="hlt">phase</span> <span class="hlt">separation</span> under illumination. This leads to iodide- and bromide-rich domains along with corresponding changes to the material’s optical/electrical response. Here, using combined spectroscopic measurements and theoretical modeling, we quantitatively rationalize all microscopic processes that occur during <span class="hlt">phase</span> <span class="hlt">separation</span>. Our model suggests that the driving force behind <span class="hlt">phase</span> <span class="hlt">separation</span> is the bandgap reduction of iodiderich <span class="hlt">phases</span>. It additionallymore » explains observed non-linear intensity dependencies, as well as self-limited growth of iodide-rich domains. Most importantly, our model reveals that mixed halide perovskites can be stabilized against <span class="hlt">phase</span> <span class="hlt">separation</span> by deliberately engineering carrier diffusion lengths and injected carrier densities.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1425742-rationalizing-light-induced-phase-separation-mixed-halide-organicinorganic-perovskites','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1425742-rationalizing-light-induced-phase-separation-mixed-halide-organicinorganic-perovskites"><span>Rationalizing the light-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> of mixed halide organic–inorganic perovskites</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Draguta, Sergiu; Sharia, Onise; Yoon, Seog Joon; ...</p> <p>2017-08-04</p> <p>Mixed halide hybrid perovskites, CH 3NH 3Pb(I 1-xBrx) 3' represent good candidates for lowcost, high efficiency photovoltaic, and light-emitting devices. Their band gaps can be tuned from 1.6 to 2.3 eV, by changing the halide anion identity. Unfortunately, mixed halide perovskites undergo <span class="hlt">phase</span> <span class="hlt">separation</span> under illumination. This leads to iodide- and bromide-rich domains along with corresponding changes to the material’s optical/electrical response. Here, using combined spectroscopic measurements and theoretical modeling, we quantitatively rationalize all microscopic processes that occur during <span class="hlt">phase</span> <span class="hlt">separation</span>. Our model suggests that the driving force behind <span class="hlt">phase</span> <span class="hlt">separation</span> is the bandgap reduction of iodiderich <span class="hlt">phases</span>. It additionallymore » explains observed non-linear intensity dependencies, as well as self-limited growth of iodide-rich domains. Most importantly, our model reveals that mixed halide perovskites can be stabilized against <span class="hlt">phase</span> <span class="hlt">separation</span> by deliberately engineering carrier diffusion lengths and injected carrier densities.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19850020200','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19850020200"><span>Water <span class="hlt">vapor</span> radiometry research and development <span class="hlt">phase</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Resch, G. M.; Chavez, M. C.; Yamane, N. L.; Barbier, K. M.; Chandlee, R. C.</p> <p>1985-01-01</p> <p>This report describes the research and development <span class="hlt">phase</span> for eight dual-channel water <span class="hlt">vapor</span> radiometers constructed for the Crustal Dynamics Project at the Goddard Space Flight Center, Greenbelt, Maryland, and for the NASA Deep Space Network. These instruments were developed to demonstrate that the variable path delay imposed on microwave radio transmissions by atmospheric water <span class="hlt">vapor</span> can be calibrated, particularly as this phenomenon affects very long baseline interferometry measurement systems. Water <span class="hlt">vapor</span> radiometry technology can also be used in systems that involve moist air meteorology and propagation studies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/872466','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/872466"><span>Combined heat and mass transfer device for improving <span class="hlt">separation</span> process</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Tran, Thanh Nhon</p> <p>1999-01-01</p> <p>A two-<span class="hlt">phase</span> small channel heat exchange matrix simultaneously provides for heat transfer and mass transfer between the liquid and <span class="hlt">vapor</span> <span class="hlt">phases</span> of a multi-component mixture at a single, predetermined location within a <span class="hlt">separation</span> column, significantly improving the thermodynamic efficiency of the <span class="hlt">separation</span> process. The small channel heat exchange matrix is composed of a series of channels having a hydraulic diameter no greater than 5.0 millimeters for conducting a two-<span class="hlt">phase</span> coolant. In operation, the matrix provides the liquid-<span class="hlt">vapor</span> contacting surfaces within the <span class="hlt">separation</span> column, such that heat and mass are transferred simultaneously between the liquid and <span class="hlt">vapor</span> <span class="hlt">phases</span>. The two-<span class="hlt">phase</span> coolant allows for a uniform heat transfer coefficient to be maintained along the length of the channels and across the surface of the matrix. Preferably, a perforated, concave sheet connects each channel to an adjacent channel to facilitate the flow of the liquid and <span class="hlt">vapor</span> <span class="hlt">phases</span> within the column and to increase the liquid-<span class="hlt">vapor</span> contacting surface area.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/678598','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/biblio/678598"><span>Combined heat and mass transfer device for improving <span class="hlt">separation</span> process</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Tran, T.N.</p> <p>1999-08-24</p> <p>A two-<span class="hlt">phase</span> small channel heat exchange matrix simultaneously provides for heat transfer and mass transfer between the liquid and <span class="hlt">vapor</span> <span class="hlt">phases</span> of a multi-component mixture at a single, predetermined location within a <span class="hlt">separation</span> column, significantly improving the thermodynamic efficiency of the <span class="hlt">separation</span> process. The small channel heat exchange matrix is composed of a series of channels having a hydraulic diameter no greater than 5.0 millimeters for conducting a two-<span class="hlt">phase</span> coolant. In operation, the matrix provides the liquid-<span class="hlt">vapor</span> contacting surfaces within the <span class="hlt">separation</span> column, such that heat and mass are transferred simultaneously between the liquid and <span class="hlt">vapor</span> <span class="hlt">phases</span>. The two-<span class="hlt">phase</span> coolant allows for a uniform heat transfer coefficient to be maintained along the length of the channels and across the surface of the matrix. Preferably, a perforated, concave sheet connects each channel to an adjacent channel to facilitate the flow of the liquid and <span class="hlt">vapor</span> <span class="hlt">phases</span> within the column and to increase the liquid-<span class="hlt">vapor</span> contacting surface area. 12 figs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013APS..DFD.E7003N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013APS..DFD.E7003N"><span>Dehydration <span class="hlt">induced</span> <span class="hlt">phase</span> transitions in a microfluidic droplet array for the <span class="hlt">separation</span> of biomolecules</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nelson, Chris; Anna, Shelley</p> <p>2013-11-01</p> <p>Droplet-based strategies for fluid manipulation have seen significant application in microfluidics due to their ability to compartmentalize solutions and facilitate highly parallelized reactions. Functioning as micro-scale reaction vessels, droplets have been used to study protein crystallization, enzyme kinetics, and to encapsulate whole cells. Recently, the mass transport out of droplets has been used to concentrate solutions and <span class="hlt">induce</span> <span class="hlt">phase</span> transitions. Here, we show that droplets trapped in a microfluidic array will spontaneously dehydrate over the course of several hours. By loading these devices with an initially dilute aqueous polymer solution, we use this slow dehydration to observe <span class="hlt">phase</span> transitions and the evolution of droplet morphology in hundreds of droplets simultaneously. As an example, we trap and dehydrate droplets of a model aqueous two-<span class="hlt">phase</span> system consisting of polyethylene glycol and dextran. Initially the drops are homogenous, then after some time the polymer concentration reaches a critical point and two <span class="hlt">phases</span> form. As water continues to leave the system, the drops transition from a microemulsion of DEX in PEG to a core-shell configuration. Eventually, changes in interfacial tension, driven by dehydration, cause the DEX core to completely de-wet from the PEG shell. Since aqueous two <span class="hlt">phase</span> systems are able to selectively <span class="hlt">separate</span> a variety of biomolecules, this core shedding behavior has the potential to provide selective, on-chip <span class="hlt">separation</span> and concentration.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/879760','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/879760"><span>Gas <span class="hlt">Separation</span> Using Organic-<span class="hlt">Vapor</span>-Resistent Membranes In Conjunctin With Organic-<span class="hlt">Vapor</span>-Selective Membranes</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Baker, Richard W.; Pinnau, Ingo; He, Zhenjie; Da Costa, Andre R.; Daniels, Ramin; Amo, Karl D.; Wijmans, Johannes G.</p> <p>2003-06-03</p> <p>A process for treating a gas mixture containing at least an organic compound gas or <span class="hlt">vapor</span> and a second gas, such as natural gas, refinery off-gas or air. The process uses two sequential membrane <span class="hlt">separation</span> steps, one using membrane selective for the organic compound over the second gas, the other selective for the second gas over the organic <span class="hlt">vapor</span>. The second-gas-selective membranes use a selective layer made from a polymer having repeating units of a fluorinated polymer, and demonstrate good resistance to plasticization by the organic components in the gas mixture under treatment, and good recovery after exposure to liquid aromatic hydrocarbons. The membrane steps can be combined in either order.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19980008385','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19980008385"><span>Application of Thioether for <span class="hlt">Vapor</span> <span class="hlt">Phase</span> Lubrication</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Graham, E. Earl</p> <p>1997-01-01</p> <p>The objective of these studies was to identify the optimal conditions for <span class="hlt">vapor</span> <span class="hlt">phase</span> lubrication using Thioether for both sliding and rolling wear. The important variable include; (1) The component materials including M50 steel, monel and silicon nitride. (2) The <span class="hlt">vapor</span> concentration and flow rate. (3) The temperature in the range of 600 F to 1500 F. (4) The loads and rolling and/or sliding speeds.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1339827-scalable-production-method-graphene-oxide-water-vapor-separation-membranes','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1339827-scalable-production-method-graphene-oxide-water-vapor-separation-membranes"><span>Scalable Production Method for Graphene Oxide Water <span class="hlt">Vapor</span> <span class="hlt">Separation</span> Membranes</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Fifield, Leonard S.; Shin, Yongsoon; Liu, Wei</p> <p></p> <p>ABSTRACT Membranes for selective water <span class="hlt">vapor</span> <span class="hlt">separation</span> were assembled from graphene oxide suspension using techniques compatible with high volume industrial production. The large-diameter graphene oxide flake suspensions were synthesized from graphite materials via relatively efficient chemical oxidation steps with attention paid to maintaining flake size and achieving high graphene oxide concentrations. Graphene oxide membranes produced using scalable casting methods exhibited water <span class="hlt">vapor</span> flux and water/nitrogen selectivity performance meeting or exceeding that of membranes produced using vacuum-assisted laboratory techniques. (PNNL-SA-117497)</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28134530','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28134530"><span>Origin of Reversible Photoinduced <span class="hlt">Phase</span> <span class="hlt">Separation</span> in Hybrid Perovskites.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bischak, Connor G; Hetherington, Craig L; Wu, Hao; Aloni, Shaul; Ogletree, D Frank; Limmer, David T; Ginsberg, Naomi S</p> <p>2017-02-08</p> <p>The distinct physical properties of hybrid organic-inorganic materials can lead to unexpected nonequilibrium phenomena that are difficult to characterize due to the broad range of length and time scales involved. For instance, mixed halide hybrid perovskites are promising materials for optoelectronics, yet bulk measurements suggest the halides reversibly <span class="hlt">phase</span> <span class="hlt">separate</span> upon photoexcitation. By combining nanoscale imaging and multiscale modeling, we find that the nature of halide demixing in these materials is distinct from macroscopic <span class="hlt">phase</span> <span class="hlt">separation</span>. We propose that the localized strain <span class="hlt">induced</span> by a single photoexcited charge interacting with the soft, ionic lattice is sufficient to promote halide <span class="hlt">phase</span> <span class="hlt">separation</span> and nucleate a light-stabilized, low-bandgap, ∼8 nm iodide-rich cluster. The limited extent of this polaron is essential to promote demixing because by contrast bulk strain would simply be relaxed. Photoinduced <span class="hlt">phase</span> <span class="hlt">separation</span> is therefore a consequence of the unique electromechanical properties of this hybrid class of materials. Exploiting photoinduced <span class="hlt">phase</span> <span class="hlt">separation</span> and other nonequilibrium phenomena in hybrid materials more generally could expand applications in sensing, switching, memory, and energy storage.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1338488-superhydrophobic-superhydrophilic-surface-enhanced-separation-performance-porous-inorganic-membranes-biomass-biofuel-conversion-applications','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1338488-superhydrophobic-superhydrophilic-surface-enhanced-separation-performance-porous-inorganic-membranes-biomass-biofuel-conversion-applications"><span>Superhydrophobic and superhydrophilic surface-enhanced <span class="hlt">separation</span> performance of porous inorganic membranes for biomass-to-biofuel conversion applications</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Hu, Michael Z.; Engtrakul, Chaiwat; Bischoff, Brian L.; ...</p> <p>2016-11-14</p> <p>A new class of inorganic-based membranes, i.e., High-Performance Architectured Surface Selective (HiPAS) membranes, is introduced to provide high perm-selective flux by exploiting unique <span class="hlt">separation</span> mechanisms <span class="hlt">induced</span> by superhydrophobic or superhydrophilic surface interactions and confined capillary condensation in enlarged membrane pores (~8 nm). The super-hydro-tunable HiPAS membranes were originally developed for the purpose of bio-oil/biofuel processing to achieve selective <span class="hlt">separations</span> at higher flux relative to size selective porous membranes (e.g., inorganic zeolite-based membranes) and better high-temperature tolerance than polymer membranes (>250 C) for hot <span class="hlt">vapor</span> processing. Due to surface-enhanced <span class="hlt">separation</span> selectivity, HiPAS membranes can thus possibly enable larger pores to facilitatemore » large-flux <span class="hlt">separations</span> by increasing from sub-nanometer pores to mesopores (2-50 nm) for <span class="hlt">vapor</span> <span class="hlt">phase</span> or micron-scale pores for liquid <span class="hlt">phase</span> <span class="hlt">separations</span>. In this paper, we describe an innovative membrane concept and a materials synthesis strategy to fabricate HiPAS membranes, and demonstrate selective permeation in both <span class="hlt">vapor</span>- and liquid-<span class="hlt">phase</span> applications. High permeability and selectivity were demonstrated using surrogate mixtures, such as ethanol-water, toluene-water, and toluene-phenol-water. The overall membrane evaluation results show promise for the future processing of biomass pyrolysis and upgraded product <span class="hlt">vapors</span> and condensed liquid bio-oil intermediates.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4703326','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4703326"><span><span class="hlt">Phase</span>-transition thresholds and <span class="hlt">vaporization</span> phenomena for ultrasound <span class="hlt">phase</span>-change nanoemulsions assessed via high speed optical microscopy</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Sheeran, Paul S.; Matsunaga, Terry O.; Dayton, Paul A.</p> <p>2015-01-01</p> <p>Ultrasonically activated <span class="hlt">phase</span>-change contrast agents (PCCAs) based on perfluorocarbon droplets have been proposed for a variety of therapeutic and diagnostic clinical applications. When generated at the nanoscale, droplets may be small enough to exit the vascular space and then be <span class="hlt">induced</span> to <span class="hlt">vaporize</span> with high spatial and temporal specificity by externally-applied ultrasound. The use of acoustical techniques for optimizing ultrasound parameters for given applications can be a significant challenge for nanoscale PCCAs due to the contributions of larger outlier droplets. Similarly, optical techniques can be a challenge due to the sub-micron size of nanodroplet agents and resolution limits of optical microscopy. In this study, an optical method for determining activation thresholds of nanoscale emulsions based on the in vitro distribution of bubbles resulting from <span class="hlt">vaporization</span> of PCCAs after single, short (<10 cycles) ultrasound pulses is evaluated. Through ultra-high-speed microscopy it is shown that the bubbles produced early in the pulse from <span class="hlt">vaporized</span> droplets are strongly affected by subsequent cycles of the <span class="hlt">vaporization</span> pulse, and these effects increase with pulse length. Results show that decafluorobutane nanoemulsions with peak diameters on the order of 200 nm can be optimally <span class="hlt">vaporized</span> with short pulses using pressures amenable to clinical diagnostic ultrasound machines. PMID:23760161</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27435379','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27435379"><span>Liquid-Desiccant <span class="hlt">Vapor</span> <span class="hlt">Separation</span> Reduces the Energy Requirements of Atmospheric Moisture Harvesting.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gido, Ben; Friedler, Eran; Broday, David M</p> <p>2016-08-02</p> <p>An innovative atmospheric moisture harvesting system is proposed, where water <span class="hlt">vapor</span> is <span class="hlt">separated</span> from the air prior to cooling and condensation. The system was studied using a model that simulates its three interconnected cycles (air, desiccant, and water) over a range of ambient conditions, and optimal configurations are reported for different operation conditions. Model results were compared to specifications of commercial atmospheric moisture harvesting systems and found to represent saving of 5-65% of the electrical energy requirements due to the <span class="hlt">vapor</span> <span class="hlt">separation</span> process. We show that the liquid desiccant <span class="hlt">separation</span> stage that is integrated into atmospheric moisture harvesting systems can work under a wide range of environmental conditions using low grade or solar heating as a supplementary energy source, and that the performance of the combined system is superior.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27116639','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27116639"><span>Role of the source to building lateral <span class="hlt">separation</span> distance in petroleum <span class="hlt">vapor</span> intrusion.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Verginelli, Iason; Capobianco, Oriana; Baciocchi, Renato</p> <p>2016-06-01</p> <p>The adoption of source to building <span class="hlt">separation</span> distances to screen sites that need further field investigation is becoming a common practice for the evaluation of the <span class="hlt">vapor</span> intrusion pathway at sites contaminated by petroleum hydrocarbons. Namely, for the source to building vertical distance, the screening criteria for petroleum <span class="hlt">vapor</span> intrusion have been deeply investigated in the recent literature and fully addressed in the recent guidelines issued by ITRC and U.S.EPA. Conversely, due to the lack of field and modeling studies, the source to building lateral distance received relatively low attention. To address this issue, in this work we present a steady-state <span class="hlt">vapor</span> intrusion analytical model incorporating a piecewise first-order aerobic biodegradation limited by oxygen availability that accounts for lateral source to building <span class="hlt">separation</span>. The developed model can be used to evaluate the role and relevance of lateral <span class="hlt">vapor</span> attenuation as well as to provide a site-specific assessment of the lateral screening distances needed to attenuate <span class="hlt">vapor</span> concentrations to risk-based values. The simulation outcomes showed to be consistent with field data and 3-D numerical modeling results reported in previous studies and, for shallow sources, with the screening criteria recommended by U.S.EPA for the vertical <span class="hlt">separation</span> distance. Indeed, although petroleum <span class="hlt">vapors</span> can cover maximum lateral distances up to 25-30m, as highlighted by the comparison of model outputs with field evidences of <span class="hlt">vapor</span> migration in the subsurface, simulation results by this new model indicated that, regardless of the source concentration and depth, 6m and 7m lateral distances are sufficient to attenuate petroleum <span class="hlt">vapors</span> below risk-based values for groundwater and soil sources, respectively. However, for deep sources (>5m) and for low to moderate source concentrations (benzene concentrations lower than 5mg/L in groundwater and 0.5mg/kg in soil) the above criteria were found extremely conservative as the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JPCM...28x4012O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JPCM...28x4012O"><span>Density functional theory of gas-liquid <span class="hlt">phase</span> <span class="hlt">separation</span> in dilute binary mixtures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Okamoto, Ryuichi; Onuki, Akira</p> <p>2016-06-01</p> <p>We examine statics and dynamics of <span class="hlt">phase-separated</span> states of dilute binary mixtures using density functional theory. In our systems, the difference of the solvation chemical potential between liquid and gas Δ {μ\\text{s}} (the Gibbs energy of transfer) is considerably larger than the thermal energy {{k}\\text{B}}T for each solute particle and the attractive interaction among the solute particles is weaker than that among the solvent particles. In these conditions, the saturated <span class="hlt">vapor</span> pressure increases by {{k}\\text{B}}Tn2\\ell\\exp ≤ft(Δ {μ\\text{s}}/{{k}\\text{B}}T\\right) , where n2\\ell is the solute density added in liquid. For \\exp ≤ft(Δ {μ\\text{s}}/{{k}\\text{B}}T\\right)\\gg 1 , <span class="hlt">phase</span> <span class="hlt">separation</span> is <span class="hlt">induced</span> at low solute densities in liquid and the new <span class="hlt">phase</span> remains in gaseous states, even when the liquid pressure is outside the coexistence curve of the solvent. This explains the widely observed formation of stable nanobubbles in ambient water with a dissolved gas. We calculate the density and stress profiles across planar and spherical interfaces, where the surface tension decreases with increasing interfacial solute adsorption. We realize stable solute-rich bubbles with radius about 30 nm, which minimize the free energy functional. We then study dynamics around such a bubble after a decompression of the surrounding liquid, where the bubble undergoes a damped oscillation. In addition, we present some exact and approximate expressions for the surface tension and the interfacial stress tensor.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1333074','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1333074"><span>Volatile particles measured by <span class="hlt">vapor</span>-particle <span class="hlt">separator</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Cheng, Meng -Dawn; Corporan, Edwin</p> <p></p> <p><span class="hlt">Vapor</span>-Particle <span class="hlt">Separator</span> (VPS) is a new technology developed for characterization of the volatile fraction of particulate matter in a combustion aerosol population. VPS incorporates a novel metallic membrane and operates in a cross-flow filtration mode for <span class="hlt">separation</span> of <span class="hlt">vapor</span> and solid (i.e. non-volatile) particles. Demonstration of the VPS technology on aircraft engine-emitted particles has led to the improvement of the technology and increased confidence on the robustness of its field performance. In this study, the performance of the VPS was evaluated against the Particle Measurement Programme (PMP) volatile particle remover (VPR), a standardized device used in heavy duty diesel enginesmore » for <span class="hlt">separation</span> and characterization of non-volatile particulate matter. Using tetracontane particles in the laboratory reveals that the VPS performed reasonably well in removing the volatile species. In the field conditions, a single-mode particle size distribution was found for emitted particles from a T63 turboshaft engine at both idle and cruise engine power conditions. Removal of the volatile T63 engine particles by the VPS was consistent with that of PMP VPR. In tests on an F117 turbofan engine, the size distribution at the idle (4% rated) engine power condition was found to be bimodal, with the first mode consisting of particles smaller than 10nm, which are believed to be mostly semi-volatile particles, while the second mode of larger size was a mixture of semi-volatile and non-volatile particles. The distribution was single modal at the 33% rated engine power with no secondary mode observed. Altogether, for particles emitted by both engines, the removal efficiency of the VPS appears to surpass that of the PMP VPR by 8-10%.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1333074-volatile-particles-measured-vapor-particle-separator','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1333074-volatile-particles-measured-vapor-particle-separator"><span>Volatile particles measured by <span class="hlt">vapor</span>-particle <span class="hlt">separator</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Cheng, Meng -Dawn; Corporan, Edwin</p> <p>2016-08-25</p> <p><span class="hlt">Vapor</span>-Particle <span class="hlt">Separator</span> (VPS) is a new technology developed for characterization of the volatile fraction of particulate matter in a combustion aerosol population. VPS incorporates a novel metallic membrane and operates in a cross-flow filtration mode for <span class="hlt">separation</span> of <span class="hlt">vapor</span> and solid (i.e. non-volatile) particles. Demonstration of the VPS technology on aircraft engine-emitted particles has led to the improvement of the technology and increased confidence on the robustness of its field performance. In this study, the performance of the VPS was evaluated against the Particle Measurement Programme (PMP) volatile particle remover (VPR), a standardized device used in heavy duty diesel enginesmore » for <span class="hlt">separation</span> and characterization of non-volatile particulate matter. Using tetracontane particles in the laboratory reveals that the VPS performed reasonably well in removing the volatile species. In the field conditions, a single-mode particle size distribution was found for emitted particles from a T63 turboshaft engine at both idle and cruise engine power conditions. Removal of the volatile T63 engine particles by the VPS was consistent with that of PMP VPR. In tests on an F117 turbofan engine, the size distribution at the idle (4% rated) engine power condition was found to be bimodal, with the first mode consisting of particles smaller than 10nm, which are believed to be mostly semi-volatile particles, while the second mode of larger size was a mixture of semi-volatile and non-volatile particles. The distribution was single modal at the 33% rated engine power with no secondary mode observed. Altogether, for particles emitted by both engines, the removal efficiency of the VPS appears to surpass that of the PMP VPR by 8-10%.« less</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_3");'>3</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li class="active"><span>5</span></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_5 --> <div id="page_6" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li class="active"><span>6</span></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="101"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018APExp..11d5502F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018APExp..11d5502F"><span>Elimination of macrostep-<span class="hlt">induced</span> current flow nonuniformity in vertical GaN PN diode using carbon-free drift layer grown by hydride <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fujikura, Hajime; Hayashi, Kentaro; Horikiri, Fumimasa; Narita, Yoshinobu; Konno, Taichiro; Yoshida, Takehiro; Ohta, Hiroshi; Mishima, Tomoyoshi</p> <p>2018-04-01</p> <p>In vertical GaN PN diodes (PNDs) grown entirely by metal–organic chemical <span class="hlt">vapor</span> deposition (MOCVD), large current nonuniformity was observed. This nonuniformity was <span class="hlt">induced</span> by macrosteps on the GaN surface through modulation of carbon incorporation into the n-GaN crystal. It was eliminated in a hybrid PND consisting of a carbon-free n-GaN layer grown by hydride <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy (HVPE) and an MOCVD-regrown p-GaN layer. The hybrid PND showed a fairly low on-resistance (2 mΩ cm2) and high breakdown voltage (2 kV) even without a field plate electrode. These results clearly indicated the strong advantages of the HVPE-grown drift layer for improving power device performance, uniformity, and yield.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29751711','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29751711"><span>[Study on essential oil <span class="hlt">separation</span> from Forsythia suspensa oil-bearing water body based on <span class="hlt">vapor</span> permeation membrane <span class="hlt">separation</span> technology].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Qian; Zhu, Hua-Xu; Tang, Zhi-Shu; Pan, Yong-Lan; Li, Bo; Fu, Ting-Ming; Yao, Wei-Wei; Liu, Hong-Bo; Pan, Lin-Mei</p> <p>2018-04-01</p> <p>To investigate the feasibility of <span class="hlt">vapor</span> permeation membrane technology in <span class="hlt">separating</span> essential oil from oil-water extract by taking the Forsythia suspensa as an example. The polydimethylsiloxane/polyvinylidene fluoride (PDMS/PVDF) composite flat membrane and a polyvinylidene fluoride (PVDF) flat membrane was collected as the membrane material respectively. Two kinds of membrane osmotic liquids were collected by self-made <span class="hlt">vapor</span> permeation device. The yield of essential oil <span class="hlt">separated</span> and enriched from two kinds of membrane materials was calculated, and the microscopic changes of membrane materials were analyzed and compared. Meanwhile, gas chromatography-mass spectrometry (GC-MS) was used to compare and analyze the differences in chemical compositions of essential oil between traditional steam distillation, PVDF membrane enriched method and PDMS/PVDF membrane enriched method. The results showed that the yield of essential oil enriched by PVDF membrane was significantly higher than that of PDMS/PVDF membrane, and the GC-MS spectrum showed that the content of main compositions was higher than that of PDMS/PVDF membrane; The GC-MS spectra showed that the components of essential oil enriched by PVDF membrane were basically the same as those obtained by traditional steam distillation. The above results showed that <span class="hlt">vapor</span> permeation membrane <span class="hlt">separation</span> technology shall be feasible for the <span class="hlt">separation</span> of Forsythia essential oil-bearing water body, and PVDF membrane was more suitable for <span class="hlt">separation</span> and enrichment of Forsythia essential oil than PDMS/PVDF membrane. Copyright© by the Chinese Pharmaceutical Association.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21080701','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21080701"><span>Bacterial chemotaxis along <span class="hlt">vapor-phase</span> gradients of naphthalene.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hanzel, Joanna; Harms, Hauke; Wick, Lukas Y</p> <p>2010-12-15</p> <p>The role of bacterial growth and translocation for the bioremediation of organic contaminants in the vadose zone is poorly understood. Whereas air-filled pores restrict the mobility of bacteria, diffusion of volatile organic compounds in air is more efficient than in water. Past research, however, has focused on chemotactic swimming of bacteria along gradients of water-dissolved chemicals. In this study we tested if and to what extent Pseudomonas putida PpG7 (NAH7) chemotactically reacts to <span class="hlt">vapor-phase</span> gradients forming above their swimming medium by the volatilization from a spot source of solid naphthalene. The development of an aqueous naphthalene gradient by air-water partitioning was largely suppressed by means of activated carbon in the agar. Surprisingly, strain PpG7 was repelled by <span class="hlt">vapor-phase</span> naphthalene although the steady state gaseous concentrations were 50-100 times lower than the aqueous concentrations that result in positive chemotaxis of the same strain. It is thus assumed that the efficient gas-<span class="hlt">phase</span> diffusion resulting in a steady, and possibly toxic, naphthalene flux to the cells controlled the chemotactic reaction rather than the concentration to which the cells were exposed. To our knowledge this is the first demonstration of apparent chemotactic behavior of bacteria in response to <span class="hlt">vapor-phase</span> effector gradients.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70029055','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70029055"><span><span class="hlt">Vapor-phase</span> exchange of perchloroethene between soil and plants</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Struckhoff, G.C.; Burken, J.G.; Schumacher, J.G.</p> <p>2005-01-01</p> <p>Tree core concentrations of tetrachloroethylene (perchloroethene, PCE) at the Riverfront Superfund Site in New Haven, MO, were found to mimic the profile of soil <span class="hlt">phase</span> concentrations. The observed soil-tree core relationship was stronger than that of groundwater PCE to tree core concentrations at the same site. Earlier research has shown a direct, linear relationship between tree core and groundwater concentrations of chlorinated solvents and other organics. Laboratory-scale experiments were performed to elucidate this phenomenon, including determining partitioning coefficients of PCE between plant tissues and air and between plant tissues and water, measured to be 8.1 and 49 L/kg, respectively. The direct relationship of soil to tree core PCE concentrations was hypothesized to be caused by diffusion between tree roots and the soil <span class="hlt">vapor</span> <span class="hlt">phase</span> in the subsurface. The central findings of this research are discovering the importance of subsurface <span class="hlt">vapor-phase</span> transfer for VOCs and uncovering a direct relationship between soil <span class="hlt">vapor-phase</span> chlorinated solvents and uptake rates that impact contaminant translocation from the subsurface and transfer into the atmosphere. ?? 2005 American Chemical Society.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28867448','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28867448"><span>Moisture-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> and recrystallization in amorphous solid dispersions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Luebbert, Christian; Sadowski, Gabriele</p> <p>2017-10-30</p> <p>Active Pharmaceutical Ingredients (APIs) are often dissolved in polymeric matrices to control the gastrointestinal dissolution and to stabilize the amorphous state of the API. During the pharmaceutical development of new formulations, stability studies via storage at certain temperature and relative humidity (RH) have to be carried out to verify the long-term thermodynamic stability of these formulations against unwanted recrystallization and moisture-<span class="hlt">induced</span> amorphous-amorphous <span class="hlt">phase</span> <span class="hlt">separation</span> (MIAPS). This study focuses on predicting the MIAPS of API/polymer formulations at elevated RH. In a first step, the <span class="hlt">phase</span> behavior of water-free formulations of ibuprofen (IBU) and felodipine (FEL) combined with the polymers poly(vinyl pyrrolidone) (PVP), poly(vinyl acetate) (PVAC) and poly (vinyl pyrrolidone-co-vinyl acetate) (PVPVA64) was determined experimentally by differential scanning calorimetry (DSC). The <span class="hlt">phase</span> behavior of these water-free formulations was modeled using the Perturbed-Chain Statistical Associating Fluid Theory (PC-SAFT). Based on this, the API solubility and MIAPS in the above-mentioned formulations at humid conditions was predicted in perfect agreement with the results of two-year lasting stability studies at 25°C/0% RH and 40°C/75% RH. MIAPS was predicted and also experimentally found for the FEL/PVP, FEL/PVPVA64 and IBU/PVP formulations, whereas MIAPS was neither predicted nor measured for the IBU/PVPVA64 system and PVAC-containing formulations. It was thus shown that the results of time-consuming long-term stability tests can be correctly predicted via thermodynamic modeling with PC-SAFT. Copyright © 2017 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011PhRvE..84d6715G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011PhRvE..84d6715G"><span><span class="hlt">Phase</span> <span class="hlt">separation</span> in thermal systems: A lattice Boltzmann study and morphological characterization</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gan, Yanbiao; Xu, Aiguo; Zhang, Guangcai; Li, Yingjun; Li, Hua</p> <p>2011-10-01</p> <p>We investigate thermal and isothermal symmetric liquid-<span class="hlt">vapor</span> <span class="hlt">separations</span> via a fast Fourier transform thermal lattice Boltzmann (FFT-TLB) model. Structure factor, domain size, and Minkowski functionals are employed to characterize the density and velocity fields, as well as to understand the configurations and the kinetic processes. Compared with the isothermal <span class="hlt">phase</span> <span class="hlt">separation</span>, the freedom in temperature prolongs the spinodal decomposition (SD) stage and <span class="hlt">induces</span> different rheological and morphological behaviors in the thermal system. After the transient procedure, both the thermal and isothermal <span class="hlt">separations</span> show power-law scalings in domain growth, while the exponent for thermal system is lower than that for isothermal system. With respect to the density field, the isothermal system presents more likely bicontinuous configurations with narrower interfaces, while the thermal system presents more likely configurations with scattered bubbles. Heat creation, conduction, and lower interfacial stresses are the main reasons for the differences in thermal system. Different from the isothermal case, the release of latent heat causes the changing of local temperature, which results in new local mechanical balance. When the Prandtl number becomes smaller, the system approaches thermodynamical equilibrium much more quickly. The increasing of mean temperature makes the interfacial stress lower in the following way: σ=σ0[(Tc-T)/(Tc-T0)]3/2, where Tc is the critical temperature and σ0 is the interfacial stress at a reference temperature T0, which is the main reason for the prolonged SD stage and the lower growth exponent in the thermal case. Besides thermodynamics, we probe how the local viscosities influence the morphology of the <span class="hlt">phase</span> <span class="hlt">separating</span> system. We find that, for both the isothermal and thermal cases, the growth exponents and local flow velocities are inversely proportional to the corresponding viscosities. Compared with the isothermal case, the local flow velocity</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29738068','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29738068"><span>Pressure-<span class="hlt">Induced</span> Dissolution and Reentrant Formation of Condensed, Liquid-Liquid <span class="hlt">Phase-Separated</span> Elastomeric α-Elastin.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cinar, Hasan; Cinar, Süleyman; Chan, Hue Sun; Winter, Roland</p> <p>2018-05-08</p> <p>We investigated the combined effects of temperature and pressure on liquid-liquid <span class="hlt">phase</span> <span class="hlt">separation</span> (LLPS) phenomena of α-elastin up to the multi-kbar regime. FT-IR spectroscopy, CD, UV/Vis absorption, <span class="hlt">phase</span>-contrast light and fluorescence microscopy techniques were employed to reveal structural changes and mesoscopic <span class="hlt">phase</span> states of the system. A novel pressure-<span class="hlt">induced</span> reentrant LLPS was observed in the intermediate temperature range. A molecular-level picture, in particular on the role of hydrophobic interactions, hydration, and void volume in controlling LLPS phenomena is presented. The potential role of the LLPS phenomena in the development of early cellular compartmentalization is discussed, which might have started in the deep sea, where pressures up to the kbar level are encountered. © 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23822250','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23822250"><span>TES buffer-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> of aqueous solutions of several water-miscible organic solvents at 298.15 K: <span class="hlt">phase</span> diagrams and molecular dynamic simulations.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Taha, Mohamed; Lee, Ming-Jer</p> <p>2013-06-28</p> <p>Water and the organic solvents tetrahydrofuran, 1,3-dioxolane, 1,4-dioxane, 1-propanol, 2-propanol, tert-butanol, acetonitrile, or acetone are completely miscible in all proportions at room temperature. Here, we present new buffering-out <span class="hlt">phase</span> <span class="hlt">separation</span> systems that the above mentioned organic aqueous solutions can be <span class="hlt">induced</span> to form two liquid <span class="hlt">phases</span> in the presence of a biological buffer 2-[[1,3-dihydroxy-2-(hydroxymethyl)propan-2-yl]amino]ethanesulfonic acid (TES). The lower liquid <span class="hlt">phase</span> is rich in water and buffer, and the upper <span class="hlt">phase</span> is organic rich. This observation has both practical and mechanistic interests. The <span class="hlt">phase</span> diagrams of these systems were constructed by experimental measurements at ambient conditions. Molecular dynamic (MD) simulations were performed for TES + water + THF system to understand the interactions between TES, water, and organic solvent at molecular level. Several composition-sets for this system, beyond and inside the liquid-liquid <span class="hlt">phase</span>-splitting region, have been simulated. Interestingly, the MD simulation for compositions inside the <span class="hlt">phase</span> <span class="hlt">separation</span> region showed that THF molecules are forced out from the water network to start forming a new liquid <span class="hlt">phase</span>. The hydrogen-bonds, hydrogen-bonds lifetimes, hydrogen-bond energies, radial distribution functions, coordination numbers, the electrostatic interactions, and the van der Waals interactions between the different pairs have been calculated. Additionally, MD simulations for TES + water + tert-butanol∕acetonitrile∕acetone <span class="hlt">phase</span> <span class="hlt">separation</span> systems were simulated. The results from MD simulations provide an explanation for the buffering-out phenomena observed in [TES + water + organic solvent] systems by a mechanism controlled by the competitive interactions of the buffer and the organic solvent with water. The molecular mechanism reported here is helpful for designing new benign <span class="hlt">separation</span> materials.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22493773-analysis-organic-vapors-laser-induced-breakdown-spectroscopy','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22493773-analysis-organic-vapors-laser-induced-breakdown-spectroscopy"><span>Analysis of organic <span class="hlt">vapors</span> with laser <span class="hlt">induced</span> breakdown spectroscopy</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Nozari, Hadi; Tavassoli, Seyed Hassan; Rezaei, Fatemeh, E-mail: fatemehrezaei@kntu.ac.ir</p> <p>2015-09-15</p> <p>In this paper, laser <span class="hlt">induced</span> breakdown spectroscopy (LIBS) is utilized in the study of acetone, ethanol, methanol, cyclohexane, and nonane <span class="hlt">vapors</span>. Carbon, hydrogen, oxygen, and nitrogen atomic emission spectra have been recorded following laser-<span class="hlt">induced</span> breakdown of the organic <span class="hlt">vapors</span> that are mixed with air inside a quartz chamber at atmospheric pressure. The plasma is generated with focused, Q-switched Nd:YAG radiation at the wavelength of 1064 nm. The effects of ignition and <span class="hlt">vapor</span> pressure are discussed in view of the appearance of the emission spectra. The recorded spectra are proportional to the <span class="hlt">vapor</span> pressure in air. The hydrogen and oxygen contributions diminishmore » gradually with consecutive laser-plasma events without gas flow. The results show that LIBS can be used to characterize organic <span class="hlt">vapor</span>.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018NatCh..10..506W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018NatCh..10..506W"><span>Control over <span class="hlt">phase</span> <span class="hlt">separation</span> and nucleation using a laser-tweezing potential</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Walton, Finlay; Wynne, Klaas</p> <p>2018-05-01</p> <p>Control over the nucleation of new <span class="hlt">phases</span> is highly desirable but elusive. Even though there is a long history of crystallization engineering by varying physicochemical parameters, controlling which polymorph crystallizes or whether a molecule crystallizes or forms an amorphous precipitate is still a poorly understood practice. Although there are now numerous examples of control using laser-<span class="hlt">induced</span> nucleation, the absence of physical understanding is preventing progress. Here we show that the proximity of a liquid-liquid critical point or the corresponding binodal line can be used by a laser-tweezing potential to <span class="hlt">induce</span> concentration gradients. A simple theoretical model shows that the stored electromagnetic energy of the laser beam produces a free-energy potential that forces <span class="hlt">phase</span> <span class="hlt">separation</span> or triggers the nucleation of a new <span class="hlt">phase</span>. Experiments in a liquid mixture using a low-power laser diode confirm the effect. <span class="hlt">Phase</span> <span class="hlt">separation</span> and nucleation using a laser-tweezing potential explains the physics behind non-photochemical laser-<span class="hlt">induced</span> nucleation and suggests new ways of manipulating matter.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010SPIE.7835E..07E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010SPIE.7835E..07E"><span>Stand-off detection of <span class="hlt">vapor</span> <span class="hlt">phase</span> explosives by resonance enhanced Raman spectroscopy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ehlerding, Anneli; Johansson, Ida; Wallin, Sara; Östmark, Henric</p> <p>2010-10-01</p> <p>Stand-off measurements on nitromethane (NM), 2,4-DNT and 2,4,6-TNT in <span class="hlt">vapor</span> <span class="hlt">phase</span> using resonance Raman spectroscopy have been performed. The Raman cross sections for NM, DNT and TNT in <span class="hlt">vapor</span> <span class="hlt">phase</span> have been measured in the wavelength range 210-300 nm under laboratory conditions, in order to estimate how large resonance enhancement factors can be achieved for these explosives. The measurements show that the signal is greatly enhanced, up to 250.000 times for 2,4-DNT and 60.000 times for 2,4,6-TNT compared to the non-resonant signal at 532 nm. For NM the resonance enhancement enabled realistic outdoor measurements in <span class="hlt">vapor</span> <span class="hlt">phase</span> at 13 m distance. This all indicate a potential for resonance Raman spectroscopy as a stand-off technique for detection of <span class="hlt">vapor</span> <span class="hlt">phase</span> explosives.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25689018','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25689018"><span>Sodium triflate decreases interaggregate repulsion and <span class="hlt">induces</span> <span class="hlt">phase</span> <span class="hlt">separation</span> in cationic micelles.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lima, Filipe S; Cuccovia, Iolanda M; Buchner, Richard; Antunes, Filipe E; Lindman, Björn; Miguel, Maria G; Horinek, Dominik; Chaimovich, Hernan</p> <p>2015-03-10</p> <p>Dodecyltrimethylammonium triflate (DTATf) micelles possess lower degree of counterion dissociation (α), lower hydration, and higher packing of monomers than other micelles of similar structure. Addition of sodium triflate ([NaTf] > 0.05 M) to DTATf solutions promotes <span class="hlt">phase</span> <span class="hlt">separation</span>. This phenomenon is commonly observed in oppositely charged surfactant mixtures, but it is rare for ionic surfactants and relatively simple counterions. While the properties of DTATf have already been reported, the driving forces for the observed <span class="hlt">phase</span> <span class="hlt">separation</span> with added salt remain unclear. Thus, we propose an interpretation for the observed <span class="hlt">phase</span> <span class="hlt">separation</span> in cationic surfactant solutions. Addition of up to 0.03 M NaTf to micellar DTATf solutions led to a limited increase of the aggregation number, to interface dehydration, and to a progressive decrease in α. The viscosity of DTATf solutions of higher concentration ([DTATf] ≥ 0.06 M) reached a maximum with increasing [NaTf], though the aggregation number slightly increased, and no shape change occurred. We hypothesize that this maximum results from a decrease in interaggregate repulsion, as a consequence of increased ion binding. This reduction in micellar repulsion without simultaneous infinite micellar growth is, probably, the major driving force for <span class="hlt">phase</span> <span class="hlt">separation</span> at higher [NaTf].</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011PhFl...23c2102K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011PhFl...23c2102K"><span>Nonlinear dynamics of confined thin liquid-<span class="hlt">vapor</span> bilayer systems with <span class="hlt">phase</span> change</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kanatani, Kentaro; Oron, Alexander</p> <p>2011-03-01</p> <p>We numerically investigate the nonlinear evolution of the interface of a thin liquid-<span class="hlt">vapor</span> bilayer system confined by rigid horizontal walls from both below and above. The lateral variation of the <span class="hlt">vapor</span> pressure arising from <span class="hlt">phase</span> change is taken into account in the present analysis. When the liquid (<span class="hlt">vapor</span>) is heated (cooled) and gravity acts toward the liquid, the deflection of the interface monotonically grows, leading to a rupture of the <span class="hlt">vapor</span> layer, whereas nonruptured stationary states are found when the liquid (<span class="hlt">vapor</span>) is cooled (heated) and gravity acts toward the <span class="hlt">vapor</span>. In the latter case, <span class="hlt">vapor</span>-flow-driven convective cells are found in the liquid <span class="hlt">phase</span> in the stationary state. The average <span class="hlt">vapor</span> pressure and interface temperature deviate from their equilibrium values once the interface departs from the flat equilibrium state. Thermocapillarity does not have a significant effect near the thermodynamic equilibrium, but becomes important if the system significantly deviates from it.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24316887','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24316887"><span>APTS and rGO co-functionalized pyrenated fluorescent nanonets for representative <span class="hlt">vapor</span> <span class="hlt">phase</span> nitroaromatic explosive detection.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Guo, Linjuan; Zu, Baiyi; Yang, Zheng; Cao, Hongyu; Zheng, Xuefang; Dou, Xincun</p> <p>2014-01-01</p> <p>For the first time, flexible PVP/pyrene/APTS/rGO fluorescent nanonets were designed and synthesized via a one-step electrospinning method to detect representative subsaturated nitroaromatic explosive <span class="hlt">vapor</span>. The functional fluorescent nanonets, which were highly stable in air, showed an 81% quenching efficiency towards TNT <span class="hlt">vapor</span> (∼10 ppb) with an exposure time of 540 s at room temperature. The nice performance of the nanonets was ascribed to the synergistic effects <span class="hlt">induced</span> by the specific adsorption properties of APTS, the fast charge transfer properties and the effective π-π interaction with pyrene and TNT of rGO. Compared to the analogues of TNT, the PVP/pyrene/APTS/rGO nanonets showed notable selectivity towards TNT and DNT <span class="hlt">vapors</span>. The explored functionalization method opens up brand new insight into sensitive and selective detection of <span class="hlt">vapor</span> <span class="hlt">phase</span> nitroaromatic explosives.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1143649','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/1143649"><span><span class="hlt">Vapor</span> <span class="hlt">phase</span> elemental sulfur amendment for sequestering mercury in contaminated soil</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Looney, Brian B.; Denham, Miles E.; Jackson, Dennis G.</p> <p>2014-07-08</p> <p>The process of treating elemental mercury within the soil is provided by introducing into the soil a heated <span class="hlt">vapor</span> <span class="hlt">phase</span> of elemental sulfur. As the <span class="hlt">vapor</span> <span class="hlt">phase</span> of elemental sulfur cools, sulfur is precipitated within the soil and then reacts with any elemental mercury thereby producing a reaction product that is less hazardous than elemental mercury.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20010121492&hterms=soaps&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dsoaps','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20010121492&hterms=soaps&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dsoaps"><span><span class="hlt">Vapor</span> <span class="hlt">Phase</span> Catalytic Ammonia Reduction</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Flynn, Michael T.; Harper, Lynn D. (Technical Monitor)</p> <p>1994-01-01</p> <p>This paper discusses the development of a <span class="hlt">Vapor</span> <span class="hlt">Phase</span> Catalytic Ammonia Reduction (VPCAR) teststand and the results of an experimental program designed to evaluate the potential of the technology as a water purification process. In the experimental program the technology is evaluated based upon product water purity, water recovery rate, and power consumption. The experimental work demonstrates that the technology produces high purity product water and attains high water recovery rates at a relatively high specific power consumption. The experimental program was conducted in 3 <span class="hlt">phases</span>. In <span class="hlt">phase</span> I an Igepon(TM) soap and water mixture was used to evaluate the performance of an innovative Wiped-Film Rotating-Disk evaporator and associated demister. In <span class="hlt">phase</span> II a phenol-water solution was used to evaluate the performance of the high temperature catalytic oxidation reactor. In <span class="hlt">phase</span> III a urine analog was used to evaluate the performance of the combined distillation/oxidation functions of the processor.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24679215','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24679215"><span>Quantitative evaluation of colloidal stability of antibody solutions using PEG-<span class="hlt">induced</span> liquid-liquid <span class="hlt">phase</span> <span class="hlt">separation</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Ying; Latypov, Ramil F; Lomakin, Aleksey; Meyer, Julie A; Kerwin, Bruce A; Vunnum, Suresh; Benedek, George B</p> <p>2014-05-05</p> <p>Colloidal stability of antibody solutions, i.e., the propensity of the folded protein to precipitate, is an important consideration in formulation development of therapeutic monoclonal antibodies. In a protein solution, different pathways including crystallization, colloidal aggregation, and liquid-liquid <span class="hlt">phase</span> <span class="hlt">separation</span> (LLPS) can lead to the formation of precipitates. The kinetics of crystallization and aggregation are often slow and vary from protein to protein. Due to the diverse mechanisms of these protein condensation processes, it is a challenge to develop a standardized test for an early evaluation of the colloidal stability of antibody solutions. LLPS would normally occur in antibody solutions at sufficiently low temperature, provided that it is not preempted by freezing of the solution. Poly(ethylene glycol) (PEG) can be used to <span class="hlt">induce</span> LLPS at temperatures above the freezing point. Here, we propose a colloidal stability test based on <span class="hlt">inducing</span> LLPS in antibody solutions and measuring the antibody concentration of the dilute <span class="hlt">phase</span>. We demonstrate experimentally that such a PEG-<span class="hlt">induced</span> LLPS test can be used to compare colloidal stability of different antibodies in different solution conditions and can be readily applied to high-throughput screening. We have derived an equation for the effects of PEG concentration and molecular weight on the results of the LLPS test. Finally, this equation defines a binding energy in the condensed <span class="hlt">phase</span>, which can be determined in the PEG-<span class="hlt">induced</span> LLPS test. This binding energy is a measure of attractive interactions between antibody molecules and can be used for quantitative characterization of the colloidal stability of antibody solutions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003JAP....93.5167S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003JAP....93.5167S"><span>Shock wave <span class="hlt">induced</span> <span class="hlt">vaporization</span> of porous solids</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shen, Andy H.; Ahrens, Thomas J.; O'Keefe, John D.</p> <p>2003-05-01</p> <p>Strong shock waves generated by hypervelocity impact can <span class="hlt">induce</span> <span class="hlt">vaporization</span> in solid materials. To pursue knowledge of the chemical species in the shock-<span class="hlt">induced</span> <span class="hlt">vapors</span>, one needs to design experiments that will drive the system to such thermodynamic states that sufficient <span class="hlt">vapor</span> can be generated for investigation. It is common to use porous media to reach high entropy, <span class="hlt">vaporized</span> states in impact experiments. We extended calculations by Ahrens [J. Appl. Phys. 43, 2443 (1972)] and Ahrens and O'Keefe [The Moon 4, 214 (1972)] to higher distentions (up to five) and improved their method with a different impedance match calculation scheme and augmented their model with recent thermodynamic and Hugoniot data of metals, minerals, and polymers. Although we reconfirmed the competing effects reported in the previous studies: (1) increase of entropy production and (2) decrease of impedance match, when impacting materials with increasing distentions, our calculations did not exhibit optimal entropy-generating distention. For different materials, very different impact velocities are needed to initiate <span class="hlt">vaporization</span>. For aluminum at distention (m)<2.2, a minimum impact velocity of 2.7 km/s is required using tungsten projectile. For ionic solids such as NaCl at distention <2.2, 2.5 km/s is needed. For carbonate and sulfate minerals, the minimum impact velocities are much lower, ranging from less than 1 to 1.5 km/s.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17237604','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17237604"><span>Sporicidal Activity of the KMT reagent in its <span class="hlt">vapor</span> <span class="hlt">phase</span> against Geobacillus stearothermophilus Spores.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kida, Nori; Mochizuki, Yasushi; Taguchi, Fumiaki</p> <p>2007-01-01</p> <p>In an investigation of the sporicidal activity of the KMT reagent, a <span class="hlt">vapor</span> <span class="hlt">phase</span> study was performed using five kinds of carriers contaminated with Geobacillus stearothermophilus spores. When 25 ml of the KMT reagent was <span class="hlt">vaporized</span> in a chamber (capacity; approximately 95 liters), the 2-step heating method (<span class="hlt">vaporization</span> by a combination of low temperature and high temperature) showed the most effective sporicidal activity in comparison with the 1-step heating method (rapid <span class="hlt">vaporization</span>). The 2-step heating method appeared to be related to the sporicidal activity of <span class="hlt">vaporized</span> KMT reagent, i.e., ethanol and iodine, which <span class="hlt">vaporized</span> mainly when heated at a low temperature such as 55 C, and acidic water, which <span class="hlt">vaporized</span> mainly when heated at a high temperature such as 300 C. We proposed that the KMT reagent can be used as a new disinfectant not only in the liquid <span class="hlt">phase</span> but also in the <span class="hlt">vapor</span> <span class="hlt">phase</span> in the same way as peracetic acid and hydrogen peroxide.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MS%26E..370a2007T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MS%26E..370a2007T"><span>A review of <span class="hlt">phase</span> <span class="hlt">separation</span> issues in aviation gasoline fuel and motor gasoline fuels in aviation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Thanikasalam, K.; Rahmat, M.; Fahmi, A. G. Mohammad; Zulkifli, A. M.; Shawal, N. Noor; Ilanchelvi, K.; Ananth, M.; Elayarasan, R.</p> <p>2018-05-01</p> <p>In an attempt to bring in sustainable energy resources into the current combustibles mix, recent European legislations make obligatory the addition of biogenic fuels into traditional fossil gasoline. The preferred biogenic fuel, for economic reasons, is predominantly ethanol. Even though likened to fossil gasoline constituents, ethanol has a dissimilar chemical formulation that may lead to a potentially hazardous physicochemical phenomenon, particularly in the presence of water. Owing to increased financially driven propensity to utilize motor vehicle gasoline as aviation gasoline fuel, this may result in potentially hazardous situations, specifically in running smaller or compact General Aviation aircraft. The potential risks posed by ethanol admixtures in aircraft are <span class="hlt">phase</span> <span class="hlt">separation</span> and carburettor icing. Gasoline mixed with ethanol is also prone to an increased vulnerability to <span class="hlt">vapor</span> lock that happens when fuel turns into <span class="hlt">vapor</span> in the fuel pumps due to high temperatures and lessened ambient pressure at high altitudes. This article provides a literature review on <span class="hlt">phase</span> <span class="hlt">separation</span> issues in aviation gasoline fuel and motor gasoline fuels in aviation.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_4");'>4</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li class="active"><span>6</span></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_6 --> <div id="page_7" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li class="active"><span>7</span></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="121"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19371050','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19371050"><span>Investigation of local evaporation flux and <span class="hlt">vapor-phase</span> pressure at an evaporative droplet interface.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Duan, Fei; Ward, C A</p> <p>2009-07-07</p> <p>In the steady-state experiments of water droplet evaporation, when the throat was heating at a stainless steel conical funnel, the interfacial liquid temperature was found to increase parabolically from the center line to the rim of the funnel with the global <span class="hlt">vapor-phase</span> pressure at around 600 Pa. The energy conservation analysis at the interface indicates that the energy required for evaporation is maintained by thermal conduction to the interface from the liquid and <span class="hlt">vapor</span> <span class="hlt">phases</span>, thermocapillary convection at interface, and the viscous dissipation globally and locally. The local evaporation flux increases from the center line to the periphery as a result of multiple effects of energy transport at the interface. The local <span class="hlt">vapor-phase</span> pressure predicted from statistical rate theory (SRT) is also found to increase monotonically toward the interface edge from the center line. However, the average value of the local <span class="hlt">vapor-phase</span> pressures is in agreement with the measured global <span class="hlt">vapor-phase</span> pressure within the measured error bar.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16904162','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16904162"><span>Speciation and quantification of <span class="hlt">vapor</span> <span class="hlt">phases</span> in soy biodiesel and waste cooking oil biodiesel.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Peng, Chiung-Yu; Lan, Cheng-Hang; Dai, Yu-Tung</p> <p>2006-12-01</p> <p>This study characterizes the compositions of two biodiesel <span class="hlt">vapors</span>, soy biodiesel and waste cooking oil biodiesel, to provide a comprehensive understanding of biodiesels. <span class="hlt">Vapor</span> <span class="hlt">phases</span> were sampled by purging oil <span class="hlt">vapors</span> through thermal desorption tubes which were then analyzed by the thermal desorption/GC/MS system. The results show that the compounds of biodiesel <span class="hlt">vapors</span> can be divided into four groups. They include methyl esters (the main biodiesel components), oxygenated chemicals, alkanes and alkenes, and aromatics. The first two chemical groups are only found in biodiesel <span class="hlt">vapors</span>, not in the diesel <span class="hlt">vapor</span> emissions. The percentages of mean concentrations for methyl esters, oxygenated chemicals, alkanes and alkenes, and aromatics are 66.1%, 22.8%, 4.8% and 6.4%, respectively for soy biodiesel, and 35.8%, 35.9%, 27.9% and 0.3%, respectively for waste cooking oil biodiesel at a temperature of 25+/-2 degrees C. These results show that biodiesels have fewer chemicals and lower concentrations in <span class="hlt">vapor</span> <span class="hlt">phase</span> than petroleum diesel, and the total emission rates are between one-sixteenth and one-sixth of that of diesel emission, corresponding to fuel evaporative emissions of loading losses of between 106 microg l(-1) and 283 microg l(-1). Although diesels generate more <span class="hlt">vapor</span> <span class="hlt">phase</span> emissions, biodiesels still generate considerable amount of <span class="hlt">vapor</span> emissions, particularly the emissions from methyl esters and oxygenated chemicals. These two chemical groups are more reactive than alkanes and aromatics. Therefore, speciation and quantification of biodiesel <span class="hlt">vapor</span> <span class="hlt">phases</span> are important.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010PhDT.........9Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010PhDT.........9Y"><span>Polymer composites and porous materials prepared by thermally <span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> and polymer-metal hybrid methods</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yoon, Joonsung</p> <p></p> <p>The primary objective of this research is to investigate the morphological and mechanical properties of composite materials and porous materials prepared by thermally <span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span>. High melting crystallizable diluents were mixed with polymers so that the <span class="hlt">phase</span> <span class="hlt">separation</span> would be <span class="hlt">induced</span> by the solidification of the diluents upon cooling. Theoretical <span class="hlt">phase</span> diagrams were calculated using Flory-Huggins solution thermodynamics which show good agreement with the experimental results. Porous materials were prepared by the extraction of the crystallized diluents after cooling the mixtures (hexamethylbenzene/polyethylene and pyrene/polyethylene). Anisotropic structures show strong dependence on the identity of the diluents and the composition of the mixtures. Anisotropic crystal growth of the diluents was studied in terms of thermodynamics and kinetics using DSC, optical microscopy and SEM. Microstructures of the porous materials were explained in terms of supercooling and dendritic solidification. Dual functionality of the crystallizable diluents for composite materials was evaluated using isotactic polypropylene (iPP) and compatible diluents that crystallize upon cooling. The selected diluents form homogeneous mixtures with iPP at high temperature and lower the viscosity (improved processability), which undergo <span class="hlt">phase</span> <span class="hlt">separation</span> upon cooling to form solid particles that function as a toughening agent at room temperature. Tensile properties and morphology of the composites showed that organic crystalline particles have the similar effect as rigid particles to increase toughness; de-wetting between the particle and iPP matrix occurs at the early stage of deformation, followed by unhindered plastic flow that consumes significant amount of fracture energy. The effect of the diluents, however, strongly depends on the identity of the diluents that interact with the iPP during solidification step, which was demonstrated by comparing tetrabromobisphenol-A and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19947706','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19947706"><span><span class="hlt">Phase</span> <span class="hlt">separations</span> in mixtures of a liquid crystal and a nanocolloidal particle.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Matsuyama, Akihiko</p> <p>2009-11-28</p> <p>We present a mean field theory to describe <span class="hlt">phase</span> <span class="hlt">separations</span> in mixtures of a liquid crystal and a nanocolloidal particle. By taking into account a nematic, a smectic A ordering of the liquid crystal, and a crystalline ordering of the nanoparticle, we calculate the <span class="hlt">phase</span> diagrams on the temperature-concentration plane. We predict various <span class="hlt">phase</span> <span class="hlt">separations</span>, such as a smectic A-crystal <span class="hlt">phase</span> <span class="hlt">separation</span> and a smectic A-isotropic-crystal triple point, etc., depending on the interactions between the liquid crystal and the colloidal surface. Inside binodal curves, we find new unstable and metastable regions, which are important in the <span class="hlt">phase</span> ordering dynamics. We also find a crystalline ordering of the nanoparticles dispersed in a smectic A <span class="hlt">phase</span> and a nematic <span class="hlt">phase</span>. The cooperative phenomena between liquid-crystalline ordering and crystalline ordering <span class="hlt">induce</span> a variety of <span class="hlt">phase</span> diagrams.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20090041646','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20090041646"><span>Development of <span class="hlt">Vapor-Phase</span> Catalytic Ammonia Removal System</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Flynn, Michael; Fisher, John; Kiss, Mark; Borchers, Bruce; Tleimat, Badawi; Tleimat, Maher; Quinn, Gregory; Fort, James; Nalette, Tim; Baker, Gale; <a style="text-decoration: none; " href="javascript:void(0); " onClick="displayelement('author_20090041646'); toggleEditAbsImage('author_20090041646_show'); toggleEditAbsImage('author_20090041646_hide'); "> <img style="display:inline; width:12px; height:12px; " src="images/arrow-up.gif" width="12" height="12" border="0" alt="hide" id="author_20090041646_show"> <img style="width:12px; height:12px; display:none; " src="images/arrow-down.gif" width="12" height="12" border="0" alt="hide" id="author_20090041646_hide"></p> <p>2007-01-01</p> <p>A report describes recent accomplishments of a continuing effort to develop the <span class="hlt">vapor-phase</span> catalytic ammonia removal (VPCAR) process for recycling wastewater for consumption by humans aboard a spacecraft in transit to Mars.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JPS...350..127H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JPS...350..127H"><span>A fracture mechanics study of the <span class="hlt">phase</span> <span class="hlt">separating</span> planar electrodes: <span class="hlt">Phase</span> field modeling and analytical results</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Haftbaradaran, H.; Maddahian, A.; Mossaiby, F.</p> <p>2017-05-01</p> <p>It is well known that <span class="hlt">phase</span> <span class="hlt">separation</span> could severely intensify mechanical degradation and expedite capacity fading in lithium-ion battery electrodes during electrochemical cycling. Experiments have frequently revealed that such degradation effects could be substantially mitigated via reducing the electrode feature size to the nanoscale. The purpose of this work is to present a fracture mechanics study of the <span class="hlt">phase</span> <span class="hlt">separating</span> planar electrodes. To this end, a <span class="hlt">phase</span> field model is utilized to predict how <span class="hlt">phase</span> <span class="hlt">separation</span> affects evolution of the solute distribution and stress profile in a planar electrode. Behavior of the preexisting flaws in the electrode in response to the diffusion <span class="hlt">induced</span> stresses is then examined via computing the time dependent stress intensity factor arising at the tip of flaws during both the insertion and extraction half-cycles. Further, adopting a sharp-interphase approximation of the system, a critical electrode thickness is derived below which the <span class="hlt">phase</span> <span class="hlt">separating</span> electrode becomes flaw tolerant. Numerical results of the <span class="hlt">phase</span> field model are also compared against analytical predictions of the sharp-interphase model. The results are further discussed with reference to the available experiments in the literature. Finally, some of the limitations of the model are cautioned.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22468544','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22468544"><span>[Experimental research of oil <span class="hlt">vapor</span> pollution control for gas station with membrane <span class="hlt">separation</span> technology].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhu, Ling; Chen, Jia-Qing; Zhang, Bao-Sheng; Wang, Jian-Hong</p> <p>2011-12-01</p> <p>Two kinds of membranes modules, <span class="hlt">vapor</span> retained glassy membrane based on PEEK hollow fiber membrane modules and <span class="hlt">vapor</span> permeated rubbery membrane system based on GMT plate-and-frame membrane modules, were used to control the oil <span class="hlt">vapor</span> pollution during the course of receiving and transferring gasoline in oil station. The efficiencies of the membrane module and the membrane system of them were evaluated and compared respectively in the facilities which were developed by ourselves. It was found that both the two kinds of membranes modules had high efficiency for the <span class="hlt">separation</span> of VOCs-air mixed gases, and the outlet <span class="hlt">vapor</span> after treatment all can meet the national standard. When the <span class="hlt">vapor</span>-enriched gas was returned to the oil tank to simulate the continuously cycle test, the concentration of VOCs in the outlet was also below 25 g x m(-3).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19860019676','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19860019676"><span>The non-Newtonian heat and mass transport of He 2 in porous media used for <span class="hlt">vapor</span>-liquid <span class="hlt">phase</span> <span class="hlt">separation</span>. Ph.D. Thesis</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Yuan, S. W. K.</p> <p>1985-01-01</p> <p>This investigation of <span class="hlt">vapor</span>-liquid <span class="hlt">phase</span> <span class="hlt">separation</span> (VLPS) of He 2 is related to long-term storage of cryogenic liquid. The VLPS system utilizes porous plugs in order to generate thermomechanical (thermo-osmotic) force which in turn prevents liquid from flowing out of the cryo-vessel (e.g., Infrared Astronomical Satellite). An apparatus was built and VLPS data were collected for a 2 and a 10 micrometer sintered stainless steel plug and a 5 to 15 micrometer sintered bronze plug. The VLPS data obtained at high temperature were in the nonlinear turbulent regime. At low temperature, the Stokes regime was approached. A turbulent flow model was developed, which provides a phenomenological description of the VLPS data. According to the model, most of the <span class="hlt">phase</span> <span class="hlt">separation</span> data are in the turbulent regime. The model is based on concepts of the Gorter-Mellink transport involving the mutual friction known from the zero net mass flow (ZNMF) studies. The latter had to be modified to obtain agreement with the present experimental VLPS evidence. In contrast to the well-known ZNMF mode, the VLPS results require a geometry dependent constant (Gorter-Mellink constant). A theoretical interpretation of the phenomenological equation for the VLPS data obtained, is based on modelling of the dynamics of quantized vortices proposed by Vinen. In extending Vinen's model to the VLPS transport of He 2 in porous media, a correlation between the K*(GM) and K(p) was obtained which permits an interpretation of the present findings. As K(p) is crucial, various methods were introduced to measure the permeability of the porous media at low temperatures. Good agreement was found between the room temperature and the low temperature K(p)-value of the plugs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018IJT....39...84G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018IJT....39...84G"><span>An Indirect Method for <span class="hlt">Vapor</span> Pressure and <span class="hlt">Phase</span> Change Enthalpy Determination by Thermogravimetry</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Giani, Samuele; Riesen, Rudolf; Schawe, Jürgen E. K.</p> <p>2018-07-01</p> <p><span class="hlt">Vapor</span> pressure is a fundamental property of a pure substance. This property is the pressure of a compound's <span class="hlt">vapor</span> in thermodynamic equilibrium with its condensed <span class="hlt">phase</span> (solid or liquid). When <span class="hlt">phase</span> equilibrium condition is met, <span class="hlt">phase</span> coexistence of a pure substance involves a continuum interplay of <span class="hlt">vaporization</span> or sublimation to gas and condensation back to their liquid or solid form, respectively. Thermogravimetric analysis (TGA) techniques are based on mass loss determination and are well suited for the study of such phenomena. In this work, it is shown that TGA method using a reference substance is a suitable technique for <span class="hlt">vapor</span> pressure determination. This method is easy and fast because it involves a series of isothermal segments. In contrast to original Knudsen's approach, where the use of high vacuum is mandatory, adopting the proposed method a given experimental setup is calibrated under ambient pressure conditions. The theoretical framework of this method is based on a generalization of Langmuir equation of free evaporation: The real strength of the proposed method is the ability to determine the <span class="hlt">vapor</span> pressure independently of the molecular mass of the <span class="hlt">vapor</span>. A demonstration of this method has been performed using the Clausius-Clapeyron equation of state to derive the working equation. This algorithm, however, is adaptive and admits the use of other equations of state. The results of a series of experiments with organic molecules indicate that the average difference of the measured and the literature <span class="hlt">vapor</span> pressure amounts to about 5 %. <span class="hlt">Vapor</span> pressure determined in this study spans from few mPa up to several kPa. Once the p versus T diagram is obtained, <span class="hlt">phase</span> transition enthalpy can additionally be calculated from the data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.usgs.gov/of/2001/ofr-01-445/pdf/ofr01445v1.pdf','USGSPUBS'); return false;" href="https://pubs.usgs.gov/of/2001/ofr-01-445/pdf/ofr01445v1.pdf"><span>Determination of methyl mercury by aqueous <span class="hlt">phase</span> Eehylation, followed by gas chromatographic <span class="hlt">separation</span> with cold <span class="hlt">vapor</span> atomic fluorescence detection</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>De Wild, John F.; Olsen, Mark L.; Olund, Shane D.</p> <p>2002-01-01</p> <p>A recent national sampling of streams in the United States revealed low methyl mercury concentrations in surface waters. The resulting median and mean concentrations, calculated from 104 samples, were 0.06 nanograms per liter (ng/L) and 0.15 ng/L, respectively. This level of methyl mercury in surface water in the United States has created a need for analytical techniques capable of detecting sub-nanogram per liter concentrations. In an attempt to create a U.S. Geological Survey approved method, the Wisconsin District Mercury Laboratory has adapted a distillation/ethylation/ gas-<span class="hlt">phase</span> <span class="hlt">separation</span> method with cold <span class="hlt">vapor</span> atomic fluorescence spectroscopy detection for the determination of methyl mercury in filtered and unfiltered waters. This method is described in this report. Based on multiple analyses of surface water and ground-water samples, a method detection limit of 0.04 ng/L was established. Precision and accuracy were evaluated for the method using both spiked and unspiked ground-water and surface-water samples. The percent relative standard deviations ranged from 10.2 to 15.6 for all analyses at all concentrations. Average recoveries obtained for the spiked matrices ranged from 88.8 to 117 percent. The precision and accuracy ranges are within the acceptable method-performance limits. Considering the demonstrated detection limit, precision, and accuracy, the method is an effective means to quantify methyl mercury in waters at or below environmentally relevant concentrations</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27875928','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27875928"><span>Combining mechanical foaming and thermally <span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> to generate chitosan scaffolds for soft tissue engineering.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Biswas, D P; Tran, P A; Tallon, C; O'Connor, A J</p> <p>2017-02-01</p> <p>In this paper, a novel foaming methodology consisting of turbulent mixing and thermally <span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> (TIPS) was used to generate scaffolds for tissue engineering. Air bubbles were mechanically introduced into a chitosan solution which forms the continuous polymer/liquid <span class="hlt">phase</span> in the foam created. The air bubbles entrained in the foam act as a template for the macroporous architecture of the final scaffolds. Wet foams were crosslinked via glutaraldehyde and frozen at -20 °C to <span class="hlt">induce</span> TIPS in order to limit film drainage, bubble coalescence and Ostwald ripening. The effects of production parameters, including mixing speed, surfactant concentration and chitosan concentration, on foaming are explored. Using this method, hydrogel scaffolds were successfully produced with up to 80% porosity, average pore sizes of 120 μm and readily tuneable compressive modulus in the range of 2.6 to 25 kPa relevant to soft tissue engineering applications. These scaffolds supported 3T3 fibroblast cell proliferation and penetration and therefore show significant potential for application in soft tissue engineering.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=129123&keyword=aviation&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50','EPA-EIMS'); return false;" href="https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=129123&keyword=aviation&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50"><span>FIELD TRAPPING OF SUBSURFACE <span class="hlt">VAPOR</span> <span class="hlt">PHASE</span> PETROLEUM HYDROCARBONS</span></a></p> <p><a target="_blank" href="http://oaspub.epa.gov/eims/query.page">EPA Science Inventory</a></p> <p></p> <p></p> <p>Soil gas samples from intact soil cores were collected on adsorbents at a field site, then thermally desorbed and analyzed by laboratory gas chromatography (GC). ertical concentration profiles of predominant <span class="hlt">vapor</span> <span class="hlt">phase</span> petroleum hydrocarbons under ambient conditions were obtaine...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19880065961&hterms=passive+transport&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dpassive%2Btransport','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19880065961&hterms=passive+transport&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dpassive%2Btransport"><span>Turbulent transport of He II in active and passive <span class="hlt">phase</span> <span class="hlt">separators</span> using slit devices and porous media</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Yuan, S. W. K.; Lee, J. M.; Frederking, T. H. K.</p> <p>1988-01-01</p> <p>The turbulent transport mode of <span class="hlt">vapor</span> liquid <span class="hlt">phase</span> <span class="hlt">separators</span> (VLPS) for He II has been investigated comparing passive porous plug <span class="hlt">separators</span> with active <span class="hlt">phase</span> <span class="hlt">separators</span> (APS) using slits of variable flow paths within a common frame of reference. It is concluded that the basic transport regimes in both devices are identical. An integrated Gorter-Mellink (1949) equation, found previously to predict VLPS results of porous plugs, is employed to analyze APS data published in the literature. It is found that the Gorter-Mellink flow rate parameter for 9-micron and 14-micron APS slit widths are relatively independent of the slit width, having a rate constant of about 9 + or - 10 percent. This agrees with the early heat flow results for He II entropy transport at zero net mass flow in wide capillaries and slits.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4664304','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4664304"><span>Conserved interdomain linker promotes <span class="hlt">phase</span> <span class="hlt">separation</span> of the multivalent adaptor protein Nck</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Banjade, Sudeep; Wu, Qiong; Mittal, Anuradha; Peeples, William B.; Pappu, Rohit V.; Rosen, Michael K.</p> <p>2015-01-01</p> <p>The organization of membranes, the cytosol, and the nucleus of eukaryotic cells can be controlled through <span class="hlt">phase</span> <span class="hlt">separation</span> of lipids, proteins, and nucleic acids. Collective interactions of multivalent molecules mediated by modular binding domains can <span class="hlt">induce</span> gelation and <span class="hlt">phase</span> <span class="hlt">separation</span> in several cytosolic and membrane-associated systems. The adaptor protein Nck has three SRC-homology 3 (SH3) domains that bind multiple proline-rich segments in the actin regulatory protein neuronal Wiskott-Aldrich syndrome protein (N-WASP) and an SH2 domain that binds to multiple phosphotyrosine sites in the adhesion protein nephrin, leading to <span class="hlt">phase</span> <span class="hlt">separation</span>. Here, we show that the 50-residue linker between the first two SH3 domains of Nck enhances <span class="hlt">phase</span> <span class="hlt">separation</span> of Nck/N-WASP/nephrin assemblies. Two linear motifs within this element, as well as its overall positively charged character, are important for this effect. The linker increases the driving force for self-assembly of Nck, likely through weak interactions with the second SH3 domain, and this effect appears to promote <span class="hlt">phase</span> <span class="hlt">separation</span>. The linker sequence is highly conserved, suggesting that the sequence determinants of the driving forces for <span class="hlt">phase</span> <span class="hlt">separation</span> may be generally important to Nck functions. Our studies demonstrate that linker regions between modular domains can contribute to the driving forces for self-assembly and <span class="hlt">phase</span> <span class="hlt">separation</span> of multivalent proteins. PMID:26553976</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19730020190','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19730020190"><span>Prediction of fluctuating pressure environments associated with plume-<span class="hlt">induced</span> <span class="hlt">separated</span> flow fields</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Plotkin, K. J.</p> <p>1973-01-01</p> <p>The <span class="hlt">separated</span> flow environment <span class="hlt">induced</span> by underexpanded rocket plumes during boost <span class="hlt">phase</span> of rocket vehicles has been investigated. A simple semi-empirical model for predicting the extent of <span class="hlt">separation</span> was developed. This model offers considerable computational economy as compared to other schemes reported in the literature, and has been shown to be in good agreement with limited flight data. The unsteady pressure field in plume-<span class="hlt">induced</span> <span class="hlt">separated</span> regions was investigated. It was found that fluctuations differed from those for a rigid flare only at low frequencies. The major difference between plume-<span class="hlt">induced</span> <span class="hlt">separation</span> and flare-<span class="hlt">induced</span> <span class="hlt">separation</span> was shown to be an increase in shock oscillation distance for the plume case. The prediction schemes were applied to PRR shuttle launch configuration. It was found that fluctuating pressures from plume-<span class="hlt">induced</span> <span class="hlt">separation</span> are not as severe as for other fluctuating environments at the critical flight condition of maximum dynamic pressure.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19980200978','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19980200978"><span><span class="hlt">Vapor-Phase</span> Stoichiometry and Heat Treatment of CdTe Starting Material for Physical <span class="hlt">Vapor</span> Transport</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Su, Ching-Hua; Sha, Yi-Gao; Lehoczky, S. L.; Liu, Hao-Chieh; Fang, Rei; Brebrick, R. F.</p> <p>1998-01-01</p> <p>Six batches of CdTe, having total amounts of material from 99 to 203 g and gross mole fraction of Te, X(sub Te), 0.499954-0.500138, were synthesized from pure Cd and Te elements. The <span class="hlt">vapor-phase</span> stoichiometry of the assynthesized CdTe batches was determined from the partial pressure of Te2, P(sub Te2) using an optical absorption technique. The measured <span class="hlt">vapor</span> compositions at 870 C were Te-rich for all of the batches with partial pressure ratios of Cd to Te2, P(sub Cd)/P(sub Te2), ranging from 0.00742 to 1.92. After the heat treatment of baking under dynamic vacuum at 870 C for 8 min, the <span class="hlt">vapor-phase</span> compositions moved toward that of the congruent sublimation, i.e. P(sub Cd)/P(sub Te2) = 2.0, with the measured P(sub Cd)/P(sub Te2) varying from 1.84 to 3.47. The partial pressure measurements on one of the heat-treated samples also showed that the sample remained close to the congruent sublimation condition over the temperature range 800-880 C.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014ExFl...55.1804Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014ExFl...55.1804Z"><span>Experimental study of flash boiling spray <span class="hlt">vaporization</span> through quantitative <span class="hlt">vapor</span> concentration and liquid temperature measurements</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhang, Gaoming; Hung, David L. S.; Xu, Min</p> <p>2014-08-01</p> <p>Flash boiling sprays of liquid injection under superheated conditions provide the novel solutions of fast <span class="hlt">vaporization</span> and better air-fuel mixture formation for internal combustion engines. However, the physical mechanisms of flash boiling spray <span class="hlt">vaporization</span> are more complicated than the droplet surface <span class="hlt">vaporization</span> due to the unique bubble generation and boiling process inside a superheated bulk liquid, which are not well understood. In this study, the <span class="hlt">vaporization</span> of flash boiling sprays was investigated experimentally through the quantitative measurements of <span class="hlt">vapor</span> concentration and liquid temperature. Specifically, the laser-<span class="hlt">induced</span> exciplex fluorescence technique was applied to distinguish the liquid and <span class="hlt">vapor</span> distributions. Quantitative <span class="hlt">vapor</span> concentration was obtained by correlating the intensity of <span class="hlt">vapor-phase</span> fluorescence with <span class="hlt">vapor</span> concentration through systematic corrections and calibrations. The intensities of two wavelengths were captured simultaneously from the liquid-<span class="hlt">phase</span> fluorescence spectra, and their intensity ratios were correlated with liquid temperature. The results show that both liquid and <span class="hlt">vapor</span> <span class="hlt">phase</span> of multi-hole sprays collapse toward the centerline of the spray with different mass distributions under the flash boiling conditions. Large amount of <span class="hlt">vapor</span> aggregates along the centerline of the spray to form a "gas jet" structure, whereas the liquid distributes more uniformly with large vortexes formed in the vicinity of the spray tip. The <span class="hlt">vaporization</span> process under the flash boiling condition is greatly enhanced due to the intense bubble generation and burst. The liquid temperature measurements show strong temperature variations inside the flash boiling sprays with hot zones present in the "gas jet" structure and vortex region. In addition, high <span class="hlt">vapor</span> concentration and closed vortex motion seem to have inhibited the heat and mass transfer in these regions. In summary, the <span class="hlt">vapor</span> concentration and liquid temperature provide detailed information</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JMMM..431...84K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JMMM..431...84K"><span>Magnetic filtration of <span class="hlt">phase</span> <span class="hlt">separating</span> ferrofluids: From basic concepts to microfluidic device</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kuzhir, P.; Magnet, C.; Ezzaier, H.; Zubarev, A.; Bossis, G.</p> <p>2017-06-01</p> <p>In this work, we briefly review magnetic <span class="hlt">separation</span> of ferrofluids composed of large magnetic particles (60 nm of the average size) possessing an <span class="hlt">induced</span> dipole moment. Such ferrofluids exhibit field-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> at relatively low particle concentrations (∼0.8 vol%) and magnetic fields (∼10 kA/m). Particle aggregates appearing during the <span class="hlt">phase</span> <span class="hlt">separation</span> are extracted from the suspending fluid by magnetic field gradients much easier than individual nanoparticles in the absence of <span class="hlt">phase</span> <span class="hlt">separation</span>. Nanoparticle capture by a single magnetized microbead and by multi-collector systems (packed bed of spheres and micro-pillar array) has been studied both experimentally and theoretically. Under flow and magnetic fields, the particle capture efficiency Λ decreases with an increasing Mason number for all considered geometries. This decrease may become stronger for aggregated magnetic particles (Λ ∝Ma-1.7) than for individual ones (Λ ∝Ma-1) if the shear fields are strong enough to provoke aggregate rupture. These results can be useful for development of new magneto-microfluidic immunoassays based on magnetic nanoparticles offering a much better sensitivity as compared to presently used magnetic microbeads.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28558907','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28558907"><span>Cold-<span class="hlt">induced</span> aqueous acetonitrile <span class="hlt">phase</span> <span class="hlt">separation</span>: A salt-free way to begin quick, easy, cheap, effective, rugged, safe.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Shao, Gang; Agar, Jeffrey; Giese, Roger W</p> <p>2017-07-14</p> <p>Cooling a 1:1 (v/v) solution of acetonitrile and water at -16° C is known to result in two clear <span class="hlt">phases</span>. We will refer to this event as "cold-<span class="hlt">induced</span> aqueous acetonitrile <span class="hlt">phase</span> <span class="hlt">separation</span> (CIPS)". On a molar basis, acetonitrile is 71.7% and 13.6% in the upper and lower <span class="hlt">phases</span>, respectively, in our study. The <span class="hlt">phase</span> <span class="hlt">separation</span> proceeds as a descending cloud of microdroplets. At the convenient temperature (typical freezer) employed here the lower <span class="hlt">phase</span> is rather resistant to solidification, although it emerges from the freezer as a solid if various insoluble matter is present at the outset. In a preliminary way, we replaced the initial (salting-out) step of a representative QuEChERS procedure with CIPS, applying this modified procedure ("CIPS-QuEChERS") to a homogenate of salmon (and partly to beef). Three <span class="hlt">phases</span> resulted, where only the upper, acetonitrile-rich <span class="hlt">phase</span> is a liquid (that is completely clear). The middle <span class="hlt">phase</span> comprises ice and precipitated lipids, while the lower <span class="hlt">phase</span> is the residual matrix of undissolved salmon or meat. Treating the upper <span class="hlt">phase</span> from salmon, after isolation, with anhydrous MgSO 4 and C18-Si (typical QuEChERS dispersive solid <span class="hlt">phase</span> extraction sorbents), and injecting into a GC-MS in a nontargeted mode, gives two-fold more preliminary hits for chemicals, and also number of spiked pesticides recovered, relative to that from a comparable QuEChERS method. In part, this is because of much higher background signals in the latter case. Further study of CIPS-QuEChERS is encouraged, including taking advantage of other QuERChERS conditions. Copyright © 2017 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ApSS..395...86D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ApSS..395...86D"><span>Relation between secondary doping and <span class="hlt">phase</span> <span class="hlt">separation</span> in PEDOT:PSS films</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Donoval, Martin; Micjan, Michal; Novota, Miroslav; Nevrela, Juraj; Kovacova, Sona; Pavuk, Milan; Juhasz, Peter; Jagelka, Martin; Kovac, Jaroslav; Jakabovic, Jan; Cigan, Marek; Weis, Martin</p> <p>2017-02-01</p> <p>Conductive copolymer poly(3,4-ethylenedioxythiophene):poly(4-styrenesulfonate) (PEDOT:PSS) has been proposed as an alternative to transparent conductive oxides because of its flexibility, transparency, and low-cost production. Four different secondary dopants, namely N,N-dimethylformamide, ethyleneglycol, sorbitol, and dimethyl sulfoxide, have been used to improve the conductivity. The relation between the structure changes and conductivity enhancement is studied in detail. Atomic force microscopy study of the thin film surface reveals the <span class="hlt">phase</span> <span class="hlt">separation</span> of PEDOT and PSS. We demonstrate that secondary doping <span class="hlt">induces</span> the <span class="hlt">phase</span> <span class="hlt">separation</span> as well as the conductivity enhancement.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_5");'>5</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li class="active"><span>7</span></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_7 --> <div id="page_8" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li class="active"><span>8</span></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="141"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29924608','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29924608"><span>Temperature-<span class="hlt">Induced</span> <span class="hlt">Phase</span> <span class="hlt">Separation</span> in Molecular Assembly of Nanotubes Comprising Amphiphilic Polypeptide with Poly( N-Ethyl Glycine) in Water by a Hydrophilic-Region Driven Type Mechanism.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hattori, Tetsuya; Itagaki, Toru; Uji, Hirotaka; Kimura, Shunsaku</p> <p>2018-06-20</p> <p>Two kinds of amphiphilic polypeptides having different types of hydrophilic polypeptoids, poly(sarcosine)-b-(L-Leu-Aib)6 (ML12) and poly(N-ethyl glycine)-b-(L-Leu-Aib)6 (EL12), were self-assembled via two paths to <span class="hlt">phase-separated</span> nanotubes. One path was via sticking ML12 nanotubes with EL12 nanotubes, and the other was a preparation from a mixture of ML12 and EL12 in solution. In either case, nanotubes showed temperature-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> along the long axis, which was observed by two methods of labeling one <span class="hlt">phase</span> with gold nanoparticles and fluorescence resonance energy transfer between the components. The <span class="hlt">phase-separation</span> was ascribed to aggregation of poly(N-ethyl glycine) blocks over the cloud point temperature. The addition of 5% trifluoroethanol was needed for the <span class="hlt">phase</span> <span class="hlt">separation</span>, because the tight association of the helices in the hydrophobic region should be loosened to allow lateral diffusion of the components to be <span class="hlt">separated</span>. The <span class="hlt">phase-separation</span> in molecular assemblies in water based on the hydrophilic-region driven type mechanism therefore requires sophisticated balances of association forces exerting among the hydrophilic and hydrophobic regions of the amphiphilic polypeptoids.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28678400','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28678400"><span>Nicotine <span class="hlt">Vapor</span> Method to <span class="hlt">Induce</span> Nicotine Dependence in Rodents.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kallupi, Marsida; George, Olivier</p> <p>2017-07-05</p> <p>Nicotine, the main addictive component of tobacco, <span class="hlt">induces</span> potentiation of brain stimulation reward, increases locomotor activity, and <span class="hlt">induces</span> conditioned place preference. Nicotine cessation produces a withdrawal syndrome that can be relieved by nicotine replacement therapy. In the last decade, the market for electronic cigarettes has flourished, especially among adolescents. The nicotine <span class="hlt">vaporizer</span> or electronic nicotine delivery system is a battery-operated device that allows the user to simulate the experience of tobacco smoking without inhaling smoke. The device is designed to be an alternative to conventional cigarettes that emits <span class="hlt">vaporized</span> nicotine inhaled by the user. This report describes a procedure to <span class="hlt">vaporize</span> nicotine in the air to produce blood nicotine levels in rodents that are clinically relevant to those that are observed in humans and produce dependence. We also describe how to construct the apparatus to deliver nicotine <span class="hlt">vapor</span> in a stable, reliable, and consistent manner, as well as how to analyze air for nicotine content. © 2017 by John Wiley & Sons, Inc. Copyright © 2017 John Wiley & Sons, Inc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19790012164','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19790012164"><span>Method and turbine for extracting kinetic energy from a stream of two-<span class="hlt">phase</span> fluid</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Elliott, D. G. (Inventor)</p> <p>1979-01-01</p> <p>An axial flow <span class="hlt">separator</span> turbine is described which includes a number of nozzles for delivering streams of a two-<span class="hlt">phase</span> fluid along linear paths. A <span class="hlt">phase</span> <span class="hlt">separator</span> which responsively <span class="hlt">separates</span> the <span class="hlt">vapor</span> and liquid is characterized by concentrically related annuli supported for rotation within the paths. The <span class="hlt">separator</span> has endless channels for confining the liquid under the influence of centrifugal forces. A <span class="hlt">vapor</span> turbine fan extracts kinetic energy from the liquid. Angular momentum of both the liquid <span class="hlt">phase</span> and the <span class="hlt">vapor</span> <span class="hlt">phase</span> of the fluid is converted to torque.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23767508','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23767508"><span><span class="hlt">Vapor</span>-liquid coexistence of the Stockmayer fluid in nonuniform external fields.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Samin, Sela; Tsori, Yoav; Holm, Christian</p> <p>2013-05-01</p> <p>We investigate the structure and <span class="hlt">phase</span> behavior of the Stockmayer fluid in the presence of nonuniform electric fields using molecular simulation. We find that an initially homogeneous <span class="hlt">vapor</span> <span class="hlt">phase</span> undergoes a local <span class="hlt">phase</span> <span class="hlt">separation</span> in a nonuniform field due to the combined effect of the field gradient and the fluid <span class="hlt">vapor</span>-liquid equilibrium. This results in a high-density fluid condensing in the strong field region. The system polarization exhibits a strong field dependence due to the fluid condensation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AGUFM.C13C0978D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AGUFM.C13C0978D"><span>The Breathing Snowpack: Pressure-<span class="hlt">induced</span> <span class="hlt">Vapor</span> Flux of Temperate Snow</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Drake, S. A.; Selker, J. S.; Higgins, C. W.</p> <p>2017-12-01</p> <p>As surface air pressure increases, hydrostatic compression of the air column forces atmospheric air into snowpack pore space. Likewise, as surface air pressure decreases, the atmospheric air column decompresses and saturated air exits the snow. Alternating influx and efflux of air can be thought of as a "breathing" process that produces an upward <span class="hlt">vapor</span> flux when air above the snow is not saturated. The impact of pressure-<span class="hlt">induced</span> <span class="hlt">vapor</span> exchange is assumed to be small and is thus ignored in model parameterizations of surface processes over snow. Rationale for disregarding this process is that large amplitude pressure changes as caused by synoptic weather patterns are too infrequent to credibly impact <span class="hlt">vapor</span> flux. The amplitude of high frequency pressure changes is assumed to be too small to affect <span class="hlt">vapor</span> flux, however, the basis for this hypothesis relies on pressure measurements collected over an agricultural field (rather than snow). Resolution of the impact of pressure changes on <span class="hlt">vapor</span> flux over seasonal cycles depends on an accurate representation of the magnitude of pressure changes caused by changes in wind as a function of the frequency of pressure changes. High precision in situ pressure measurements in a temperature snowpack allowed us to compute the spectra of pressure changes vs. wind forcing. Using a simplified model for <span class="hlt">vapor</span> exchange we then computed the frequency of pressure changes that maximize <span class="hlt">vapor</span> exchange. We examine and evaluate the seasonal impact of pressure-<span class="hlt">induced</span> <span class="hlt">vapor</span> exchange relative to other snow ablation processes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011JChPh.135w4503T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011JChPh.135w4503T"><span>Kinetics of <span class="hlt">phase</span> <span class="hlt">separation</span> and coarsening in dilute surfactant pentaethylene glycol monododecyl ether solutions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tanaka, S.; Kubo, Y.; Yokoyama, Y.; Toda, A.; Taguchi, K.; Kajioka, H.</p> <p>2011-12-01</p> <p>We investigated the <span class="hlt">phase</span> <span class="hlt">separation</span> phenomena in dilute surfactant pentaethylene glycol monodedecyl ether (C12E5) solutions focusing on the growth law of <span class="hlt">separated</span> domains. The solutions confined between two glass plates were found to exhibit the <span class="hlt">phase</span> inversion, characteristic of the viscoelastic <span class="hlt">phase</span> <span class="hlt">separation</span>; the majority <span class="hlt">phase</span> (water-rich <span class="hlt">phase</span>) nucleated as droplets and the minority <span class="hlt">phase</span> (micelle-rich <span class="hlt">phase</span>) formed a network temporarily, then they collapsed into an usual sea-island pattern where minority <span class="hlt">phase</span> formed islands. We found from the real-space microscopic imaging that the dynamic scaling hypothesis did not hold throughout the coarsening process. The power law growth of the domains with the exponent close to 1/3 was observed even though the coarsening was <span class="hlt">induced</span> mainly by hydrodynamic flow, which was explained by Darcy's law of laminar flow.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19990047142','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19990047142"><span>A Preliminary Study on the <span class="hlt">Vapor</span>/Mist <span class="hlt">Phase</span> Lubrication of a Spur Gearbox</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Morales, Wilfredo; Handschuh, Robert F.</p> <p>1999-01-01</p> <p>Organophosphates have been the primary compounds used in <span class="hlt">vapor</span>/mist <span class="hlt">phase</span> lubrication studies involving ferrous bearing material. Experimental results have indicated that the initial formation of an iron phosphate film on a rubbing ferrous surface, followed by the growth (by cationic diffusion) of a lubricious pyrophosphate-type coating over the iron phosphate, is the reason organophosphates work well as <span class="hlt">vapor</span>/mist <span class="hlt">phase</span> lubricants. Recent work, however, has shown that this mechanism leads to the depletion of surface iron atoms and to eventual lubrication failure. A new organophosphate formulation was developed which circumvents surface iron depletion. This formulation was tested by generating an iron phosphate coating on an aluminum surface. The new formulation was then used to <span class="hlt">vapor</span>/mist <span class="hlt">phase</span> lubricate a spur gearbox in a preliminary study.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2553125','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=2553125"><span><span class="hlt">Phase</span> <span class="hlt">Separation</span> and Crystallization of Hemoglobin C in Transgenic Mouse and Human Erythrocytes</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Canterino, Joseph E.; Galkin, Oleg; Vekilov, Peter G.; Hirsch, Rhoda Elison</p> <p>2008-01-01</p> <p>Individuals expressing hemoglobin C (β6 Glu→Lys) present red blood cells (RBC) with intraerythrocytic crystals that form when hemoglobin (Hb) is oxygenated. Our earlier in vitro liquid-liquid (L-L) <span class="hlt">phase</span> <span class="hlt">separation</span> studies demonstrated that liganded HbC exhibits a stronger net intermolecular attraction with a longer range than liganded HbS or HbA, and that L-L <span class="hlt">phase</span> <span class="hlt">separation</span> preceded and enhanced crystallization. We now present evidence for the role of <span class="hlt">phase</span> <span class="hlt">separation</span> in HbC crystallization in the RBC, and the role of the RBC membrane as a nucleation center. RBC obtained from both human homozygous HbC patients and transgenic mice expressing only human HbC were studied by bright-field and differential interference contrast video-enhanced microscopy. RBC were exposed to hypertonic NaCl solution (1.5–3%) to <span class="hlt">induce</span> crystallization within an appropriate experimental time frame. L-L <span class="hlt">phase</span> <span class="hlt">separation</span> occurred inside the RBC, which in turn enhanced the formation of intraerythrocytic crystals. RBC L-L <span class="hlt">phase</span> <span class="hlt">separation</span> and crystallization comply with the thermodynamic and kinetics laws established through in vitro studies of <span class="hlt">phase</span> transformations. This is the first report, to the best of our knowledge, to capture a temporal view of intraerythrocytic HbC <span class="hlt">phase</span> <span class="hlt">separation</span>, crystal formation, and dissolution. PMID:18621841</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24089800','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24089800"><span>Effect of temperature gradient on liquid-liquid <span class="hlt">phase</span> <span class="hlt">separation</span> in a polyolefin blend.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jiang, Hua; Dou, Nannan; Fan, Guoqiang; Yang, Zhaohui; Zhang, Xiaohua</p> <p>2013-09-28</p> <p>We have investigated experimentally the structure formation processes during <span class="hlt">phase</span> <span class="hlt">separation</span> via spinodal decomposition above and below the spinodal line in a binary polymer blend system exposed to in-plane stationary thermal gradients using <span class="hlt">phase</span> contrast optical microscopy and temperature gradient hot stage. Below the spinodal line there is a coupling of concentration fluctuations and thermal gradient imposed by the temperature gradient hot stage. Also under the thermal gradient annealing <span class="hlt">phase-separated</span> domains grow faster compared with the system under homogeneous temperature annealing on a zero-gradient or a conventional hot stage. We suggest that the in-plane thermal gradient accelerates <span class="hlt">phase</span> <span class="hlt">separation</span> through the enhancement in concentration fluctuations in the early and intermediate stages of spinodal decomposition. In a thermal gradient field, the strength of concentration fluctuation close to the critical point (above the spinodal line) is strong enough to <span class="hlt">induce</span> <span class="hlt">phase</span> <span class="hlt">separation</span> even in one-<span class="hlt">phase</span> regime of the <span class="hlt">phase</span> diagram. In the presence of a temperature gradient the equilibrium <span class="hlt">phase</span> diagrams are no longer valid, and the systems with an upper critical solution temperature can be quenched into <span class="hlt">phase</span> <span class="hlt">separation</span> by applying the stationary temperature gradient. The in-plane temperature gradient drives enhanced concentration fluctuations in a binary polymer blend system above and below the spinodal line.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28382158','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28382158"><span><span class="hlt">Phase</span>-transitional Fe3O4/perfluorohexane Microspheres for Magnetic Droplet <span class="hlt">Vaporization</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Ronghui; Zhou, Yang; Zhang, Ping; Chen, Yu; Gao, Wei; Xu, Jinshun; Chen, Hangrong; Cai, Xiaojun; Zhang, Kun; Li, Pan; Wang, Zhigang; Hu, Bing; Ying, Tao; Zheng, Yuanyi</p> <p>2017-01-01</p> <p>Activating droplets <span class="hlt">vaporization</span> has become an attractive strategy for ultrasound imaging and physical therapy due to the significant increase in ultrasound backscatter signals and its ability to physically damage the tumor cells. However, the current two types of transitional droplets named after their activation methods have their respective limitations. To circumvent the limitations of these activation methods, here we report the concept of magnetic droplet <span class="hlt">vaporization</span> (MDV) for stimuli-responsive cancer theranostics by a magnetic-responsive <span class="hlt">phase</span>-transitional agent. This magnetic-sensitive <span class="hlt">phase</span>-transitional agent-perfluorohexane (PFH)-loaded porous magnetic microspheres (PFH-PMMs), with high magnetic-thermal energy-transfer capability, could quickly respond to external alternating current (AC) magnetic fields to produce thermal energy and trigger the <span class="hlt">vaporization</span> of the liquid PFH. We systematically demonstrated MDV both in vitro and in vivo. This novel trigger method with deep penetration can penetrate the air-filled viscera and trigger the <span class="hlt">vaporization</span> of the <span class="hlt">phase</span>-transitional agent without the need of pre-focusing lesion. This unique MDV strategy is expected to substantially broaden the biomedical applications of nanotechnology and promote the clinical treatment of tumors that are not responsive to chemical therapies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1337626','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/1337626"><span>Hybrid <span class="hlt">vapor</span> <span class="hlt">phase</span>-solution <span class="hlt">phase</span> growth techniques for improved CZT(S,Se) photovoltaic device performance</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Chang, Liang-Yi; Gershon, Talia S.; Haight, Richard A.; Lee, Yun Seog</p> <p>2016-12-27</p> <p>A hybrid <span class="hlt">vapor</span> <span class="hlt">phase</span>-solution <span class="hlt">phase</span> CZT(S,Se) growth technique is provided. In one aspect, a method of forming a kesterite absorber material on a substrate includes the steps of: depositing a layer of a first kesterite material on the substrate using a <span class="hlt">vapor</span> <span class="hlt">phase</span> deposition process, wherein the first kesterite material includes Cu, Zn, Sn, and at least one of S and Se; annealing the first kesterite material to crystallize the first kesterite material; and depositing a layer of a second kesterite material on a side of the first kesterite material opposite the substrate using a solution <span class="hlt">phase</span> deposition process, wherein the second kesterite material includes Cu, Zn, Sn, and at least one of S and Se, wherein the first kesterite material and the second kesterite material form a multi-layer stack of the absorber material on the substrate. A photovoltaic device and method of formation thereof are also provided.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JHyd..549..452H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JHyd..549..452H"><span>Analytical solutions for a soil <span class="hlt">vapor</span> extraction model that incorporates gas <span class="hlt">phase</span> dispersion and molecular diffusion</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Huang, Junqi; Goltz, Mark N.</p> <p>2017-06-01</p> <p>To greatly simplify their solution, the equations describing radial advective/dispersive transport to an extraction well in a porous medium typically neglect molecular diffusion. While this simplification is appropriate to simulate transport in the saturated zone, it can result in significant errors when modeling gas <span class="hlt">phase</span> transport in the vadose zone, as might be applied when simulating a soil <span class="hlt">vapor</span> extraction (SVE) system to remediate vadose zone contamination. A new analytical solution for the equations describing radial gas <span class="hlt">phase</span> transport of a sorbing contaminant to an extraction well is presented. The equations model advection, dispersion (including both mechanical dispersion and molecular diffusion), and rate-limited mass transfer of dissolved, <span class="hlt">separate</span> <span class="hlt">phase</span>, and sorbed contaminants into the gas <span class="hlt">phase</span>. The model equations are analytically solved by using the Laplace transform with respect to time. The solutions are represented by confluent hypergeometric functions in the Laplace domain. The Laplace domain solutions are then evaluated using a numerical Laplace inversion algorithm. The solutions can be used to simulate the spatial distribution and the temporal evolution of contaminant concentrations during operation of a soil <span class="hlt">vapor</span> extraction well. Results of model simulations show that the effect of gas <span class="hlt">phase</span> molecular diffusion upon concentrations at the extraction well is relatively small, although the effect upon the distribution of concentrations in space is significant. This study provides a tool that can be useful in designing SVE remediation strategies, as well as verifying numerical models used to simulate SVE system performance.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013ACP....13.4681S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013ACP....13.4681S"><span>Heterogeneous ice nucleation on <span class="hlt">phase-separated</span> organic-sulfate particles: effect of liquid vs. glassy coatings</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schill, G. P.; Tolbert, M. A.</p> <p>2013-05-01</p> <p>Atmospheric ice nucleation on aerosol particles relevant to cirrus clouds remains one of the least understood processes in the atmosphere. Upper tropospheric aerosols as well as sub-visible cirrus residues are known to be enhanced in both sulfates and organics. The hygroscopic <span class="hlt">phase</span> transitions of organic-sulfate particles can have an impact on both the cirrus cloud formation mechanism and resulting cloud microphysical properties. In addition to deliquescence and efflorescence, organic-sulfate particles are known to undergo another <span class="hlt">phase</span> transition known as liquid-liquid <span class="hlt">phase</span> <span class="hlt">separation</span>. The ice nucleation properties of particles that have undergone liquid-liquid <span class="hlt">phase</span> <span class="hlt">separation</span> are unknown. Here, Raman microscopy coupled with an environmental cell was used to study the low temperature deliquescence, efflorescence, and liquid-liquid <span class="hlt">phase</span> <span class="hlt">separation</span> behavior of 2 : 1 mixtures of organic polyols (1,2,6-hexanetriol and 1 : 1 1,2,6-hexanetriol + 2,2,6,6-tetrakis(hydroxymethyl)cyclohexanol) and ammonium sulfate from 240-265 K. Further, the ice nucleation efficiency of these organic-sulfate systems after liquid-liquid <span class="hlt">phase</span> <span class="hlt">separation</span> and efflorescence was investigated from 210-235 K. Raman mapping and volume-geometry analysis indicate that these particles contain solid ammonium sulfate cores fully engulfed in organic shells. For the ice nucleation experiments, we find that if the organic coatings are liquid, water <span class="hlt">vapor</span> diffuses through the shell and ice nucleates on the ammonium sulfate core. In this case, the coatings minimally affect the ice nucleation efficiency of ammonium sulfate. In contrast, if the coatings become semi-solid or glassy, ice instead nucleates on the organic shell. Consistent with recent findings that glasses can be efficient ice nuclei, the <span class="hlt">phase-separated</span> particles are nearly as efficient at ice nucleation as pure crystalline ammonium sulfate.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012ACPD...1230951S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012ACPD...1230951S"><span>Heterogeneous ice nucleation on <span class="hlt">phase-separated</span> organic-sulfate particles: effect of liquid vs. glassy coatings</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schill, G. P.; Tolbert, M. A.</p> <p>2012-12-01</p> <p>Atmospheric ice nucleation on aerosol particles relevant to cirrus clouds remains one of the least understood processes in the atmosphere. Upper tropospheric aerosols as well as sub-visible cirrus residues are known to be enhanced in both sulfates and organics. The hygroscopic <span class="hlt">phase</span> transitions of organic-sulfate particles can have an impact on both the cirrus cloud formation mechanism and resulting cloud microphysical properties. In addition to deliquescence and efflorescence, organic-sulfate particles are known to undergo another <span class="hlt">phase</span> transition known as liquid-liquid <span class="hlt">phase</span> <span class="hlt">separation</span>. The ice nucleation properties of particles that have undergone liquid-liquid <span class="hlt">phase</span> <span class="hlt">separation</span> are unknown. Here, Raman microscopy coupled with an environmental cell was used to study the low temperature deliquescence, efflorescence, and liquid-liquid <span class="hlt">phase</span> <span class="hlt">separation</span> behavior of 2:1 mixtures of organic polyols (1,2,6-hexanetriol, and 1:1 1,2,6-hexanetriol +2,2,6,6-tetrakis(hydroxymethyl)cycohexanol) and ammonium sulfate from 240-265 K. Further, the ice nucleation efficiency of these organic-sulfate systems after liquid-liquid <span class="hlt">phase</span> <span class="hlt">separation</span> and efflorescence was investigated from 210-235 K. Raman mapping and volume-geometry analysis indicates that these particles contain solid ammonium sulfate cores fully engulfed in organic shells. For the ice nucleation experiments, we find that if the organic coatings are liquid, water <span class="hlt">vapor</span> diffuses through the shell and ice nucleates on the ammonium sulfate core. In this case, the coatings minimally affect the ice nucleation efficiency of ammonium sulfate. In contrast, if the coatings become semi-solid or glassy, ice instead nucleates on the organic shell. Consistent with recent findings that glasses can be efficient ice nuclei, the <span class="hlt">phase</span> <span class="hlt">separated</span> particles are nearly as efficient at ice nucleation as pure crystalline ammonium sulfate.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19800020334','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19800020334"><span>Microwave radiometry as a tool to calibrate tropospheric water-<span class="hlt">vapor</span> delay</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Resch, G. M.; Claflin, E. S.</p> <p>1980-01-01</p> <p>Microwave radiometers were used to measure the emission line due to the water <span class="hlt">vapor</span> molecules of atmospheric emission. Four <span class="hlt">separate</span> field tests were completed which compared radiometers to other techniques which measure water <span class="hlt">vapor</span>. It is shown that water <span class="hlt">vapor</span> <span class="hlt">induced</span> delay can be estimated with an accuracy of plus or minus 2 cm for elevation angles above 17 degrees.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..SHK.T6005S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..SHK.T6005S"><span>Shock-and-Release to the Liquid-<span class="hlt">Vapor</span> <span class="hlt">Phase</span> Boundary: Experiments and Applications to Planetary Science</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Stewart, Sarah</p> <p>2017-06-01</p> <p>Shock-<span class="hlt">induced</span> <span class="hlt">vaporization</span> was a common process during the end stages of terrestrial planet formation and transient features in extra-solar systems are attributed to recent giant impacts. At the Sandia Z Machine, my collaborators and I are conducting experiments to study the shock Hugoniot and release to the liquid-<span class="hlt">vapor</span> <span class="hlt">phase</span> boundary of major minerals in rocky planets. Current work on forsterite, enstatite and bronzite and previous results on silica, iron and periclase demonstrate that shock-<span class="hlt">induced</span> <span class="hlt">vaporization</span> played a larger role during planet formation than previously thought. I will provide an overview of the experimental results and describe how the data have changed our views of planetary impact events in our solar system and beyond. Sandia National Laboratories is a multiprogram laboratory managed and operated by Sandia Corporation, a wholly owned subsidiary of Lockheed Martin Corporation, for the U.S. Department of Energy's National Nuclear Security Administration under Contract No. DE-AC04-94AL85000. This work was performed under the auspices of the U.S. Department of Energy by Lawrence Livermore National Laboratory under Contract DE-AC52-07NA27344. This work is supported by the Z Fundamental Science Program at Sandia National Laboratories, DOE-NNSA Grant DE- NA0002937, NASA Grant # NNX15AH54G, and UC Multicampus-National Lab Collaborative Research and Training Grant #LFR-17-449059.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhSS...59.2418A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhSS...59.2418A"><span>Optical properties of bulk gallium nitride single crystals grown by chloride-hydride <span class="hlt">vapor-phase</span> epitaxy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Agyekyan, V. F.; Borisov, E. V.; Serov, A. Yu.; Filosofov, N. G.</p> <p>2017-12-01</p> <p>A gallium nitride crystal 5 mm in thickness was grown by chloride-hydride <span class="hlt">vapor-phase</span> epitaxy on a sapphire substrate, from which the crystal <span class="hlt">separated</span> during cooling. At an early stage, a three-dimensional growth mode was implemented, followed by a switch to a two-dimensional mode. Spectra of exciton reflection, exciton luminescence, and Raman scattering are studied in several regions characteristic of the sample. Analysis of these spectra and comparison with previously obtained data for thin epitaxial GaN layers with a wide range of silicon doping enabled conclusions about the quality of the crystal lattice in these characteristic regions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2001APS..MARQ24008G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2001APS..MARQ24008G"><span>Field <span class="hlt">induced</span> ferromagnetic fraction enlargement in <span class="hlt">phase</span> <span class="hlt">separated</span> La_0.5Ca_0.5MnO_3</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ghivelder, Luis; Freitas, R. S.; Sacanel, J.; Parisi, F.; Levy, P.</p> <p>2001-03-01</p> <p>A systematic study of the magnetic and transport properties of a series of <span class="hlt">phase</span> <span class="hlt">separated</span> La_0.5Ca_0.5MnO3 compounds is reported. The investigated samples all have the same composition but different grain sizes, which modifies the volume fraction of the coexisting ferromagnetic (FM) and antiferromagnetic charge-ordered (AFM-CO) <span class="hlt">phases</span>. Magnetoresistance and magnetization measurements were performed with two different experimental procedures: a standard field-cooled cooling (FC) mode, and a second method in which the field is turned on only while measuring each data point, and switched off while cooling the samples. Magnetization and magnetoresistance measurements display big differences when comparing the data obtained with the different procedures. The overall results are interpret in terms of a field <span class="hlt">induced</span> FM fraction enlargement. In transport measurements this effect yield a percolative transition. Magnetization data shows evidence for the formation of AFM-CO regions within the FM <span class="hlt">phase</span>. * e-mail: luisghiv@if.ufrj.br</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20060054441&hterms=ammonia&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3Dammonia','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20060054441&hterms=ammonia&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3Dammonia"><span>An Evaluation of the <span class="hlt">Vapor</span> <span class="hlt">Phase</span> Catalytic Ammonia Removal Process for Use in a Mars Transit Vehicle</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Flynn, Michael; Borchers, Bruce</p> <p>1998-01-01</p> <p>An experimental program has been developed to evaluate the potential of the <span class="hlt">Vapor</span> <span class="hlt">Phase</span> Catalytic Ammonia Reduction (VPCAR) technology for use as a Mars Transit Vehicle water purification system. Design modifications which will be required to ensure proper operation of the VPCAR system in reduced gravity are also evaluated. The VPCAR system is an integrated wastewater treatment technology that combines a distillation process with high temperature catalytic oxidation. The distillation portion of the system utilizes a <span class="hlt">vapor</span> compression distillation process to provide an energy efficient <span class="hlt">phase</span> change <span class="hlt">separation</span>. This portion of the system removes any inorganic salts and large molecular weight, organic contaminates, i.e., non-volatile, from the product water stream and concentrates these contaminates into a byproduct stream. To oxidize the volatile organic compounds and ammonia, a <span class="hlt">vapor</span> <span class="hlt">phase</span>, high temperature catalytic oxidizer is used. This catalytic system converts these compounds along with the aqueous product into CO2, H2O, and N2O. A secondary catalytic bed can then be used to reduce the N2O to nitrogen and oxygen (although not evaluated in this study). This paper describes the design specification of the VPCAR process, the relative benefits of its utilization in a Mars Transit Vehicle, and the design modification which will be required to ensure its proper operation in reduced gravity. In addition, the results of an experimental evaluation of the processors is presented. This evaluation presents the processors performance based upon product water purity, water recovery rates, and power.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4034005','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4034005"><span>Vertical <span class="hlt">phase</span> <span class="hlt">separation</span> in bulk heterojunction solar cells formed by in situ polymerization of fulleride</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Zhang, Lipei; Xing, Xing; Zheng, Lingling; Chen, Zhijian; Xiao, Lixin; Qu, Bo; Gong, Qihuang</p> <p>2014-01-01</p> <p>Vertical <span class="hlt">phase</span> <span class="hlt">separation</span> of the donor and the acceptor in organic bulk heterojunction solar cells is crucial to improve the exciton dissociation and charge transport efficiencies. This is because whilst the exciton diffusion length is limited, the organic film must be thick enough to absorb sufficient light. However, it is still a challenge to control the <span class="hlt">phase</span> <span class="hlt">separation</span> of a binary blend in a bulk heterojunction device architecture. Here we report the realization of vertical <span class="hlt">phase</span> <span class="hlt">separation</span> <span class="hlt">induced</span> by in situ photo-polymerization of the acrylate-based fulleride. The power conversion efficiency of the devices with vertical <span class="hlt">phase</span> <span class="hlt">separation</span> increased by 20%. By optimising the device architecture, the power conversion efficiency of the single junction device reached 8.47%. We believe that in situ photo-polymerization of acrylate-based fulleride is a universal and controllable way to realise vertical <span class="hlt">phase</span> <span class="hlt">separation</span> in organic blends. PMID:24861168</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_6");'>6</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li class="active"><span>8</span></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_8 --> <div id="page_9" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li class="active"><span>9</span></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="161"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25218062','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25218062"><span>Current-<span class="hlt">induced</span> transition from particle-by-particle to concurrent intercalation in <span class="hlt">phase-separating</span> battery electrodes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Li, Yiyang; El Gabaly, Farid; Ferguson, Todd R; Smith, Raymond B; Bartelt, Norman C; Sugar, Joshua D; Fenton, Kyle R; Cogswell, Daniel A; Kilcoyne, A L David; Tyliszczak, Tolek; Bazant, Martin Z; Chueh, William C</p> <p>2014-12-01</p> <p>Many battery electrodes contain ensembles of nanoparticles that <span class="hlt">phase-separate</span> on (de)intercalation. In such electrodes, the fraction of actively intercalating particles directly impacts cycle life: a vanishing population concentrates the current in a small number of particles, leading to current hotspots. Reports of the active particle population in the <span class="hlt">phase-separating</span> electrode lithium iron phosphate (LiFePO4; LFP) vary widely, ranging from near 0% (particle-by-particle) to 100% (concurrent intercalation). Using synchrotron-based X-ray microscopy, we probed the individual state-of-charge for over 3,000 LFP particles. We observed that the active population depends strongly on the cycling current, exhibiting particle-by-particle-like behaviour at low rates and increasingly concurrent behaviour at high rates, consistent with our <span class="hlt">phase</span>-field porous electrode simulations. Contrary to intuition, the current density, or current per active internal surface area, is nearly invariant with the global electrode cycling rate. Rather, the electrode accommodates higher current by increasing the active particle population. This behaviour results from thermodynamic transformation barriers in LFP, and such a phenomenon probably extends to other <span class="hlt">phase-separating</span> battery materials. We propose that modifying the transformation barrier and exchange current density can increase the active population and thus the current homogeneity. This could introduce new paradigms to enhance the cycle life of <span class="hlt">phase-separating</span> battery electrodes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014JTePh..59..449B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014JTePh..59..449B"><span><span class="hlt">Vapor</span>-crystal <span class="hlt">phase</span> transition in synthesis of paracetamol films by vacuum evaporation and condensation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Belyaev, A. P.; Rubets, V. P.; Antipov, V. V.; Bordei, N. S.; Zarembo, V. I.</p> <p>2014-03-01</p> <p>We report on the structural and technological investigations of the <span class="hlt">vapor</span>-crystal <span class="hlt">phase</span> transition during synthesis of paracetamol films of the monoclinic system by vacuum evaporation and condensation in the temperature range 220-320 K. The complex nature of the transformation accompanied by the formation of a gel-like <span class="hlt">phase</span> is revealed. The results are interpreted using a model according to which the <span class="hlt">vapor</span>-crystal <span class="hlt">phase</span> transition is not a simple first-order <span class="hlt">phase</span> transition, but is a nonlinear superposition of two <span class="hlt">phase</span> transitions: a first-order transition with a change in density and a second-order <span class="hlt">phase</span> transition with a change in ordering. Micrographs of the surface of the films are obtained at different <span class="hlt">phases</span> of formation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/12033313','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12033313"><span>Gas-liquid chromatography with a volatile "stationary" liquid <span class="hlt">phase</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wells, P S; Zhou, S; Parcher, J F</p> <p>2002-05-01</p> <p>A unique type of gas-liquid chromatography is described in which both mobile and "stationary" <span class="hlt">phases</span> are composed of synthetic mixtures of helium and carbon dioxide. At temperatures below the critical point of the binary mixture and pressures above the <span class="hlt">vapor</span> pressure of pure liquid carbon dioxide, helium and carbon dioxide can form two immiscible <span class="hlt">phases</span> over extended composition ranges. A binary <span class="hlt">vapor</span> <span class="hlt">phase</span> enriched in helium can act as the mobile <span class="hlt">phase</span> for chromatographic <span class="hlt">separations</span>, whereas a CO2-rich liquid in equilibrium with the <span class="hlt">vapor</span> <span class="hlt">phase</span>, but condensed on the column wall, can act as a pseudostationary <span class="hlt">phase</span>. Several examples of chromatographic <span class="hlt">separations</span> obtained in "empty" capillary columns with no ordinary stationary liquid <span class="hlt">phase</span> illustrate the range of conditions that produce such <span class="hlt">separations</span>. In addition, several experiments are reported that confirm the proposed two-<span class="hlt">phase</span> hypothesis. The possible consequences of the observed chromatographic phenomenon in the field of supercritical fluid chromatography with helium headspace carbon dioxide are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017OptCo.405..127Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017OptCo.405..127Z"><span>Wave-mixing-<span class="hlt">induced</span> transparency with zero <span class="hlt">phase</span> shift in atomic <span class="hlt">vapors</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhou, F.; Zhu, C. J.; Li, Y.</p> <p>2017-12-01</p> <p>We present a wave-mixing <span class="hlt">induced</span> transparency that can lead to a hyper-Raman gain-clamping effect. This new type of transparency is originated from a dynamic gain cancellation effect in a multiphoton process where a highly efficient light field of new frequency is generated and amplified. We further show that this novel dynamic gain cancellation effect not only makes the medium transparent to a probe light field at appropriate frequency but also eliminates the probe field propagation <span class="hlt">phase</span> shift. This gain-cancellation-based <span class="hlt">induced</span> transparency holds for many potential applications on optical communication and may lead to effective suppression of parasitic Raman/hyper-Raman noise field generated in high intensity optical fiber transmissions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21203643','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21203643"><span>Cd(1-x)Zn(x)O [0.05 ≤x≤ 0.26] synthesized by <span class="hlt">vapor</span>-diffusion <span class="hlt">induced</span> hydrolysis and co-nucleation from aqueous metal salt solutions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Schwenzer, Birgit; Neilson, James R; Jeffries, Stacie M; Morse, Daniel E</p> <p>2011-02-14</p> <p>Nanoparticulate Cd(1-x)Zn(x)O (x = 0, 0.05-0.26, 1) is synthesized in a simple two-step synthesis approach. <span class="hlt">Vapor</span>-diffusion <span class="hlt">induced</span> catalytic hydrolysis of two molecular precursors at low temperature <span class="hlt">induces</span> co-nucleation and polycondensation to produce bimetallic layered hydroxide salts (M = Cd, Zn) as precursor materials which are subsequently converted to Cd(1-x)Zn(x)O at 400 °C. Unlike ternary materials prepared by standard co-precipitation procedures, all products presented here containing < 30 mol% Zn(2+) ions are homogeneous in elemental composition on the micrometre scale. This measured compositional homogeneity within the samples, as determined by energy dispersive spectroscopy and inductively coupled plasma spectroscopy, is a testimony to the kinetic control achieved by employing slow hydrolysis conditions. In agreement with this observation, the optical properties of the materials obey Vegard's Law for a homogeneous solid solution of Cd(1-x)Zn(x)O, where x corresponds to the values determined by inductively coupled plasma analysis, even though powder X-ray diffraction shows <span class="hlt">phase</span> <span class="hlt">separation</span> into a cubic mixed metal oxide <span class="hlt">phase</span> and a hexagonal ZnO <span class="hlt">phase</span> at all doping levels.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4237215','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4237215"><span>Stress reduction in <span class="hlt">phase-separated</span>, cross-linked networks: influence of <span class="hlt">phase</span> structure and kinetics of reaction</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Szczepanski, Caroline R.; Stansbury, Jeffrey W.</p> <p>2014-01-01</p> <p>A mechanism for polymerization shrinkage and stress reduction was developed for heterogeneous networks formed via ambient, photo-initiated polymerization-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> (PIPS). The material system used consists of a bulk homopolymer matrix of triethylene glycol dimethacrylate (TEGDMA) modified with one of three non-reactive, linear prepolymers (poly-methyl, ethyl and butyl methacrylate). At higher prepolymer loading levels (10–20 wt%) an enhanced reduction in both shrinkage and polymerization stress is observed. The onset of gelation in these materials is delayed to a higher degree of methacrylate conversion (~15–25%), providing more time for <span class="hlt">phase</span> structure evolution by thermodynamically driven monomer diffusion between immiscible <span class="hlt">phases</span> prior to network macro-gelation. The resulting <span class="hlt">phase</span> structure was probed by introducing a fluorescently tagged prepolymer into the matrix. The <span class="hlt">phase</span> structure evolves from a dispersion of prepolymer at low loading levels to a fully co-continuous heterogeneous network at higher loadings. The bulk modulus in <span class="hlt">phase</span> <span class="hlt">separated</span> networks is equivalent or greater than that of poly(TEGDMA), despite a reduced polymerization rate and cross-link density in the prepolymer-rich domains. PMID:25418999</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010EGUGA..1214235B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010EGUGA..1214235B"><span>The solubility of gallium oxide in <span class="hlt">vapor</span> and two-<span class="hlt">phase</span> fluid filtration in hydrothermal systems</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bychkov, Andrew; Matveeva, Svetlana; Nekrasov, Stanislav</p> <p>2010-05-01</p> <p>The solubility of gallium and aluminum oxides in gas <span class="hlt">phase</span> in the system Ga2O3 (Al2O3)-HCl-H2O was studied at 150-350°C and pressure up to saturated <span class="hlt">vapor</span>. The concentration of gallium increases with the increasing of HCl pressure. The formulae of gallium gaseous specie was determined as GaOHCl2. The constant of gallium oxide solubility reaction was calculated at 150, 200, 250, 300 and 350°C. The concentration of aluminum in gas <span class="hlt">phase</span> is insignificant in the same conditions. The possibility of gallium transportation in gas <span class="hlt">phase</span> with small quantity of Al allow to divide this elements in hydrothermal processes with gas <span class="hlt">phase</span>. The Ga/Al ratio in muscovite can be used as the indicator of gas <span class="hlt">phase</span> <span class="hlt">separation</span> and condensation. This indicator was not considered in the geochemical literature earlier. The <span class="hlt">separation</span> of gas and liquid <span class="hlt">phases</span> was determined in Akchatau (Kazahstan) and Spokoinoe (Russia) greisen W deposit by carbon isotope fractionation of carbon dioxide in fluid inclusion. The important feature of both ore mains is heterogenization and boiling of ore-forming fluids. Greisen ore bodies are formed as a result of strongly focused solution flow in the T-P gradient fields. It is possible to divide ore bodies of Akchatau in two types: muscovite and quartz. Muscovite type veins are thin and have small metasyntactic zone. Quartz type veins are localized in fault with large vertical extent (500 m) and content the large quantity of wolframite. These veins formed in condition of significant pressure decreasing from 2.5 to 0.5 kbar with fluid boiling. Gas and liquid <span class="hlt">phase</span> <span class="hlt">separation</span> specifies the vertical zonality of quartz type veins. The gas <span class="hlt">phase</span> with the high gallium concentration is <span class="hlt">separated</span> from a flow of liquid <span class="hlt">phase</span>. Liquid <span class="hlt">phase</span> react with the granites forming greisen metasomatites. Condensation of the gas <span class="hlt">phase</span> in upper parts of massive produces the increasing of Ga/Al ratio in muscovite 3-5 times more, then in granites and bottom part of vein (from 2×10</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70016290','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70016290"><span><span class="hlt">Vapor-phase</span> interactions and diffusion of organic solvents in the unsaturated zone</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Roy, W.R.; Griffin, R.A.</p> <p>1990-01-01</p> <p>This article presents an analysis of the interactions and static movement of 37 organic solvents as <span class="hlt">vapors</span> through the unsaturated soil zone. The physicochemical interactions of the organic <span class="hlt">vapors</span> with unsaturated soil materials were emphasized with focus on diffusive, and adsorptive interactions. Fick's Law and porous media diffusion coefficients for most of the solvent <span class="hlt">vapors</span> were either compiled or estimated; coefficients were not available for some of the fluorinated solvents. The adsorption of some of the solvent <span class="hlt">vapors</span> by silica was concluded to be due to hydrogen bond formation with surface silanol groups. Heats of adsorption data for different adsorbents were also compiled. There were very few data on the adsorption of these solvent <span class="hlt">vapors</span> by soils, but it appears that the magnitude of adsorption of nonpolar solvents is reduced as the relative humidity of the <span class="hlt">vapor</span>-solid system is increased. Consequently, the interaction of the <span class="hlt">vapors</span> may then <span class="hlt">separated</span> into two processes; (1) gas-water partitioning described by Henry's Law constants, and (2) solid-water adsorption coefficients which may be estimated from liquid-solid partition coefficients (Kd values). ?? 1990 Springer-Verlag New York Inc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29932899','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29932899"><span>Functional Domains of NEAT1 Architectural lncRNA <span class="hlt">Induce</span> Paraspeckle Assembly through <span class="hlt">Phase</span> <span class="hlt">Separation</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yamazaki, Tomohiro; Souquere, Sylvie; Chujo, Takeshi; Kobelke, Simon; Chong, Yee Seng; Fox, Archa H; Bond, Charles S; Nakagawa, Shinichi; Pierron, Gerard; Hirose, Tetsuro</p> <p>2018-06-21</p> <p>A class of long noncoding RNAs (lncRNAs) has architectural functions in nuclear body construction; however, specific RNA domains dictating their architectural functions remain uninvestigated. Here, we identified the domains of the architectural NEAT1 lncRNA that construct paraspeckles. Systematic deletion of NEAT1 portions using CRISPR/Cas9 in haploid cells revealed modular domains of NEAT1 important for RNA stability, isoform switching, and paraspeckle assembly. The middle domain, containing functionally redundant subdomains, was responsible for paraspeckle assembly. Artificial tethering of the NONO protein to a NEAT1_2 mutant lacking the functional subdomains rescued paraspeckle assembly, and this required the NOPS dimerization domain of NONO. Paraspeckles exhibit <span class="hlt">phase-separated</span> properties including susceptibility to 1,6-hexanediol treatment. RNA fragments of the NEAT1_2 subdomains preferentially bound NONO/SFPQ, leading to <span class="hlt">phase-separated</span> aggregates in vitro. Thus, we demonstrate that the enrichment of NONO dimers on the redundant NEAT1_2 subdomains initiates construction of <span class="hlt">phase-separated</span> paraspeckles, providing mechanistic insights into lncRNA-based nuclear body formation. Copyright © 2018 Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19890000240&hterms=Glass+bubble&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3DGlass%2Bbubble','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19890000240&hterms=Glass+bubble&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3DGlass%2Bbubble"><span>Study Of <span class="hlt">Phase</span> <span class="hlt">Separation</span> In Glass</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Neilson, George F.; Weinberg, Michael C.; Smith, Gary L.</p> <p>1989-01-01</p> <p>Report describes an experimental study of effect of hydroxide content on <span class="hlt">phase</span> <span class="hlt">separation</span> in soda/silica glasses. Ordinary and gel glasses melted at 1,565 degree C, and melts stirred periodically. "Wet" glasses produced by passing bubbles of N2 saturated with water through melts; "dry" glasses prepared in similar manner, except N2 dried before passage through melts. Analyses of compositions of glasses performed by atomic-absorption and index-of-refraction measurements. Authors conclude hydroxide speeds up <span class="hlt">phase</span> <span class="hlt">separation</span>, regardless of method (gel or ordinary) by which glass prepared. Eventually helps material scientists to find ways to control morphology of <span class="hlt">phase</span> <span class="hlt">separation</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AcSpA.195....1C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AcSpA.195....1C"><span>Investigation of <span class="hlt">phase</span> <span class="hlt">separated</span> polyimide blend films containing boron nitride using FTIR imaging</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chae, Boknam; Hong, Deok Gi; Jung, Young Mee; Won, Jong Chan; Lee, Seung Woo</p> <p>2018-04-01</p> <p>Immiscible aromatic polyimide (PI) blend films and a PI blend film incorporated with thermally conductive boron nitride (BN) were prepared, and their <span class="hlt">phase</span> <span class="hlt">separation</span> behaviors were examined by optical microscopy and FTIR imaging. The 2,2‧-bis(trifluoromethyl)benzidine (TFMB)-containing and 4,4‧-thiodianiline (TDA)-containing aromatic PI blend films and a PI blend/BN composite film show two clearly <span class="hlt">separated</span> regions; one region is the TFMB-rich <span class="hlt">phase</span>, and the other region is the TDA-rich <span class="hlt">phase</span>. The introduction of BN <span class="hlt">induces</span> morphological changes in the immiscible aromatic PI blend film without altering the composition of either domain. In particular, the BN is selectively incorporated into the TDA-rich <span class="hlt">phase</span> in this study.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/873519','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/873519"><span>Method for the generation of variable density metal <span class="hlt">vapors</span> which bypasses the liquidus <span class="hlt">phase</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Kunnmann, Walter; Larese, John Z.</p> <p>2001-01-01</p> <p>The present invention provides a method for producing a metal <span class="hlt">vapor</span> that includes the steps of combining a metal and graphite in a vessel to form a mixture; heating the mixture to a first temperature in an argon gas atmosphere to form a metal carbide; maintaining the first temperature for a period of time; heating the metal carbide to a second temperature to form a metal <span class="hlt">vapor</span>; withdrawing the metal <span class="hlt">vapor</span> and the argon gas from the vessel; and <span class="hlt">separating</span> the metal <span class="hlt">vapor</span> from the argon gas. Metal <span class="hlt">vapors</span> made using this method can be used to produce uniform powders of the metal oxide that have narrow size distribution and high purity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70012579','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70012579"><span>Transient-pressure analysis in geothermal steam reservoirs with an immobile <span class="hlt">vaporizing</span> liquid <span class="hlt">phase</span></span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Moench, A.F.; Atkinson, P.G.</p> <p>1978-01-01</p> <p>A finite-difference model for the radial horizontal flow of steam through a porous medium is used to evaluate transient-pressure behavior in the presence of an immobile <span class="hlt">vaporizing</span> or condensing liquid <span class="hlt">phase</span>. Graphs of pressure drawdown and buildup in terms of dimensionless pressure and time are obtained for a well discharging steam at a constant mass flow rate for a specified time. The assumptions are made that the steam is in local thermal equilibrium with the reservoir rocks, that temperature changes are due only to <span class="hlt">phase</span> change, and that effects of <span class="hlt">vapor</span>-pressure lowering are negligible. Computations show that when a <span class="hlt">vaporizing</span> liquid <span class="hlt">phase</span> is present the pressure drawdown exhibits behavior similar to that observed in noncondensable gas reservoirs, but delayed in time. A theoretical analysis allows for the computation of this delay and demonstrates that it is independent of flow geometry. The response that occurs upon pressure buildup is markedly different from that in a noncondensable gas system. This result may provide a diagnostic tool for establishing the existence of <span class="hlt">phase</span>-change phenomena within a reservoir. ?? 1979.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19720010486','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19720010486"><span>Vacuum distillation: <span class="hlt">vapor</span> filtered-catalytic oxidation water reclamation system utilizing radioisotopes</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Honegger, R. J.; Remus, G. A.; Kurg, E. K.</p> <p>1971-01-01</p> <p>The development of a functional model water reclamation system is discussed. The system produces potable water by distillation from the urine and respiration-perspiration condensate at the normal rate generated by four men. Basic processes employed are vacuum distillation, <span class="hlt">vapor</span> filtration, <span class="hlt">vapor</span> <span class="hlt">phase</span> catalytic oxidation, and condensation. The system is designed to use four 75-watt isotope heaters for distillation thermal input, and one 45-watt isotope for the catalytic oxidation unit. The system is capable of collecting and storing urine, and provides for stabilizing the urine by chemical pretreatment. The functional model system is designed for operation in a weightless condition with liquid-<span class="hlt">vapor</span> <span class="hlt">phase</span> <span class="hlt">separators</span> for the evaporator still, and centrifugal <span class="hlt">separators</span> for urine collection and <span class="hlt">vapor</span> condensation. The system provides for storing and dispensing reclaimed potable water. The system operates in a batch mode for 40 days, with urine residues accumulating in the evaporator. The evaporator still and residue are removed to storage and replaced with a fresh still for the next 40-day period.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20040142361','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20040142361"><span>Gas-Liquid Flows and <span class="hlt">Phase</span> <span class="hlt">Separation</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>McQuillen, John</p> <p>2004-01-01</p> <p>Common issues for space system designers include:Ability to Verify Performance in Normal Gravity prior to Deployment; System Stability; <span class="hlt">Phase</span> Accumulation & Shedding; <span class="hlt">Phase</span> <span class="hlt">Separation</span>; Flow Distribution through Tees & Manifolds Boiling Crisis; Heat Transfer Coefficient; and Pressure Drop.The report concludes:Guidance similar to "A design that operates in a single <span class="hlt">phase</span> is less complex than a design that has two-<span class="hlt">phase</span> flow" is not always true considering the amount of effort spent on pressurizing, subcooling and <span class="hlt">phase</span> <span class="hlt">separators</span> to ensure single <span class="hlt">phase</span> operation. While there is still much to learn about two-<span class="hlt">phase</span> flow in reduced gravity, we have a good start. Focus now needs to be directed more towards system level problems .</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhRvE..96f3303L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhRvE..96f3303L"><span>Improved thermal lattice Boltzmann model for simulation of liquid-<span class="hlt">vapor</span> <span class="hlt">phase</span> change</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Qing; Zhou, P.; Yan, H. J.</p> <p>2017-12-01</p> <p>In this paper, an improved thermal lattice Boltzmann (LB) model is proposed for simulating liquid-<span class="hlt">vapor</span> <span class="hlt">phase</span> change, which is aimed at improving an existing thermal LB model for liquid-<span class="hlt">vapor</span> <span class="hlt">phase</span> change [S. Gong and P. Cheng, Int. J. Heat Mass Transfer 55, 4923 (2012), 10.1016/j.ijheatmasstransfer.2012.04.037]. First, we emphasize that the replacement of ∇ .(λ ∇ T ) /∇.(λ ∇ T ) ρ cV </mml:mpadded></mml:mphantom> ρ cV with ∇ .(χ ∇ T ) is an inappropriate treatment for diffuse interface modeling of liquid-<span class="hlt">vapor</span> <span class="hlt">phase</span> change. Furthermore, the error terms ∂t 0(T v ) +∇ .(T vv ) , which exist in the macroscopic temperature equation recovered from the previous model, are eliminated in the present model through a way that is consistent with the philosophy of the LB method. Moreover, the discrete effect of the source term is also eliminated in the present model. Numerical simulations are performed for droplet evaporation and bubble nucleation to validate the capability of the model for simulating liquid-<span class="hlt">vapor</span> <span class="hlt">phase</span> change. It is shown that the numerical results of the improved model agree well with those of a finite-difference scheme. Meanwhile, it is found that the replacement of ∇ .(λ ∇ T ) /∇ .(λ ∇ T ) ρ cV </mml:mpadded></mml:mphantom> ρ cV with ∇ .(χ ∇ T ) leads to significant numerical errors and the error terms in the recovered macroscopic temperature equation also result in considerable errors.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70160653','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70160653"><span>Irradiation of fish fillets: Relation of <span class="hlt">vapor</span> <span class="hlt">phase</span> reactions to storage quality</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Spinelli, J.; Dollar, A.M.; Wedemeyer, G.A.; Gallagher, E.C.</p> <p>1969-01-01</p> <p>Fish fillets irradiated under air, nitrogen, oxygen, or carbon dioxide atmospheres developed rancidlike flavors when they were stored at refrigerated temperatures. Packing and irradiating under vacuum or helium prevented development of off-flavors during storage.Significant quantities of nitrate and oxidizing substances were formed when oxygen, nitrogen, or air were present in the <span class="hlt">vapor</span> or liquid <span class="hlt">phases</span> contained in a Pyrex glass model system exposed to ionizing radiation supplied by a 60Co source. It was demonstrated that the delayed flavor changes that occur in stored fish fillets result from the reaction of <span class="hlt">vapor</span> <span class="hlt">phase</span> radiolysis products and the fish tissue substrates.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17461649','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17461649"><span>Computer simulation of <span class="hlt">phase</span> <span class="hlt">separation</span> under a double temperature quench.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Podariu, Iulia; Chakrabarti, Amitabha</p> <p>2007-04-21</p> <p>The authors numerically study a two-step quench process in an asymmetric binary mixture. The mixture is first quenched to an unstable state in the two-<span class="hlt">phase</span> region. After a large <span class="hlt">phase-separated</span> structure is formed, the authors again quench the system deeper. The second quench <span class="hlt">induces</span> the formation of small secondary droplets inside the large domains created by the first quench. The authors characterize this secondary droplet growth in terms of the temperature of the first quench as well as the depth of the second one.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17772911','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17772911"><span>Evidence for extreme partitioning of copper into a magmatic <span class="hlt">vapor</span> <span class="hlt">phase</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lowenstern, J B; Mahood, G A; Rivers, M L; Sutton, S R</p> <p>1991-06-07</p> <p>The discovery of copper sulfides in carbon dioxide- and chlorine-bearing bubbles in phenocryst-hosted melt inclusions shows that copper resides in a <span class="hlt">vapor</span> <span class="hlt">phase</span> in some shallow magma chambers. Copper is several hundred times more concentrated in magmatic <span class="hlt">vapor</span> than in coexisting pantellerite melt. The volatile behavior of copper should be considered when modeling the volcanogenic contribution of metals to the atmosphere and may be important in the formation of copper porphyry ore deposits.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29735679','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29735679"><span>Flow-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> of active particles is controlled by boundary conditions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Thutupalli, Shashi; Geyer, Delphine; Singh, Rajesh; Adhikari, Ronojoy; Stone, Howard A</p> <p>2018-05-22</p> <p>Active particles, including swimming microorganisms, autophoretic colloids, and droplets, are known to self-organize into ordered structures at fluid-solid boundaries. The entrainment of particles in the attractive parts of their spontaneous flows has been postulated as a possible mechanism underlying this phenomenon. Here, combining experiments, theory, and numerical simulations, we demonstrate the validity of this flow-<span class="hlt">induced</span> ordering mechanism in a suspension of active emulsion droplets. We show that the mechanism can be controlled, with a variety of resultant ordered structures, by simply altering hydrodynamic boundary conditions. Thus, for flow in Hele-Shaw cells, metastable lines or stable traveling bands can be obtained by varying the cell height. Similarly, for flow bounded by a plane, dynamic crystallites are formed. At a no-slip wall, the crystallites are characterized by a continuous out-of-plane flux of particles that circulate and re-enter at the crystallite edges, thereby stabilizing them. At an interface where the tangential stress vanishes, the crystallites are strictly 2D, with no out-of-plane flux. We rationalize these experimental results by calculating, in each case, the slow viscous flow produced by the droplets and the long-ranged, many-body active forces and torques between them. The results of numerical simulations of motion under the action of the active forces and torques are in excellent agreement with experiments. Our work elucidates the mechanism of flow-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> in active fluids, particularly active colloidal suspensions, and demonstrates its control by boundaries, suggesting routes to geometric and topological phenomena in an active matter.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_7");'>7</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li class="active"><span>9</span></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_9 --> <div id="page_10" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li class="active"><span>10</span></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="181"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4516965','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4516965"><span>Potentiometric detection of chemical <span class="hlt">vapors</span> using molecularly imprinted polymers as receptors</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Liang, Rongning; Chen, Lusi; Qin, Wei</p> <p>2015-01-01</p> <p>Ion-selective electrode (ISE) based potentiometric gas sensors have shown to be promising analytical tools for detection of chemical <span class="hlt">vapors</span>. However, such sensors are only capable of detecting those <span class="hlt">vapors</span> which can be converted into ionic species in solution. This paper describes for the first time a polymer membrane ISE based potentiometric sensing system for sensitive and selective determination of neutral <span class="hlt">vapors</span> in the gas <span class="hlt">phase</span>. A molecularly imprinted polymer (MIP) is incorporated into the ISE membrane and used as the receptor for selective adsorption of the analyte <span class="hlt">vapor</span> from the gas <span class="hlt">phase</span> into the sensing membrane <span class="hlt">phase</span>. An indicator ion with a structure similar to that of the <span class="hlt">vapor</span> molecule is employed to indicate the change in the MIP binding sites in the membrane <span class="hlt">induced</span> by the molecular recognition of the <span class="hlt">vapor</span>. The toluene <span class="hlt">vapor</span> is used as a model and benzoic acid is chosen as its indicator. Coupled to an apparatus manifold for preparation of <span class="hlt">vapor</span> samples, the proposed ISE can be utilized to determine volatile toluene in the gas <span class="hlt">phase</span> and allows potentiometric detection down to parts per million levels. This work demonstrates the possibility of developing a general sensing principle for detection of neutral <span class="hlt">vapors</span> using ISEs. PMID:26215887</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1413902-improving-biomass-pyrolysis-economics-integrating-vapor-liquid-phase-upgrading','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1413902-improving-biomass-pyrolysis-economics-integrating-vapor-liquid-phase-upgrading"><span>Improving biomass pyrolysis economics by integrating <span class="hlt">vapor</span> and liquid <span class="hlt">phase</span> upgrading</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Iisa, Kristiina; Robichaud, David J.; Watson, Michael J.; ...</p> <p>2017-11-24</p> <p>Partial deoxygenation of bio-oil by catalytic fast pyrolysis with subsequent coupling and hydrotreating can lead to improved economics and will aid commercial deployment of pyrolytic conversion of biomass technologies. Biomass pyrolysis efficiently depolymerizes and deconstructs solid plant matter into carbonaceous molecules that, upon catalytic upgrading, can be used for fuels and chemicals. Upgrading strategies include catalytic deoxygenation of the <span class="hlt">vapors</span> before they are condensed (in situ and ex situ catalytic fast pyrolysis), or hydrotreating following condensation of the bio-oil. In general, deoxygenation carbon efficiencies, one of the most important cost drivers, are typically higher for hydrotreating when compared to catalyticmore » fast pyrolysis alone. However, using catalytic fast pyrolysis as the primary conversion step can benefit the entire process chain by: (1) reducing the reactivity of the bio-oil, thereby mitigating issues with aging and transport and eliminating need for multi-stage hydroprocessing configurations; (2) producing a bio-oil that can be fractionated through distillation, which could lead to more efficient use of hydrogen during hydrotreating and facilitate integration in existing petroleum refineries; and (3) allowing for the <span class="hlt">separation</span> of the aqueous <span class="hlt">phase</span>. In this perspective, we investigate in detail a combination of these approaches, where some oxygen is removed during catalytic fast pyrolysis and the remainder removed by downstream hydrotreating, accompanied by carbon–carbon coupling reactions in either the <span class="hlt">vapor</span> or liquid <span class="hlt">phase</span> to maximize carbon efficiency toward value-driven products (e.g. fuels or chemicals). The economic impact of partial deoxygenation by catalytic fast pyrolysis will be explored in the context of an integrated two-stage process. In conclusion, improving the overall pyrolysis-based biorefinery economics by inclusion of production of high-value co-products will be examined.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1413902-improving-biomass-pyrolysis-economics-integrating-vapor-liquid-phase-upgrading','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1413902-improving-biomass-pyrolysis-economics-integrating-vapor-liquid-phase-upgrading"><span>Improving biomass pyrolysis economics by integrating <span class="hlt">vapor</span> and liquid <span class="hlt">phase</span> upgrading</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Iisa, Kristiina; Robichaud, David J.; Watson, Michael J.</p> <p></p> <p>Partial deoxygenation of bio-oil by catalytic fast pyrolysis with subsequent coupling and hydrotreating can lead to improved economics and will aid commercial deployment of pyrolytic conversion of biomass technologies. Biomass pyrolysis efficiently depolymerizes and deconstructs solid plant matter into carbonaceous molecules that, upon catalytic upgrading, can be used for fuels and chemicals. Upgrading strategies include catalytic deoxygenation of the <span class="hlt">vapors</span> before they are condensed (in situ and ex situ catalytic fast pyrolysis), or hydrotreating following condensation of the bio-oil. In general, deoxygenation carbon efficiencies, one of the most important cost drivers, are typically higher for hydrotreating when compared to catalyticmore » fast pyrolysis alone. However, using catalytic fast pyrolysis as the primary conversion step can benefit the entire process chain by: (1) reducing the reactivity of the bio-oil, thereby mitigating issues with aging and transport and eliminating need for multi-stage hydroprocessing configurations; (2) producing a bio-oil that can be fractionated through distillation, which could lead to more efficient use of hydrogen during hydrotreating and facilitate integration in existing petroleum refineries; and (3) allowing for the <span class="hlt">separation</span> of the aqueous <span class="hlt">phase</span>. In this perspective, we investigate in detail a combination of these approaches, where some oxygen is removed during catalytic fast pyrolysis and the remainder removed by downstream hydrotreating, accompanied by carbon–carbon coupling reactions in either the <span class="hlt">vapor</span> or liquid <span class="hlt">phase</span> to maximize carbon efficiency toward value-driven products (e.g. fuels or chemicals). The economic impact of partial deoxygenation by catalytic fast pyrolysis will be explored in the context of an integrated two-stage process. In conclusion, improving the overall pyrolysis-based biorefinery economics by inclusion of production of high-value co-products will be examined.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23586696','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23586696"><span>Additive-free size-controlled synthesis of gold square nanoplates using photochemical reaction in dynamic <span class="hlt">phase-separating</span> media.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kajimoto, Shinji; Shirasawa, Daisuke; Horimoto, Noriko Nishizawa; Fukumura, Hiroshi</p> <p>2013-05-14</p> <p>Ultrafast <span class="hlt">phase</span> <span class="hlt">separation</span> of water and 2-butoxyethanol mixture was <span class="hlt">induced</span> by nanosecond IR laser pulse irradiation. After a certain delay time, a UV laser pulse was introduced to <span class="hlt">induce</span> photoreduction of aurate ions, which led to the formation of gold nanoparticles in dynamic <span class="hlt">phase-separating</span> media. The structure and size of the nanoparticles varied depending on the delay time between the IR and UV pulses. For a delay time of 5 and 6 μs, gold square plates having edge lengths of 150 and 100 nm were selectively obtained, respectively. With a delay time of 3 μs, on the other hand, the size of the square plates varied widely from 100 nm to a few micrometers. The size of the gold square plates was also varied by varying the total irradiation time of the IR and UV pulses. The size distribution of the square plates obtained under different conditions suggests that the growth process of the square plates was affected by the size of the nanophases during <span class="hlt">phase</span> <span class="hlt">separation</span>. Electron diffraction patterns of the synthesized square plates showed that the square plates were highly crystalline with a Au(100) surface. These results showed that the nanophases formed during laser-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> can provide detergent-free reaction fields for size-controlled nanomaterial synthesis.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JChPh.146x4504V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JChPh.146x4504V"><span>Predicting <span class="hlt">vapor</span>-liquid <span class="hlt">phase</span> equilibria with augmented ab initio interatomic potentials</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Vlasiuk, Maryna; Sadus, Richard J.</p> <p>2017-06-01</p> <p>The ability of ab initio interatomic potentials to accurately predict <span class="hlt">vapor</span>-liquid <span class="hlt">phase</span> equilibria is investigated. Monte Carlo simulations are reported for the <span class="hlt">vapor</span>-liquid equilibria of argon and krypton using recently developed accurate ab initio interatomic potentials. Seventeen interatomic potentials are studied, formulated from different combinations of two-body plus three-body terms. The simulation results are compared to either experimental or reference data for conditions ranging from the triple point to the critical point. It is demonstrated that the use of ab initio potentials enables systematic improvements to the accuracy of predictions via the addition of theoretically based terms. The contribution of three-body interactions is accounted for using the Axilrod-Teller-Muto plus other multipole contributions and the effective Marcelli-Wang-Sadus potentials. The results indicate that the predictive ability of recent interatomic potentials, obtained from quantum chemical calculations, is comparable to that of accurate empirical models. It is demonstrated that the Marcelli-Wang-Sadus potential can be used in combination with accurate two-body ab initio models for the computationally inexpensive and accurate estimation of <span class="hlt">vapor</span>-liquid <span class="hlt">phase</span> equilibria.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28668034','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28668034"><span>Predicting <span class="hlt">vapor</span>-liquid <span class="hlt">phase</span> equilibria with augmented ab initio interatomic potentials.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Vlasiuk, Maryna; Sadus, Richard J</p> <p>2017-06-28</p> <p>The ability of ab initio interatomic potentials to accurately predict <span class="hlt">vapor</span>-liquid <span class="hlt">phase</span> equilibria is investigated. Monte Carlo simulations are reported for the <span class="hlt">vapor</span>-liquid equilibria of argon and krypton using recently developed accurate ab initio interatomic potentials. Seventeen interatomic potentials are studied, formulated from different combinations of two-body plus three-body terms. The simulation results are compared to either experimental or reference data for conditions ranging from the triple point to the critical point. It is demonstrated that the use of ab initio potentials enables systematic improvements to the accuracy of predictions via the addition of theoretically based terms. The contribution of three-body interactions is accounted for using the Axilrod-Teller-Muto plus other multipole contributions and the effective Marcelli-Wang-Sadus potentials. The results indicate that the predictive ability of recent interatomic potentials, obtained from quantum chemical calculations, is comparable to that of accurate empirical models. It is demonstrated that the Marcelli-Wang-Sadus potential can be used in combination with accurate two-body ab initio models for the computationally inexpensive and accurate estimation of <span class="hlt">vapor</span>-liquid <span class="hlt">phase</span> equilibria.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20377200','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20377200"><span>Performance enhancement of hybrid solar cells through chemical <span class="hlt">vapor</span> annealing.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wu, Yue; Zhang, Genqiang</p> <p>2010-05-12</p> <p>Improvement in power conversion efficiency has been observed in cadmium selenide nanorods/poly(3-hexylthiophene) hybrid solar cells through benzene-1,3-dithiol chemical <span class="hlt">vapor</span> annealing. Phosphor NMR studies of the nanorods and TEM/AFM characterizations of the morphology of the blended film showed that the ligand exchange reaction and related <span class="hlt">phase</span> <span class="hlt">separation</span> happening during the chemical <span class="hlt">vapor</span> annealing are responsible for the performance enhancement.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19910068816&hterms=educacion&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Deducacion','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19910068816&hterms=educacion&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Deducacion"><span>Influence of mass diffusion on the stability of thermophoretic growth of a solid from the <span class="hlt">vapor</span> <span class="hlt">phase</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Castillo, J. L.; Garcia-Ybarra, P. L.; Rosner, D. E.</p> <p>1991-01-01</p> <p>The stability of solid planar growth from a binary <span class="hlt">vapor</span> <span class="hlt">phase</span> with a condensing species dilute in a carrier gas is examined when the ratio of depositing to carrier species molecular mass is large and the main diffusive transport mechanism is thermal diffusion. It is shown that a deformation of the solid-gas interface <span class="hlt">induces</span> a deformation of the gas <span class="hlt">phase</span> isotherms that increases the thermal gradients and thereby the local mass deposition rate at the crests and reduces them at the valleys. The initial surface deformation is enhanced by the modified deposition rates in the absence of appreciable Fick/Brownian diffusion and interfacial energy effects.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27846698','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27846698"><span>Evaluation of effects of pH and ionic strength on colloidal stability of IgG solutions by PEG-<span class="hlt">induced</span> liquid-liquid <span class="hlt">phase</span> <span class="hlt">separation</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Thompson, Ronald W; Latypov, Ramil F; Wang, Ying; Lomakin, Aleksey; Meyer, Julie A; Vunnum, Suresh; Benedek, George B</p> <p>2016-11-14</p> <p>Colloidal stability of IgG antibody solutions is important for pharmaceutical and medicinal applications. Solution pH and ionic strength are two key factors that affect the colloidal stability of protein solutions. In this work, we use a method based on the PEG-<span class="hlt">induced</span> liquid-liquid <span class="hlt">phase</span> <span class="hlt">separation</span> to examine the effects of pH and ionic strength on the colloidal stability of IgG solutions. We found that at high ionic strength (≥0.25M), the colloidal stability of most of our IgGs is insensitive to pH, and at low ionic strength (≤0.15M), all IgG solutions are much more stable at pH 5 than at pH 7. In addition, the PEG-<span class="hlt">induced</span> depletion force is less efficient in causing <span class="hlt">phase</span> <span class="hlt">separation</span> at pH 5 than at pH 7. In contrast to the native inter-protein interaction of IgGs, the effect of depletion force on <span class="hlt">phase</span> <span class="hlt">separation</span> of the antibody solutions is insensitive to ionic strength. Our results suggest that the long-range electrostatic inter-protein repulsion at low ionic strength stabilizes the IgG solutions at low pH. At high ionic strength, the short-range electrostatic interactions do not make a significant contribution to the colloidal stability for most IgGs with a few exceptions. The weaker effect of depletion force at lower pH indicates a reduction of protein concentration in the condensed <span class="hlt">phase</span>. This work advances our basic understanding of the colloidal stability of IgG solutions and also introduces a practical approach to measuring protein colloidal stability under various solution conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19860049534&hterms=binary+search&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dbinary%2Bsearch','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19860049534&hterms=binary+search&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dbinary%2Bsearch"><span>A search for the prewetting line. [in binary liquid system at <span class="hlt">vapor</span>-liquid interface</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Schmidt, J. W.; Moldover, M. R.</p> <p>1986-01-01</p> <p>This paper describes efforts to locate the prewetting line in a binary liquid system (isopropanol-perfluoromethylcyclohexane) at the <span class="hlt">vapor</span>-liquid interface. Tight upper bounds were placed on the temperature <span class="hlt">separation</span> (0.2 K) between the prewetting line and the line of bulk liquid <span class="hlt">phase</span> <span class="hlt">separation</span>. The prewetting line in systems at equilibrium was not detected. Experimental signatures indicative of the prewetting line occurred only in nonequilibrium situations. Several theories predict that the adsorption of one of the components (the fluorocarbon, in this case) at the liquid-<span class="hlt">vapor</span> interface should increase abruptly, at a temperature sightly above the temperature at which the mixture <span class="hlt">separates</span> into two liquid <span class="hlt">phases</span>. A regular solution calculation indicates that this prewetting line should have been easily detectable with the instruments used in this experiment. Significant features of the experiment are: (1) low-gradient thermostatting, (2) in situ stirring, (3) precision ellipsometry from the <span class="hlt">vapor</span>-liquid interface, (4) high resolution differential index of refraction measurements using a novel cell design, and (5) computer control.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21164923','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21164923"><span>Atomic <span class="hlt">vapor</span> quantum memory for a photonic polarization qubit.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cho, Young-Wook; Kim, Yoon-Ho</p> <p>2010-12-06</p> <p>We report an experimental realization of an atomic <span class="hlt">vapor</span> quantum memory for the photonic polarization qubit. The performance of the quantum memory for the polarization qubit, realized with electromagnetically-<span class="hlt">induced</span> transparency in two spatially <span class="hlt">separated</span> ensembles of warm Rubidium atoms in a single <span class="hlt">vapor</span> cell, has been characterized with quantum process tomography. The process fidelity better than 0.91 for up to 16 μs of storage time has been achieved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28140362','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28140362"><span>Self-catalyzed GaAs nanowires on silicon by hydride <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Dong, Zhenning; André, Yamina; Dubrovskii, Vladimir G; Bougerol, Catherine; Leroux, Christine; Ramdani, Mohammed R; Monier, Guillaume; Trassoudaine, Agnès; Castelluci, Dominique; Gil, Evelyne</p> <p>2017-03-24</p> <p>Gold-free GaAs nanowires on silicon substrates can pave the way for monolithic integration of photonic nanodevices with silicon electronic platforms. It is extensively documented that the self-catalyzed approach works well in molecular beam epitaxy but is much more difficult to implement in <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxies. Here, we report the first gallium-catalyzed hydride <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy growth of long (more than 10 μm) GaAs nanowires on Si(111) substrates with a high integrated growth rate up to 60 μm h -1 and pure zincblende crystal structure. The growth is achieved by combining a low temperature of 600 °C with high gaseous GaCl/As flow ratios to enable dechlorination and formation of gallium droplets. GaAs nanowires exhibit an interesting bottle-like shape with strongly tapered bases, followed by straight tops with radii as small as 5 nm. We present a model that explains the peculiar growth mechanism in which the gallium droplets nucleate and rapidly swell on the silicon surface but then are gradually consumed to reach a stationary size. Our results unravel the necessary conditions for obtaining gallium-catalyzed GaAs nanowires by <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy techniques.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3894580','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3894580"><span>Comparative Study of Solution <span class="hlt">Phase</span> and <span class="hlt">Vapor</span> <span class="hlt">Phase</span> Deposition of Aminosilanes on Silicon Dioxide Surfaces</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Yadav, Amrita R.; Sriram, Rashmi; Carter, Jared A.; Miller, Benjamin L.</p> <p>2014-01-01</p> <p>The uniformity of aminosilane layers typically used for the modification of hydroxyl bearing surfaces such as silicon dioxide is critical for a wide variety of applications, including biosensors. However, in spite of many studies that have been undertaken on surface silanization, there remains a paucity of easy-to-implement deposition methods reproducibly yielding smooth aminosilane monolayers. In this study, solution- and <span class="hlt">vapor-phase</span> deposition methods for three aminoalkoxysilanes differing in the number of reactive groups (3-aminopropyl triethoxysilane (APTES), 3-aminopropyl methyl diethoxysilane (APMDES) and 3-aminopropyl dimethyl ethoxysilane (APDMES)) were assessed with the aim of identifying methods that yield highly uniform and reproducible silane layers that are resistant to minor procedural variations. Silane film quality was characterized based on measured thickness, hydrophilicity and surface roughness. Additionally, hydrolytic stability of the films was assessed via these thickness and contact angle values following desorption in water. We found that two simple solution-<span class="hlt">phase</span> methods, an aqueous deposition of APTES and a toluene based deposition of APDMES, yielded high quality silane layers that exhibit comparable characteristics to those deposited via <span class="hlt">vapor-phase</span> methods. PMID:24411379</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvE..97d2609B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvE..97d2609B"><span><span class="hlt">Phase</span> <span class="hlt">separation</span> of self-propelled ballistic particles</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bruss, Isaac R.; Glotzer, Sharon C.</p> <p>2018-04-01</p> <p>Self-propelled particles <span class="hlt">phase-separate</span> into coexisting dense and dilute regions above a critical density. The statistical nature of their stochastic motion lends itself to various theories that predict the onset of <span class="hlt">phase</span> <span class="hlt">separation</span>. However, these theories are ill-equipped to describe such behavior when noise becomes negligible. To overcome this limitation, we present a predictive model that relies on two density-dependent timescales: τF, the mean time particles spend between collisions; and τC, the mean lifetime of a collision. We show that only when τF<τC do collisions last long enough to develop a growing cluster and initiate <span class="hlt">phase</span> <span class="hlt">separation</span>. Using both analytical calculations and active particle simulations, we measure these timescales and determine the critical density for <span class="hlt">phase</span> <span class="hlt">separation</span> in both two and three dimensions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29588464','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29588464"><span>Rydberg interaction <span class="hlt">induced</span> enhanced excitation in thermal atomic <span class="hlt">vapor</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kara, Dushmanta; Bhowmick, Arup; Mohapatra, Ashok K</p> <p>2018-03-27</p> <p>We present the experimental demonstration of interaction <span class="hlt">induced</span> enhancement in Rydberg excitation or Rydberg anti-blockade in thermal atomic <span class="hlt">vapor</span>. We have used optical heterodyne detection technique to measure Rydberg population due to two-photon excitation to the Rydberg state. The anti-blockade peak which doesn't satisfy the two-photon resonant condition is observed along with the usual two-photon resonant peak which can't be explained using the model with non-interacting three-level atomic system. A model involving two interacting atoms is formulated for thermal atomic <span class="hlt">vapor</span> using the dressed states of three-level atomic system to explain the experimental observations. A non-linear dependence of <span class="hlt">vapor</span> density is observed for the anti-blockade peak which also increases with increase in principal quantum number of the Rydberg state. A good agreement is found between the experimental observations and the proposed interacting model. Our result implies possible applications towards quantum logic gates using Rydberg anti-blockade in thermal atomic <span class="hlt">vapor</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..DFD.E7001P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..DFD.E7001P"><span>The influence of liquid/<span class="hlt">vapor</span> <span class="hlt">phase</span> change onto the Nusselt number</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Popescu, Elena-Roxana; Colin, Catherine; Tanguy, Sebastien</p> <p>2017-11-01</p> <p>In spite of its significant interest in various fields, there is currently a very few information on how an external flow will modify the evaporation or the condensation of a liquid surface. Although most applications involve turbulent flows, the simpler configuration where a laminar superheated or subcooled <span class="hlt">vapor</span> flow is shearing a saturated liquid interface has still never been solved. Based on a numerical approach, we propose to characterize the interaction between a laminar boundary layer of a superheated or subcooled <span class="hlt">vapor</span> flow and a static liquid pool at saturation temperature. By performing a full set of simulations sweeping the parameters space, correlations are proposed for the first time on the Nusselt number depending on the dimensionless numbers that characterize both <span class="hlt">vaporization</span> and condensation. As attended, the Nusselt number decreases or increases in the configurations involving respectively <span class="hlt">vaporization</span> or condensation. More unexpected is the behaviour of the friction of the <span class="hlt">vapor</span> flow on the liquid pool, for which we report that it is weakly affected by the <span class="hlt">phase</span> change, despite the important variation of the local flow structure due to evaporation or condensation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20233090','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20233090"><span>Systemic molecular and cellular changes <span class="hlt">induced</span> in rats upon inhalation of JP-8 petroleum fuel <span class="hlt">vapor</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hanas, Jay S; Bruce Briggs, G; Lerner, Megan R; Lightfoot, Stan A; Larabee, Jason L; Karsies, Todd J; Epstein, Robert B; Hanas, Rushie J; Brackett, Daniel J; Hocker, James R</p> <p>2010-05-01</p> <p>Limited information is available regarding systemic changes in mammals associated with exposures to petroleum/hydrocarbon fuels. In this study, systemic toxicity of JP-8 jet fuel was observed in a rat inhalation model at different JP-8 fuel <span class="hlt">vapor</span> concentrations (250, 500, or 1000 mg/m(3), for 91 days). Gel electrophoresis and mass spectrometry sequencing identified the alpha-2 microglobulin protein to be elevated in rat kidney in a JP-8 dose-dependent manner. Western blot analysis of kidney and lung tissue extracts revealed JP-8 dependent elevation of <span class="hlt">inducible</span> heat shock protein 70 (HSP70). Tissue changes were observed histologically (hematoxylin and eosin staining) in liver, kidney, lung, bone marrow, and heart, and more prevalently at medium or high JP-8 <span class="hlt">vapor</span> <span class="hlt">phase</span> exposures (500-1000 mg/m(3)) than at low <span class="hlt">vapor</span> <span class="hlt">phase</span> exposure (250 mg/m(3)) or non-JP-8 controls. JP-8 fuel-<span class="hlt">induced</span> liver alterations included dilated sinusoids, cytoplasmic clumping, and fat cell deposition. Changes to the kidneys included reduced numbers of nuclei, and cytoplasmic dumping in the lumen of proximal convoluted tubules. JP-8 dependent lung alterations were edema and dilated alveolar capillaries, which allowed clumping of red blood cells (RBCs). Changes in the bone marrow in response to JP-8 included reduction of fat cells and fat globules, and cellular proliferation (RBCs, white blood cells-WBCs, and megakaryocytes). Heart tissue from JP-8 exposed animals contained increased numbers of inflammatory and fibroblast cells, as well as myofibril scarring. cDNA array analysis of heart tissue revealed a JP-8 dependent increase in atrial natriuretic peptide precursor mRNA and a decrease in voltage-gated potassium (K+) ion channel mRNA.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017OptLT..89...75T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017OptLT..89...75T"><span>Carrier-<span class="hlt">separating</span> demodulation of <span class="hlt">phase</span> shifting self-mixing interferometry</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tao, Yufeng; Wang, Ming; Xia, Wei</p> <p>2017-03-01</p> <p>A carrier <span class="hlt">separating</span> method associated with noise-elimination had been introduced into a sinusoidal <span class="hlt">phase</span>-shifting self-mixing interferometer. The conventional sinusoidal <span class="hlt">phase</span> shifting self-mixing interferometry was developed into a more competitive instrument with high computing efficiency and nanometer accuracy of λ / 100 in dynamical vibration measurement. The high slew rate electro-optic modulator <span class="hlt">induced</span> a sinusoidal <span class="hlt">phase</span> carrier with ultralow insertion loss in this paper. In order to extract <span class="hlt">phase</span>-shift quickly and precisely, this paper employed the carrier-<span class="hlt">separating</span> to directly generate quadrature signals without complicated frequency domain transforms. Moreover, most noises were evaluated and suppressed by a noise-elimination technology synthesizing empirical mode decomposition with wavelet transform. The overall laser system was described and inherent advantages such as high computational efficiency and decreased nonlinear errors of the established system were demonstrated. The experiment implemented on a high precision PZT (positioning accuracy was better than 1 nm) and compared with laser Doppler velocity meter. The good agreement of two instruments shown that the short-term resolution had improved from 10 nm to 1.5 nm in dynamic vibration measurement with reduced time expense. This was useful in precision measurement to improve the SMI with same sampling rate. The proposed signal processing was performed in pure time-domain requiring no preprocessing electronic circuits.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20538708','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20538708"><span>Vitamins A and E reverse gasoline <span class="hlt">vapors-induced</span> hematotoxicity and weight loss in female rats.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Uboh, F E; Eteng, M U; Ebong, P E; Umoh, I B</p> <p>2010-10-01</p> <p>In this study, gasoline <span class="hlt">vapors-induced</span> hematotoxicity, growth-depression and weight-loss reversal effect of vitamins A (retinol) and E (α-tocopherol) was assessed in female Wistar albino rats. The rats were exposed to gasoline <span class="hlt">vapors</span> (17.8 ± 2.6 cm(3)/h/m(3)/day), 6 hours/day, 6 days/week, for 20 weeks. Vitamins A and E at prophylactic dosage (400 and 200 IU/kg/day, respectively) were orally administered to the rats, <span class="hlt">separately</span>, in the last 2 weeks of exposure. The levels of hemoglobin (Hb), hematocrit or packed cell volume (PCV), red blood cells (RBC), growth rate and weight gain in the rats exposed to the <span class="hlt">vapors</span> were significantly lower (p < 0.05) compared, respectively, to the levels obtained for control rats. On the other hand, the levels of white blood cells (WBCs) in the test rats were significantly higher (p < 0.05) compared, respectively, with the level obtained for female control rats. These observations indicated that exposure to gasoline <span class="hlt">vapors</span> may cause hematotoxicity, growth depression and weight loss in female rats. However, administration of vitamins A and E was observed to produce a significant recovery (p < 0.05) in hematotoxicity, growth depression and weight loss observed to be associated with exposure to gasoline <span class="hlt">vapors</span>, although the rats administered with vitamin E were noted to respond more favorably than those administered with vitamin A. This suggests that although retinol and α-tocopherol may be used to reverse or prevent hematotoxicity, growth depression and weight loss in subjects exposed to gasoline <span class="hlt">vapors</span>, the reversal potency of α-tocopherol is higher than that of retinol.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5600379','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5600379"><span>Cooling <span class="hlt">induces</span> <span class="hlt">phase</span> <span class="hlt">separation</span> in membranes derived from isolated CNS myelin</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Pusterla, Julio M.; Schneck, Emanuel; Funari, Sérgio S.; Démé, Bruno; Tanaka, Motomu</p> <p>2017-01-01</p> <p>Purified myelin membranes (PMMs) are the starting material for biochemical analyses such as the isolation of detergent-insoluble glycosphingolipid-rich domains (DIGs), which are believed to be representatives of functional lipid rafts. The normal DIGs isolation protocol involves the extraction of lipids under moderate cooling. Here, we thus address the influence of cooling on the structure of PMMs and its sub-fractions. Thermodynamic and structural aspects of periodic, multilamellar PMMs are examined between 4°C and 45°C and in various biologically relevant aqueous solutions. The <span class="hlt">phase</span> behavior is investigated by small-angle X-ray scattering (SAXS) and differential scanning calorimetry (DSC). Complementary neutron diffraction (ND) experiments with solid-supported myelin multilayers confirm that the <span class="hlt">phase</span> behavior is unaffected by planar confinement. SAXS and ND consistently show that multilamellar PMMs in pure water become heterogeneous when cooled by more than 10–15°C below physiological temperature, as during the DIGs isolation procedure. The heterogeneous state of PMMs is stabilized in physiological solution, where <span class="hlt">phase</span> coexistence persists up to near the physiological temperature. This result supports the general view that membranes under physiological conditions are close to critical points for <span class="hlt">phase</span> <span class="hlt">separation</span>. In presence of elevated Ca2+ concentrations (> 10 mM), <span class="hlt">phase</span> coexistence is found even far above physiological temperatures. The relative fractions of the two <span class="hlt">phases</span>, and thus presumably also their compositions, are found to vary with temperature. Depending on the conditions, an “expanded” <span class="hlt">phase</span> with larger lamellar period or a “compacted” <span class="hlt">phase</span> with smaller lamellar period coexists with the native <span class="hlt">phase</span>. Both expanded and compacted periods are also observed in DIGs under the respective conditions. The observed subtle temperature-dependence of the <span class="hlt">phase</span> behavior of PMMs suggests that the composition of DIGs is sensitive to the details of</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_8");'>8</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li class="active"><span>10</span></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_10 --> <div id="page_11" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li class="active"><span>11</span></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="201"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012APS..MARA40002B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012APS..MARA40002B"><span><span class="hlt">Phase</span> <span class="hlt">Separation</span> in Solutions of Monoclonal Antibodies</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Benedek, George; Wang, Ying; Lomakin, Aleksey; Latypov, Ramil</p> <p>2012-02-01</p> <p>We report the observation of liquid-liquid <span class="hlt">phase</span> <span class="hlt">separation</span> (LLPS) in a solution of humanized monoclonal antibodies, IgG2, and the effects of human serum albumin, a major blood protein, on this <span class="hlt">phase</span> <span class="hlt">separation</span>. We find a significant reduction of <span class="hlt">phase</span> <span class="hlt">separation</span> temperature in the presence of albumin, and a preferential partitioning of the albumin into the antibody-rich <span class="hlt">phase</span>. We provide a general thermodynamic analysis of the antibody-albumin mixture <span class="hlt">phase</span> diagram and relate its features to the magnitude of the effective inter-protein interactions. Our analysis suggests that additives (HSA in this report), which have moderate attraction with antibody molecules, may be used to forestall undesirable protein condensation in antibody solutions. Our findings are relevant to understanding the stability of pharmaceutical solutions of antibodies and the mechanisms of cryoglobulinemia.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20090014794','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20090014794"><span>Binary Colloidal Alloy Test-5: <span class="hlt">Phase</span> <span class="hlt">Separation</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Lynch, Matthew; Weitz, David A.; Lu, Peter J.</p> <p>2008-01-01</p> <p>The Binary Colloidal Alloy Test - 5: <span class="hlt">Phase</span> <span class="hlt">Separation</span> (BCAT-5-<span class="hlt">Phase</span>Sep) experiment will photograph initially randomized colloidal samples onboard the ISS to determine their resulting structure over time. This allows the scientists to capture the kinetics (evolution) of their samples, as well as the final equilibrium state of each sample. BCAT-5-<span class="hlt">Phase</span>Sep studies collapse (<span class="hlt">phase</span> <span class="hlt">separation</span> rates that impact product shelf-life); in microgravity the physics of collapse is not masked by being reduced to a simple top and bottom <span class="hlt">phase</span> as it is on Earth.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19860000779&hterms=dextran&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Ddextran','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19860000779&hterms=dextran&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Ddextran"><span><span class="hlt">Separation</span> of aqueous two-<span class="hlt">phase</span> polymer systems in microgravity</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Vanalstine, J. M.; Harris, J. M.; Synder, S.; Curreri, P. A.; Bamberger, S. B.; Brooks, D. E.</p> <p>1984-01-01</p> <p><span class="hlt">Phase</span> <span class="hlt">separation</span> of polymer systems in microgravity is studied in aircraft flights to prepare shuttle experiments. Short duration (20 sec) experiments demonstrate that <span class="hlt">phase</span> <span class="hlt">separation</span> proceeds rapidly in low gravity despite appreciable <span class="hlt">phase</span> viscosities and low liquid interfacial tensions (i.e., 50 cP, 10 micro N/m). Ostwald ripening does not appear to be a satisfactory model for the <span class="hlt">phase</span> <span class="hlt">separation</span> mechanism. Polymer coated surfaces are evaluated as a means to localize <span class="hlt">phases</span> <span class="hlt">separated</span> in low gravity. Contact angle measurements demonstrate that covalently coupling dextran or PEG to glass drastically alters the 1-g wall wetting behavior of the <span class="hlt">phases</span> in dextran-PEG two <span class="hlt">phase</span> systems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/3838252','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/3838252"><span>Ca2+-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> in black lipid membranes and its effect on the transport of a hydrophobic ion.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Miller, A; Schmidt, G; Eibl, H; Knoll, W</p> <p>1985-03-14</p> <p>Voltage jump-current relaxation studies have been performed with dipicrylamine-doped black membranes of binary lipid mixtures. As in the case of the carrier-mediated ion transport (Schmidt, G., Eibl, H. and Knoll, W. (1982) J. Membrane Biol. 70, 147-155) no evidence was found that the neutral lipid phosphatidylcholine (DPMPC) and the charged phosphatidic acid (DPMPA) are heterogeneously distributed in the membrane over the whole range of composition. However, besides a continuous dilution of the surface charges of DPMPA by the addition of DPMPC molecules, different structural properties of mixed membranes influence the kinetics of the dipicrylamine transport. The addition of Ca2+ to the electrolyte <span class="hlt">induces</span> a lipid <span class="hlt">phase</span> <span class="hlt">separation</span> within the membrane into two fluid <span class="hlt">phases</span> of distinctly different characteristics of the translocation of hydrophobic ions. Thus, it is possible to determine a preliminary composition <span class="hlt">phase</span> diagram for the DPMPA/DPMPC mixtures as a function of the Ca2+ concentration.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24411379','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24411379"><span>Comparative study of solution-<span class="hlt">phase</span> and <span class="hlt">vapor-phase</span> deposition of aminosilanes on silicon dioxide surfaces.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yadav, Amrita R; Sriram, Rashmi; Carter, Jared A; Miller, Benjamin L</p> <p>2014-02-01</p> <p>The uniformity of aminosilane layers typically used for the modification of hydroxyl bearing surfaces such as silicon dioxide is critical for a wide variety of applications, including biosensors. However, in spite of many studies that have been undertaken on surface silanization, there remains a paucity of easy-to-implement deposition methods reproducibly yielding smooth aminosilane monolayers. In this study, solution- and <span class="hlt">vapor-phase</span> deposition methods for three aminoalkoxysilanes differing in the number of reactive groups (3-aminopropyl triethoxysilane (APTES), 3-aminopropyl methyl diethoxysilane (APMDES) and 3-aminopropyl dimethyl ethoxysilane (APDMES)) were assessed with the aim of identifying methods that yield highly uniform and reproducible silane layers that are resistant to minor procedural variations. Silane film quality was characterized based on measured thickness, hydrophilicity and surface roughness. Additionally, hydrolytic stability of the films was assessed via these thickness and contact angle values following desorption in water. We found that two simple solution-<span class="hlt">phase</span> methods, an aqueous deposition of APTES and a toluene based deposition of APDMES, yielded high quality silane layers that exhibit comparable characteristics to those deposited via <span class="hlt">vapor-phase</span> methods. Copyright © 2013 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26831764','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26831764"><span>Selective Template Wetting Routes to Hierarchical Polymer Films: Polymer Nanotubes from <span class="hlt">Phase-Separated</span> Films via Solvent Annealing.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ko, Hao-Wen; Cheng, Ming-Hsiang; Chi, Mu-Huan; Chang, Chun-Wei; Chen, Jiun-Tai</p> <p>2016-03-01</p> <p>We demonstrate a novel wetting method to prepare hierarchical polymer films with polymer nanotubes on selective regions. This strategy is based on the selective wetting abilities of polymer chains, annealed in different solvent <span class="hlt">vapors</span>, into the nanopores of porous templates. <span class="hlt">Phase-separated</span> films of polystyrene (PS) and poly(methyl methacrylate) (PMMA), two commonly used polymers, are prepared as a model system. After anodic aluminum oxide (AAO) templates are placed on the films, the samples are annealed in <span class="hlt">vapors</span> of acetic acid, in which the PMMA chains are swollen and wet the nanopores of the AAO templates selectively. As a result, hierarchical polymer films containing PMMA nanotubes can be obtained after the AAO templates are removed. The distribution of the PMMA nanotubes of the hierarchical polymer films can also be controlled by changing the compositions of the polymer blends. This work not only presents a novel method to fabricate hierarchical polymer films with polymer nanotubes on selective regions, but also gives a deeper understanding in the selective wetting ability of polymer chains in solvent <span class="hlt">vapors</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20000080605&hterms=ambient&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dambient','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20000080605&hterms=ambient&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dambient"><span>Spectroscopic Observation of Chemical Interaction Between Impact-<span class="hlt">induced</span> <span class="hlt">Vapor</span> Clouds and the Ambient Atmosphere</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Sugita, S.; Heineck, J. T.; Schultz, P. H.</p> <p>2000-01-01</p> <p>Chemical reactions within impact-<span class="hlt">induced</span> <span class="hlt">vapor</span> clouds were observed in laboratory experiments using a spectroscopic method. The results indicate that projectile-derived carbon-rich <span class="hlt">vapor</span> reacts intensively with atmospheric nitrogen.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017E%26ES...93a2072S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017E%26ES...93a2072S"><span><span class="hlt">Phase</span> <span class="hlt">separation</span> of bio-oil produced by co-pyrolysis of corn cobs and polypropylene</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Supramono, D.; Julianto; Haqqyana; Setiadi, H.; Nasikin, M.</p> <p>2017-11-01</p> <p>In co-pyrolysis of biomass-plastics, bio-oil produced contains both oxygenated and non-oxygenated compounds. High oxygen composition is responsible for instability and low heating value of bio-oil and high acid content for corrosiveness. Aims of the present work are to evaluate possibilities of achieving <span class="hlt">phase</span> <span class="hlt">separation</span> between oxygenated and non-oxygenated compounds in bio-oil using a proposed stirred tank reactor and to achieve synergistic effects on bio-oil yield and non-oxygenated compound layer yield. <span class="hlt">Separation</span> of bio-oil into two layers, i.e. that containing oxygenated compounds (polar <span class="hlt">phase</span>) and non-oxygenated compounds (non-polar <span class="hlt">phase</span>) is important to obtain pure non-polar <span class="hlt">phase</span> ready for the next processing of hydrogenation and used directly as bio-fuel. There has been no research work on co-pyrolysis of biomass-plastic considering possibility of <span class="hlt">phase</span> <span class="hlt">separation</span> of bio-oil. The present work is proposing a stirred tank reactor for co-pyrolysis with nitrogen injection, which is capable of tailoring co-pyrolysis conditions leading to low viscosity and viscosity asymmetry, which <span class="hlt">induce</span> <span class="hlt">phase</span> <span class="hlt">separation</span> between polar <span class="hlt">phase</span> and non-polar <span class="hlt">phase</span>. The proposed reactor is capable of generating synergistic effect on bio-oil and non-polar yields as the composition of PP in feed is more than 25% weight in which non-polar layers contain only alkanes, alkenes, cycloalkanes and cycloalkenes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015PhRvL.114o5501K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015PhRvL.114o5501K"><span>New Density Functional Approach for Solid-Liquid-<span class="hlt">Vapor</span> Transitions in Pure Materials</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kocher, Gabriel; Provatas, Nikolas</p> <p>2015-04-01</p> <p>A new <span class="hlt">phase</span> field crystal (PFC) type theory is presented, which accounts for the full spectrum of solid-liquid-<span class="hlt">vapor</span> <span class="hlt">phase</span> transitions within the framework of a single density order parameter. Its equilibrium properties show the most quantitative features to date in PFC modeling of pure substances, and full consistency with thermodynamics in pressure-volume-temperature space is demonstrated. A method to control either the volume or the pressure of the system is also introduced. Nonequilibrium simulations show that 2- and 3-<span class="hlt">phase</span> growth of solid, <span class="hlt">vapor</span>, and liquid can be achieved, while our formalism also allows for a full range of pressure-<span class="hlt">induced</span> transformations. This model opens up a new window for the study of pressure driven interactions of condensed <span class="hlt">phases</span> with <span class="hlt">vapor</span>, an experimentally relevant paradigm previously missing from <span class="hlt">phase</span> field crystal theories.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011JTST...20..736V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011JTST...20..736V"><span>Plasma Spray-PVD: A New Thermal Spray Process to Deposit Out of the <span class="hlt">Vapor</span> <span class="hlt">Phase</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>von Niessen, Konstantin; Gindrat, Malko</p> <p>2011-06-01</p> <p>Plasma spray-physical <span class="hlt">vapor</span> deposition (PS-PVD) is a low pressure plasma spray technology recently developed by Sulzer Metco AG (Switzerland). Even though it is a thermal spray process, it can deposit coatings out of the <span class="hlt">vapor</span> <span class="hlt">phase</span>. The basis of PS-PVD is the low pressure plasma spraying (LPPS) technology that has been well established in industry for several years. In comparison to conventional vacuum plasma spraying (VPS) or low pressure plasma spraying (LPPS), the new proposed process uses a high energy plasma gun operated at a reduced work pressure of 0.1 kPa (1 mbar). Owing to the high energy plasma and further reduced work pressure, PS-PVD is able to deposit a coating not only by melting the feed stock material which builds up a layer from liquid splats but also by <span class="hlt">vaporizing</span> the injected material. Therefore, the PS-PVD process fills the gap between the conventional physical <span class="hlt">vapor</span> deposition (PVD) technologies and standard thermal spray processes. The possibility to <span class="hlt">vaporize</span> feedstock material and to produce layers out of the <span class="hlt">vapor</span> <span class="hlt">phase</span> results in new and unique coating microstructures. The properties of such coatings are superior to those of thermal spray and electron beam-physical <span class="hlt">vapor</span> deposition (EB-PVD) coatings. In contrast to EB-PVD, PS-PVD incorporates the <span class="hlt">vaporized</span> coating material into a supersonic plasma plume. Owing to the forced gas stream of the plasma jet, complex shaped parts such as multi-airfoil turbine vanes can be coated with columnar thermal barrier coatings using PS-PVD. Even shadowed areas and areas which are not in the line of sight of the coating source can be coated homogeneously. This article reports on the progress made by Sulzer Metco in developing a thermal spray process to produce coatings out of the <span class="hlt">vapor</span> <span class="hlt">phase</span>. Columnar thermal barrier coatings made of Yttria-stabilized Zircona (YSZ) are optimized to serve in a turbine engine. This process includes not only preferable coating properties such as strain tolerance and erosion</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18222437','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18222437"><span>Comparison of cryopreserved human sperm in <span class="hlt">vapor</span> and liquid <span class="hlt">phases</span> of liquid nitrogen: effect on motility parameters, morphology, and sperm function.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Punyatanasakchai, Piyaphan; Sophonsritsuk, Areephan; Weerakiet, Sawaek; Wansumrit, Surapee; Chompurat, Deonthip</p> <p>2008-11-01</p> <p>To compare the effects of cryopreserved sperm in <span class="hlt">vapor</span> and liquid <span class="hlt">phases</span> of liquid nitrogen on sperm motility, morphology, and sperm function. Experimental study. Andrology laboratory at Ramathibodi Hospital, Thailand. Thirty-eight semen samples with normal motility and sperm count were collected from 38 men who were either patients of an infertility clinic or had donated sperm for research. Each semen sample was divided into two aliquots. Samples were frozen with static-<span class="hlt">phase</span> <span class="hlt">vapor</span> cooling. One aliquot was plunged into liquid nitrogen (-196 degrees C), and the other was stored in <span class="hlt">vapor-phase</span> nitrogen (-179 degrees C) for 3 days. Thawing was performed at room temperature. Motility was determined by using computer-assisted semen analysis, sperm morphology was determined by using eosin-methylene blue staining, and sperm function was determined by using a hemizona binding test. Most of the motility parameters of sperm stored in the <span class="hlt">vapor</span> <span class="hlt">phase</span> were not significantly different from those stored in the liquid <span class="hlt">phase</span> of liquid nitrogen, except in amplitude of lateral head displacement. The percentages of normal sperm morphology in both <span class="hlt">vapor</span> and liquid <span class="hlt">phases</span> also were not significantly different. There was no significant difference in the number of bound sperm in hemizona between sperm cryopreserved in both <span class="hlt">vapor</span> and liquid <span class="hlt">phases</span> of liquid nitrogen. Cryopreservation of human sperm in a <span class="hlt">vapor</span> <span class="hlt">phase</span> of liquid nitrogen was comparable to cryopreservation in a liquid <span class="hlt">phase</span> of liquid nitrogen.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22274165-solgel-synthesis-mcm-silicas-selective-vapor-phase-modification-surface','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22274165-solgel-synthesis-mcm-silicas-selective-vapor-phase-modification-surface"><span>Sol–gel synthesis of MCM-41 silicas and selective <span class="hlt">vapor-phase</span> modification of their surface</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Roik, N.V., E-mail: roik_nadya@ukr.net; Belyakova, L.A.</p> <p>2013-11-15</p> <p>Silica particles with uniform hexagonal mesopore architecture were synthesized by template directed sol–gel condensation of tetraethoxysilane or mixture of tetraethoxysilane and (3-chloropropyl)triethoxysilane in a water–ethanol–ammonia solution. Selective functionalization of exterior surface of parent materials was carried out by postsynthetic treatment of template-filled MCM-41 and Cl-MCM-41 with <span class="hlt">vapors</span> of (3-chloropropyl)triethoxysilane and 1,2-ethylenediamine in vacuum. The chemical composition of obtained mesoporous silicas was estimated by IR spectroscopy and chemical analysis of surface products of reactions. Characteristics of porous structure of resulting materials were determined from the data of X-ray, low-temperature nitrogen ad-desorption and transmission electron microscopy measurements. Obtained results confirm invariability ofmore » highly ordered mesoporous structure of MCM-41 and Cl-MCM-41 after their selective postsynthetic modification in <span class="hlt">vapor</span> <span class="hlt">phase</span>. It was proved that proposed method of <span class="hlt">vapor-phase</span> functionalization of template-filled starting materials is not accompanied by dissolution of the template and chemical modification of pores surface. This provides preferential localization of grafted functional groups onto the exterior surface of mesoporous silicas. - Graphical abstract: Sol–gel synthesis and postsynthetic chemical modification of template-filled MCM-41 and Cl-MCM-41 with (3-chloropropyl)triethoxysilane and 1,2-ethylenediamine in <span class="hlt">vapor</span> <span class="hlt">phase</span>. Display Omitted - Highlights: • Synthesis of MCM-41 silica by template directed sol–gel condensation. • Selective <span class="hlt">vapor-phase</span> functionalization of template-filled silica particles. • Preferential localization of grafted groups onto the exterior surface of mesoporous silicas.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29335709','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29335709"><span>Elongated <span class="hlt">phase</span> <span class="hlt">separation</span> domains in spin-cast polymer blend thin films characterized using a panoramic image.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Hong; Okamura, Yosuke</p> <p>2018-02-14</p> <p>Polymer thin films with micro/nano-structures can be prepared by a solvent evaporation <span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> process via spin-casting a polymer blend, where the elongated <span class="hlt">phase</span> <span class="hlt">separation</span> domains are always inevitable. The striation defect, as a thickness nonunifomity in spin-cast films, is generally coexistent with the elongated domains. Herein, the morphologies of polymer blend thin films are recorded from the spin-cast center to the edge in a panoramic view. The elongated domains are inclined to appear at the ridge regions of striations with increasing radial distance and align radially, exhibiting a coupling between the <span class="hlt">phase</span> <span class="hlt">separation</span> morphology and the striation defect that may exist. We demonstrate that the formation of elongated domains is not attributed to shape deformation, but is accomplished in situ. A possible model to describe the initiation and evolution of the polymer blend <span class="hlt">phase</span> <span class="hlt">separation</span> morphology during spin-casting is proposed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1999JChEd..76.1710B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1999JChEd..76.1710B"><span>Chromatographic <span class="hlt">Separations</span> Using Solid-<span class="hlt">Phase</span> Extraction Cartridges: <span class="hlt">Separation</span> of Wine Phenolics</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Brenneman, Charles A.; Ebeler, Susan E.</p> <p>1999-12-01</p> <p>We describe a simple laboratory experiment that demonstrates the principles of chromatographic <span class="hlt">separation</span> using solid-<span class="hlt">phase</span> extraction columns and red wine. By adjusting pH and mobile <span class="hlt">phase</span> composition, the wine is <span class="hlt">separated</span> into three fractions of differing polarity. The content of each fraction can be monitored by UV-vis spectroscopy. When the experiment is combined with experiments involving HPLC or GC <span class="hlt">separations</span>, students gain a greater appreciation for and understanding of the highly automated instrumental systems currently available. In addition, they learn about the chemistry of polyphenolic compounds, which are present in many foods and beverages and which are receiving much attention for their potentially beneficial health effects.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5389134','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5389134"><span>Role of Co-<span class="hlt">Vapors</span> in <span class="hlt">Vapor</span> Deposition Polymerization</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Lee, Ji Eun; Lee, Younghee; Ahn, Ki-Jin; Huh, Jinyoung; Shim, Hyeon Woo; Sampath, Gayathri; Im, Won Bin; Huh, Yang–Il; Yoon, Hyeonseok</p> <p>2015-01-01</p> <p>Polypyrrole (PPy)/cellulose (PPCL) composite papers were fabricated by <span class="hlt">vapor</span> <span class="hlt">phase</span> polymerization. Importantly, the <span class="hlt">vapor-phase</span> deposition of PPy onto cellulose was assisted by employing different co-<span class="hlt">vapors</span> namely methanol, ethanol, benzene, water, toluene and hexane, in addition to pyrrole. The resulting PPCL papers possessed high mechanical flexibility, large surface-to-volume ratio, and good redox properties. Their main properties were highly influenced by the nature of the co-<span class="hlt">vaporized</span> solvent. The morphology and oxidation level of deposited PPy were tuned by employing co-<span class="hlt">vapors</span> during the polymerization, which in turn led to change in the electrochemical properties of the PPCL papers. When methanol and ethanol were used as co-<span class="hlt">vapors</span>, the conductivities of PPCL papers were found to have improved five times, which was likely due to the enhanced orientation of PPy chain by the polar co-<span class="hlt">vapors</span> with high dipole moment. The specific capacitance of PPCL papers obtained using benzene, toluene, water and hexane co-<span class="hlt">vapors</span> was higher than those of the others, which is attributed to the enlarged effective surface area of the electrode material. The results indicate that the judicious choice and combination of co-<span class="hlt">vapors</span> in <span class="hlt">vapor</span>-deposition polymerization (VDP) offers the possibility of tuning the morphological, electrical, and electrochemical properties of deposited conducting polymers. PMID:25673422</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25314443','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25314443"><span>Hydrodynamic suppression of <span class="hlt">phase</span> <span class="hlt">separation</span> in active suspensions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Matas-Navarro, Ricard; Golestanian, Ramin; Liverpool, Tanniemola B; Fielding, Suzanne M</p> <p>2014-09-01</p> <p>We simulate with hydrodynamics a suspension of active disks squirming through a Newtonian fluid. We explore numerically the full range of squirmer area fractions from dilute to close packed and show that "motility <span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span>," which was recently proposed to arise generically in active matter, and which has been seen in simulations of active Brownian disks, is strongly suppressed by hydrodynamic interactions. We give an argument for why this should be the case and support it with counterpart simulations of active Brownian disks in a parameter regime that provides a closer counterpart to hydrodynamic suspensions than in previous studies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015EGUGA..17.2915P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015EGUGA..17.2915P"><span>A novel mechanical model for <span class="hlt">phase-separation</span> in debris flows</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pudasaini, Shiva P.</p> <p>2015-04-01</p> <p>Understanding the physics of <span class="hlt">phase-separation</span> between solid and fluid <span class="hlt">phases</span> as a two-<span class="hlt">phase</span> mass moves down slope is a long-standing challenge. Here, I propose a fundamentally new mechanism, called '<span class="hlt">separation</span>-flux', that leads to strong <span class="hlt">phase-separation</span> in avalanche and debris flows. This new model extends the general two-<span class="hlt">phase</span> debris flow model (Pudasaini, 2012) to include a <span class="hlt">separation</span>-flux mechanism. The new flux <span class="hlt">separation</span> mechanism is capable of describing and controlling the dynamically evolving <span class="hlt">phase-separation</span>, segregation, and/or levee formation in a real two-<span class="hlt">phase</span>, geometrically three-dimensional debris flow motion and deposition. These are often observed phenomena in natural debris flows and industrial processes that involve the transportation of particulate solid-fluid mixture material. The novel <span class="hlt">separation</span>-flux model includes several dominant physical and mechanical aspects that result in strong <span class="hlt">phase-separation</span> (segregation). These include pressure gradients, volume fractions of solid and fluid <span class="hlt">phases</span> and their gradients, shear-rates, flow depth, material friction, viscosity, material densities, boundary structures, gravity and topographic constraints, grain shape, size, etc. Due to the inherent <span class="hlt">separation</span> mechanism, as the mass moves down slope, more and more solid particles are brought to the front, resulting in a solid-rich and mechanically strong frontal surge head followed by a weak tail largely consisting of the viscous fluid. The primary frontal surge head followed by secondary surge is the consequence of the <span class="hlt">phase-separation</span>. Such typical and dominant <span class="hlt">phase-separation</span> phenomena are revealed here for the first time in real two-<span class="hlt">phase</span> debris flow modeling and simulations. However, these phenomena may depend on the bulk material composition and the applied forces. Reference: Pudasaini, Shiva P. (2012): A general two-<span class="hlt">phase</span> debris flow model. J. Geophys. Res., 117, F03010, doi: 10.1029/2011JF002186.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/404509-phase-equilibrium-measurements-nine-binary-mixtures','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/404509-phase-equilibrium-measurements-nine-binary-mixtures"><span><span class="hlt">Phase</span> equilibrium measurements on nine binary mixtures</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Wilding, W.V.; Giles, N.F.; Wilson, L.C.</p> <p>1996-11-01</p> <p><span class="hlt">Phase</span> equilibrium measurements have been performed on nine binary mixtures. The PTx method was used to obtain <span class="hlt">vapor</span>-liquid equilibrium data for the following systems at two temperatures each: (aminoethyl)piperazine + diethylenetriamine; 2-butoxyethyl acetate + 2-butoxyethanol; 2-methyl-2-propanol + 2-methylbutane; 2-methyl-2-propanol + 2-methyl-2-butene; methacrylonitrile + methanol; 1-chloro-1,1-difluoroethane + hydrogen chloride; 2-(hexyloxy)ethanol + ethylene glycol; butane + ammonia; propionaldehyde + butane. Equilibrium <span class="hlt">vapor</span> and liquid <span class="hlt">phase</span> compositions were derived form the PTx data using the Soave equation of state to represent the <span class="hlt">vapor</span> <span class="hlt">phase</span> and the Wilson or the NRTL activity coefficient model to represent the liquid <span class="hlt">phase</span>. A large immiscibility region existsmore » in the butane + ammonia system at 0 C. Therefore, <span class="hlt">separate</span> <span class="hlt">vapor</span>-liquid-liquid equilibrium measurements were performed on this system to more precisely determine the miscibility limits and the composition of the <span class="hlt">vapor</span> <span class="hlt">phase</span> in equilibrium with the two liquid <span class="hlt">phases</span>.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24967848','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24967848"><span>Ultrahigh-density sub-10 nm nanowire array formation via surface-controlled <span class="hlt">phase</span> <span class="hlt">separation</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Tian, Yuan; Mukherjee, Pinaki; Jayaraman, Tanjore V; Xu, Zhanping; Yu, Yongsheng; Tan, Li; Sellmyer, David J; Shield, Jeffrey E</p> <p>2014-08-13</p> <p>We present simple, self-assembled, and robust fabrication of ultrahigh density cobalt nanowire arrays. The binary Co-Al and Co-Si systems <span class="hlt">phase-separate</span> during physical <span class="hlt">vapor</span> deposition, resulting in Co nanowire arrays with average diameter as small as 4.9 nm and nanowire density on the order of 10(16)/m(2). The nanowire diameters were controlled by moderating the surface diffusivity, which affected the lateral diffusion lengths. High resolution transmission electron microscopy reveals that the Co nanowires formed in the face-centered cubic structure. Elemental mapping showed that in both systems the nanowires consisted of Co with undetectable Al or Si and that the matrix consisted of Al with no distinguishable Co in the Co-Al system and a mixture of Si and Co in the Co-Si system. Magnetic measurements clearly indicate anisotropic behavior consistent with shape anisotropy. The dynamics of nanowire growth, simulated using an Ising model, is consistent with the experimental <span class="hlt">phase</span> and geometry of the nanowires.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28558250','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28558250"><span>Direct NMR Monitoring of <span class="hlt">Phase</span> <span class="hlt">Separation</span> Behavior of Highly Supersaturated Nifedipine Solution Stabilized with Hypromellose Derivatives.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ueda, Keisuke; Higashi, Kenjirou; Moribe, Kunikazu</p> <p>2017-07-03</p> <p>We investigated the <span class="hlt">phase</span> <span class="hlt">separation</span> behavior and maintenance mechanism of the supersaturated state of poorly water-soluble nifedipine (NIF) in hypromellose (HPMC) derivative solutions. Highly supersaturated NIF formed NIF-rich nanodroplets through <span class="hlt">phase</span> <span class="hlt">separation</span> from aqueous solution containing HPMC derivative. Dissolvable NIF concentration in the bulk water <span class="hlt">phase</span> was limited by the <span class="hlt">phase</span> <span class="hlt">separation</span> of NIF from the aqueous solution. HPMC derivatives stabilized the NIF-rich nanodroplets and maintained the NIF supersaturation with <span class="hlt">phase-separated</span> NIF for several hours. The size of the NIF-rich <span class="hlt">phase</span> was different depending on the HPMC derivatives dissolved in aqueous solution, although the droplet size had no correlation with the time for which NIF supersaturation was maintained without NIF crystallization. HPMC acetate and HPMC acetate succinate (HPMC-AS) effectively maintained the NIF supersaturation containing <span class="hlt">phase-separated</span> NIF compared with HPMC. Furthermore, HPMC-AS stabilized NIF supersaturation more effectively in acidic conditions. Solution 1 H NMR measurements of NIF-supersaturated solution revealed that HPMC derivatives distributed into the NIF-rich <span class="hlt">phase</span> during the <span class="hlt">phase</span> <span class="hlt">separation</span> of NIF from the aqueous solution. The hydrophobicity of HPMC derivative strongly affected its distribution into the NIF-rich <span class="hlt">phase</span>. Moreover, the distribution of HPMC-AS into the NIF-rich <span class="hlt">phase</span> was promoted at lower pH due to the lower aqueous solubility of HPMC-AS. The distribution of a large amount of HPMC derivatives into NIF-rich <span class="hlt">phase</span> <span class="hlt">induced</span> the strong inhibition of NIF crystallization from the NIF-rich <span class="hlt">phase</span>. Polymer distribution into the drug-rich <span class="hlt">phase</span> directly monitored by solution NMR technique can be a useful index for the stabilization efficiency of drug-supersaturated solution containing a drug-rich <span class="hlt">phase</span>.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_9");'>9</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li class="active"><span>11</span></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_11 --> <div id="page_12" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li class="active"><span>12</span></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="221"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015EL....10917006S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015EL....10917006S"><span><span class="hlt">Phase</span> <span class="hlt">separation</span> and long-wavelength charge instabilities in spin-orbit coupled systems</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Seibold, G.; Bucheli, D.; Caprara, S.; Grilli, M.</p> <p>2015-01-01</p> <p>We investigate a two-dimensional electron model with Rashba spin-orbit interaction where the coupling constant g=g(n) depends on the electronic density. It is shown that this dependence may drive the system unstable towards a long-wavelength charge density wave (CDW) where the associated second-order instability occurs in close vicinity to global <span class="hlt">phase</span> <span class="hlt">separation</span>. For very low electron densities the CDW instability is nesting-<span class="hlt">induced</span> and the modulation follows the Fermi momentum kF. At higher density the instability criterion becomes independent of kF and the system may become unstable in a broad momentum range. Finally, upon filling the upper spin-orbit split band, finite momentum instabilities disappear in favor of <span class="hlt">phase</span> <span class="hlt">separation</span> alone. We discuss our results with regard to the inhomogeneous <span class="hlt">phases</span> observed at the LaAlO3/SrTiO3 or LaTiO3/SrTiO3 interfaces.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28799364','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28799364"><span>Overview: Homogeneous nucleation from the <span class="hlt">vapor</span> <span class="hlt">phase</span>-The experimental science.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wyslouzil, Barbara E; Wölk, Judith</p> <p>2016-12-07</p> <p>Homogeneous nucleation from the <span class="hlt">vapor</span> <span class="hlt">phase</span> has been a well-defined area of research for ∼120 yr. In this paper, we present an overview of the key experimental and theoretical developments that have made it possible to address some of the fundamental questions first delineated and investigated in C. T. R. Wilson's pioneering paper of 1897 [C. T. R. Wilson, Philos. Trans. R. Soc., A 189, 265-307 (1897)]. We review the principles behind the standard experimental techniques currently used to measure isothermal nucleation rates, and discuss the molecular level information that can be extracted from these measurements. We then highlight recent approaches that interrogate the <span class="hlt">vapor</span> and intermediate clusters leading to particle formation, more directly.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012APS..MARV44008L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012APS..MARV44008L"><span>Self-assembly of Nano-rods in Photosensitive <span class="hlt">Phase</span> <span class="hlt">Separation</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Liu, Ya; Kuksenok, Olga; Maresov, Egor; Balazs, Anna</p> <p>2012-02-01</p> <p>Computer simulations reveal how photo-<span class="hlt">induced</span> chemical reactions in polymeric mixtures can be exploited to create long-range order in materials whose features range from the sub-micron to the nanoscale. The process is initiated by shining a spatially uniform light on a photosensitive AB binary blend, which thereby undergoes both a reversible chemical reaction and <span class="hlt">phase</span> <span class="hlt">separation</span>. When a well-collimated, higher intensity light is rastered over the sample, the system forms defect-free, spatially periodic structures. We now build on this approach by introducing nanorods that have a preferential affinity for one the <span class="hlt">phases</span> in a binary mixture. By rastering over the sample with the higher intensity light, we can create ordered arrays of rods within periodically ordered materials in essentially one processing step.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29658886','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29658886"><span>Swelling kinetics and electrical charge transport in PEDOT:PSS thin films exposed to water <span class="hlt">vapor</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sarkar, Biporjoy; Jaiswal, Manu; Satapathy, Dillip K</p> <p>2018-06-06</p> <p>We report the swelling kinetics and evolution of the electrical charge transport in poly(3,4-ethylene dioxythiophene) polystyrene sulfonate (PEDOT:PSS) thin films subjected to water <span class="hlt">vapor</span>. Polymer films swell by the diffusion of water <span class="hlt">vapor</span> and are found to undergo structural relaxations. Upon exposure to water <span class="hlt">vapor</span>, primarily the hygroscopic PSS shell, which surrounds the conducting PEDOT-rich cores, takes up water <span class="hlt">vapor</span> and subsequently swells. We found that the degree of swelling largely depends on the PEDOT to PSS ratio. Swelling driven microscopic rearrangement of the conducting PEDOT-rich cores in the PSS matrix strongly influences the electrical charge transport of the polymer film. Swelling <span class="hlt">induced</span> increase as well as decrease of electrical resistance are observed in polymer films having different PEDOT to PSS ratio. This anomalous charge transport behavior in PEDOT:PSS films is reconciled by taking into account the contrasting swelling behavior of the PSS and the conducting PEDOT-rich cores leading to spatial segregation of PSS in films with PSS as a minority <span class="hlt">phase</span> and by a net increase in mean <span class="hlt">separation</span> between conducting PEDOT-rich cores for films having abundance of PSS.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JPCM...30v5101S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JPCM...30v5101S"><span>Swelling kinetics and electrical charge transport in PEDOT:PSS thin films exposed to water <span class="hlt">vapor</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sarkar, Biporjoy; Jaiswal, Manu; Satapathy, Dillip K.</p> <p>2018-06-01</p> <p>We report the swelling kinetics and evolution of the electrical charge transport in poly(3,4-ethylene dioxythiophene) polystyrene sulfonate (PEDOT:PSS) thin films subjected to water <span class="hlt">vapor</span>. Polymer films swell by the diffusion of water <span class="hlt">vapor</span> and are found to undergo structural relaxations. Upon exposure to water <span class="hlt">vapor</span>, primarily the hygroscopic PSS shell, which surrounds the conducting PEDOT-rich cores, takes up water <span class="hlt">vapor</span> and subsequently swells. We found that the degree of swelling largely depends on the PEDOT to PSS ratio. Swelling driven microscopic rearrangement of the conducting PEDOT-rich cores in the PSS matrix strongly influences the electrical charge transport of the polymer film. Swelling <span class="hlt">induced</span> increase as well as decrease of electrical resistance are observed in polymer films having different PEDOT to PSS ratio. This anomalous charge transport behavior in PEDOT:PSS films is reconciled by taking into account the contrasting swelling behavior of the PSS and the conducting PEDOT-rich cores leading to spatial segregation of PSS in films with PSS as a minority <span class="hlt">phase</span> and by a net increase in mean <span class="hlt">separation</span> between conducting PEDOT-rich cores for films having abundance of PSS.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1197673-oscillatory-phase-separation-giant-lipid-vesicles-induced-transmembrane-osmotic-differentials','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1197673-oscillatory-phase-separation-giant-lipid-vesicles-induced-transmembrane-osmotic-differentials"><span>Oscillatory <span class="hlt">phase</span> <span class="hlt">separation</span> in giant lipid vesicles <span class="hlt">induced</span> by transmembrane osmotic differentials</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Oglęcka, Kamila; Rangamani, Padmini; Liedberg, Bo</p> <p></p> <p>Giant lipid vesicles are closed compartments consisting of semi-permeable shells, which isolate femto- to pico-liter quantities of aqueous core from the bulk. Although water permeates readily across vesicular walls, passive permeation of solutes is hindered. In this study, we show that, when subject to a hypotonic bath, giant vesicles consisting of <span class="hlt">phase</span> <span class="hlt">separating</span> lipid mixtures undergo osmotic relaxation exhibiting damped oscillations in <span class="hlt">phase</span> behavior, which is synchronized with swell–burst lytic cycles: in the swelled state, osmotic pressure and elevated membrane tension due to the influx of water promote domain formation. During bursting, solute leakage through transient pores relaxes the pressuremore » and tension, replacing the domain texture by a uniform one. This isothermal <span class="hlt">phase</span> transition—resulting from a well-coordinated sequence of mechanochemical events—suggests a complex emergent behavior allowing synthetic vesicles produced from simple components, namely, water, osmolytes, and lipids to sense and regulate their micro-environment.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1197673-oscillatory-phase-separation-giant-lipid-vesicles-induced-transmembrane-osmotic-differentials','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1197673-oscillatory-phase-separation-giant-lipid-vesicles-induced-transmembrane-osmotic-differentials"><span>Oscillatory <span class="hlt">phase</span> <span class="hlt">separation</span> in giant lipid vesicles <span class="hlt">induced</span> by transmembrane osmotic differentials</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Oglęcka, Kamila; Rangamani, Padmini; Liedberg, Bo; ...</p> <p>2014-10-15</p> <p>Giant lipid vesicles are closed compartments consisting of semi-permeable shells, which isolate femto- to pico-liter quantities of aqueous core from the bulk. Although water permeates readily across vesicular walls, passive permeation of solutes is hindered. In this study, we show that, when subject to a hypotonic bath, giant vesicles consisting of <span class="hlt">phase</span> <span class="hlt">separating</span> lipid mixtures undergo osmotic relaxation exhibiting damped oscillations in <span class="hlt">phase</span> behavior, which is synchronized with swell–burst lytic cycles: in the swelled state, osmotic pressure and elevated membrane tension due to the influx of water promote domain formation. During bursting, solute leakage through transient pores relaxes the pressuremore » and tension, replacing the domain texture by a uniform one. This isothermal <span class="hlt">phase</span> transition—resulting from a well-coordinated sequence of mechanochemical events—suggests a complex emergent behavior allowing synthetic vesicles produced from simple components, namely, water, osmolytes, and lipids to sense and regulate their micro-environment.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4197780','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4197780"><span>Oscillatory <span class="hlt">phase</span> <span class="hlt">separation</span> in giant lipid vesicles <span class="hlt">induced</span> by transmembrane osmotic differentials</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Oglęcka, Kamila; Rangamani, Padmini; Liedberg, Bo; Kraut, Rachel S; Parikh, Atul N</p> <p>2014-01-01</p> <p>Giant lipid vesicles are closed compartments consisting of semi-permeable shells, which isolate femto- to pico-liter quantities of aqueous core from the bulk. Although water permeates readily across vesicular walls, passive permeation of solutes is hindered. In this study, we show that, when subject to a hypotonic bath, giant vesicles consisting of <span class="hlt">phase</span> <span class="hlt">separating</span> lipid mixtures undergo osmotic relaxation exhibiting damped oscillations in <span class="hlt">phase</span> behavior, which is synchronized with swell–burst lytic cycles: in the swelled state, osmotic pressure and elevated membrane tension due to the influx of water promote domain formation. During bursting, solute leakage through transient pores relaxes the pressure and tension, replacing the domain texture by a uniform one. This isothermal <span class="hlt">phase</span> transition—resulting from a well-coordinated sequence of mechanochemical events—suggests a complex emergent behavior allowing synthetic vesicles produced from simple components, namely, water, osmolytes, and lipids to sense and regulate their micro-environment. DOI: http://dx.doi.org/10.7554/eLife.03695.001 PMID:25318069</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=knock&pg=2&id=EJ853489','ERIC'); return false;" href="https://eric.ed.gov/?q=knock&pg=2&id=EJ853489"><span>A Classroom Demonstration of Water-<span class="hlt">Induced</span> <span class="hlt">Phase</span> <span class="hlt">Separation</span> of Alcohol-Gasoline Biofuel Blends</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Mueller, Sherry A.; Anderson, James E.; Wallington, Timothy J.</p> <p>2009-01-01</p> <p>A significant issue associated with ethanol-gasoline blends is the <span class="hlt">phase</span> <span class="hlt">separation</span> that occurs with the addition of small volumes of water, producing an ethanol-deficient gasoline layer and an ethanol-rich aqueous layer. The gasoline layer may have a lower-than-desired octane rating due to the decrease in ethanol content, resulting in engine…</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/2628','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/2628"><span>Enhanced <span class="hlt">Vapor-Phase</span> Diffusion in Porous Media - LDRD Final Report</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Ho, C.K.; Webb, S.W.</p> <p>1999-01-01</p> <p>As part of the Laboratory-Directed Research and Development (LDRD) Program at Sandia National Laboratories, an investigation into the existence of enhanced <span class="hlt">vapor-phase</span> diffusion (EVD) in porous media has been conducted. A thorough literature review was initially performed across multiple disciplines (soil science and engineering), and based on this review, the existence of EVD was found to be questionable. As a result, modeling and experiments were initiated to investigate the existence of EVD. In this LDRD, the first mechanistic model of EVD was developed which demonstrated the mechanisms responsible for EVD. The first direct measurements of EVD have also been conductedmore » at multiple scales. Measurements have been made at the pore scale, in a two- dimensional network as represented by a fracture aperture, and in a porous medium. Significant enhancement of <span class="hlt">vapor-phase</span> transport relative to Fickian diffusion was measured in all cases. The modeling and experimental results provide additional mechanisms for EVD beyond those presented by the generally accepted model of Philip and deVries (1957), which required a thermal gradient for EVD to exist. Modeling and experimental results show significant enhancement under isothermal conditions. Application of EVD to <span class="hlt">vapor</span> transport in the near-surface vadose zone show a significant variation between no enhancement, the model of Philip and deVries, and the present results. Based on this information, the model of Philip and deVries may need to be modified, and additional studies are recommended.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003PhRvE..68d1601M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003PhRvE..68d1601M"><span>Interfacial nonequilibrium and Bénard-Marangoni instability of a liquid-<span class="hlt">vapor</span> system</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Margerit, J.; Colinet, P.; Lebon, G.; Iorio, C. S.; Legros, J. C.</p> <p>2003-10-01</p> <p>We study Bénard-Marangoni instability in a system formed by a horizontal liquid layer and its overlying <span class="hlt">vapor</span>. The liquid is lying on a hot rigid plate and the <span class="hlt">vapor</span> is bounded by a cold parallel plate. A pump maintains a reduced pressure in the <span class="hlt">vapor</span> layer and evacuates the <span class="hlt">vapor</span>. This investigation is undertaken within the classical quasisteady approximation for both the <span class="hlt">vapor</span> and the liquid <span class="hlt">phases</span>. The two layers are <span class="hlt">separated</span> by a deformable interface. Temporarily frozen temperature and velocity distributions are employed at each instant for the stability analysis, limited to infinitesimal disturbances (linear regime). We use irreversible thermodynamics to model the <span class="hlt">phase</span> change under interfacial nonequilibrium. Within this description, the interface appears as a barrier for transport of both heat and mass. Hence, in contrast with previous studies, we consider the possibility of a temperature jump across the interface, as recently measured experimentally. The stability analysis shows that the interfacial resistances to heat and mass transfer have a destabilizing influence compared to an interface that is in thermodynamic equilibrium. The role of the fluctuations in the <span class="hlt">vapor</span> <span class="hlt">phase</span> on the onset of instability is discussed. The conditions to reduce the system to a one <span class="hlt">phase</span> model are also established. Finally, the influence of the evaporation parameters and of the presence of an inert gas on the marginal stability curves is discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/909584','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/909584"><span>Nanostructures produced by <span class="hlt">phase-separation</span> during growth of (III-V).sub.1-x(IV.sub.2).sub.x alloys</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Norman, Andrew G [Evergreen, CO; Olson, Jerry M [Lakewood, CO</p> <p>2007-06-12</p> <p>Nanostructures (18) and methods for production thereof by <span class="hlt">phase</span> <span class="hlt">separation</span> during metal organic <span class="hlt">vapor-phase</span> epitaxy (MOVPE). An embodiment of one of the methods may comprise providing a growth surface in a reaction chamber and introducing a first mixture of precursor materials into the reaction chamber to form a buffer layer (12) thereon. A second mixture of precursor materials may be provided into the reaction chamber to form an active region (14) on the buffer layer (12), wherein the nanostructure (18) is embedded in a matrix (16) in the active region (14). Additional steps are also disclosed for preparing the nanostructure (18) product for various applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JThSc..27..230T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JThSc..27..230T"><span>Thermodynamic Analysis of a Mixed Refrigerant Ejector Refrigeration Cycle Operating with Two <span class="hlt">Vapor</span>-liquid <span class="hlt">Separators</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tan, Yingying; Chen, Youming; Wang, Lin</p> <p>2018-06-01</p> <p>A mixed refrigerant ejector refrigeration cycle operating with two-stage <span class="hlt">vapor</span>-liquid <span class="hlt">separators</span> (MRERC2) is proposed to obtain refrigeration temperature at -40°C. The thermodynamic investigations on performance of MRERC2 using zeotropic mixture refrigerant R23/R134a are performed, and the comparisons of cycle performance between MRERC2 and MRERC1 (MRERC with one-stage <span class="hlt">vapor</span>-liquid <span class="hlt">separator</span>) are conducted. The results show that MRERC2 can achieve refrigeration temperature varying between -23.9°C and -42.0°C when ejector pressure ratio ranges from 1.6 to 2.3 at the generation temperature of 57.3-84.9°C. The parametric analysis indicates that increasing condensing temperature decreases coefficient of performance ( COP) of MRERC2, and increasing ejector pressure ratio and mass fraction of the low boiling point component in the mixed refrigerant can improve COP of MRERC2. The MRERC2 shows its potential in utilizing low grade thermal energy as driving power to obtain low refrigeration temperature for the ejector refrigeration cycle.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4021964','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4021964"><span>Carbon Nanotube- and Carbon Fiber-Reinforcement of Ethylene-Octene Copolymer Membranes for Gas and <span class="hlt">Vapor</span> <span class="hlt">Separation</span></span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Sedláková, Zuzana; Clarizia, Gabriele; Bernardo, Paola; Jansen, Johannes Carolus; Slobodian, Petr; Svoboda, Petr; Kárászová, Magda; Friess, Karel; Izak, Pavel</p> <p>2014-01-01</p> <p>Gas and <span class="hlt">vapor</span> transport properties were studied in mixed matrix membranes containing elastomeric ethylene-octene copolymer (EOC or poly(ethylene-co-octene)) with three types of carbon fillers: virgin or oxidized multi-walled carbon nanotubes (CNTs) and carbon fibers (CFs). Helium, hydrogen, nitrogen, oxygen, methane, and carbon dioxide were used for gas permeation rate measurements. <span class="hlt">Vapor</span> transport properties were studied for the aliphatic hydrocarbon (hexane), aromatic compound (toluene), alcohol (ethanol), as well as water for the representative samples. The mechanical properties and homogeneity of samples was checked by stress-strain tests. The addition of virgin CNTs and CFs improve mechanical properties. Gas permeability of EOC lies between that of the more permeable PDMS and the less permeable semi-crystalline polyethylene and polypropylene. Organic <span class="hlt">vapors</span> are more permeable than permanent gases in the composite membranes, with toluene and hexane permeabilities being about two orders of magnitude higher than permanent gas permeability. The results of the carbon-filled membranes offer perspectives for application in gas/<span class="hlt">vapor</span> <span class="hlt">separation</span> with improved mechanical resistance. PMID:24957119</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19780004447','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19780004447"><span>Low gravity <span class="hlt">phase</span> <span class="hlt">separator</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Smoot, G. F.; Pope, W. L.; Smith, L. (Inventor)</p> <p>1977-01-01</p> <p>An apparatus is described for <span class="hlt">phase</span> <span class="hlt">separating</span> a gas-liquid mixture as might exist in a subcritical cryogenic helium vessel for cooling a superconducting magnet at low gravity such as in planetary orbit, permitting conservation of the liquid and extended service life of the superconducting magnet.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015APS..DFDL37001B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015APS..DFDL37001B"><span>Bioeffects due to acoustic droplet <span class="hlt">vaporization</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bull, Joseph</p> <p>2015-11-01</p> <p>Encapsulated micro- and nano-droplets can be <span class="hlt">vaporized</span> via ultrasound, a process termed acoustic droplet <span class="hlt">vaporization</span>. Our interest is primarily motivated by a developmental gas embolotherapy technique for cancer treatment. In this methodology, infarction of tumors is <span class="hlt">induced</span> by selectively formed vascular gas bubbles that arise from the acoustic <span class="hlt">vaporization</span> of vascular microdroplets. Additionally, the microdroplets may be used as vehicles for localized drug delivery, with or without flow occlusion. In this talk, we examine the dynamics of acoustic droplet <span class="hlt">vaporization</span> through experiments and theoretical/computational fluid mechanics models, and investigate the bioeffects of acoustic droplet <span class="hlt">vaporization</span> on endothelial cells and in vivo. Early timescale <span class="hlt">vaporization</span> events, including <span class="hlt">phase</span> change, are directly visualized using ultra-high speed imaging, and the influence of acoustic parameters on droplet/bubble dynamics is discussed. Acoustic and fluid mechanics parameters affecting the severity of endothelial cell bioeffects are explored. These findings suggest parameter spaces for which bioeffects may be reduced or enhanced, depending on the objective of the therapy. This work was supported by NIH grant R01EB006476.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29701714','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29701714"><span>Doxorubicin Release Controlled by <span class="hlt">Induced</span> <span class="hlt">Phase</span> <span class="hlt">Separation</span> and Use of a Co-Solvent.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Park, Seok Chan; Yuan, Yue; Choi, Kyoungju; Choi, Seong-O; Kim, Jooyoun</p> <p>2018-04-26</p> <p>Electrospun-based drug delivery is emerging as a versatile means of localized therapy; however, controlling the release rates of active agents still remains as a key question. We propose a facile strategy to control the drug release behavior from electrospun fibers by a simple modification of polymer matrices. Polylactic acid (PLA) was used as a major component of the drug-carrier, and doxorubicin hydrochloride (Dox) was used as a model drug. The influences of a polar co-solvent, dimethyl sulfoxide (DMSO), and a hydrophilic polymer additive, polyvinylpyrrolidone (PVP), on the drug miscibility, loading efficiency and release behavior were investigated. The use of DMSO enabled the homogeneous internalization of the drug as well as higher drug loading efficiency within the electrospun fibers. The PVP additive <span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> in the PLA matrix and acted as a porogen. Preferable partitioning of Dox into the PVP domain resulted in increased drug loading efficiency in the PLA/PVP fiber. Fast dissolution of PVP domains created pores in the fibers, facilitating the release of internalized Dox. The novelty of this study lies in the detailed experimental investigation of the effect of additives in pre-spinning formulations, such as co-solvents and polymeric porogens, on the drug release behavior of nanofibers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5978058','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5978058"><span>Doxorubicin Release Controlled by <span class="hlt">Induced</span> <span class="hlt">Phase</span> <span class="hlt">Separation</span> and Use of a Co-Solvent</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Park, Seok Chan; Choi, Kyoungju; Choi, Seong-O</p> <p>2018-01-01</p> <p>Electrospun-based drug delivery is emerging as a versatile means of localized therapy; however, controlling the release rates of active agents still remains as a key question. We propose a facile strategy to control the drug release behavior from electrospun fibers by a simple modification of polymer matrices. Polylactic acid (PLA) was used as a major component of the drug-carrier, and doxorubicin hydrochloride (Dox) was used as a model drug. The influences of a polar co-solvent, dimethyl sulfoxide (DMSO), and a hydrophilic polymer additive, polyvinylpyrrolidone (PVP), on the drug miscibility, loading efficiency and release behavior were investigated. The use of DMSO enabled the homogeneous internalization of the drug as well as higher drug loading efficiency within the electrospun fibers. The PVP additive <span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> in the PLA matrix and acted as a porogen. Preferable partitioning of Dox into the PVP domain resulted in increased drug loading efficiency in the PLA/PVP fiber. Fast dissolution of PVP domains created pores in the fibers, facilitating the release of internalized Dox. The novelty of this study lies in the detailed experimental investigation of the effect of additives in pre-spinning formulations, such as co-solvents and polymeric porogens, on the drug release behavior of nanofibers. PMID:29701714</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28759860','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28759860"><span>Super-hydrophobic coatings based on non-solvent <span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> during electro-spraying.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gao, Jiefeng; Huang, Xuewu; Wang, Ling; Zheng, Nan; Li, Wan; Xue, Huaiguo; Li, Robert K Y; Mai, Yiu-Wing</p> <p>2017-11-15</p> <p>The polymer solution concentration determines whether electrospinning or electro-spraying occurs, while the addition of the non-solvent into the polymer solution strongly influences the surface morphology of the obtained products. Both smooth and porous surfaces of the electro-sprayed microspheres can be harvested by choosing different non-solvent and its amount as well as incorporating polymeric additives. The influences of the solution concentration, weight ratio between the non-solvent and the copolymer, and the polymeric additives on the surface morphology and the wettability of the electro-sprayed products were systematically studied. Surface pores and/or asperities on the microsphere surface were mainly caused by the non-solvent <span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> (NIPS) and subsequent evaporation of the non-solvent during electro-spraying. With increasing polymer solution concentration, the microsphere was gradually changed to the bead-on-string geometry and finally to a nanofiber form, leading to a sustained decrease of the contact angle (CA). It was found that the substrate coatings derived from the microspheres possessing hierarchical surface pores or dense asperities had high surface roughness and super-hydrophobicity with CAs larger than 150° while sliding angles smaller than 10°; but coatings composed of microspheres with smooth surfaces gave relatively low CAs. Copyright © 2017 Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017EPJP..132..483H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017EPJP..132..483H"><span>Effect of the Hartmann number on <span class="hlt">phase</span> <span class="hlt">separation</span> controlled by magnetic field for binary mixture system with large component ratio</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Heping, Wang; Xiaoguang, Li; Duyang, Zang; Rui, Hu; Xingguo, Geng</p> <p>2017-11-01</p> <p>This paper presents an exploration for <span class="hlt">phase</span> <span class="hlt">separation</span> in a magnetic field using a coupled lattice Boltzmann method (LBM) with magnetohydrodynamics (MHD). The left vertical wall was kept at a constant magnetic field. Simulations were conducted by the strong magnetic field to enhance <span class="hlt">phase</span> <span class="hlt">separation</span> and increase the size of <span class="hlt">separated</span> <span class="hlt">phases</span>. The focus was on the effect of magnetic intensity by defining the Hartmann number (Ha) on the <span class="hlt">phase</span> <span class="hlt">separation</span> properties. The numerical investigation was carried out for different governing parameters, namely Ha and the component ratio of the mixed liquid. The effective morphological evolutions of <span class="hlt">phase</span> <span class="hlt">separation</span> in different magnetic fields were demonstrated. The patterns showed that the slant elliptical <span class="hlt">phases</span> were created by increasing Ha, due to the formation and increase of magnetic torque and force. The dataset was rearranged for growth kinetics of magnetic <span class="hlt">phase</span> <span class="hlt">separation</span> in a plot by spherically averaged structure factor and the ratio of <span class="hlt">separated</span> <span class="hlt">phases</span> and total system. The results indicate that the increase in Ha can increase the average size of <span class="hlt">separated</span> <span class="hlt">phases</span> and accelerate the spinodal decomposition and domain growth stages. Specially for the larger component ratio of mixed <span class="hlt">phases</span>, the <span class="hlt">separation</span> degree was also significantly improved by increasing magnetic intensity. These numerical results provide guidance for setting the optimum condition for the <span class="hlt">phase</span> <span class="hlt">separation</span> <span class="hlt">induced</span> by magnetic field.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_10");'>10</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li class="active"><span>12</span></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_12 --> <div id="page_13" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li class="active"><span>13</span></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="241"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26774779','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26774779"><span>Biodegradation of <span class="hlt">vapor-phase</span> toluene in unsaturated porous media: Column experiments.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Khan, Ali M; Wick, Lukas Y; Harms, Hauke; Thullner, Martin</p> <p>2016-04-01</p> <p>Biodegradation of organic chemicals in the <span class="hlt">vapor</span> <span class="hlt">phase</span> of soils and vertical flow filters has gained attention as promising approach to clean up volatile organic compounds (VOC). The drivers of VOC biodegradation in unsaturated systems however still remain poorly understood. Here, we analyzed the processes controlling aerobic VOC biodegradation in a laboratory setup mimicking the unsaturated zone above a shallow aquifer. The setup allowed for diffusive <span class="hlt">vapor-phase</span> transport and biodegradation of three VOC: non-deuterated and deuterated toluene as two compounds of highly differing biodegradability but (nearly) identical physical and chemical properties, and MTBE as (at the applied experimental conditions) non-biodegradable tracer and internal control. Our results showed for toluene an effective microbial degradation within centimeter VOC transport distances despite high gas-<span class="hlt">phase</span> diffusivity. Degradation rates were controlled by the reactivity of the compounds while oxic conditions were found everywhere in the system. This confirms hypotheses that vadose zone biodegradation rates can be extremely high and are able to prevent the outgassing of VOC to the atmosphere within a centimeter range if compound properties and site conditions allow for sufficiently high degradation rates. Copyright © 2016 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24496120','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24496120"><span>Lower critical solution temperature (LCST) <span class="hlt">phase</span> <span class="hlt">separation</span> of glycol ethers for forward osmotic control.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Nakayama, Daichi; Mok, Yeongbong; Noh, Minwoo; Park, Jeongseon; Kang, Sunyoung; Lee, Yan</p> <p>2014-03-21</p> <p>Lower critical solution temperature (LCST) <span class="hlt">phase</span> transition of glycol ether (GE)-water mixtures <span class="hlt">induces</span> an abrupt change in osmotic pressure driven by a mild temperature change. The temperature-controlled osmotic change was applied for the forward osmosis (FO) desalination. Among three GEs evaluated, di(ethylene glycol) n-hexyl ether (DEH) was selected as a potential FO draw solute. A DEH-water mixture with a high osmotic pressure could draw fresh water from a high-salt feed solution such as seawater through a semipermeable membrane at around 10 °C. The water-drawn DEH-water mixture was <span class="hlt">phase-separated</span> into a water-rich <span class="hlt">phase</span> and a DEH-rich <span class="hlt">phase</span> at around 30 °C. The water-rich <span class="hlt">phase</span> with a much reduced osmotic pressure released water into a low-salt solution, and the DEH-rich <span class="hlt">phase</span> was recovered into the initial DEH-water mixture. The <span class="hlt">phase</span> <span class="hlt">separation</span> behaviour, the residual GE concentration in the water-rich <span class="hlt">phase</span>, the osmotic pressure of the DEH-water mixture, and the osmotic flux between the DEH-water mixture and salt solutions were carefully analysed for FO desalination. The liquid-liquid <span class="hlt">phase</span> <span class="hlt">separation</span> of the GE-water mixture driven by the mild temperature change between 10 °C and 30 °C is very attractive for the development of an ideal draw solute for future practical FO desalination.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28671747','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28671747"><span>Tailorable Surface Morphology of 3D Scaffolds by Combining Additive Manufacturing with Thermally <span class="hlt">Induced</span> <span class="hlt">Phase</span> <span class="hlt">Separation</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Di Luca, Andrea; de Wijn, Joost R; van Blitterswijk, Clemens A; Camarero-Espinosa, Sandra; Moroni, Lorenzo</p> <p>2017-08-01</p> <p>The functionalization of biomaterials substrates used for cell culture is gearing towards an increasing control over cell activity. Although a number of biomaterials have been successfully modified by different strategies to display tailored physical and chemical surface properties, it is still challenging to step from 2D substrates to 3D scaffolds with instructive surface properties for cell culture and tissue regeneration. In this study, additive manufacturing and thermally <span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> are combined to create 3D scaffolds with tunable surface morphology from polymer gels. Surface features vary depending on the gel concentration, the exchanging temperature, and the nonsolvent used. When preosteoblasts (MC-3T3 cells) are cultured on these scaffolds, a significant increase in alkaline phosphatase activity is measured for submicron surface topography, suggesting a potential role on early cell differentiation. © 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JPSJ...85d4708J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JPSJ...85d4708J"><span><span class="hlt">Phase</span> <span class="hlt">Separation</span> of Superconducting <span class="hlt">Phases</span> in the Penson-Kolb-Hubbard Model</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jerzy Kapcia, Konrad; Czart, Wojciech Robert; Ptok, Andrzej</p> <p>2016-04-01</p> <p>In this paper, we determine the <span class="hlt">phase</span> diagrams (for T = 0 as well as T > 0) of the Penson-Kolb-Hubbard model for two dimensional square lattice within Hartree-Fock mean-field theory focusing on an investigation of superconducting <span class="hlt">phases</span> and on a possibility of the occurrence of the <span class="hlt">phase</span> <span class="hlt">separation</span>. We obtain that the <span class="hlt">phase</span> <span class="hlt">separation</span>, which is a state of coexistence of two different superconducting <span class="hlt">phases</span> (with s- and η-wave symmetries), occurs in definite ranges of the electron concentration. In addition, increasing temperature can change the symmetry of the superconducting order parameter (from η-wave into s-wave). The system considered exhibits also an interesting multicritical behaviour including bicritical points. The relevance of the results to experiments for real materials is also discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012JFS....34..190S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012JFS....34..190S"><span>Cavitating flow during water hammer using a generalized interface <span class="hlt">vaporous</span> cavitation model</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sadafi, Mohamadhosein; Riasi, Alireza; Nourbakhsh, Seyed Ahmad</p> <p>2012-10-01</p> <p>In a transient flow simulation, column <span class="hlt">separation</span> may occur when the calculated pressure head decreases to the saturated <span class="hlt">vapor</span> pressure head in a computational grid. Abrupt valve closure or pump failure can result in a fast transient flow with column <span class="hlt">separation</span>, potentially causing problems such as pipe failure, hydraulic equipment damage, cavitation or corrosion. This paper reports a numerical study of water hammer with column <span class="hlt">separation</span> in a simple reservoir-pipeline-valve system and pumping station. The governing equations for two-<span class="hlt">phase</span> transient flow in pipes are solved based on the method of characteristics (MOC) using a generalized interface <span class="hlt">vaporous</span> cavitating model (GIVCM). The numerical results were compared with the experimental data for validation purposes, and the comparison indicated that the GIVCM describes the experimental results more accurately than the discrete <span class="hlt">vapor</span> cavity model (DVCM). In particular, the GIVCM correlated better with the experimental data than the DVCM in terms of timing and pressure magnitude. The effects of geometric and hydraulic parameters on flow behavior in a pumping station with column <span class="hlt">separation</span> were also investigated in this study.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016AGUFMPP51C2324W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016AGUFMPP51C2324W"><span>The effect of heated <span class="hlt">vapor-phase</span> acidification on organic carbon concentrations and isotopic values in geologic rock samples</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, R. Z.; West, A. J.; Yager, J. A.; Rollins, N.; Li, G.; Berelson, W.</p> <p>2016-12-01</p> <p>Carbon signatures recorded in the modern and geologic rock record can give insight on the Earth's carbon cycle through time. This is especially true for organic carbon (OC), which can help us understand how the biosphere has evolved over Earth's history. However, carbon recorded in rocks is a combination of OC and inorganic carbon (IC) mostly in the form of carbonate minerals. To measure OC, IC must therefore first be removed through a process called "decarbonation." This is often done through a leaching process with hydrochloric acid (HCl). However, three well known problems exist for the decarbonation process: 1) Incomplete removal of IC, 2) Unintentional removal of OC, and 3) Addition of false carbon blank. Currently, <span class="hlt">vapor</span> (gas) <span class="hlt">phase</span> removal of OC is preferred to liquid <span class="hlt">phase</span> treatment because it has been shown that OC is lost to solubilization during liquid <span class="hlt">phase</span> acidification. <span class="hlt">Vapor</span> <span class="hlt">phase</span> treatment is largely thought to avoid the problem of OC loss, but this has not yet been rigorously investigated. This study investigates that assumption and shows that <span class="hlt">vapor</span> <span class="hlt">phase</span> treatment can cause unintentional OC loss. We show that <span class="hlt">vapor</span> <span class="hlt">phase</span> treatment must be sensitive to rock type and treatment length to produce robust OC isotopic measurements and concentrations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016RJPCA..90.1596S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016RJPCA..90.1596S"><span>Liquid-<span class="hlt">vapor</span> <span class="hlt">phase</span> equilibria and the thermodynamic properties of 2-methylpropanol- n-alkyl propanoate solutions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Suntsov, Yu. K.; Goryunov, V. A.; Chuikov, A. M.; Meshcheryakov, A. V.</p> <p>2016-08-01</p> <p>The boiling points of solutions of five binary systems are measured via ebulliometry in the pressure range of 2.05-103.3 kPa. Equilibrium <span class="hlt">vapor</span> <span class="hlt">phase</span> compositions, the values of the excess Gibbs energies, enthalpies, and entropies of solution of these systems are calculated. Patterns in the changes of <span class="hlt">phase</span> equilibria and thermodynamic properties of solutions are established, depending on the compositions and temperatures of the systems. Liquid-<span class="hlt">vapor</span> equilibria in the systems are described using the equations of Wilson and the NRTL (Non-Random Two-Liquid Model).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19870014058','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19870014058"><span>Impact <span class="hlt">vaporization</span>: Late time phenomena from experiments</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Schultz, P. H.; Gault, D. E.</p> <p>1987-01-01</p> <p>While simple airflow produced by the outward movement of the ejecta curtain can be scaled to large dimensions, the interaction between an impact-<span class="hlt">vaporized</span> component and the ejecta curtain is more complicated. The goal of these experiments was to examine such interaction in a real system involving crater growth, ejection of material, two <span class="hlt">phased</span> mixtures of gas and dust, and strong pressure gradients. The results will be complemented by theoretical studies at laboratory scales in order to <span class="hlt">separate</span> the various parameters for planetary scale processes. These experiments prompt, however, the following conclusions that may have relevance at broader scales. First, under near vacuum or low atmospheric pressures, an expanding <span class="hlt">vapor</span> cloud scours the surrounding surface in advance of arriving ejecta. Second, the effect of early-time <span class="hlt">vaporization</span> is relatively unimportant at late-times. Third, the overpressure created within the crater cavity by significant <span class="hlt">vaporization</span> results in increased cratering efficiency and larger aspect ratios.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19800016957','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19800016957"><span>External fuel <span class="hlt">vaporization</span> study, <span class="hlt">phase</span> 1</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Szetela, E. J.; Chiappetta, L.</p> <p>1980-01-01</p> <p>A conceptual design study was conducted to devise and evaluate techniques for the external <span class="hlt">vaporization</span> of fuel for use in an aircraft gas turbine with characteristics similar to the Energy Efficient Engine (E(3)). Three <span class="hlt">vaporizer</span> concepts were selected and they were analyzed from the standpoint of fuel thermal stability, integration of the <span class="hlt">vaporizer</span> system into the aircraft engine, engine and <span class="hlt">vaporizer</span> dynamic response, startup and altitude restart, engine performance, control requirements, safety, and maintenance. One of the concepts was found to improve the performance of the baseline E(3) engine without seriously compromising engine startup and power change response. Increased maintenance is required because of the need for frequent pyrolytic cleaning of the surfaces in contact with hot fuel.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20020049818&hterms=gallium+vapor&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dgallium%2Bvapor','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20020049818&hterms=gallium+vapor&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Dgallium%2Bvapor"><span>Modeling of Gallium Nitride Hydride <span class="hlt">Vapor</span> <span class="hlt">Phase</span> Epitaxy</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Meyyappan, Meyya; Arnold, James O. (Technical Monitor)</p> <p>1997-01-01</p> <p>A reactor model for the hydride <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy of GaN is presented. The governing flow, energy, and species conservation equations are solved in two dimensions to examine the growth characteristics as a function of process variables and reactor geometry. The growth rate varies with GaCl composition but independent of NH3 and H2 flow rates. A change in carrier gas for Ga source from H2 to N2 affects the growth rate and uniformity for a fixed reactor configuration. The model predictions are in general agreement with observed experimental behavior.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016PhFl...28h2108Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016PhFl...28h2108Z"><span>Hydrodynamic effects on <span class="hlt">phase</span> <span class="hlt">separation</span> morphologies in evaporating thin films of polymer solutions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zoumpouli, Garyfalia A.; Yiantsios, Stergios G.</p> <p>2016-08-01</p> <p>We examine effects of hydrodynamics on <span class="hlt">phase</span> <span class="hlt">separation</span> morphologies developed during drying of thin films containing a volatile solvent and two dissolved polymers. Cahn-Hilliard and Flory-Huggins theories are used to describe the free energy of the <span class="hlt">phase</span> <span class="hlt">separating</span> systems. The thin films, considered as Newtonian fluids, flow in response to Korteweg stresses arising due to concentration non-uniformities that develop during solvent evaporation. Numerical simulations are employed to investigate the effects of a Peclet number, defined in terms of system physical properties, as well as the effects of parameters characterizing the speed of evaporation and preferential wetting of the solutes at the gas interface. For systems exhibiting preferential wetting, diffusion alone is known to favor lamellar configurations for the <span class="hlt">separated</span> <span class="hlt">phases</span> in the dried film. However, a mechanism of hydrodynamic instability of a short length scale is revealed, which beyond a threshold Peclet number may deform and break the lamellae. The critical Peclet number tends to decrease as the evaporation rate increases and to increase with the tendency of the polymers to selectively wet the gas interface. As the Peclet number increases, the instability moves closer to the gas interface and <span class="hlt">induces</span> the formation of a lateral segregation template that guides the subsequent evolution of the <span class="hlt">phase</span> <span class="hlt">separation</span> process. On the other hand, for systems with no preferential wetting or any other property asymmetries between the two polymers, diffusion alone favors the formation of laterally <span class="hlt">separated</span> configurations. In this case, concentration perturbation modes that lead to enhanced Korteweg stresses may be favored for sufficiently large Peclet numbers. For such modes, a second mechanism is revealed, which is similar to the solutocapillary Marangoni instability observed in evaporating solutions when interfacial tension increases with the concentration of the non-volatile component. This mechanism may lead</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/442608-experimental-study-phase-separation-dividing-two-phase-flow','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/442608-experimental-study-phase-separation-dividing-two-phase-flow"><span>Experimental study of <span class="hlt">phase</span> <span class="hlt">separation</span> in dividing two <span class="hlt">phase</span> flow</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Qian Yong; Yang Zhilin; Xu Jijun</p> <p>1996-12-31</p> <p>Experimental study of <span class="hlt">phase</span> <span class="hlt">separation</span> of air-water two <span class="hlt">phase</span> bubbly, slug flow in the horizontal T-junction is carried out. The influences of the inlet mass quality X1, mass extraction rate G3/G1, and fraction of extracted liquid QL3/QL1 on <span class="hlt">phase</span> <span class="hlt">separation</span> characteristics are analyzed. For the first time, the authors have found and defined pulsating run effect by the visual experiments, which show that under certain conditions, the down stream flow of the T-junction has strangely affected the <span class="hlt">phase</span> redistribution of the junction, and firstly point out that the downstream geometric condition is very important to the study of <span class="hlt">phase</span> separationmore » phenomenon of two-<span class="hlt">phase</span> flow in a T-junction. This kind of phenomenon has many applications in the field of energy, power, petroleum and chemical industries, such as the loss of coolant accident (LOCA) caused by a small break in a horizontal coolant pipe in nuclear reactor, and the flip-flop effect in the natural gas transportation pipeline system, etc.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24919675','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24919675"><span>Tube radial distribution phenomenon with a two-<span class="hlt">phase</span> <span class="hlt">separation</span> solution of a fluorocarbon and hydrocarbon organic solvent mixture in a capillary tube and metal compounds <span class="hlt">separation</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kitaguchi, Koichi; Hanamura, Naoya; Murata, Masaharu; Hashimoto, Masahiko; Tsukagoshi, Kazuhiko</p> <p>2014-01-01</p> <p>A fluorocarbon and hydrocarbon organic solvent mixture is known as a temperature-<span class="hlt">induced</span> <span class="hlt">phase-separation</span> solution. When a mixed solution of tetradecafluorohexane as a fluorocarbon organic solvent and hexane as a hydrocarbon organic solvent (e.g., 71:29 volume ratio) was delivered in a capillary tube that was controlled at 10°C, the tube radial distribution phenomenon (TRDP) of the solvents was clearly observed through fluorescence images of the dye, perylene, dissolved in the mixed solution. The homogeneous mixed solution (single <span class="hlt">phase</span>) changed to a heterogeneous solution (two <span class="hlt">phases</span>) with inner tetradecafluorohexane and outer hexane <span class="hlt">phases</span> in the tube under laminar flow conditions, generating the dynamic liquid-liquid interface. We also tried to apply TRDP to a <span class="hlt">separation</span> technique for metal compounds. A model analyte mixture, copper(II) and hematin, was <span class="hlt">separated</span> through the capillary tube, and detected with a chemiluminescence detector in this order within 4 min.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29057594','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29057594"><span><span class="hlt">Phase</span> Equilibrium of TiO2 Nanocrystals in Flame-Assisted Chemical <span class="hlt">Vapor</span> Deposition.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Liu, Changran; Camacho, Joaquin; Wang, Hai</p> <p>2018-01-19</p> <p>Nano-scale titanium oxide (TiO 2 ) is a material useful for a wide range of applications. In a previous study, we showed that TiO 2 nanoparticles of both rutile and anatase crystal <span class="hlt">phases</span> could be synthesized over the size range of 5 to 20 nm in flame-assisted chemical <span class="hlt">vapor</span> deposition. Rutile was unexpectedly dominant in oxygen-lean synthesis conditions, whereas anatase is the preferred <span class="hlt">phase</span> in oxygen-rich gases. The observation is in contrast to the 14 nm rutile-anatase crossover size derived from the existing crystal-<span class="hlt">phase</span> equilibrium model. In the present work, we made additional measurements over a wider range of synthesis conditions; the results confirm the earlier observations. We propose an improved model for the surface energy that considers the role of oxygen desorption at high temperatures. The model successfully explains the observations made in the current and previous work. The current results provide a useful path to designing flame-assisted chemical <span class="hlt">vapor</span> deposition of TiO 2 nanocrystals with controllable crystal <span class="hlt">phases</span>. © 2018 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvB..97a4505F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvB..97a4505F"><span>Direct visualization of <span class="hlt">phase</span> <span class="hlt">separation</span> between superconducting and nematic domains in Co-doped CaFe2As2 close to a first-order <span class="hlt">phase</span> transition</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fente, Antón; Correa-Orellana, Alexandre; Böhmer, Anna E.; Kreyssig, Andreas; Ran, S.; Bud'ko, Sergey L.; Canfield, Paul C.; Mompean, Federico J.; García-Hernández, Mar; Munuera, Carmen; Guillamón, Isabel; Suderow, Hermann</p> <p>2018-01-01</p> <p>We show that biaxial strain <span class="hlt">induces</span> alternating tetragonal superconducting and orthorhombic nematic domains in Co-substituted CaFe2As2 . We use atomic force, magnetic force, and scanning tunneling microscopy to identify the domains and characterize their properties, finding in particular that tetragonal superconducting domains are very elongated, more than several tens of micrometers long and about 30 nm wide; have the same Tc as unstrained samples; and hold vortices in a magnetic field. Thus, biaxial strain produces a <span class="hlt">phase-separated</span> state, where each <span class="hlt">phase</span> is equivalent to what is found on either side of the first-order <span class="hlt">phase</span> transition between antiferromagnetic orthorhombic and superconducting tetragonal <span class="hlt">phases</span> found in unstrained samples when changing Co concentration. Having such alternating superconducting domains <span class="hlt">separated</span> by normal conducting domains with sizes of the order of the coherence length opens opportunities to build Josephson junction networks or vortex pinning arrays and suggests that first-order quantum <span class="hlt">phase</span> transitions lead to nanometric-size <span class="hlt">phase</span> <span class="hlt">separation</span> under the influence of strain.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23977837','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23977837"><span>Integrated lab-in-syringe platform incorporating a membraneless gas-liquid <span class="hlt">separator</span> for automatic cold <span class="hlt">vapor</span> atomic absorption spectrometry.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Giakisikli, Georgia; Miró, Manuel; Anthemidis, Aristidis</p> <p>2013-10-01</p> <p>This manuscript reports the proof-of-concept of a novel integrated lab-in-syringe/gas-liquid <span class="hlt">separation</span> (LIS/GLS) batch-flow system based on a programmable flow for automatic cold <span class="hlt">vapor</span> atomic absorption spectrometric assays. Homogeneous mixing of metered volumes of sample and reagent solutions drawn up in a sandwich-type mode along with in situ <span class="hlt">vapor</span> generation are accomplished inside the microsyringe in a closed manner, while the <span class="hlt">separation</span> of <span class="hlt">vapor</span> species is achieved via the membraneless GLS located at the top of the syringe's valve in the upright position. The potentials of the proposed manifold were demonstrated for trace inorganic mercury determination in drinking waters and seawater. For a 3.0 mL sample, the limit of detection and repeatability (RSD) were found to be 0.03 μg L(-1) Hg(II) and 3.1% (at the 2.0 μg L(-1) concentration level), respectively, with a dynamic range extending up to 10.0 μg L(-1). The proposed system fulfills the requirements of US-EPA, WHO, and EU Council Directives for measurements of the maximum allowed concentrations of inorganic mercury in drinking water.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29677515','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29677515"><span>FUS <span class="hlt">Phase</span> <span class="hlt">Separation</span> Is Modulated by a Molecular Chaperone and Methylation of Arginine Cation-π Interactions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Qamar, Seema; Wang, GuoZhen; Randle, Suzanne J; Ruggeri, Francesco Simone; Varela, Juan A; Lin, Julie Qiaojin; Phillips, Emma C; Miyashita, Akinori; Williams, Declan; Ströhl, Florian; Meadows, William; Ferry, Rodylyn; Dardov, Victoria J; Tartaglia, Gian G; Farrer, Lindsay A; Kaminski Schierle, Gabriele S; Kaminski, Clemens F; Holt, Christine E; Fraser, Paul E; Schmitt-Ulms, Gerold; Klenerman, David; Knowles, Tuomas; Vendruscolo, Michele; St George-Hyslop, Peter</p> <p>2018-04-19</p> <p>Reversible <span class="hlt">phase</span> <span class="hlt">separation</span> underpins the role of FUS in ribonucleoprotein granules and other membrane-free organelles and is, in part, driven by the intrinsically disordered low-complexity (LC) domain of FUS. Here, we report that cooperative cation-π interactions between tyrosines in the LC domain and arginines in structured C-terminal domains also contribute to <span class="hlt">phase</span> <span class="hlt">separation</span>. These interactions are modulated by post-translational arginine methylation, wherein arginine hypomethylation strongly promotes <span class="hlt">phase</span> <span class="hlt">separation</span> and gelation. Indeed, significant hypomethylation, which occurs in FUS-associated frontotemporal lobar degeneration (FTLD), <span class="hlt">induces</span> FUS condensation into stable intermolecular β-sheet-rich hydrogels that disrupt RNP granule function and impair new protein synthesis in neuron terminals. We show that transportin acts as a physiological molecular chaperone of FUS in neuron terminals, reducing <span class="hlt">phase</span> <span class="hlt">separation</span> and gelation of methylated and hypomethylated FUS and rescuing protein synthesis. These results demonstrate how FUS condensation is physiologically regulated and how perturbations in these mechanisms can lead to disease. Copyright © 2018 The Authors. Published by Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28603980','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28603980"><span>Sericin Promotes Fibroin Silk I Stabilization Across a <span class="hlt">Phase-Separation</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kwak, Hyo Won; Ju, Ji Eun; Shin, Munju; Holland, Chris; Lee, Ki Hoon</p> <p>2017-08-14</p> <p>Natural silk spinning offers several advantages over the synthetic fiber spinning, although the underlying mechanisms of this process are yet to be fully elucidated. Silkworm silks, specifically B. mori, comprise two main proteins: fibroin, which forms the fiber, and sericin, a coextruded coating that acts as a matrix in the resulting nonwoven composite cocoon. To date, most studies have focused on fibroin's self-assembly and gelation, with the influence of sericin during spinning receiving little to no attention. This study investigates sericin's effects on the self-assembly of fibroin via their natural <span class="hlt">phase-separation</span>. Through changes in sample opacity, FTIR, and XRD, we report that increasing sericin concentration retards the time to gelation and β-sheet formation of fibroin, causing it to adopt a Silk I conformation. Such findings have important implications for both the natural silk spinning process and any future industrial applications, suggesting that sericin may be able to <span class="hlt">induce</span> long-range conformational and stability control in silk fibroin, while being in a <span class="hlt">separate</span> <span class="hlt">phase</span>, a factor that would facilitate long-term storage or silk feedstocks.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29753780','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29753780"><span>The Role of RNA in Biological <span class="hlt">Phase</span> <span class="hlt">Separations</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Fay, Marta M; Anderson, Paul J</p> <p>2018-05-10</p> <p><span class="hlt">Phase</span> transitions that alter the physical state of ribonucleoprotein particles contribute to the spacial and temporal organization of the densely packed intracellular environment. This allows cells to organize biologically coupled processes as well as respond to environmental stimuli. RNA plays a key role in <span class="hlt">phase</span> <span class="hlt">separation</span> events that modulate various aspects of RNA metabolism. Here, we review the role that RNA plays in ribonucleoprotein <span class="hlt">phase</span> <span class="hlt">separations</span>. Copyright © 2018 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19760016013&hterms=tellurium&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dtellurium','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19760016013&hterms=tellurium&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dtellurium"><span>Crystal growth from the <span class="hlt">vapor</span> <span class="hlt">phase</span> experiment MA-085</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Wiedemeir, H.; Sadeek, H.; Klaessig, F. C.; Norek, M.</p> <p>1976-01-01</p> <p>Three <span class="hlt">vapor</span> transport experiments on multicomponent systems were performed during the Apollo Soyuz mission to determine the effects of microgravity forces on crystal morphology and mass transport rates. The mixed systems used germanium selenide, tellurium, germanium tetraiodide (transport agent), germanium monosulfide, germanium tetrachloride (transport agent), and argon (inert atmosphere). The materials were enclosed in evacuated sealed ampoules of fused silica and were transported in a temperature gradient of the multipurpose electric furnace onboard the Apollo Soyuz spacecraft. Preliminary evaluation of 2 systems shows improved quality of space grown crystals in terms of growth morphology and bulk perfection. This conclusion is based on a direct comparison of space grown and ground based crystals by means of X-ray diffraction, microscopic, and chemical etching techniques. The observation of greater mass transport rates than predicted for a microgravity environment by existing <span class="hlt">vapor</span> transport models indicates the existence of nongravity caused transport effects in a reactive solid/gas <span class="hlt">phase</span> system.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_11");'>11</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li class="active"><span>13</span></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_13 --> <div id="page_14" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li class="active"><span>14</span></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="261"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016PhRvL.116b6804S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016PhRvL.116b6804S"><span><span class="hlt">Phase</span> <span class="hlt">Separation</span> from Electron Confinement at Oxide Interfaces</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Scopigno, N.; Bucheli, D.; Caprara, S.; Biscaras, J.; Bergeal, N.; Lesueur, J.; Grilli, M.</p> <p>2016-01-01</p> <p>Oxide heterostructures are of great interest for both fundamental and applicative reasons. In particular, the two-dimensional electron gas at the LaAlO3/SrTiO3 or LaTiO3/SrTiO3 interfaces displays many different properties and functionalities. However, there are clear experimental indications that the interface electronic state is strongly inhomogeneous and therefore it is crucial to investigate possible intrinsic mechanisms underlying this inhomogeneity. Here, the electrostatic potential confining the electron gas at the interface is calculated self-consistently, finding that such confinement may <span class="hlt">induce</span> <span class="hlt">phase</span> <span class="hlt">separation</span>, to avoid a thermodynamically unstable state with a negative compressibility. This provides a robust mechanism for the inhomogeneous character of these interfaces.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/3986837','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/3986837"><span>Glass fibers and <span class="hlt">vapor</span> <span class="hlt">phase</span> components of cigarette smoke as cofactors in experimental respiratory tract carcinogenesis.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Feron, V J; Kuper, C F; Spit, B J; Reuzel, P G; Woutersen, R A</p> <p>1985-01-01</p> <p>Syrian golden hamsters were given intratracheal instillations of glass fibers with or without BP suspended in saline, once a fortnight for 52 weeks; the experiment was terminated at week 85. No tumors of the respiratory tract were observed in hamsters treated with glass fibers alone. There was no indication that glass fibers enhanced the development of respiratory tract tumors <span class="hlt">induced</span> by BP. In another study Syrian golden hamsters were exposed to fresh air or to a mixture of 4 major <span class="hlt">vapor</span> <span class="hlt">phase</span> components of cigarette smoke, viz. isoprene (800----700 ppm), methyl chloride (1000----900 ppm), methyl nitrite (200----190 ppm) and acetaldehyde (1400----1200 ppm) for a period of at most 23 months. Some of the animals were also given repeated intratracheal instillations of BP or norharman in saline. Laryngeal tumors were found in 7/31 male and 6/32 female hamsters exposed only to the <span class="hlt">vapor</span> mixture, whereas no laryngeal tumors occurred in controls. The tumor response of the larynx most probably has to be ascribed entirely to the action of acetaldehyde. Simultaneous treatment with norharman or BP did not affect the tumor response of the larynx. Acetaldehyde may occur in the <span class="hlt">vapor</span> <span class="hlt">phase</span> of cigarette smoke at levels up to 2000 ppm. Chronic inhalation exposure of rats to acetaldehyde at levels of 0 (controls), 750, 1500 or 3000----1000 ppm resulted in a high incidence of nasal carcinomas, both squamous cell carcinomas of the respiratory epithelium and adenocarcinomas of the olfactory epithelium. It was discussed that acetaldehyde may significantly contribute to the induction of bronchogenic cancer by cigarette smoke in man. No evidence was obtained for a role of isoprene, methyl chloride or methyl nitrite in the induction of lung cancer by cigarette smoke.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/869565','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/869565"><span>Enhanced quality thin film Cu(In,Ga)Se.sub.2 for semiconductor device applications by <span class="hlt">vapor-phase</span> recrystallization</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Tuttle, John R.; Contreras, Miguel A.; Noufi, Rommel; Albin, David S.</p> <p>1994-01-01</p> <p>Enhanced quality thin films of Cu.sub.w (In,Ga.sub.y)Se.sub.z for semiconductor device applications are fabricated by initially forming a Cu-rich, <span class="hlt">phase-separated</span> compound mixture comprising Cu(In,Ga):Cu.sub.x Se on a substrate to form a large-grain precursor and then converting the excess Cu.sub.x Se to Cu(In,Ga)Se.sub.2 by exposing it to an activity of In and/or Ga, either in <span class="hlt">vapor</span> In and/or Ga form or in solid (In,Ga).sub.y Se.sub.z. Alternatively, the conversion can be made by sequential deposition of In and/or Ga and Se onto the <span class="hlt">phase-separated</span> precursor. The conversion process is preferably performed in the temperature range of about 300.degree.-600.degree. C., where the Cu(In,Ga)Se.sub.2 remains solid, while the excess Cu.sub.x Se is in a liquid flux. The characteristic of the resulting Cu.sub.w (In,Ga).sub.y Se.sub.z can be controlled by the temperature. Higher temperatures, such as 500.degree.-600.degree. C., result in a nearly stoichiometric Cu(In,Ga)Se.sub.2, whereas lower temperatures, such as 300.degree.-400.degree. C., result in a more Cu-poor compound, such as the Cu.sub.z (In,Ga).sub.4 Se.sub.7 <span class="hlt">phase</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/6893589','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/biblio/6893589"><span>Enhanced quality thin film Cu(In,Ga)Se[sub 2] for semiconductor device applications by <span class="hlt">vapor-phase</span> recrystallization</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Tuttle, J.R.; Contreras, M.A.; Noufi, R.; Albin, D.S.</p> <p>1994-10-18</p> <p>Enhanced quality thin films of Cu[sub w](In,Ga[sub y])Se[sub z] for semiconductor device applications are fabricated by initially forming a Cu-rich, <span class="hlt">phase-separated</span> compound mixture comprising Cu(In,Ga):Cu[sub x]Se on a substrate to form a large-grain precursor and then converting the excess Cu[sub x]Se to Cu(In,Ga)Se[sub 2] by exposing it to an activity of In and/or Ga, either in <span class="hlt">vapor</span> In and/or Ga form or in solid (In,Ga)[sub y]Se[sub z]. Alternatively, the conversion can be made by sequential deposition of In and/or Ga and Se onto the <span class="hlt">phase-separated</span> precursor. The conversion process is preferably performed in the temperature range of about 300--600 C, where the Cu(In,Ga)Se[sub 2] remains solid, while the excess Cu[sub x]Se is in a liquid flux. The characteristic of the resulting Cu[sub w](In,Ga)[sub y]Se[sub z] can be controlled by the temperature. Higher temperatures, such as 500--600 C, result in a nearly stoichiometric Cu(In,Ga)Se[sub 2], whereas lower temperatures, such as 300--400 C, result in a more Cu-poor compound, such as the Cu[sub z](In,Ga)[sub 4]Se[sub 7] <span class="hlt">phase</span>. 7 figs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014JChPh.140w4506S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014JChPh.140w4506S"><span>Curvature <span class="hlt">induced</span> <span class="hlt">phase</span> stability of an intensely heated liquid</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sasikumar, Kiran; Liang, Zhi; Cahill, David G.; Keblinski, Pawel</p> <p>2014-06-01</p> <p>We use non-equilibrium molecular dynamics simulations to study the heat transfer around intensely heated solid nanoparticles immersed in a model Lennard-Jones fluid. We focus our studies on the role of the nanoparticle curvature on the liquid <span class="hlt">phase</span> stability under steady-state heating. For small nanoparticles we observe a stable liquid <span class="hlt">phase</span> near the nanoparticle surface, which can be at a temperature well above the boiling point. Furthermore, for particles with radius smaller than a critical radius of 2 nm we do not observe formation of <span class="hlt">vapor</span> even above the critical temperature. Instead, we report the existence of a stable fluid region with a density much larger than that of the <span class="hlt">vapor</span> <span class="hlt">phase</span>. We explain the stability in terms of the Laplace pressure associated with the formation of a <span class="hlt">vapor</span> nanocavity and the associated effect on the Gibbs free energy.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=75832&keyword=heat+AND+exchanger&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50','EPA-EIMS'); return false;" href="https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=75832&keyword=heat+AND+exchanger&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50"><span><span class="hlt">SEPARATION</span> OF <span class="hlt">VAPOR-PHASE</span> ALCOHOL/WATER MIXTURES VIA FRACTIONAL CONDENSATION USING A PILOT-SCALE DEPHLEGMATOR: ENHANCEMENT OF THE PREVAPORATION PROCESS <span class="hlt">SEPARATION</span> FACTOR</span></a></p> <p><a target="_blank" href="http://oaspub.epa.gov/eims/query.page">EPA Science Inventory</a></p> <p></p> <p></p> <p>In prevaporation, a liquid mixture contacts a membrane surface that preferentially permeates one of the liquid components as a <span class="hlt">vapor</span>. Our approach to improving pervaporation performance is to replace the one-stage condenser traditionally used to condense the permeate with a frac...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/6928809-reduction-degradation-vapor-phase-transported-inp-ingaasp-mushroom-stripe-lasers','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/6928809-reduction-degradation-vapor-phase-transported-inp-ingaasp-mushroom-stripe-lasers"><span>Reduction of degradation in <span class="hlt">vapor</span> <span class="hlt">phase</span> transported InP/InGaAsP mushroom stripe lasers</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Jung, H.; Burkhardt, E.G.; Pfister, W.</p> <p>1988-10-03</p> <p>The rapid degradation rate generally observed in InP/InGaAsP mushroom stripe lasers can be considerably decreased by regrowing the open sidewalls of the active stripe with low-doped InP in a second epitaxial step using the hydride <span class="hlt">vapor</span> <span class="hlt">phase</span> transport technique. This technique does not change the fundamental laser parameters like light-current and current-voltage characteristics. Because of this drastic reduction in degradation, the <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy regrown InP/InGaAsP mushroom laser seems to be an interesting candidate for application in optical communication.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApSS..427...92L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApSS..427...92L"><span><span class="hlt">Vapor</span>-liquid interfacial reaction to fabricate superhydrophilic and underwater superoleophobic thiol-ene/silica hybrid decorated fabric for oil/water <span class="hlt">separation</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Li, Hongqiang; Liang, Tao; Lai, Xuejun; Su, Xiaojing; Zhang, Lin; Zeng, Xingrong</p> <p>2018-01-01</p> <p>With oil spill accidents and oil industrial wastewater increasing, oil/water <span class="hlt">separation</span> has attracted much attention in recent years. Herein, we report the fabrication of superhydrophilic and underwater superoleophobic thiol-ene/silica hybrid decorated fabrics for oil/water <span class="hlt">separation</span> via <span class="hlt">vapor</span>-liquid interfacial reaction. It is based on sol-gel reaction of tetraethyl orthosilicate (TEOS) to generate silica and thiol-ene reaction between poly(ethylene glycol) dimethacrylate (PEGDMA) and trimethylolpropane tris(3-mercaptopropionate) (TTMP) to form crosslinked hydrophilic polymer on polyester fabric under the catalysis of butylamine/ammonia <span class="hlt">vapor</span>. The chemical structure of the surfaces on thiol-ene/silica hybrid decorated fabric was confirmed by FTIR and XPS, and obvious micro-nano morphology and roughness were observed with SEM and AFM. The water contact angle of the fabric attained 0° in 0.36 s, and the underwater oil contact angle reached up to 160°. Importantly, the fabric exhibited high <span class="hlt">separation</span> efficiency at 99.5%, fast water flux above 71600 Lm-2h-1 and excellent recyclability in oil/water <span class="hlt">separation</span>. Our findings open a new strategy to fabricate organic-inorganic hybrid superhydrophobic and underwater superoleophobic materials for oil/water <span class="hlt">separation</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JPCRD..45c3101A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JPCRD..45c3101A"><span><span class="hlt">Phase</span> Transition Enthalpy Measurements of Organic and Organometallic Compounds. Sublimation, <span class="hlt">Vaporization</span> and Fusion Enthalpies From 1880 to 2015. Part 1. C1 - C10</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Acree, William; Chickos, James S.</p> <p>2016-09-01</p> <p>A compendium of <span class="hlt">phase</span> change enthalpies published in 2010 is updated to include the period 1880-2015. <span class="hlt">Phase</span> change enthalpies including fusion, <span class="hlt">vaporization</span>, and sublimation enthalpies are included for organic, organometallic, and a few inorganic compounds. Part 1 of this compendium includes organic compounds from C1 to C10. Part 2 of this compendium, to be published <span class="hlt">separately</span>, will include organic and organometallic compounds from C11 to C192. Sufficient data are presently available to permit thermodynamic cycles to be constructed as an independent means of evaluating the reliability of the data. Temperature adjustments of <span class="hlt">phase</span> change enthalpies from the temperature of measurement to the standard reference temperature, T = 298.15 K, and a protocol for doing so are briefly discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1417992-direct-visualization-phase-separation-between-superconducting-nematic-domains-co-doped-cafe2as2-close-first-order-phase-transition','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1417992-direct-visualization-phase-separation-between-superconducting-nematic-domains-co-doped-cafe2as2-close-first-order-phase-transition"><span>Direct visualization of <span class="hlt">phase</span> <span class="hlt">separation</span> between superconducting and nematic domains in Co-doped CaFe 2 As 2 close to a first-order <span class="hlt">phase</span> transition</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Fente, Antón; Correa-Orellana, Alexandre; Böhmer, Anna E.; ...</p> <p>2018-01-09</p> <p>We show that biaxial strain <span class="hlt">induces</span> alternating tetragonal superconducting and orthorhombic nematic domains in Co substituted CaFe 2As 2. We use Atomic Force, Magnetic Force and Scanning Tunneling Microscopy (AFM, MFM and STM) to identify the domains and characterize their properties, nding in particular that tetragonal superconducting domains are very elongated, more than several tens of μm long and about 30 nm wide, have the same Tc than unstrained samples and hold vortices in a magnetic eld. Thus, biaxial strain produces a <span class="hlt">phase</span> <span class="hlt">separated</span> state, where each <span class="hlt">phase</span> is equivalent to what is found at either side of the rstmore » order <span class="hlt">phase</span> transition between antiferromagnetic orthorhombic and superconducting tetragonal <span class="hlt">phases</span> found in unstrained samples when changing Co concentration. Having such alternating superconducting domains <span class="hlt">separated</span> by normal conducting domains with sizes of order of the coherence length opens opportunities to build Josephson junction networks or vortex pinning arrays and suggests that first order quantum <span class="hlt">phase</span> transitions lead to nanometric size <span class="hlt">phase</span> <span class="hlt">separation</span> under the influence of strain.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25227572','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25227572"><span>Anisotropic imprint of amorphization and <span class="hlt">phase</span> <span class="hlt">separation</span> in manganite thin films via laser interference irradiation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ding, Junfeng; Lin, Zhipeng; Wu, Jianchun; Dong, Zhili; Wu, Tom</p> <p>2015-02-04</p> <p>Materials with mesoscopic structural and electronic <span class="hlt">phase</span> <span class="hlt">separation</span>, either inherent from synthesis or created via external means, are known to exhibit functionalities absent in the homogeneous counterparts. One of the most notable examples is the colossal magnetoresistance discovered in mixed-valence manganites, where the coexistence of nano- to micrometer-sized <span class="hlt">phase-separated</span> domains dictates the magnetotransport. However, it remains challenging to pattern and process such materials into predesigned structures and devices. In this work, a direct laser interference irradiation (LII) method is employed to produce periodic stripes in thin films of a prototypical <span class="hlt">phase-separated</span> manganite Pr0.65 (Ca0.75 Sr0.25 )0.35 MnO3 (PCSMO). LII <span class="hlt">induces</span> selective structural amorphization within the crystalline PCSMO matrix, forming arrays with dimensions commensurate with the laser wavelength. Furthermore, because the length scale of LII modification is compatible to that of <span class="hlt">phase</span> <span class="hlt">separation</span> in PCSMO, three orders of magnitude of increase in magnetoresistance and significant in-plane transport anisotropy are observed in treated PCSMO thin films. Our results show that LII is a rapid, cost-effective and contamination-free technique to tailor and improve the physical properties of manganite thin films, and it is promising to be generalized to other functional materials. © 2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70014455','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70014455"><span><span class="hlt">Vapor</span>-dominated zones within hydrothermal systems: evolution and natural state</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Ingebritsen, S.E.; Sorey, M.L.</p> <p>1988-01-01</p> <p>Three conceptual models illustrate the range of hydrothermal systems in which <span class="hlt">vapor</span>-dominated conditions are found. The first model (model I) represents a system with an extensive near-vaporstatic <span class="hlt">vapor</span>-dominated zone and limited liquid throughflow and is analogous to systems such as The Geysers, California. Models II and III represent systems with significant liquid throughflow and include steam-heated discharge features at higher elevations and high-chloride springs at lower elevations connected to and fed by a single circulation system at depth. In model II, as in model I, the <span class="hlt">vapor</span>-dominated zone has a near-vaporstatic vertical pressure gradient and is generally underpressured with respect to local hydrostatic pressure. The <span class="hlt">vapor</span>-dominated zone in model III is quite different, in that <span class="hlt">phase</span> <span class="hlt">separation</span> takes place at pressures close to local hydrostatic and the overall pressure gradient is near hydrostatic. -from Authors</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016APS..MARA37007X','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016APS..MARA37007X"><span><span class="hlt">Phase</span> <span class="hlt">separation</span> and the formation of cellular bodies</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Xu, Bin; Broedersz, Chase P.; Meir, Yigal; Wingreen, Ned S.</p> <p></p> <p>Cellular bodies in eukaryotic cells spontaneously assemble to form cellular compartments. Among other functions, these bodies carry out essential biochemical reactions. Cellular bodies form micron-sized structures, which, unlike canonical cell organelles, are not surrounded by membranes. A recent in vitro experiment has shown that <span class="hlt">phase</span> <span class="hlt">separation</span> of polymers in solution can explain the formation of cellular bodies. We constructed a lattice-polymer model to capture the essential mechanism leading to this <span class="hlt">phase</span> <span class="hlt">separation</span>. We used both analytical and numerical tools to predict the <span class="hlt">phase</span> diagram of a system of two interacting polymers, including the concentration of each polymer type in the condensed and dilute <span class="hlt">phase</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4724432','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4724432"><span>Widely Tunable Morphologies in Block Copolymer Thin Films Through Solvent <span class="hlt">Vapor</span> Annealing Using Mixtures of Selective Solvents</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Chavis, Michelle A.; Smilgies, Detlef-M.; Wiesner, Ulrich B.; Ober, Christopher K.</p> <p>2015-01-01</p> <p>Thin films of block copolymers are extremely attractive for nanofabrication because of their ability to form uniform and periodic nanoscale structures by microphase <span class="hlt">separation</span>. One shortcoming of this approach is that to date the design of a desired equilibrium structure requires synthesis of a block copolymer de novo within the corresponding volume ratio of the blocks. In this work, we investigated solvent <span class="hlt">vapor</span> annealing in supported thin films of poly(2-hydroxyethyl methacrylate)-block-poly(methyl methacrylate) [PHEMA-b-PMMA] by means of grazing incidence small angle X–ray scattering (GISAXS). A spin-coated thin film of lamellar block copolymer was solvent <span class="hlt">vapor</span> annealed to <span class="hlt">induce</span> microphase <span class="hlt">separation</span> and improve the long-range order of the self-assembled pattern. Annealing in a mixture of solvent <span class="hlt">vapors</span> using a controlled volume ratio of solvents (methanol, MeOH, and tetrahydrofuran, THF), which are chosen to be preferential for each block, enabled selective formation of ordered lamellae, gyroid, hexagonal or spherical morphologies from a single block copolymer with a fixed volume fraction. The selected microstructure was then kinetically trapped in the dry film by rapid drying. To our knowledge, this paper describes the first reported case where in-situ methods are used to study the transition of block copolymer films from one initial disordered morphology to four different ordered morphologies, covering much of the theoretical diblock copolymer <span class="hlt">phase</span> diagram. PMID:26819574</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018SuMi..117..293W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018SuMi..117..293W"><span>Preparation of freestanding GaN wafer by hydride <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy on porous silicon</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wu, Xian; Li, Peng; Liang, Renrong; Xiao, Lei; Xu, Jun; Wang, Jing</p> <p>2018-05-01</p> <p>A freestanding GaN wafer was prepared on porous Si (111) substrate using hydride <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy (HVPE). To avoid undesirable effects of the porous surface on the crystallinity of the GaN, a GaN seed layer was first grown on the Si (111) bare wafer. A pattern with many apertures was fabricated in the GaN seed layer using lithography and etching processes. A porous layer was formed in the Si substrate immediately adjacent to the GaN seed layer by an anodic etching process. A 500-μm-thick GaN film was then grown on the patterned GaN seed layer using HVPE. The GaN film was <span class="hlt">separated</span> from the Si substrate through the formation of cracks in the porous layer caused by thermal mismatch stress during the cooling stage of the HVPE. Finally, the GaN film was polished to obtain a freestanding GaN wafer.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19970028025','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19970028025"><span>Removal of Oxygen from Electronic Materials by <span class="hlt">Vapor-Phase</span> Processes</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Palosz, Witold</p> <p>1997-01-01</p> <p>Thermochemical analyses of equilibrium partial pressures over oxides with and without the presence of the respective element condensed <span class="hlt">phase</span>, and hydrogen, chalcogens, hydrogen chalcogenides, and graphite are presented. Theoretical calculations are supplemented with experimental results on the rate of decomposition and/or sublimation/<span class="hlt">vaporization</span> of the oxides under dynamic vacuum, and on the rate of reaction with hydrogen, graphite, and chalcogens. Procedures of removal of a number of oxides under different conditions are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009JChPh.130t4905G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009JChPh.130t4905G"><span>Polymer depletion-driven cluster aggregation and initial <span class="hlt">phase</span> <span class="hlt">separation</span> in charged nanosized colloids</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gögelein, Christoph; Nägele, Gerhard; Buitenhuis, Johan; Tuinier, Remco; Dhont, Jan K. G.</p> <p>2009-05-01</p> <p>We study polymer depletion-driven cluster aggregation and initial <span class="hlt">phase</span> <span class="hlt">separation</span> in aqueous dispersions of charge-stabilized silica spheres, where the ionic strength and polymer (dextran) concentration are systematically varied, using dynamic light scattering and visual observation. Without polymers and for increasing salt and colloid content, the dispersions become increasingly unstable against irreversible cluster formation. By adding nonadsorbing polymers, a depletion-driven attraction is <span class="hlt">induced</span>, which lowers the stabilizing Coulomb barrier and enhances the cluster growth rate. The initial growth rate increases with increasing polymer concentration and decreases with increasing polymer molar mass. These observations can be quantitatively understood by an irreversible dimer formation theory based on the classical Derjaguin, Landau, Verwey, and Overbeek pair potential, with the depletion attraction modeled by the Asakura-Oosawa-Vrij potential. At low colloid concentration, we observe an exponential cluster growth rate for all polymer concentrations considered, indicating a reaction-limited aggregation mechanism. At sufficiently high polymer and colloid concentrations, and lower salt content, a gas-liquidlike demixing is observed initially. Later on, the system <span class="hlt">separates</span> into a gel and fluidlike <span class="hlt">phase</span>. The experimental time-dependent state diagram is compared to the theoretical equilibrium <span class="hlt">phase</span> diagram obtained from a generalized free-volume theory and is discussed in terms of an initial reversible <span class="hlt">phase</span> <span class="hlt">separation</span> process in combination with irreversible aggregation at later times.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19485479','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19485479"><span>Polymer depletion-driven cluster aggregation and initial <span class="hlt">phase</span> <span class="hlt">separation</span> in charged nanosized colloids.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gögelein, Christoph; Nägele, Gerhard; Buitenhuis, Johan; Tuinier, Remco; Dhont, Jan K G</p> <p>2009-05-28</p> <p>We study polymer depletion-driven cluster aggregation and initial <span class="hlt">phase</span> <span class="hlt">separation</span> in aqueous dispersions of charge-stabilized silica spheres, where the ionic strength and polymer (dextran) concentration are systematically varied, using dynamic light scattering and visual observation. Without polymers and for increasing salt and colloid content, the dispersions become increasingly unstable against irreversible cluster formation. By adding nonadsorbing polymers, a depletion-driven attraction is <span class="hlt">induced</span>, which lowers the stabilizing Coulomb barrier and enhances the cluster growth rate. The initial growth rate increases with increasing polymer concentration and decreases with increasing polymer molar mass. These observations can be quantitatively understood by an irreversible dimer formation theory based on the classical Derjaguin, Landau, Verwey, and Overbeek pair potential, with the depletion attraction modeled by the Asakura-Oosawa-Vrij potential. At low colloid concentration, we observe an exponential cluster growth rate for all polymer concentrations considered, indicating a reaction-limited aggregation mechanism. At sufficiently high polymer and colloid concentrations, and lower salt content, a gas-liquidlike demixing is observed initially. Later on, the system <span class="hlt">separates</span> into a gel and fluidlike <span class="hlt">phase</span>. The experimental time-dependent state diagram is compared to the theoretical equilibrium <span class="hlt">phase</span> diagram obtained from a generalized free-volume theory and is discussed in terms of an initial reversible <span class="hlt">phase</span> <span class="hlt">separation</span> process in combination with irreversible aggregation at later times.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25358016','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25358016"><span>Simultaneous visualization of water and hydrogen peroxide <span class="hlt">vapor</span> using two-photon laser-<span class="hlt">induced</span> fluorescence and photofragmentation laser-<span class="hlt">induced</span> fluorescence.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Larsson, Kajsa; Johansson, Olof; Aldén, Marcus; Bood, Joakim</p> <p>2014-01-01</p> <p>A concept based on a combination of photofragmentation laser-<span class="hlt">induced</span> fluorescence (PF-LIF) and two-photon laser-<span class="hlt">induced</span> fluorescence (LIF) is for the first time demonstrated for simultaneous detection of hydrogen peroxide (H2O2) and water (H2O) <span class="hlt">vapor</span>. Water detection is based on two-photon excitation by an injection-locked krypton fluoride (KrF) excimer laser (248.28 nm), which <span class="hlt">induces</span> broadband fluorescence (400-500 nm) from water. The same laser simultaneously photodissociates H2O2, whereupon the generated OH fragments are probed by LIF after a time delay of typically 50 ns, by a frequency-doubled dye laser (281.91 nm). Experiments in six different H2O2/H2O mixtures of known compositions show that both signals are linearly dependent on respective species concentration. For the H2O2 detection there is a minor interfering signal contribution from OH fragments created by two-photon photodissociation of H2O. Since the PF-LIF signal yield from H2O2 is found to be at least ∼24,000 times higher than the PF-LIF signal yield from H2O at room temperature, this interference is negligible for most H2O/H2O2 mixtures of practical interest. Simultaneous single-shot imaging of both species was demonstrated in a slightly turbulent flow. For single-shot imaging the minimum detectable H2O2 and H2O concentration is 10 ppm and 0.5%, respectively. The proposed measurement concept could be a valuable asset in several areas, for example, in atmospheric and combustion science and research on <span class="hlt">vapor-phase</span> H2O2 sterilization in the pharmaceutical and aseptic food-packaging industries.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012PhRvE..86a1404M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012PhRvE..86a1404M"><span>Haloing in bimodal magnetic colloids: The role of field-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Magnet, C.; Kuzhir, P.; Bossis, G.; Meunier, A.; Suloeva, L.; Zubarev, A.</p> <p>2012-07-01</p> <p>If a suspension of magnetic micrometer-sized and nanosized particles is subjected to a homogeneous magnetic field, the nanoparticles are attracted to the microparticles and form thick anisotropic halos (clouds) around them. Such clouds can hinder the approach of microparticles and result in effective repulsion between them [M. T. López-López, A. Yu. Zubarev, and G. Bossis, Soft Matter10.1039/c0sm00261e 6, 4346 (2010)]. In this paper, we present detailed experimental and theoretical studies of nanoparticle concentration profiles and of the equilibrium shapes of nanoparticle clouds around a single magnetized microsphere, taking into account interactions between nanoparticles. We show that at a strong enough magnetic field, the ensemble of nanoparticles experiences a gas-liquid <span class="hlt">phase</span> transition such that a dense liquid <span class="hlt">phase</span> is condensed around the magnetic poles of a microsphere while a dilute gas <span class="hlt">phase</span> occupies the rest of the suspension volume. Nanoparticle accumulation around a microsphere is governed by two dimensionless parameters—the initial nanoparticle concentration (φ0) and the magnetic-to-thermal energy ratio (α)—and the three accumulation regimes are mapped onto a α-φ0 <span class="hlt">phase</span> diagram. Our local thermodynamic equilibrium approach gives a semiquantitative agreement with the experiments on the equilibrium shapes of nanoparticle clouds. The results of this work could be useful for the development of the bimodal magnetorheological fluids and of the magnetic <span class="hlt">separation</span> technologies used in bioanalysis and water purification systems.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_12");'>12</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li class="active"><span>14</span></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_14 --> <div id="page_15" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li class="active"><span>15</span></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="281"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011IJT....32..706Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011IJT....32..706Y"><span>Distillation <span class="hlt">Separation</span> of Hydrofluoric Acid and Nitric Acid from Acid Waste Using the Salt Effect on <span class="hlt">Vapor</span>-Liquid Equilibrium</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yamamoto, Hideki; Sumoge, Iwao</p> <p>2011-03-01</p> <p>This study presents the distillation <span class="hlt">separation</span> of hydrofluoric acid with use of the salt effect on the <span class="hlt">vapor</span>-liquid equilibrium for acid aqueous solutions and acid mixtures. The <span class="hlt">vapor</span>-liquid equilibrium of hydrofluoric acid + salt systems (fluorite, potassium nitrate, cesium nitrate) was measured using an apparatus made of perfluoro alkylvinylether. Cesium nitrate showed a salting-out effect on the <span class="hlt">vapor</span>-liquid equilibrium of the hydrofluoric acid-water system. Fluorite and potassium nitrate showed a salting-in effect on the hydrofluoric acid-water system. <span class="hlt">Separation</span> of hydrofluoric acid from an acid mixture containing nitric acid and hydrofluoric acid was tested by the simple distillation treatment using the salt effect of cesium nitrate (45 mass%). An acid mixture of nitric acid (5.0 mol · dm-3) and hydrofluoric acid (5.0 mol · dm-3) was prepared as a sample solution for distillation tests. The concentration of nitric acid in the first distillate decreased from 5.0 mol · dm-3 to 1.13 mol · dm-3, and the concentration of hydrofluoric acid increased to 5.41 mol · dm-3. This first distillate was further distilled without the addition of salt. The concentrations of hydrofluoric acid and nitric acid in the second distillate were 7.21 mol · dm-3 and 0.46 mol · dm-3, respectively. It was thus found that the salt effect on <span class="hlt">vapor</span>-liquid equilibrium of acid mixtures was effective for the recycling of acids from acid mixture wastes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1016129','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/1016129"><span>Condensed <span class="hlt">phase</span> conversion and growth of nanorods and other materials instead of from <span class="hlt">vapor</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Geohegan, David B.; Seals, Roland D.; Puretzky, Alex A.; Fan, Xudong</p> <p>2010-10-19</p> <p>Compositions, systems and methods are described for condensed <span class="hlt">phase</span> conversion and growth of nanorods and other materials. A method includes providing a condensed <span class="hlt">phase</span> matrix material; and activating the condensed <span class="hlt">phase</span> matrix material to produce a plurality of nanorods by condensed <span class="hlt">phase</span> conversion and growth from the condensed <span class="hlt">phase</span> matrix material instead of from <span class="hlt">vapor</span>. The compositions are very strong. The compositions and methods provide advantages because they allow (1) formation rates of nanostructures necessary for reasonable production rates, and (2) the near net shaped production of component structures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22415964-dynamical-mean-field-theory-weakly-non-linear-analysis-phase-separation-active-brownian-particles','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22415964-dynamical-mean-field-theory-weakly-non-linear-analysis-phase-separation-active-brownian-particles"><span>Dynamical mean-field theory and weakly non-linear analysis for the <span class="hlt">phase</span> <span class="hlt">separation</span> of active Brownian particles</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Speck, Thomas; Menzel, Andreas M.; Bialké, Julian</p> <p>2015-06-14</p> <p>Recently, we have derived an effective Cahn-Hilliard equation for the <span class="hlt">phase</span> <span class="hlt">separation</span> dynamics of active Brownian particles by performing a weakly non-linear analysis of the effective hydrodynamic equations for density and polarization [Speck et al., Phys. Rev. Lett. 112, 218304 (2014)]. Here, we develop and explore this strategy in more detail and show explicitly how to get to such a large-scale, mean-field description starting from the microscopic dynamics. The effective free energy emerging from this approach has the form of a conventional Ginzburg-Landau function. On the coarsest scale, our results thus agree with the mapping of active <span class="hlt">phase</span> <span class="hlt">separation</span> ontomore » that of passive fluids with attractive interactions through a global effective free energy (motility-<span class="hlt">induced</span> <span class="hlt">phase</span> transition). Particular attention is paid to the square-gradient term necessary for the <span class="hlt">phase</span> <span class="hlt">separation</span> kinetics. We finally discuss results from numerical simulations corroborating the analytical results.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014MSMSE..22e5005W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014MSMSE..22e5005W"><span>A three-dimensional <span class="hlt">phase</span> field model for nanowire growth by the <span class="hlt">vapor</span>-liquid-solid mechanism</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, Yanming; Ryu, Seunghwa; McIntyre, Paul C.; Cai, Wei</p> <p>2014-07-01</p> <p>We present a three-dimensional multi-<span class="hlt">phase</span> field model for catalyzed nanowire (NW) growth by the <span class="hlt">vapor</span>-liquid-solid (VLS) mechanism. The equation of motion contains both a Ginzburg-Landau term for deposition and a diffusion (Cahn-Hilliard) term for interface relaxation without deposition. Direct deposition from <span class="hlt">vapor</span> to solid, which competes with NW crystal growth through the molten catalyst droplet, is suppressed by assigning a very small kinetic coefficient at the solid-<span class="hlt">vapor</span> interface. The thermodynamic self-consistency of the model is demonstrated by its ability to reproduce the equilibrium contact angles at the VLS junction. The incorporation of orientation dependent gradient energy leads to faceting of the solid-liquid and solid-<span class="hlt">vapor</span> interfaces. The model successfully captures the curved shape of the NW base and the Gibbs-Thomson effect on growth velocity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AIPC.1817b0007A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AIPC.1817b0007A"><span>Effect of freezing temperature in thermally <span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> method in hydroxyapatite/chitosan-based bone scaffold biomaterial</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Albab, Muh Fadhil; Yuwono, Akhmad Herman; Sofyan, Nofrijon; Ramahdita, Ghiska</p> <p>2017-02-01</p> <p>In the current study, hydroxyapatite (HA)/chitosan-based bone scaffold has been fabricated using Thermally <span class="hlt">Induced</span> <span class="hlt">Phase</span> <span class="hlt">Separation</span> (TIPS) method under freezing temperature variation of -20, -30, -40 and -80 °C. The samples with weight percent ratio of 70% HA and 30% chitosan were homogeneously mixed and subsequently dissolved in 2% acetic acid. The synthesized samples were further characterized using Fourier transform infrared (FTIR), compressive test and scanning electron microscope (SEM). The investigation results showed that low freezing temperature reduced the pore size and increased the compressive strength of the scaffold. In the freezing temperature of -20 °C, the pore size was 133.93 µm with the compressive strength of 5.9 KPa, while for -80 °C, the pore size declined to 60.55 µm with the compressive strength 29.8 KPa. Considering the obtained characteristics, HA/chitosan obtained in this work has potential to be applied as a bone scaffold.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15267695','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15267695"><span>Scaling behavior of nonisothermal <span class="hlt">phase</span> <span class="hlt">separation</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Rüllmann, Max; Alig, Ingo</p> <p>2004-04-22</p> <p>The <span class="hlt">phase</span> <span class="hlt">separation</span> process in a critical mixture of polydimethylsiloxane and polyethylmethylsiloxane (PDMS/PEMS, a system with an upper critical solution temperature) was investigated by time-resolved light scattering during continuous quenches from the one-<span class="hlt">phase</span> into the two-<span class="hlt">phase</span> region. Continuous quenches were realized by cooling ramps with different cooling rates kappa. <span class="hlt">Phase</span> <span class="hlt">separation</span> kinetics is studied by means of the temporal evolution of the scattering vector qm and the intensity Im at the scattering peak. The curves qm(t) for different cooling rates can be shifted onto a single mastercurve. The curves Im(t) show similar behavior. As shift factors, a characteristic length Lc and a characteristic time tc are introduced. Both characteristic quantities depend on the cooling rate through power laws: Lc approximately kappa(-delta) and tc approximately kappa(-rho). Scaling behavior in isothermal critical demixing is well known. There the temporal evolutions of qm and Im for different quench depths DeltaT can be scaled with the correlation length xi and the interdiffusion coefficient D, both depending on DeltaT through critical power laws. We show in this paper that the cooling rate scaling in nonisothermal demixing is a consequence of the quench depth scaling in the isothermal case. The exponents delta and rho are related to the critical exponents nu and nu* of xi and D, respectively. The structure growth during nonisothermal demixing can be described with a semiempirical model based on the hydrodynamic coarsening mechanism well known in the isothermal case. In very late stages of nonisothermal <span class="hlt">phase</span> <span class="hlt">separation</span> a secondary scattering maximum appears. This is due to secondary demixing. We explain the onset of secondary demixing by a competition between interdiffusion and coarsening. (c) 2004 American Institute of Physics</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011ApPhB.102..345P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011ApPhB.102..345P"><span>Mid-infrared laser-absorption diagnostic for <span class="hlt">vapor-phase</span> fuel mole fraction and liquid fuel film thickness</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Porter, J. M.; Jeffries, J. B.; Hanson, R. K.</p> <p>2011-02-01</p> <p>A novel two-wavelength mid-infrared laser-absorption diagnostic has been developed for simultaneous measurements of <span class="hlt">vapor-phase</span> fuel mole fraction and liquid fuel film thickness. The diagnostic was demonstrated for time-resolved measurements of n-dodecane liquid films in the absence and presence of n-decane <span class="hlt">vapor</span> at 25°C and 1 atm. Laser wavelengths were selected from FTIR measurements of the C-H stretching band of <span class="hlt">vapor</span> n-decane and liquid n-dodecane near 3.4 μm (3000 cm-1). n-Dodecane film thicknesses <20 μm were accurately measured in the absence of <span class="hlt">vapor</span>, and simultaneous measurements of n-dodecane liquid film thickness and n-decane <span class="hlt">vapor</span> mole fraction (300 ppm) were measured with <10% uncertainty for film thicknesses <10 μm. A potential application of the measurement technique is to provide accurate values of <span class="hlt">vapor</span> mole fraction in combustion environments where strong absorption by liquid fuel or oil films on windows make conventional direct absorption measurements of the gas problematic.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1400296-stress-triggered-phase-separation-adaptive-evolutionarily-tuned-response','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1400296-stress-triggered-phase-separation-adaptive-evolutionarily-tuned-response"><span>Stress-Triggered <span class="hlt">Phase</span> <span class="hlt">Separation</span> Is an Adaptive, Evolutionarily Tuned Response</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Riback, Joshua A.; Katanski, Christopher D.; Kear-Scott, Jamie L.</p> <p></p> <p>In eukaryotic cells, diverse stresses trigger coalescence of RNA-binding proteins into stress granules. In vitro, stress-granule-associated proteins can demix to form liquids, hydrogels, and other assemblies lacking fixed stoichiometry. Observing these phenomena has generally required conditions far removed from physiological stresses. We show that poly(A)-binding protein (Pab1 in yeast), a defining marker of stress granules, <span class="hlt">phase</span> <span class="hlt">separates</span> and forms hydrogels in vitro upon exposure to physiological stress conditions. Other RNA-binding proteins depend upon low-complexity regions (LCRs) or RNA for <span class="hlt">phase</span> <span class="hlt">separation</span>, whereas Pab1’s LCR is not required for demixing, and RNA inhibits it. Based on unique evolutionary patterns, we createmore » LCR mutations, which systematically tune its biophysical properties and Pab1 <span class="hlt">phase</span> <span class="hlt">separation</span> in vitro and in vivo. Mutations that impede <span class="hlt">phase</span> <span class="hlt">separation</span> reduce organism fitness during prolonged stress. Poly(A)-binding protein thus acts as a physiological stress sensor, exploiting <span class="hlt">phase</span> <span class="hlt">separation</span> to precisely mark stress onset, a broadly generalizable mechanism.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19850002944','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19850002944"><span>External fuel <span class="hlt">vaporization</span> study, <span class="hlt">phase</span> 2</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Szetela, E. J.; Chiappetta, L.</p> <p>1981-01-01</p> <p>An analytical study was conducted to evaluate the effect of variations in fuel properties on the design of an external fuel vaporizaton system. The fuel properties that were considered included thermal stability, critical temperature, enthalpy a critical conditions, volatility, and viscosity. The design parameters that were evaluated included <span class="hlt">vaporizer</span> weight and the impact on engine requirement such as maintenance, transient response, performance, and altitude relight. The baseline fuel properties were those of Jet A. The variation in thermal stability was taken as the thermal stability variation for Experimental Referee Broad Specification (ERBS) fuel. The results of the analysis indicate that a change in thermal stability equivalent to that of ERBS would increase the <span class="hlt">vaporization</span> system weight by 20 percent, decrease oprating time between cleaning by 40 percent and make altitude relight more difficult. An increase in fuel critical temperature of 39 K would require a 40 percent increase in <span class="hlt">vaporization</span> system weight. The assumed increase in enthalpy and volatility would also increase <span class="hlt">vaporizer</span> weight by 40 percent and make altitude relight extremely difficult. The variation in fuel viscosity would have a negligible effect on the design parameters.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1484149','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1484149"><span>Water Sorption and <span class="hlt">Vapor-Phase</span> Deuterium Exchange Studies on Methemoglobin CC, SC, SS, AS, and AA</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Killion, Philip J.; Cameron, Bruce F.</p> <p>1972-01-01</p> <p>Five hemoglobins whose genetic relationship to one another involves one set of alleles, hemoglobins CC, SC, SS, AS, and AA, were studied in the Met form. Two different investigations were conducted at 28°C on these methemoglobins within a McBain gravimetric sorption system: sorption of H2O <span class="hlt">vapor</span> and <span class="hlt">vapor-phase</span> deuterium-hydrogen exchange. For each of the five samples there was close agreement between the per cent hydration of polar sites as determined from sorption studies and the maximum per cent of labile hydrogens that were exchanged during the <span class="hlt">vapor-phase</span> deuterium exchange study. Both studies measured a slight increase in the number of polar sites accessible to H2O or D2O <span class="hlt">vapor</span> for those samples in which the substituent in the sixth position from the N-terminus of the two β-chains had a positively charged side chain and a slight decrease for those in which the substituent had a negatively charged side chain. The in-exchange of deuterium for hydrogen occurred at a faster observed rate than the out-exchange of hydrogen for deuterium. PMID:5030563</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19800057777&hterms=Thermodynamic+parameter&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3DThermodynamic%2Bparameter','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19800057777&hterms=Thermodynamic+parameter&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3DThermodynamic%2Bparameter"><span>Methods of calculating engineering parameters for gas <span class="hlt">separations</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Lawson, D. D.</p> <p>1980-01-01</p> <p>A group additivity method has been generated which makes it possible to estimate, from the structural formulas alone, the energy of <span class="hlt">vaporization</span> and the molar volume at 25 C of many nonpolar organic liquids. From these two parameters and appropriate thermodynamic relationships it is then possible to predict the <span class="hlt">vapor</span> pressure of the liquid <span class="hlt">phase</span> and the solubility of various gases in nonpolar organic liquids. The data are then used to evaluate organic and some inorganic liquids for use in gas <span class="hlt">separation</span> stages or as heat exchange fluids in prospective thermochemical cycles for hydrogen production.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhRvF...2f3902K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhRvF...2f3902K"><span>Influence of <span class="hlt">phase</span> transition on the instability of a liquid-<span class="hlt">vapor</span> interface in a gravitational field</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Konovalov, V. V.; Lyubimov, D. V.; Lyubimova, T. P.</p> <p>2017-06-01</p> <p>This study is concerned with the linear stability of the horizontal interface between thick layers of a viscous heat-conducting liquid and its <span class="hlt">vapor</span> in a gravitational field subject to <span class="hlt">phase</span> transition. We consider the case when the hydrostatic base state is consistent with a balanced heat flux at the liquid-<span class="hlt">vapor</span> interface. The corrections to the growth rate of the most dangerous perturbations and cutoff wave number, characterizing the influence of <span class="hlt">phase</span> transition on the Rayleigh-Taylor instability, are found to be different from the data in the literature. Most of the previous results were obtained in the framework of a quasiequilibrium approximation, which had been shown to conform to the limit of thin media layers under equality of the interface temperature to a saturation temperature. The main difference from the results obtained with the quasiequilibrium approach is new values of the proportionality coefficients that correlate our corrections with the intensity of weak heating. Moreover, at large values of the heat flux rate, when deviations from the approximate linear law are important, the effect of <span class="hlt">phase</span> transition is limited and does not exceed the size of the <span class="hlt">vapor</span> viscosity effect.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25233236','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25233236"><span>The mechanism of <span class="hlt">vapor</span> <span class="hlt">phase</span> hydration of calcium oxide: implications for CO2 capture.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kudłacz, Krzysztof; Rodriguez-Navarro, Carlos</p> <p>2014-10-21</p> <p>Lime-based sorbents are used for fuel- and flue-gas capture, thereby representing an economic and effective way to reduce CO2 emissions. Their use involves cyclic carbonation/calcination which results in a significant conversion reduction with increasing number of cycles. To reactivate spent CaO, <span class="hlt">vapor</span> <span class="hlt">phase</span> hydration is typically performed. However, little is known about the ultimate mechanism of such a hydration process. Here, we show that the <span class="hlt">vapor</span> <span class="hlt">phase</span> hydration of CaO formed after calcination of calcite (CaCO3) single crystals is a pseudomorphic, topotactic process, which progresses via an intermediate disordered <span class="hlt">phase</span> prior to the final formation of oriented Ca(OH)2 nanocrystals. The strong structural control during this solid-state <span class="hlt">phase</span> transition implies that the microstructural features of the CaO parent <span class="hlt">phase</span> predetermine the final structural and physicochemical (reactivity and attrition) features of the product hydroxide. The higher molar volume of the product can create an impervious shell around unreacted CaO, thereby limiting the efficiency of the reactivation process. However, in the case of compact, sintered CaO structures, volume expansion cannot be accommodated in the reduced pore volume, and stress generation leads to pervasive cracking. This favors complete hydration but also detrimental attrition. Implications of these results in carbon capture and storage (CCS) are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1174920','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/1174920"><span><span class="hlt">Separation</span> process using pervaporation and dephlegmation</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Vane, Leland M.; Mairal, Anurag P.; Ng, Alvin; Alvarez, Franklin R.; Baker, Richard W.</p> <p>2004-06-29</p> <p>A process for treating liquids containing organic compounds and water. The process includes a pervaporation step in conjunction with a dephlegmation step to treat at least a portion of the permeate <span class="hlt">vapor</span> from the pervaporation step. The process yields a membrane residue stream, a stream enriched in the more volatile component (usually the organic) as the overhead stream from the dephlegmator and a condensate stream enriched in the less volatile component (usually the water) as a bottoms stream from the dephlegmator. Any of these may be the principal product of the process. The membrane <span class="hlt">separation</span> step may also be performed in the <span class="hlt">vapor</span> <span class="hlt">phase</span>, or by membrane distillation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JPCM...30r5801G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JPCM...30r5801G"><span>Effect of atomic disorder on the magnetic <span class="hlt">phase</span> <span class="hlt">separation</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Groshev, A. G.; Arzhnikov, A. K.</p> <p>2018-05-01</p> <p>The effect of disorder on the magnetic <span class="hlt">phase</span> <span class="hlt">separation</span> between the antiferromagnetic and incommensurate helical and <span class="hlt">phases</span> is investigated. The study is based on the quasi-two-dimensional single-band Hubbard model in the presence of atomic disorder (the Anderson–Hubbard model). A model of binary alloy disorder is considered, in which the disorder is determined by the difference in energy between the host and impurity atomic levels at a fixed impurity concentration. The problem is solved within the theory of functional integration in static approximation. Magnetic <span class="hlt">phase</span> diagrams are obtained as functions of the temperature, the number of electrons and impurity concentration with allowance for <span class="hlt">phase</span> <span class="hlt">separation</span>. It is shown that for the model parameters chosen, the disorder caused by impurities whose atomic-level energy is greater than that of the host atomic levels, leads to qualitative changes in the <span class="hlt">phase</span> diagram of the impurity-free system. In the opposite case, only quantitative changes occur. The peculiarities of the effect of disorder on the <span class="hlt">phase</span> <span class="hlt">separation</span> regions of the quasi-two-dimensional Hubbard model are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017RuMet2017...24U','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017RuMet2017...24U"><span>Ordering-<span class="hlt">separation</span> <span class="hlt">phase</span> transitions in a Co3V alloy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ustinovshchikov, Yu. I.</p> <p>2017-01-01</p> <p>The microstructure of the Co3V alloy formed by heat treatment at various temperatures is studied by transmission electron microscopy. Two ordering-<span class="hlt">separation</span> <span class="hlt">phase</span> transitions are revealed at temperatures of 400-450 and 800°C. At the high-temperature <span class="hlt">phase</span> <span class="hlt">separation</span>, the microstructure consists of bcc vanadium particles and an fcc solid solution; at the low-temperature <span class="hlt">phase</span> <span class="hlt">separation</span>, the microstructure is cellular. In the ordering range, the microstructure consists of chemical compound Co3V particles chaotically arranged in the solid solution. The structure of the Co3V alloy is shown not to correspond to the structures indicated in the Co-V <span class="hlt">phase</span> diagram at any temperatures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5437812','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5437812"><span>Adsorptive Water Removal from Dichloromethane and <span class="hlt">Vapor-Phase</span> Regeneration of a Molecular Sieve 3A Packed Bed</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p></p> <p>2017-01-01</p> <p>The drying of dichloromethane with a molecular sieve 3A packed bed process is modeled and experimentally verified. In the process, the dichloromethane is dried in the liquid <span class="hlt">phase</span> and the adsorbent is regenerated by water desorption with dried dichloromethane product in the <span class="hlt">vapor</span> <span class="hlt">phase</span>. Adsorption equilibrium experiments show that dichloromethane does not compete with water adsorption, because of size exclusion; the pure water <span class="hlt">vapor</span> isotherm from literature provides an accurate representation of the experiments. The breakthrough curves are adequately described by a mathematical model that includes external mass transfer, pore diffusion, and surface diffusion. During the desorption step, the main heat transfer mechanism is the condensation of the superheated dichloromethane <span class="hlt">vapor</span>. The regeneration time is shortened significantly by external bed heating. Cyclic steady-state experiments demonstrate the feasibility of this novel, zero-emission drying process. PMID:28539701</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25660716','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25660716"><span>Integration of <span class="hlt">phase</span> <span class="hlt">separation</span> with ultrasound-assisted salt-<span class="hlt">induced</span> liquid-liquid microextraction for analyzing the fluoroquinones in human body fluids by liquid chromatography.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Huili; Gao, Ming; Wang, Mei; Zhang, Rongbo; Wang, Wenwei; Dahlgren, Randy A; Wang, Xuedong</p> <p>2015-03-15</p> <p>Herein, we developed a novel integrated device to perform <span class="hlt">phase</span> <span class="hlt">separation</span> based on ultrasound-assisted salt-<span class="hlt">induced</span> liquid-liquid microextraction for determination of five fluoroquinones (FQs) in human body fluids. The integrated device consisted of three simple HDPE components used to <span class="hlt">separate</span> the extraction solvent from the aqueous <span class="hlt">phase</span> prior to retrieving the extractant. A series of extraction parameters were optimized using the response surface method based on central composite design. Optimal conditions consisted of 945μL acetone extraction solvent, pH 2.1, 4.1min stir time, 5.9g Na2SO4, and 4.0min centrifugation. Under optimized conditions, the limits of detection (at S/N=3) were 0.12-0.66μgL(-1), the linear range was 0.5-500μgL(-1) and recoveries were 92.6-110.9% for the five FQs extracted from plasma and urine. The proposed method has several advantages, such as easy construction from inexpensive materials, high extraction efficiency, short extraction time, and compatibility with HPLC analysis. Thus, this method shows excellent prospects for sample pretreatment and analysis of FQs in human body fluids. Copyright © 2015 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18232788','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18232788"><span>Preventing kinetic roughening in physical <span class="hlt">vapor-phase</span>-deposited films.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Vasco, E; Polop, C; Sacedón, J L</p> <p>2008-01-11</p> <p>The growth kinetics of the mostly used physical <span class="hlt">vapor-phase</span> deposition techniques -molecular beam epitaxy, sputtering, flash evaporation, and pulsed laser deposition-is investigated by rate equations with the aim of testing their suitability for the preparation of ultraflat ultrathin films. The techniques are studied in regard to the roughness and morphology during early stages of growth. We demonstrate that pulsed laser deposition is the best technique for preparing the flattest films due to two key features [use of (i) a supersaturated pulsed flux of (ii) hyperthermal species] that promote a kinetically limited Ostwald ripening mechanism.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012SPIE.8553E..1YG','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012SPIE.8553E..1YG"><span>The threshold of <span class="hlt">vapor</span> channel formation in water <span class="hlt">induced</span> by pulsed CO2 laser</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Guo, Wenqing; Zhang, Xianzeng; Zhan, Zhenlin; Xie, Shusen</p> <p>2012-12-01</p> <p>Water plays an important role in laser ablation. There are two main interpretations of laser-water interaction: hydrokinetic effect and <span class="hlt">vapor</span> phenomenon. The two explanations are reasonable in some way, but they can't explain the mechanism of laser-water interaction completely. In this study, the dynamic process of <span class="hlt">vapor</span> channel formation <span class="hlt">induced</span> by pulsed CO2 laser in static water layer was monitored by high-speed camera. The wavelength of pulsed CO2 laser is 10.64 um, and pulse repetition rate is 60 Hz. The laser power ranged from 1 to 7 W with a step of 0.5 W. The frame rate of high-speed camera used in the experiment was 80025 fps. Based on high-speed camera pictures, the dynamic process of <span class="hlt">vapor</span> channel formation was examined, and the threshold of <span class="hlt">vapor</span> channel formation, pulsation period, the volume, the maximum depth and corresponding width of <span class="hlt">vapor</span> channel were determined. The results showed that the threshold of <span class="hlt">vapor</span> channel formation was about 2.5 W. Moreover, pulsation period, the maximum depth and corresponding width of <span class="hlt">vapor</span> channel increased with the increasing of the laser power.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_13");'>13</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li class="active"><span>15</span></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_15 --> <div id="page_16" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li class="active"><span>16</span></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="301"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1988QuEle..18.1412J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1988QuEle..18.1412J"><span>INTERNATIONAL CONFERENCE ON SEMICONDUCTOR INJECTION LASERS SELCO-87: Metal-organic <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy of (GaAl)As for 0.85-μm laser diodes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jacobs, K.; Bugge, F.; Butzke, G.; Lehmann, L.; Schimko, R.</p> <p>1988-11-01</p> <p>Metal-organic <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy was used to grow stripe heterolaser diodes that were hitherto fabricated by liquid <span class="hlt">phase</span> epitaxy. The main relationships between the growth parameters (partial input pressures, temperatures) and the properties of materials (thicknesses, solid-solution compositions, carrier densities) were investigated. The results were in full agreement with the mechanism of growth controlled by a <span class="hlt">vapor-phase</span> diffusion. The results achieved routinely in the growth of GaAs are reported. It is shown that double heterostructure laser diodes fabricated by metal-organic <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy compete favorably with those grown so far by liquid <span class="hlt">phase</span> epitaxy, including their degradation and reliability.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009ApPhB..97..215P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009ApPhB..97..215P"><span>Mid-infrared laser-absorption diagnostic for <span class="hlt">vapor-phase</span> measurements in an evaporating n-decane aerosol</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Porter, J. M.; Jeffries, J. B.; Hanson, R. K.</p> <p>2009-09-01</p> <p>A novel three-wavelength mid-infrared laser-based absorption/extinction diagnostic has been developed for simultaneous measurement of temperature and <span class="hlt">vapor-phase</span> mole fraction in an evaporating hydrocarbon fuel aerosol (<span class="hlt">vapor</span> and liquid droplets). The measurement technique was demonstrated for an n-decane aerosol with D 50˜3 μ m in steady and shock-heated flows with a measurement bandwidth of 125 kHz. Laser wavelengths were selected from FTIR measurements of the C-H stretching band of <span class="hlt">vapor</span> and liquid n-decane near 3.4 μm (3000 cm -1), and from modeled light scattering from droplets. Measurements were made for <span class="hlt">vapor</span> mole fractions below 2.3 percent with errors less than 10 percent, and simultaneous temperature measurements over the range 300 K< T<900 K were made with errors less than 3 percent. The measurement technique is designed to provide accurate values of temperature and <span class="hlt">vapor</span> mole fraction in evaporating polydispersed aerosols with small mean diameters ( D 50<10 μ m), where near-infrared laser-based scattering corrections are prone to error.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007JMMM..310..870M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007JMMM..310..870M"><span>Nanoscale ferromagnetism in <span class="hlt">phase-separated</span> manganites</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mori, S.; Horibe, Y.; Asaka, T.; Matsui, Y.; Chen, C. H.; Cheong, S. W.</p> <p>2007-03-01</p> <p>Magnetic domain structures in <span class="hlt">phase-separated</span> manganites were investigated by low-temperature Lorentz electron microscopy, in order to understand some unusual physical properties such as a colossal magnetoresistance (CMR) effect and a metal-to-insulator transition. In particular, we examined a spatial distribution of the charge/orbital-ordered (CO/OO) insulator state and the ferromagnetic (FM) metallic one in <span class="hlt">phase-separated</span> manganites; Cr-doped Nd0.5Ca0.5MnO3 and ( La1-xPrx)CaMnO3 with x=0.375, by obtaining both the dark-field images and Lorentz electron microscopic ones. It is found that an unusual coexistence of the CO/OO and FM metallic states below a FM transition temperature in the two compounds. The present experimental results clearly demonstrated the coexisting state of the two distinct ground states in manganites.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=69777&keyword=enzyme+AND+function&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50','EPA-EIMS'); return false;" href="https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=69777&keyword=enzyme+AND+function&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50"><span>EFFECT OF <span class="hlt">VAPOR-PHASE</span> BIOREACTOR OPERATION ON BIOMASS ACCUMULATION, DISTRIBUTION, AND ACTIVITY. (R826168)</span></a></p> <p><a target="_blank" href="http://oaspub.epa.gov/eims/query.page">EPA Science Inventory</a></p> <p></p> <p></p> <p>Excess biomass accumulation and activity loss in <span class="hlt">vapor-phase</span> bioreactors (VPBs) can lead to unreliable long-term operation. In this study, temporal and spatial variations in biomass accumulation, distribution and activity in VPBs treating toluene-contaminated air were monitored o...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015PhRvA..92e3803M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015PhRvA..92e3803M"><span>Time-dependent <span class="hlt">phase</span> shift of a retrieved pulse in off-resonant electromagnetically-<span class="hlt">induced</span>-transparency-based light storage</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Maynard, M.-A.; Bouchez, R.; Lugani, J.; Bretenaker, F.; Goldfarb, F.; Brion, E.</p> <p>2015-11-01</p> <p>We report measurements of the time-dependent <span class="hlt">phases</span> of the leak and retrieved pulses obtained in electromagnetically-<span class="hlt">induced</span>-transparency storage experiments with metastable helium <span class="hlt">vapor</span> at room temperature. In particular, we investigate the influence of the optical detuning at two-photon resonance and provide numerical simulations of the full dynamical Maxwell-Bloch equations, which allow us to account for the experimental results.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70015397','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70015397"><span>Removing volatile contaminants from the unsaturated zone by <span class="hlt">inducing</span> advective air-<span class="hlt">phase</span> transport</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Baehr, A.L.; Hoag, G.E.; Marley, M.C.</p> <p>1989-01-01</p> <p>Organic liquids inadvertently spilled and then distributed in the unsaturated zone can pose a long-term threat to ground water. Many of these substances have significant volatility, and thereby establish a premise for contaminant removal from the unsaturated zone by <span class="hlt">inducing</span> advective air-<span class="hlt">phase</span> transport with wells screened in the unsaturated zone. In order to focus attention on the rates of mass transfer from liquid to vapour <span class="hlt">phases</span>, sand columns were partially saturated with gasoline and vented under steady air-flow conditions. The ability of an equilibrium-based transport model to predict the hydrocarbon <span class="hlt">vapor</span> flux from the columns implies an efficient rate of local <span class="hlt">phase</span> transfer for reasonably high air-<span class="hlt">phase</span> velocities. Thus the success of venting remediations will depend primarily on the ability to <span class="hlt">induce</span> an air-flow field in a heterogeneous unsaturated zone that will intersect the distributed contaminant. To analyze this aspect of the technique, a mathematical model was developed to predict radially symmetric air flow <span class="hlt">induced</span> by venting from a single well. This model allows for in-situ determinations of air-<span class="hlt">phase</span> permeability, which is the fundamental design parameter, and for the analysis of the limitations of a single well design. A successful application of the technique at a site once contaminated by gasoline supports the optimism derived from the experimental and modeliing <span class="hlt">phases</span> of this study, and illustrates the well construction and field methods used to document the volatile contaminant recovery. ?? 1989.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19820054876&hterms=gallium+vapor&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dgallium%2Bvapor','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19820054876&hterms=gallium+vapor&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dgallium%2Bvapor"><span>OM-VPE growth of Mg-doped GaAs. [OrganoMetallic-<span class="hlt">Vapor</span> <span class="hlt">Phase</span> Epitaxy</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Lewis, C. R.; Dietze, W. T.; Ludowise, M. J.</p> <p>1982-01-01</p> <p>The epitaxial growth of Mg-doped GaAs by the organometallic <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxial process (OM-VPE) has been achieved for the first time. The doping is controllable over a wide range of input fluxes of bis (cyclopentadienyl) magnesium, (C5H5)2Mg, the organometallic precursor to Mg.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..MAR.L6010G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..MAR.L6010G"><span>Modeling <span class="hlt">phase</span> <span class="hlt">separation</span> in mixtures of intrinsically-disordered proteins</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gu, Chad; Zilman, Anton</p> <p></p> <p><span class="hlt">Phase</span> <span class="hlt">separation</span> in a pure or mixed solution of intrinsically-disordered proteins (IDPs) and its role in various biological processes has generated interest from the theoretical biophysics community. <span class="hlt">Phase</span> <span class="hlt">separation</span> of IDPs has been implicated in the formation of membrane-less organelles such as nucleoli, as well as in a mechanism of selectivity in transport through the nuclear pore complex. Based on a lattice model of polymers, we study the <span class="hlt">phase</span> diagram of IDPs in a mixture and describe the selective exclusion of soluble proteins from the dense-<span class="hlt">phase</span> IDP aggregates. The model captures the essential behaviour of <span class="hlt">phase</span> <span class="hlt">separation</span> by a minimal set of coarse-grained parameters, corresponding to the average monomer-monomer and monomer-protein attraction strength, as well as the protein-to-monomer size ratio. Contrary to the intuition that strong monomer-monomer interaction increases exclusion of soluble proteins from the dense IDP aggregates, our model predicts that the concentration of soluble proteins in the aggregate <span class="hlt">phase</span> as a function of monomer-monomer attraction is non-monotonic. We corroborate the predictions of the lattice model using Langevin dynamics simulations of grafted polymers in planar and cylindrical geometries, mimicking various in-vivo and in-vitro conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29583123','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29583123"><span>Effect of atomic disorder on the magnetic <span class="hlt">phase</span> <span class="hlt">separation</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Groshev, A G; Arzhnikov, A K</p> <p>2018-05-10</p> <p>The effect of disorder on the magnetic <span class="hlt">phase</span> <span class="hlt">separation</span> between the antiferromagnetic and incommensurate helical [Formula: see text] and [Formula: see text] <span class="hlt">phases</span> is investigated. The study is based on the quasi-two-dimensional single-band Hubbard model in the presence of atomic disorder (the [Formula: see text] Anderson-Hubbard model). A model of binary alloy disorder is considered, in which the disorder is determined by the difference in energy between the host and impurity atomic levels at a fixed impurity concentration. The problem is solved within the theory of functional integration in static approximation. Magnetic <span class="hlt">phase</span> diagrams are obtained as functions of the temperature, the number of electrons and impurity concentration with allowance for <span class="hlt">phase</span> <span class="hlt">separation</span>. It is shown that for the model parameters chosen, the disorder caused by impurities whose atomic-level energy is greater than that of the host atomic levels, leads to qualitative changes in the <span class="hlt">phase</span> diagram of the impurity-free system. In the opposite case, only quantitative changes occur. The peculiarities of the effect of disorder on the <span class="hlt">phase</span> <span class="hlt">separation</span> regions of the quasi-two-dimensional Hubbard model are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MSMSE..26c5015Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MSMSE..26c5015Z"><span>Effect of applied strain on <span class="hlt">phase</span> <span class="hlt">separation</span> of Fe-28 at.% Cr alloy: 3D <span class="hlt">phase</span>-field simulation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhu, Lihui; Li, Yongsheng; Liu, Chengwei; Chen, Shi; Shi, Shujing; Jin, Shengshun</p> <p>2018-04-01</p> <p>A quantitative simulation of the <span class="hlt">separation</span> of the α‧ <span class="hlt">phase</span> in Fe-28 at.% Cr alloy under the effects of applied strain is performed by utilizing a three-dimensional <span class="hlt">phase</span>-field model. The elongation of the Cr-enriched α‧ <span class="hlt">phase</span> becomes obvious with the influence of applied uniaxial strain for the <span class="hlt">phase</span> <span class="hlt">separation</span> transforms from spinodal decomposition of 700 K to nucleation and growth of 773 K. The applied strain shows a significant influence on the early stage <span class="hlt">phase</span> <span class="hlt">separation</span>, and the influence is enlarged with the elevated temperature. The steady-state coarsening with the mechanism of spinodal decomposition is substantially affected by the applied strain for low-temperature aging, while the influence is reduced as the temperature increases and as the <span class="hlt">phase</span> <span class="hlt">separation</span> mechanism changes to nucleation and growth. The peak value of particle size distribution decreases, and the PSD for 773 K becomes more widely influenced by the applied strain. The simulation results of <span class="hlt">separation</span> of the Cr-enriched α‧ <span class="hlt">phase</span> with the applied strain provide a further understanding of the strain effect on the <span class="hlt">phase</span> <span class="hlt">separation</span> of Fe-Cr alloys from the metastable region to spinodal regions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19740045739&hterms=nude&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dnude','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19740045739&hterms=nude&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Dnude"><span>A new mass spectrometer system for investigating laser-<span class="hlt">induced</span> <span class="hlt">vaporization</span> phenomena</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Lincoln, K. A.</p> <p>1974-01-01</p> <p>A laser has been combined with a mass spectrometer in a new configuration developed for studies of high-temperature materials. A vacuum-lock, solid-sample inlet is mounted at one end of a cylindrical, high-vacuum chamber one meter in length with a nude ion-source, time-of-flight mass spectrometer at the opposite end. The samples are positioned along the axis of the chamber at distances up to one meter from the ion source, and their surfaces are <span class="hlt">vaporized</span> by a pulsed laser beam entering via windows on one side of the chamber. The instrumentation along with its capabilities is described, and results from laser-<span class="hlt">induced</span> <span class="hlt">vaporization</span> of several graphites are presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016SPIE.9906E..5UA','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016SPIE.9906E..5UA"><span>ALMA long baseline <span class="hlt">phase</span> calibration using <span class="hlt">phase</span> referencing</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Asaki, Yoshiharu; Matsushita, Satoki; Fomalont, Edward B.; Corder, Stuartt A.; Nyman, Lars-Åke; Dent, William R. F.; Philips, Neil M.; Hirota, Akihiko; Takahashi, Satoko; Vila-Vilaro, Baltasar; Nikolic, Bojan; Hunter, Todd R.; Remijan, Anthony; Vlahakis, Catherine</p> <p>2016-08-01</p> <p>The Atacama Large Millimeter/submillimeter Array (ALMA) is the world's largest millimeter/submillimeter telescope and provides unprecedented sensitivities and spatial resolutions. To achieve the highest imaging capabilities, interferometric <span class="hlt">phase</span> calibration for the long baselines is one of the most important subjects: The longer the baselines, the worse the <span class="hlt">phase</span> stability becomes because of turbulent motions of the Earth's atmosphere, especially, the water <span class="hlt">vapor</span> in the troposphere. To overcome this subject, ALMA adopts a <span class="hlt">phase</span> correction scheme using a Water <span class="hlt">Vapor</span> Radiometer (WVR) to estimate the amount of water <span class="hlt">vapor</span> content along the antenna line of sight. An additional technique is <span class="hlt">phase</span> referencing, in which a science target and a nearby calibrator are observed by turn by quickly changing the antenna pointing. We conducted feasibility studies of the hybrid technique with the WVR <span class="hlt">phase</span> correction and the antenna Fast Switching (FS) <span class="hlt">phase</span> referencing (WVR+FS <span class="hlt">phase</span> correction) for the ALMA 16 km longest baselines in cases that (1) the same observing frequency both for a target and calibrator is used, and (2) higher and lower frequencies for a target and calibrator, respectively, with a typical switching cycle time of 20 s. It was found that the <span class="hlt">phase</span> correction performance of the hybrid technique is promising where a nearby calibrator is located within roughly 3◦ from a science target, and that the <span class="hlt">phase</span> correction with 20 s switching cycle time significantly improves the performance with the above <span class="hlt">separation</span> angle criterion comparing to the 120 s switching cycle time. The currently trial <span class="hlt">phase</span> calibration method shows the same performance independent of the observing frequencies. This result is especially important for the higher frequency observations because it becomes difficult to find a bright calibrator close to an arbitrary sky position. In the series of our experiments, it is also found that <span class="hlt">phase</span> errors affecting the image quality come from not only</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26979283','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26979283"><span>Superhigh moduli and tension-<span class="hlt">induced</span> <span class="hlt">phase</span> transition of monolayer gamma-boron at finite temperatures.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhao, Junhua; Yang, Zhaoyao; Wei, Ning; Kou, Liangzhi</p> <p>2016-03-16</p> <p>Two dimensional (2D) gamma-boron (γ-B28) thin films have been firstly reported by the experiments of the chemical <span class="hlt">vapor</span> deposition in the latest study. However, their mechanical properties are still not clear. Here we predict the superhigh moduli (785 ± 42 GPa at 300 K) and the tension-<span class="hlt">induced</span> <span class="hlt">phase</span> transition of monolayer γ-B28 along a zigzag direction for large deformations at finite temperatures using molecular dynamics (MD) simulations. The new <span class="hlt">phase</span> can be kept stable after unloading process at these temperatures. The predicted mechanical properties are reasonable when compared with our results from density functional theory. This study provides physical insights into the origins of the new <span class="hlt">phase</span> transition of monolayer γ-B28 at finite temperatures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26910479','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26910479"><span>Delivery of Epinephrine in the <span class="hlt">Vapor</span> <span class="hlt">Phase</span> for the Treatment of Croup.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Leung, Kitty; Newth, Christopher J L; Hotz, Justin C; O'Brien, Kevin C; Fink, James B; Coates, Allan L</p> <p>2016-04-01</p> <p>The Vapotherm system delivers high humidity to the airway of patients by using semipermeable tubules where heated liquid water is in contact with air. The humidified air is conducted to the patient via a heated tube. Preliminary clinical observations in infants with croup suggested that epinephrine added to the water supplying the humidity was delivered successfully in the <span class="hlt">vapor</span> <span class="hlt">phase</span>. The purpose of this study was to evaluate the efficiency of the delivery of epinephrine in the <span class="hlt">vapor</span> <span class="hlt">phase</span> and to develop the feasibility criteria for a clinical pilot study. Thirty milligrams of epinephrine in a 1-L bag of sterile water was used as the humidification source for a Vapotherm 2000i. The output of the heated circuit was condensed and collected into a small Erlenmeyer flask via a metal coil while the whole collection system was submerged in an ice slurry to maintain the outflow temperature from the flask between 0°C and 2°C. The in vitro system was tested at 40°C with flows of 5, 10, and 15 L/min and L-epinephrine concentrations of 15, 30, and 60 mg/L. Each test was duplicated at each of the six conditions. Academic children's hospital research laboratory. None. None. The system recovered more than 90% of the water <span class="hlt">vapor</span> from the fully saturated air at 40°C. The epinephrine concentration recovery quantified by ultraviolet-visible spectrophotometry was 23.9% (27.5-20.4%) (mean and range) of the initial concentration. At flows of 5, 10, and 15 L/min, the delivery of epinephrine would be 1.8, 3.6, and 4.2 μg/min, respectively, which is in the therapeutic range used for parenteral infusion in young children. The Vapotherm system can be used to deliver epinephrine in pharmacological doses to the respiratory system as a <span class="hlt">vapor</span> and thus as an alternative to droplets by conventional nebulization.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20080006653','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20080006653"><span>Feasibility Study of <span class="hlt">Vapor</span>-Mist <span class="hlt">Phase</span> Reaction Lubrication Using a Thioether Liquid</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Morales, Wilfredo; Handschuh, Robert F.; Krantz, Timothy L.</p> <p>2007-01-01</p> <p>A primary technology barrier preventing the operation of gas turbine engines and aircraft gearboxes at higher temperatures is the inability of currently used liquid lubricants to survive at the desired operating conditions over an extended time period. Current state-of-the-art organic liquid lubricants rapidly degrade at temperatures above 300 C; hence, another form of lubrication is necessary. <span class="hlt">Vapor</span> or mist <span class="hlt">phase</span> reaction lubrication is a unique, alternative technology for high temperature lubrication. The majority of past studies have employed a liquid phosphate ester that was <span class="hlt">vaporized</span> or misted, and delivered to bearings or gears where the phosphate ester reacted with the metal surfaces generating a solid lubricious film. This method resulted in acceptable operating temperatures suggesting some good lubrication properties, but the continuous reaction between the phosphate ester and the iron surfaces led to wear rates unacceptable for gas turbine engine or aircraft gearbox applications. In this study, an alternative non-phosphate liquid was used to mist <span class="hlt">phase</span> lubricate a spur gearbox rig operating at 10,000 rpm under highly loaded conditions. After 21 million shaft revolutions of operation the gears exhibited only minor wear.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015APS..MARM43012A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015APS..MARM43012A"><span>Liquid filament instability due to stretch-<span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> in polymer solutions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Arinstein, Arkadii; Kulichikhin, Valery; Malkin, Alexander; Technion-Israel Institute of Technology Collaboration; Institute of Petrochemical Synthesis, Russian Academy of Sciences Team</p> <p>2015-03-01</p> <p>The instability in a jet of a viscoelastic semi-dilute entangled polymer solution under high stretching is discussed. Initially, the variation in osmotic pressure can compensate for decrease in the capillary force, and the jet is stable. The further evolution of the polymer solution along the jet results in formation of a filament in the jet center and of a near-surface solvent layer. Such a redistribution of polymer seems like a ``<span class="hlt">phase</span> <span class="hlt">separation</span>'', but it is related to stretching of the jet. The viscous liquid shell demonstrates Raleigh-type instability resulting in the formation of individual droplets on the oriented filament. Experimental observations showed that this <span class="hlt">separation</span> is starting during few first seconds, and continues of about 10 -15 seconds. The modeling shows that a jet stretching results in a radial gradient in the polymer distribution: the polymer is concentrated in the jet center, whereas the solvent is remaining near the surface. The key point of this model is that a large longitudinal stretching of a polymer network results in its lateral contraction, so a solvent is pressed out of this polymer network because of the decrease in its volume. V.K. and A.M. acknowledge the financial support of the Russian Scientific Foundation (Grant 4-23-00003).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25688836','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25688836"><span>Direct analysis of intact biological macromolecules by low-energy, fiber-based femtosecond laser <span class="hlt">vaporization</span> at 1042 nm wavelength with nanospray postionization mass spectrometry.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Shi, Fengjian; Flanigan, Paul M; Archer, Jieutonne J; Levis, Robert J</p> <p>2015-03-17</p> <p>A fiber-based laser with a pulse duration of 435 fs and a wavelength of 1042 nm was used to <span class="hlt">vaporize</span> biological macromolecules intact from the condensed <span class="hlt">phase</span> into the gas <span class="hlt">phase</span> for nanospray postionization and mass analysis. Laser <span class="hlt">vaporization</span> of dried standard protein samples from a glass substrate by 10 Hz bursts of 20 pulses having 10 μs pulse <span class="hlt">separation</span> and <50 μJ pulse energy resulted in signal comparable to a metal substrate. The protein signal observed from an aqueous droplet on a glass substrate was negligible compared to either a droplet on metal or a thin film on glass. The mass spectra generated from dried and aqueous protein samples by the low-energy, fiber laser were similar to the results from high-energy (500 μJ), 45-fs, 800-nm Ti:sapphire-based femtosecond laser electrospray mass spectrometry (LEMS) experiments, suggesting that the fiber-based femtosecond laser desorption mechanism involves a nonresonant, multiphoton process, rather than thermal- or photoacoustic-<span class="hlt">induced</span> desorption. Direct analysis of whole blood performed without any pretreatment resulted in features corresponding to hemoglobin subunit-heme complex ions. The observation of intact molecular ions with low charge states from protein, and the tentatively assigned hemoglobin α subunit-heme complex from blood suggests that fiber-based femtosecond laser <span class="hlt">vaporization</span> is a "soft" desorption source at a laser intensity of 2.39 × 10(12) W/cm(2). The low-energy, turnkey fiber laser demonstrates the potential of a more robust and affordable laser for femtosecond laser <span class="hlt">vaporization</span> to deliver biological macromolecules into the gas <span class="hlt">phase</span> for mass analysis.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27533107','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27533107"><span>Nanoscopy of <span class="hlt">Phase</span> <span class="hlt">Separation</span> in InxGa1-xN Alloys.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Abate, Yohannes; Seidlitz, Daniel; Fali, Alireza; Gamage, Sampath; Babicheva, Viktoriia; Yakovlev, Vladislav S; Stockman, Mark I; Collazo, Ramon; Alden, Dorian; Dietz, Nikolaus</p> <p>2016-09-07</p> <p><span class="hlt">Phase</span> <span class="hlt">separations</span> in ternary/multinary semiconductor alloys is a major challenge that limits optical and electronic internal device efficiency. We have found ubiquitous local <span class="hlt">phase</span> <span class="hlt">separation</span> in In1-xGaxN alloys that persists to nanoscale spatial extent by employing high-resolution nanoimaging technique. We lithographically patterned InN/sapphire substrates with nanolayers of In1-xGaxN down to few atomic layers thick that enabled us to calibrate the near-field infrared response of the semiconductor nanolayers as a function of composition and thickness. We also developed an advanced theoretical approach that considers the full geometry of the probe tip and all the sample and substrate layers. Combining experiment and theory, we identified and quantified <span class="hlt">phase</span> <span class="hlt">separation</span> in epitaxially grown individual nanoalloys. We found that the scale of the <span class="hlt">phase</span> <span class="hlt">separation</span> varies widely from particle to particle ranging from all Ga- to all In-rich regions and covering everything in between. We have found that between 20 and 25% of particles show some level of Ga-rich <span class="hlt">phase</span> <span class="hlt">separation</span> over the entire sample region, which is in qualitative agreement with the known <span class="hlt">phase</span> diagram of In1-xGaxN system.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4907611','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4907611"><span>Melting and <span class="hlt">Vaporization</span> of the 1223 <span class="hlt">Phase</span> in the System (Tl-Pb-Ba-Sr-Ca-Cu-O)</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Cook, L. P.; Wong-Ng, W.; Paranthaman, P.</p> <p>1996-01-01</p> <p>The melting and <span class="hlt">vaporization</span> of the 1223 [(Tl,Pb):(Ba,Sr):Ca:Cu] oxide <span class="hlt">phase</span> in the system (Tl-Pb-Ba-Sr-Ca-Cu-O) have been investigated using a combination of dynamic methods (differential thermal analysis, thermogravimetry, effusion) and post-quenching characterization techniques (powder x-ray diffraction, scanning electron microscopy, energy dispersive x-ray spectrometry). <span class="hlt">Vaporization</span> rates, thermal events, and melt compositions were followed as a function of thallia loss from a 1223 stoichiometry. Melting and <span class="hlt">vaporization</span> equilibria of the 1223 <span class="hlt">phase</span> are complex, with as many as seven <span class="hlt">phases</span> participating simultaneously. At a total pressure of 0.1 MPa the 1223 <span class="hlt">phase</span> was found to melt completely at (980 ± 5) °C in oxygen, at a thallia partial pressure (pTl2O) of (4.6 ± 0.5) kPa, where the quoted uncertainties are standard uncertainties, i.e., 1 estimated standard deviation. The melting reaction involves five other solids and a liquid, nominally as follows: 1223→1212+(Ca,Sr)2CuO3+(Sr,Ca)CuO2+BaPbO3+(Ca,Sr)O+Liquid Stoichiometries of the participating <span class="hlt">phases</span> have been determined from microchemical analysis, and substantial elemental substitution on the 1212 and 1223 crystallographic sites is indicated. The 1223 <span class="hlt">phase</span> occurs in equilibrium with liquids from its melting point down to at least 935 °C. The composition of the lowest melting liquid detected for the bulk compositions of this study has been measured using microchemical analysis. Applications to the processing of superconducting wires and tapes are discussed. PMID:27805086</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23133341','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23133341"><span>Incorporating <span class="hlt">Phase</span>-Dependent Polarizability in Non-Additive Electrostatic Models for Molecular Dynamics Simulations of the Aqueous Liquid-<span class="hlt">Vapor</span> Interface.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bauer, Brad A; Warren, G Lee; Patel, Sandeep</p> <p>2009-02-10</p> <p> anticipated in regions with both liquid and <span class="hlt">vapor</span> character, interfacial simulations of TIP4P-QDP were performed and compared to TIP4P-FQ, a static polarizability analog. Despite similar features in density profiles such as the position of the GDS and interfacial width, enhanced dipole moments are observed for the TIP4P-QDP interface and onset of the <span class="hlt">vapor</span> <span class="hlt">phase</span>. Water orientational profiles show an increased preference (over TIP4P-FQ) in the orientation of the permanent dipole vector of the molecule within the interface; an enhanced z-<span class="hlt">induced</span> dipole moment directly results from this preference. Hydrogen bond formation is lower, on average, in the bulk for TIP4P-QDP than TIP4P-FQ. However, the average number of hydrogen bonds formed by TIP4P-QDP in the interface exceeds that of TIP4P-FQ, and observed hydrogen bond networks extend further into the gaseous region. The TIP4P-QDP interfacial potential, calculated to be -11.98(±0.08) kcal/mol, is less favorable than that for TIP4P-FQ by approximately 2% as a result of a diminished quadrupole contribution. Surface tension is calculated within a 1.3% reduction from the experimental value. Results reported demonstrate TIP4P-QDP as a model comparable to the popular TIP4P-FQ while accounting for a physical effect previously neglected by other water models. Further refinements to this model, as well as future applications are discussed.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_14");'>14</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li class="active"><span>16</span></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_16 --> <div id="page_17" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li class="active"><span>17</span></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="321"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3488353','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3488353"><span>Incorporating <span class="hlt">Phase</span>-Dependent Polarizability in Non-Additive Electrostatic Models for Molecular Dynamics Simulations of the Aqueous Liquid-<span class="hlt">Vapor</span> Interface</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Bauer, Brad A.; Warren, G. Lee; Patel, Sandeep</p> <p>2012-01-01</p> <p> anticipated in regions with both liquid and <span class="hlt">vapor</span> character, interfacial simulations of TIP4P-QDP were performed and compared to TIP4P-FQ, a static polarizability analog. Despite similar features in density profiles such as the position of the GDS and interfacial width, enhanced dipole moments are observed for the TIP4P-QDP interface and onset of the <span class="hlt">vapor</span> <span class="hlt">phase</span>. Water orientational profiles show an increased preference (over TIP4P-FQ) in the orientation of the permanent dipole vector of the molecule within the interface; an enhanced z-<span class="hlt">induced</span> dipole moment directly results from this preference. Hydrogen bond formation is lower, on average, in the bulk for TIP4P-QDP than TIP4P-FQ. However, the average number of hydrogen bonds formed by TIP4P-QDP in the interface exceeds that of TIP4P-FQ, and observed hydrogen bond networks extend further into the gaseous region. The TIP4P-QDP interfacial potential, calculated to be -11.98(±0.08) kcal/mol, is less favorable than that for TIP4P-FQ by approximately 2% as a result of a diminished quadrupole contribution. Surface tension is calculated within a 1.3% reduction from the experimental value. Results reported demonstrate TIP4P-QDP as a model comparable to the popular TIP4P-FQ while accounting for a physical effect previously neglected by other water models. Further refinements to this model, as well as future applications are discussed. PMID:23133341</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3421221','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3421221"><span>Images reveal that atmospheric particles can undergo liquid–liquid <span class="hlt">phase</span> <span class="hlt">separations</span></span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>You, Yuan; Renbaum-Wolff, Lindsay; Carreras-Sospedra, Marc; Hanna, Sarah J.; Hiranuma, Naruki; Kamal, Saeid; Smith, Mackenzie L.; Zhang, Xiaolu; Weber, Rodney J.; Shilling, John E.; Dabdub, Donald; Martin, Scot T.; Bertram, Allan K.</p> <p>2012-01-01</p> <p>A large fraction of submicron atmospheric aerosol particles contains both organic material and inorganic salts. As the relative humidity cycles in the atmosphere and the water content of the particles correspondingly changes, these mixed particles can undergo a range of <span class="hlt">phase</span> transitions, possibly including liquid–liquid <span class="hlt">phase</span> <span class="hlt">separation</span>. If liquid–liquid <span class="hlt">phase</span> <span class="hlt">separation</span> occurs, the gas-particle partitioning of atmospheric semivolatile organic compounds, the scattering and absorption of solar radiation, and the reactive uptake of gas species on atmospheric particles may be affected, with important implications for climate predictions. The actual occurrence of liquid–liquid <span class="hlt">phase</span> <span class="hlt">separation</span> within individual atmospheric particles has been considered uncertain, in large part because of the absence of observations for real-world samples. Here, using optical and fluorescence microscopy, we present images that show the coexistence of two noncrystalline <span class="hlt">phases</span> for real-world samples collected on multiple days in Atlanta, GA as well as for laboratory-generated samples under simulated atmospheric conditions. These results reveal that atmospheric particles can undergo liquid–liquid <span class="hlt">phase</span> <span class="hlt">separations</span>. To explore the implications of these findings, we carried out simulations of the Atlanta urban environment and found that liquid–liquid <span class="hlt">phase</span> <span class="hlt">separation</span> can result in increased concentrations of gas-<span class="hlt">phase</span> NO3 and N2O5 due to decreased particle uptake of N2O5. PMID:22847443</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22847443','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22847443"><span>Images reveal that atmospheric particles can undergo liquid-liquid <span class="hlt">phase</span> <span class="hlt">separations</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>You, Yuan; Renbaum-Wolff, Lindsay; Carreras-Sospedra, Marc; Hanna, Sarah J; Hiranuma, Naruki; Kamal, Saeid; Smith, Mackenzie L; Zhang, Xiaolu; Weber, Rodney J; Shilling, John E; Dabdub, Donald; Martin, Scot T; Bertram, Allan K</p> <p>2012-08-14</p> <p>A large fraction of submicron atmospheric aerosol particles contains both organic material and inorganic salts. As the relative humidity cycles in the atmosphere and the water content of the particles correspondingly changes, these mixed particles can undergo a range of <span class="hlt">phase</span> transitions, possibly including liquid-liquid <span class="hlt">phase</span> <span class="hlt">separation</span>. If liquid-liquid <span class="hlt">phase</span> <span class="hlt">separation</span> occurs, the gas-particle partitioning of atmospheric semivolatile organic compounds, the scattering and absorption of solar radiation, and the reactive uptake of gas species on atmospheric particles may be affected, with important implications for climate predictions. The actual occurrence of liquid-liquid <span class="hlt">phase</span> <span class="hlt">separation</span> within individual atmospheric particles has been considered uncertain, in large part because of the absence of observations for real-world samples. Here, using optical and fluorescence microscopy, we present images that show the coexistence of two noncrystalline <span class="hlt">phases</span> for real-world samples collected on multiple days in Atlanta, GA as well as for laboratory-generated samples under simulated atmospheric conditions. These results reveal that atmospheric particles can undergo liquid-liquid <span class="hlt">phase</span> <span class="hlt">separations</span>. To explore the implications of these findings, we carried out simulations of the Atlanta urban environment and found that liquid-liquid <span class="hlt">phase</span> <span class="hlt">separation</span> can result in increased concentrations of gas-<span class="hlt">phase</span> NO(3) and N(2)O(5) due to decreased particle uptake of N(2)O(5).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28758733','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28758733"><span>Conductive Textiles via <span class="hlt">Vapor-Phase</span> Polymerization of 3,4-Ethylenedioxythiophene.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ala, Okan; Hu, Bin; Li, Dapeng; Yang, Chen-Lu; Calvert, Paul; Fan, Qinguo</p> <p>2017-08-30</p> <p>We fabricated electrically conductive textiles via <span class="hlt">vapor-phase</span> polymerization of poly(3,4-ethylenedioxythiophene) (PEDOT) layers on cotton, cotton/poly(ethylene terephthalate) (PET), cotton/Lycra, and PET fabrics. We then measured the electrical resistivity values of such PEDOT-coated textiles and analyzed the effect of water treatment on the electrical resistivity. Additionally, we tested the change in the electrical resistance of the conductive textiles under cyclic stretching and relaxation. Last, we characterized the uniformity and morphology of the conductive layer formed on the fabrics using scanning electron microscopy and electron-dispersive X-ray spectroscopy.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFM.P11A3750T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFM.P11A3750T"><span>CRYOCHEM, Thermodynamic Model for Cryogenic Chemical Systems: Solid-<span class="hlt">Vapor</span> and Solid-Liquid-<span class="hlt">Vapor</span> <span class="hlt">Phase</span> Equilibria Toward Applications on Titan and Pluto</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tan, S. P.; Kargel, J. S.; Adidharma, H.; Marion, G. M.</p> <p>2014-12-01</p> <p>Until in-situ measurements can be made regularly on extraterrestrial bodies, thermodynamic models are the only tools to investigate the properties and behavior of chemical systems on those bodies. The resulting findings are often critical in describing physicochemical processes in the atmosphere, surface, and subsurface in planetary geochemistry and climate studies. The extremely cold conditions on Triton, Pluto and other Kuiper Belt Objects, and Titan introduce huge non-ideality that prevents conventional models from performing adequately. At such conditions, atmospheres as a whole—not components individually—are subject to <span class="hlt">phase</span> equilibria with their equilibrium solid <span class="hlt">phases</span> or liquid <span class="hlt">phases</span> or both. A molecular-based thermodynamic model for cryogenic chemical systems, referred to as CRYOCHEM, the development of which is still in progress, was shown to reproduce the vertical composition profile of Titan's atmospheric methane measured by the Huygens probe (Tan et al., Icarus 2013, 222, 53). Recently, the model was also used to describe Titan's global circulation where the calculated composition of liquid in Ligeia Mare is consistent with the bathymetry and microwave absorption analysis of T91 Cassini fly-by data (Tan et al., 2014, submitted). Its capability to deal with equilibria involving solid <span class="hlt">phases</span> has also been demonstrated (Tan et al., Fluid <span class="hlt">Phase</span> Equilib. 2013, 360, 320). With all those previous works done, our attention is now shifting to the lower temperatures in Titan's tropopause and on Pluto's surface, where much technical development remains for CRYOCHEM to assure adequate performance at low temperatures. In these conditions, solid-<span class="hlt">vapor</span> equilibrium (SVE) is the dominant <span class="hlt">phase</span> behavior that determines the composition of the atmosphere and the existing ices. Another potential application is for the subsurface <span class="hlt">phase</span> equilibrium, which also involves liquid, thus three-<span class="hlt">phase</span> equilibrium: solid-liquid-<span class="hlt">vapor</span> (SLV). This presentation will discuss the</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ArRMA.225..177H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ArRMA.225..177H"><span>A Temperature-Dependent <span class="hlt">Phase</span>-Field Model for <span class="hlt">Phase</span> <span class="hlt">Separation</span> and Damage</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Heinemann, Christian; Kraus, Christiane; Rocca, Elisabetta; Rossi, Riccarda</p> <p>2017-07-01</p> <p>In this paper we study a model for <span class="hlt">phase</span> <span class="hlt">separation</span> and damage in thermoviscoelastic materials. The main novelty of the paper consists in the fact that, in contrast with previous works in the literature concerning <span class="hlt">phase</span> <span class="hlt">separation</span> and damage processes in elastic media, in our model we encompass thermal processes, nonlinearly coupled with the damage, concentration and displacement evolutions. More particularly, we prove the existence of "entropic weak solutions", resorting to a solvability concept first introduced in Feireisl (Comput Math Appl 53:461-490, 2007) in the framework of Fourier-Navier-Stokes systems and then recently employed in Feireisl et al. (Math Methods Appl Sci 32:1345-1369, 2009) and Rocca and Rossi (Math Models Methods Appl Sci 24:1265-1341, 2014) for the study of PDE systems for <span class="hlt">phase</span> transition and damage. Our global-in-time existence result is obtained by passing to the limit in a carefully devised time-discretization scheme.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhRvE..96f2804K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhRvE..96f2804K"><span>Coarsening and pattern formation during true morphological <span class="hlt">phase</span> <span class="hlt">separation</span> in unstable thin films under gravity</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kumar, Avanish; Narayanam, Chaitanya; Khanna, Rajesh; Puri, Sanjay</p> <p>2017-12-01</p> <p>We address in detail the problem of true morphological <span class="hlt">phase</span> <span class="hlt">separation</span> (MPS) in three-dimensional or (2 +1 )-dimensional unstable thin liquid films (>100 nm) under the influence of gravity. The free-energy functionals of these films are asymmetric and show two points of common tangency, which facilitates the formation of two equilibrium <span class="hlt">phases</span>. Three distinct patterns formed by relative preponderance of these <span class="hlt">phases</span> are clearly identified in "true MPS". Asymmetricity <span class="hlt">induces</span> two different pathways of pattern formation, viz., defect and direct pathway for true MPS. The pattern formation and <span class="hlt">phase</span>-ordering dynamics have been studied using statistical measures such as structure factor, correlation function, and growth laws. In the late stage of coarsening, the system reaches into a scaling regime for both pathways, and the characteristic domain size follows the Lifshitz-Slyozov growth law [L (t ) ˜t1 /3] . However, for the defect pathway, there is a crossover of domain growth behavior from L (t ) ˜t1 /4→t1 /3 in the dynamical scaling regime. We also underline the analogies and differences behind the mechanisms of MPS and true MPS in thin liquid films and generic spinodal <span class="hlt">phase</span> <span class="hlt">separation</span> in binary mixtures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1174877','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/1174877"><span>Monitoring of <span class="hlt">vapor</span> <span class="hlt">phase</span> polycyclic aromatic hydrocarbons</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Vo-Dinh, Tuan; Hajaligol, Mohammad R.</p> <p>2004-06-01</p> <p>An apparatus for monitoring <span class="hlt">vapor</span> <span class="hlt">phase</span> polycyclic aromatic hydrocarbons in a high-temperature environment has an excitation source producing electromagnetic radiation, an optical path having an optical probe optically communicating the electromagnetic radiation received at a proximal end to a distal end, a spectrometer or polychromator, a detector, and a positioner coupled to the first optical path. The positioner can slidably move the distal end of the optical probe to maintain the distal end position with respect to an area of a material undergoing combustion. The emitted wavelength can be directed to a detector in a single optical probe 180.degree. backscattered configuration, in a dual optical probe 180.degree. backscattered configuration or in a dual optical probe 90.degree. side scattered configuration. The apparatus can be used to monitor an emitted wavelength of energy from a polycyclic aromatic hydrocarbon as it fluoresces in a high temperature environment.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JAP...123q3903S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JAP...123q3903S"><span>Radial elemental and <span class="hlt">phase</span> <span class="hlt">separation</span> in Ni-Mn-Ga glass-coated microwires</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shevyrtalov, S.; Zhukov, A.; Medvedeva, S.; Lyatun, I.; Zhukova, V.; Rodionova, V.</p> <p>2018-05-01</p> <p>In this manuscript, radial elemental and <span class="hlt">phase</span> <span class="hlt">separation</span> in Ni-Mn-Ga glass-coated microwires with high excess Ni as a result of high-temperature annealing was observed. Partial manganese evaporation from the outer part of the metallic nucleus and glass melting results in the formation of manganese oxide at the surface. The lack of manganese due to its evaporation <span class="hlt">induces</span> Ni3Ga formation in the intermediate part, while in the middle part of the metallic nucleus, the residual L21 <span class="hlt">phase</span> with an average chemical composition of Ni60Mn9Ga31 remains. The layered structure exhibits soft ferromagnetic behavior below 270 K. The results were discussed taking into account the chemical composition, arising internal stresses, recrystallization, and atomic ordering.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22484603','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22484603"><span>Rose oil (from Rosa × damascena Mill.) <span class="hlt">vapor</span> attenuates depression-<span class="hlt">induced</span> oxidative toxicity in rat brain.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Nazıroğlu, Mustafa; Kozlu, Süleyman; Yorgancıgil, Emre; Uğuz, Abdülhadi Cihangir; Karakuş, Kadir</p> <p>2013-01-01</p> <p>Oxidative stress is a critical route of damage in various physiological stress-<span class="hlt">induced</span> disorders, including depression. Rose oil may be a useful treatment for depression because it contains flavonoids which include free radical antioxidant compounds such as rutin and quercetin. We investigated the effects of absolute rose oil (from Rosa × damascena Mill.) and experimental depression on lipid peroxidation and antioxidant levels in the cerebral cortex of rats. Thirty-two male rats were randomly divided into four groups. The first group was used as control, while depression was <span class="hlt">induced</span> in the second group using chronic mild stress (CMS). Oral (1.5 ml/kg) and <span class="hlt">vapor</span> (0.15 ml/kg) rose oil were given for 28 days to CMS depression-<span class="hlt">induced</span> rats, constituting the third and fourth groups, respectively. The sucrose preference test was used weekly to identify depression-like phenotypes during the experiment. At the end of the experiment, cerebral cortex samples were taken from all groups. The lipid peroxidation levels in the cerebral cortex in the CMS group were higher than in control whereas their levels were decreased by rose oil <span class="hlt">vapor</span> exposure. The vitamin A, vitamin E, vitamin C and β-carotene concentrations in the cerebral cortex were lower in the CMS group than in the control group whereas their concentrations were higher in the rose oil <span class="hlt">vapor</span> plus CMS group. The CMS-<span class="hlt">induced</span> antioxidant vitamin changes were not modulated by oral treatment. Glutathione peroxidase activity and reduced glutathione did not change statistically in the four groups following CMS or either treatment. In conclusion, experimental depression is associated with elevated oxidative stress while treatment with rose oil <span class="hlt">vapor</span> <span class="hlt">induced</span> protective effects on oxidative stress in depression.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017MS%26E..209a2013T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017MS%26E..209a2013T"><span>Characterization of Polysulfone Membranes Prepared with Thermally <span class="hlt">Induced</span> <span class="hlt">Phase</span> <span class="hlt">Separation</span> Technique</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tiron, L. G.; Pintilie, Ș C.; Vlad, M.; Birsan, I. G.; Baltă, Ș</p> <p>2017-06-01</p> <p>Abstract Membrane technology is one of the most used water treatment technology because of its high removal efficiency and cost effectiveness. Preparation techniques for polymer membranes show an important aspect of membrane properties. Generally, polysulfone (PSf) and polyethersulfone (PES) are used for the preparation of ultrafiltration (UF) membranes. Polysulfone (PSf) membranes have been widely used for <span class="hlt">separation</span> and purification of different solutions because of their excellent chemical and thermal stability. Polymeric membranes were obtained by <span class="hlt">phase</span> inversion method. The polymer solution introduced in the nonsolvent bath (distilled water) initiate the evaporation of the solvent from the solution, this phenomenon has a strong influence on the transport properties. The effect of the coagulation bath temperature on the membrane properties is of interest for this study. Membranes are characterized by pure water flux, permeability, porosity and retention of methylene blue. The low temperature of coagulation bath improve the membrane’s rejection and its influence was most notable.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27043009','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27043009"><span>Steric Pressure among Membrane-Bound Polymers Opposes Lipid <span class="hlt">Phase</span> <span class="hlt">Separation</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Imam, Zachary I; Kenyon, Laura E; Carrillo, Adelita; Espinoza, Isai; Nagib, Fatema; Stachowiak, Jeanne C</p> <p>2016-04-19</p> <p>Lipid rafts are thought to be key organizers of membrane-protein complexes in cells. Many proteins that interact with rafts have bulky polymeric components such as intrinsically disordered protein domains and polysaccharide chains. Therefore, understanding the interaction between membrane domains and membrane-bound polymers provides insights into the roles rafts play in cells. Multiple studies have demonstrated that high concentrations of membrane-bound polymeric domains create significant lateral steric pressure at membrane surfaces. Furthermore, our recent work has shown that lateral steric pressure at membrane surfaces opposes the assembly of membrane domains. Building on these findings, here we report that membrane-bound polymers are potent suppressors of membrane <span class="hlt">phase</span> <span class="hlt">separation</span>, which can destabilize lipid domains with substantially greater efficiency than globular domains such as membrane-bound proteins. Specifically, we created giant vesicles with a ternary lipid composition, which <span class="hlt">separated</span> into coexisting liquid ordered and disordered <span class="hlt">phases</span>. Lipids with saturated tails and poly(ethylene glycol) (PEG) chains conjugated to their head groups were included at increasing molar concentrations. When these lipids were sparse on the membrane surface they partitioned to the liquid ordered <span class="hlt">phase</span>. However, as they became more concentrated, the fraction of GUVs that were <span class="hlt">phase-separated</span> decreased dramatically, ultimately yielding a population of homogeneous membrane vesicles. Experiments and physical modeling using compositions of increasing PEG molecular weight and lipid miscibility <span class="hlt">phase</span> transition temperature demonstrate that longer polymers are the most efficient suppressors of membrane <span class="hlt">phase</span> <span class="hlt">separation</span> when the energetic barrier to lipid mixing is low. In contrast, as the miscibility transition temperature increases, longer polymers are more readily driven out of domains by the increased steric pressure. Therefore, the concentration of shorter polymers required</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhRvB..95x5108M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhRvB..95x5108M"><span>Symmetry-protected topological <span class="hlt">phases</span> of one-dimensional interacting fermions with spin-charge <span class="hlt">separation</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Montorsi, Arianna; Dolcini, Fabrizio; Iotti, Rita C.; Rossi, Fausto</p> <p>2017-06-01</p> <p>The low energy behavior of a huge variety of one-dimensional interacting spinful fermionic systems exhibits spin-charge <span class="hlt">separation</span>, described in the continuum limit by two sine-Gordon models decoupled in the charge and spin channels. Interaction is known to <span class="hlt">induce</span>, besides the gapless Luttinger liquid <span class="hlt">phase</span>, eight possible gapped <span class="hlt">phases</span>, among which are the Mott, Haldane, charge-/spin-density, and bond-ordered wave insulators, and the Luther Emery liquid. Here we prove that some of these physically distinct <span class="hlt">phases</span> have nontrivial topological properties, notably the presence of degenerate protected edge modes with fractionalized charge/spin. Moreover, we show that the eight gapped <span class="hlt">phases</span> are in one-to-one correspondence with the symmetry-protected topological (SPT) <span class="hlt">phases</span> classified by group cohomology theory in the presence of particle-hole symmetry P. The latter result is also exploited to characterize SPT <span class="hlt">phases</span> by measurable nonlocal order parameters which follow the system evolution to the quantum <span class="hlt">phase</span> transition. The implications on the appearance of exotic orders in the class of microscopic Hubbard Hamiltonians, possibly without P symmetry at higher energies, are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29424691','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29424691"><span>Pi-Pi contacts are an overlooked protein feature relevant to <span class="hlt">phase</span> <span class="hlt">separation</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Vernon, Robert McCoy; Chong, Paul Andrew; Tsang, Brian; Kim, Tae Hun; Bah, Alaji; Farber, Patrick; Lin, Hong; Forman-Kay, Julie Deborah</p> <p>2018-02-09</p> <p>Protein <span class="hlt">phase</span> <span class="hlt">separation</span> is implicated in formation of membraneless organelles, signaling puncta and the nuclear pore. Multivalent interactions of modular binding domains and their target motifs can drive <span class="hlt">phase</span> <span class="hlt">separation</span>. However, forces promoting the more common <span class="hlt">phase</span> <span class="hlt">separation</span> of intrinsically disordered regions are less understood, with suggested roles for multivalent cation-pi, pi-pi, and charge interactions and the hydrophobic effect. Known <span class="hlt">phase-separating</span> proteins are enriched in pi-orbital containing residues and thus we analyzed pi-interactions in folded proteins. We found that pi-pi interactions involving non-aromatic groups are widespread, underestimated by force-fields used in structure calculations and correlated with solvation and lack of regular secondary structure, properties associated with disordered regions. We present a <span class="hlt">phase</span> <span class="hlt">separation</span> predictive algorithm based on pi interaction frequency, highlighting proteins involved in biomaterials and RNA processing. © 2018, Vernon et al.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5847340','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5847340"><span>Pi-Pi contacts are an overlooked protein feature relevant to <span class="hlt">phase</span> <span class="hlt">separation</span></span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Vernon, Robert McCoy; Chong, Paul Andrew; Tsang, Brian; Kim, Tae Hun; Bah, Alaji; Farber, Patrick; Lin, Hong</p> <p>2018-01-01</p> <p>Protein <span class="hlt">phase</span> <span class="hlt">separation</span> is implicated in formation of membraneless organelles, signaling puncta and the nuclear pore. Multivalent interactions of modular binding domains and their target motifs can drive <span class="hlt">phase</span> <span class="hlt">separation</span>. However, forces promoting the more common <span class="hlt">phase</span> <span class="hlt">separation</span> of intrinsically disordered regions are less understood, with suggested roles for multivalent cation-pi, pi-pi, and charge interactions and the hydrophobic effect. Known <span class="hlt">phase-separating</span> proteins are enriched in pi-orbital containing residues and thus we analyzed pi-interactions in folded proteins. We found that pi-pi interactions involving non-aromatic groups are widespread, underestimated by force-fields used in structure calculations and correlated with solvation and lack of regular secondary structure, properties associated with disordered regions. We present a <span class="hlt">phase</span> <span class="hlt">separation</span> predictive algorithm based on pi interaction frequency, highlighting proteins involved in biomaterials and RNA processing. PMID:29424691</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24425207','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24425207"><span>Effects of processing parameters in thermally <span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> technique on porous architecture of scaffolds for bone tissue engineering.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Akbarzadeh, Rosa; Yousefi, Azizeh-Mitra</p> <p>2014-08-01</p> <p>Tissue engineering makes use of 3D scaffolds to sustain three-dimensional growth of cells and guide new tissue formation. To meet the multiple requirements for regeneration of biological tissues and organs, a wide range of scaffold fabrication techniques have been developed, aiming to produce porous constructs with the desired pore size range and pore morphology. Among different scaffold fabrication techniques, thermally <span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> (TIPS) method has been widely used in recent years because of its potential to produce highly porous scaffolds with interconnected pore morphology. The scaffold architecture can be closely controlled by adjusting the process parameters, including polymer type and concentration, solvent composition, quenching temperature and time, coarsening process, and incorporation of inorganic particles. The objective of this review is to provide information pertaining to the effect of these parameters on the architecture and properties of the scaffolds fabricated by the TIPS technique. © 2014 Wiley Periodicals, Inc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JPCRD..46a3104A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JPCRD..46a3104A"><span><span class="hlt">Phase</span> Transition Enthalpy Measurements of Organic and Organometallic Compounds and Ionic Liquids. Sublimation, <span class="hlt">Vaporization</span>, and Fusion Enthalpies from 1880 to 2015. Part 2. C11-C192</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Acree, William; Chickos, James S.</p> <p>2017-03-01</p> <p>The second part of this compendium concludes with a collection of <span class="hlt">phase</span> change enthalpies of organic molecules inclusive of C11-C192 reported over the period 1880-2015. Also included are <span class="hlt">phase</span> change enthalpies including fusion, <span class="hlt">vaporization</span>, and sublimation enthalpies for organometallic, ionic liquids, and a few inorganic compounds. Paper I of this compendium, published <span class="hlt">separately</span>, includes organic compounds from C1 to C10 and describes a group additivity method for evaluating solid, liquid, and gas <span class="hlt">phase</span> heat capacities as well as temperature adjustments of <span class="hlt">phase</span> changes. Paper II of this compendium also includes an updated version of a group additivity method for evaluating total <span class="hlt">phase</span> change entropies which together with the fusion temperature can be useful in estimating total <span class="hlt">phase</span> change enthalpies. Other uses include application in identifying potential substances that either form liquid or plastic crystals or exhibit additional <span class="hlt">phase</span> changes such as undetected solid-solid transitions or behave anisotropically in the liquid state.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1987uta..rept.....D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1987uta..rept.....D"><span>Physical mechanisms in shock-<span class="hlt">induced</span> turbulent <span class="hlt">separated</span> flow</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dolling, D. S.</p> <p>1987-12-01</p> <p>It has been demonstrated that the flow downstream of the moving shock is <span class="hlt">separated</span> and that the foot of the shock is effectively the instantaneous <span class="hlt">separation</span> point. The shock <span class="hlt">induced</span> turbulent <span class="hlt">separation</span> is an intermittant process and the <span class="hlt">separation</span> line indicated by surface tracer methods, such as kerosene-lampblack, is a downstream boundary of a region of intermittent <span class="hlt">separation</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20010082954&hterms=ammonia&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dammonia','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20010082954&hterms=ammonia&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dammonia"><span>An Assessment of the Technical Readiness of the <span class="hlt">Vapor</span> <span class="hlt">Phase</span> Catalytic Ammonia Removal Process (VPCAR) Technology</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Flynn, Michael</p> <p>2000-01-01</p> <p>This poster provides an assessment of the technical readiness of the <span class="hlt">Vapor</span> <span class="hlt">Phase</span> Catalytic Ammonia Removal Process (VPCAR). The VPCAR technology is a fully regenerative water recycling technology designed specifically for applications such as a near term Mars exploration mission. The VPCAR technology is a highly integrated distillation/catalytic oxidation based water processor. It is designed to accept a combined wastewater stream (urine, condensate, and hygiene) and produces potable water in a single process step which requires -no regularly scheduled re-supply or maintenance for a 3 year mission. The technology is designed to be modular and to fit into a volume comparable to a single International Space Station Rack (when sized for a crew of 6). This poster provides a description of the VPCAR technology and a summary of the current performance of the technology. Also provided are the results of two <span class="hlt">separate</span> NASA sponsored system trade studies which investigated the potential payback of further development of the VPCAR technology.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JPhCS.916a2035Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JPhCS.916a2035Z"><span>3D CFD simulation of Multi-<span class="hlt">phase</span> flow <span class="hlt">separators</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhu, Zhiying</p> <p>2017-10-01</p> <p>During the exploitation of natural gas, some water and sands are contained. It will be better to <span class="hlt">separate</span> water and sands from natural gas to insure favourable transportation and storage. In this study, we use CFD to analyse the effect of multi-<span class="hlt">phase</span> flow <span class="hlt">separator</span>, whose detailed geometrical parameters are designed in advanced. VOF model and DPM are used here. From the results of CFD, we can draw a conclusion that <span class="hlt">separated</span> effect of multi-<span class="hlt">phase</span> flow achieves better results. No solid and water is carried out from gas outlet. CFD simulation provides an economical and efficient approach to shed more light on details of the flow behaviour.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_15");'>15</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li class="active"><span>17</span></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_17 --> <div id="page_18" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li class="active"><span>18</span></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="341"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25063098','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25063098"><span>A novel bio-safe <span class="hlt">phase</span> <span class="hlt">separation</span> process for preparing open-pore biodegradable polycaprolactone microparticles.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Salerno, Aurelio; Domingo, Concepción</p> <p>2014-09-01</p> <p>Open-pore biodegradable microparticles are object of considerable interest for biomedical applications, particularly as cell and drug delivery carriers in tissue engineering and health care treatments. Furthermore, the engineering of microparticles with well definite size distribution and pore architecture by bio-safe fabrication routes is crucial to avoid the use of toxic compounds potentially harmful to cells and biological tissues. To achieve this important issue, in the present study a straightforward and bio-safe approach for fabricating porous biodegradable microparticles with controlled morphological and structural features down to the nanometer scale is developed. In particular, ethyl lactate is used as a non-toxic solvent for polycaprolactone particles fabrication via a thermal <span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> technique. The used approach allows achieving open-pore particles with mean particle size in the 150-250 μm range and a 3.5-7.9 m(2)/g specific surface area. Finally, the combination of thermal <span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> and porogen leaching techniques is employed for the first time to obtain multi-scaled porous microparticles with large external and internal pore sizes and potential improved characteristics for cell culture and tissue engineering. Samples were characterized to assess their thermal properties, morphology and crystalline structure features and textural properties. Copyright © 2014 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JGP...104...30K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JGP...104...30K"><span>On the <span class="hlt">phase</span> form of a deformation quantization with <span class="hlt">separation</span> of variables</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Karabegov, Alexander</p> <p>2016-06-01</p> <p>Given a star product with <span class="hlt">separation</span> of variables on a pseudo-Kähler manifold, we obtain a new formal (1, 1)-form from its classifying form and call it the <span class="hlt">phase</span> form of the star product. The cohomology class of a star product with <span class="hlt">separation</span> of variables equals the class of its <span class="hlt">phase</span> form. We show that the <span class="hlt">phase</span> forms can be arbitrary and they bijectively parametrize the star products with <span class="hlt">separation</span> of variables. We also describe the action of a change of the formal parameter on a star product with <span class="hlt">separation</span> of variables, its formal Berezin transform, classifying form, <span class="hlt">phase</span> form, and canonical trace density.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20090028674','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20090028674"><span>Assessment of the <span class="hlt">Vapor</span> <span class="hlt">Phase</span> Catalytic Ammonia Removal (VPCAR) Technology at the MSFC ECLS Test Facility</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Tomes, Kristin; Long, David; Carter, Layne; Flynn, Michael</p> <p>2007-01-01</p> <p>The <span class="hlt">Vapor</span> <span class="hlt">Phase</span> Catalytic Ammonia. Removal (VPCAR) technology has been previously discussed as a viable option for. the Exploration Water Recovery System. This technology integrates a <span class="hlt">phase</span> change process with catalytic oxidation in the <span class="hlt">vapor</span> <span class="hlt">phase</span> to produce potable water from exploration mission wastewaters. A developmental prototype VPCAR was designed, built and tested under funding provided by a National Research. Announcement (NRA) project. The core technology, a Wiped Film Rotating Device (WFRD) was provided by Water Reuse Technologies under the NRA, whereas Hamilton Sundstrand Space Systems International performed the hardware integration and acceptance test. of the system. Personnel at the-Ames Research Center performed initial systems test of the VPCAR using ersatz solutions. To assess the viability of this hardware for Exploration. Life Support (ELS) applications, the hardware has been modified and tested at the MSFC ECLS Test facility. This paper summarizes the hardware modifications and test results and provides an assessment of this technology for the ELS application.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19830065390&hterms=gallium+vapor&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dgallium%2Bvapor','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19830065390&hterms=gallium+vapor&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3Dgallium%2Bvapor"><span>Use of column V alkyls in organometallic <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy (OMVPE)</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Ludowise, M. J.; Cooper, C. B., III</p> <p>1982-01-01</p> <p>The use of the column V-trialkyls trimethylarsenic (TMAs) and trimethylantimony (TMSb) for the organometallic <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy (OM-VPE) of III-V compound semiconductors is reviewed. A general discussion of the interaction chemistry of common Group III and Group V reactants is presented. The practical application of TMSb and TMAs for OM-VPE is demonstrated using the growth of GaSb, GaAs(1-y)Sb(y), Al(x)Ga(1-x)Sb, and Ga(1-x)In(x)As as examples.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015SPIE.9654E..1JD','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015SPIE.9654E..1JD"><span>Chromium doped nano-<span class="hlt">phase</span> <span class="hlt">separated</span> yttria-alumina-silica glass based optical fiber preform: fabrication and characterization</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dutta, Debjit; Dhar, Anirban; Das, Shyamal; Bysakh, Sandip; Kir'yanov, Alexandar; Paul, Mukul Chandra</p> <p>2015-06-01</p> <p>Transition metal (TM) doping in silica core optical fiber is one of the research area which has been studied for long time and Chromium (Cr) doping specially attracts a lot of research interest due to their broad emission band covering U, C and L band with many potential application such as saturable absorber or broadband amplifier etc. This paper present fabrication of Cr doped nano-<span class="hlt">phase</span> <span class="hlt">separated</span> silica fiber within yttria-alumina-silica core glass through conventional Modified Chemical <span class="hlt">Vapor</span> Deposition (MCVD) process coupled with solution doping technique along with different material and optical characterization. For the first time scanning electron microscope (SEM) / energy dispersive X-ray (EDX) analysis of porous soot sample and final preform has been utilized to investigate incorporation mechanism of Crions with special emphasis on Cr-species evaporation at different stages of fabrication. We also report that optimized annealing condition of our fabricated preform exhibited enhanced fluorescence emission and a broad band within 550- 800 nm wavelength region under pumping at 532 nm wavelength due to nano-<span class="hlt">phase</span> restructuration.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003JChPh.118.9915E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003JChPh.118.9915E"><span>Direct calculation of liquid-<span class="hlt">vapor</span> <span class="hlt">phase</span> equilibria from transition matrix Monte Carlo simulation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Errington, Jeffrey R.</p> <p>2003-06-01</p> <p>An approach for directly determining the liquid-<span class="hlt">vapor</span> <span class="hlt">phase</span> equilibrium of a model system at any temperature along the coexistence line is described. The method relies on transition matrix Monte Carlo ideas developed by Fitzgerald, Picard, and Silver [Europhys. Lett. 46, 282 (1999)]. During a Monte Carlo simulation attempted transitions between states along the Markov chain are monitored as opposed to tracking the number of times the chain visits a given state as is done in conventional simulations. Data collection is highly efficient and very precise results are obtained. The method is implemented in both the grand canonical and isothermal-isobaric ensemble. The main result from a simulation conducted at a given temperature is a density probability distribution for a range of densities that includes both liquid and <span class="hlt">vapor</span> states. <span class="hlt">Vapor</span> pressures and coexisting densities are calculated in a straightforward manner from the probability distribution. The approach is demonstrated with the Lennard-Jones fluid. Coexistence properties are directly calculated at temperatures spanning from the triple point to the critical point.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017MMTA...48.3130K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017MMTA...48.3130K"><span>Prediction of <span class="hlt">Phase</span> <span class="hlt">Separation</span> of Immiscible Ga-Tl Alloys</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kim, Yunkyum; Kim, Han Gyeol; Kang, Youn-Bae; Kaptay, George; Lee, Joonho</p> <p>2017-06-01</p> <p><span class="hlt">Phase</span> <span class="hlt">separation</span> temperature of Ga-Tl liquid alloys was investigated using the constrained drop method. With this method, density and surface tension were investigated together. Despite strong repulsive interactions, molar volume showed ideal mixing behavior, whereas surface tension of the alloy was close to that of pure Tl due to preferential adsorption of Tl. <span class="hlt">Phase</span> <span class="hlt">separation</span> temperatures and surface tension values obtained with this method were close to the theoretically calculated values using three different thermodynamic models.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19720046854&hterms=keefe&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAuthor-Name%26N%3D0%26No%3D40%26Ntt%3Dkeefe','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19720046854&hterms=keefe&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAuthor-Name%26N%3D0%26No%3D40%26Ntt%3Dkeefe"><span>Shock melting and <span class="hlt">vaporization</span> of lunar rocks and minerals.</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Ahrens, T. J.; O'Keefe, J. D.</p> <p>1972-01-01</p> <p>The entropy associated with the thermodynamic states produced by hypervelocity meteoroid impacts at various velocities are calculated for a series of lunar rocks and minerals and compared with the entropy values required for melting and <span class="hlt">vaporization</span>. Taking into account shock-<span class="hlt">induced</span> <span class="hlt">phase</span> changes in the silicates, we calculate that iron meteorites impacting at speeds varying from 4 to 6 km/sec will produce shock melting in quartz, plagioclase, olivine, and pyroxene. Although calculated with less certainty, impact speeds required for incipient <span class="hlt">vaporization</span> vary from 7 to 11 km/sec for the range of minerals going from quartz to periclase for aluminum (silicate-like) projectiles. The impact velocities, which are required to <span class="hlt">induce</span> melting in a soil, are calculated to be in the range of 3 to 4 km/sec, provided thermal equilibrium is achieved in the shock state.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26752658','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26752658"><span>Polymer scaffolds with no skin-effect for tissue engineering applications fabricated by thermally <span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kasoju, Naresh; Kubies, Dana; Sedlačík, Tomáš; Janoušková, Olga; Koubková, Jana; Kumorek, Marta M; Rypáček, František</p> <p>2016-01-11</p> <p>Thermally <span class="hlt">induced</span> <span class="hlt">phase</span> <span class="hlt">separation</span> (TIPS) based methods are widely used for the fabrication of porous scaffolds for tissue engineering and related applications. However, formation of a less-/non-porous layer at the scaffold's outer surface at the air-liquid interface, often known as the skin-effect, restricts the cell infiltration inside the scaffold and therefore limits its efficacy. To this end, we demonstrate a TIPS-based process involving the exposure of the just quenched poly(lactide-co-caprolactone):dioxane <span class="hlt">phases</span> to the pure dioxane for a short time while still being under the quenching strength, herein after termed as the second quenching (2Q). Scanning electron microscopy, mercury intrusion porosimetry and contact angle analysis revealed a direct correlation between the time of 2Q and the gradual disappearance of the skin, followed by the widening of the outer pores and the formation of the fibrous filaments over the surface, with no effect on the internal pore architecture and the overall porosity of scaffolds. The experiments at various quenching temperatures and polymer concentrations revealed the versatility of 2Q in removing the skin. In addition, the in vitro cell culture studies with the human primary fibroblasts showed that the scaffolds prepared by the TIPS based 2Q process, with the optimal exposure time, resulted in a higher cell seeding and viability in contrast to the scaffolds prepared by the regular TIPS. Thus, TIPS including the 2Q step is a facile, versatile and innovative approach to fabricate the polymer scaffolds with a skin-free and fully open porous surface morphology for achieving a better cell response in tissue engineering and related applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/12209931','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12209931"><span>Adhesive <span class="hlt">phase</span> <span class="hlt">separation</span> at the dentin interface under wet bonding conditions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Spencer, Paulette; Wang, Yong</p> <p>2002-12-05</p> <p>Under in vivo conditions, there is little control over the amount of water left on the tooth and, thus, there is the danger of leaving the dentin surface so wet that the bonding resin undergoes physical <span class="hlt">separation</span> into hydrophobic and hydrophilic-rich <span class="hlt">phases</span>. The purpose of this study was to investigate <span class="hlt">phase</span> <span class="hlt">separation</span> in 2,2-bis[4(2-hydroxy-3-methacryloyloxy-propyloxy)-phenyl] propane (BisGMA)-based adhesive using molecular microanalysis and to examine the effect of <span class="hlt">phase</span> <span class="hlt">separation</span> on the structural characteristics of the hybrid layer. Model BisGMA/HEMA (hydroxyethl methacrylate) mixtures with/without ethanol and commercial BisGMA-based adhesive (Single Bond) were combined with water at concentrations from 0 to 50 vol%. Macrophase <span class="hlt">separation</span> in the BisGMA/HEMA/water mixtures was detected using cloud point measurements. In parallel with these measurements, the BisGMA/HEMA and adhesive/water mixtures were cast as films and polymerized. Molecular structure was recorded from the distinct features in the <span class="hlt">phase-separated</span> adhesive using confocal Raman microspectroscopy (CRM). Human dentin specimens treated with Single Bond were analyzed with scanning electron microscopy (SEM) and CRM mapping across the dentin/adhesive interface. The model BisGMA/HEMA mixtures with ethanol and the commercial BisGMA-based adhesive experienced <span class="hlt">phase</span> <span class="hlt">separation</span> at approximately 25 vol% water. Raman spectra collected from the <span class="hlt">phase-separated</span> adhesive indicated that the composition of the particles and surrounding matrix material was primarily BisGMA and HEMA, respectively. Based on SEM analysis, there was substantial porosity at the adhesive interface with dentin. Micro-Raman spectral analysis of the dentin/adhesive interface indicates that the contribution from the BisGMA component decreases by nearly 50% within the first micrometer. The morphologic results in corroboration with the spectroscopic data suggest that as a result of adhesive <span class="hlt">phase</span> <span class="hlt">separation</span> the hybrid layer is not an</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23723185','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23723185"><span>Preparation of mesoporous nanofibers by <span class="hlt">vapor</span> <span class="hlt">phase</span> synthesis: control of mesopore structures with the aid of co-surfactants.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Min, Sa Hoon; Bae, Joonwon; Jang, Jyongsik; Lee, Kyung Jin</p> <p>2013-06-28</p> <p>Mesoporous nanofibers (MSNFs) can be fabricated in the pores of anodic aluminum oxide (AAO) membrane using diverse methods. Among them <span class="hlt">vapor</span> <span class="hlt">phase</span> synthesis (VPS) provides several advantages over sol-gel or evaporation-<span class="hlt">induced</span> self-assembly (EISA) based methods. One powerful advantage is that we can employ multiple surfactants as structural directing agents (SDAs) simultaneously. By adopting diverse pairs of SDAs, we can control the mesopore structures, i.e. pore size, surface area, and even the morphology of mesostructures. Here, we used F127 as a main SDA, which is relatively robust (thus, difficult to change the mesopore structures), and added a series of cationic co-surfactants to observe the systematical changes in their mesostructure with respect to the chain length of the co-surfactant.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013Nanot..24y5602M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013Nanot..24y5602M"><span>Preparation of mesoporous nanofibers by <span class="hlt">vapor</span> <span class="hlt">phase</span> synthesis: control of mesopore structures with the aid of co-surfactants</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Min, Sa Hoon; Bae, Joonwon; Jang, Jyongsik; Lee, Kyung Jin</p> <p>2013-06-01</p> <p>Mesoporous nanofibers (MSNFs) can be fabricated in the pores of anodic aluminum oxide (AAO) membrane using diverse methods. Among them <span class="hlt">vapor</span> <span class="hlt">phase</span> synthesis (VPS) provides several advantages over sol-gel or evaporation-<span class="hlt">induced</span> self-assembly (EISA) based methods. One powerful advantage is that we can employ multiple surfactants as structural directing agents (SDAs) simultaneously. By adopting diverse pairs of SDAs, we can control the mesopore structures, i.e. pore size, surface area, and even the morphology of mesostructures. Here, we used F127 as a main SDA, which is relatively robust (thus, difficult to change the mesopore structures), and added a series of cationic co-surfactants to observe the systematical changes in their mesostructure with respect to the chain length of the co-surfactant.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/11788138','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/11788138"><span>Eicosanoids modulate hyperpnea-<span class="hlt">induced</span> late <span class="hlt">phase</span> airway obstruction and hyperreactivity in dogs.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Davis, Michael S; McCulloch, Sharron; Myers, Teresa; Freed, Arthur N</p> <p>2002-01-01</p> <p>A canine model of exercise-<span class="hlt">induced</span> asthma was used to test the hypothesis that the development of a late <span class="hlt">phase</span> response to hyperventilation depends on the acute production of pro-inflammatory mediators. Peripheral airway resistance, reactivity to hypocapnia and aerosol histamine, and bronchoalveolar lavage fluid (BALF) cell and eicosanoid content were measured in dogs approximately 5 h after dry air challenge (DAC). DAC resulted in late <span class="hlt">phase</span> obstruction, hyperreactivity to histamine, and neutrophilic inflammation. Both cyclooxygenase and lipoxygenase inhibitors administered in <span class="hlt">separate</span> experiments attenuated the late <span class="hlt">phase</span> airway obstruction and hyperreactivity to histamine. Neither drug affected the late <span class="hlt">phase</span> inflammation nor the concentrations of eicosanoids in the BALF obtained 5 h after DAC. This study confirms that hyperventilation of peripheral airways with unconditioned air causes late <span class="hlt">phase</span> neutrophilia, airway obstruction, and hyperreactivity. The late <span class="hlt">phase</span> changes in airway mechanics are related to the hyperventilation-<span class="hlt">induced</span> release of both prostaglandins and leukotrienes, and appear to be independent of the late <span class="hlt">phase</span> infiltration of inflammatory cells.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApSS..435..609Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApSS..435..609Y"><span>Cellulose acetate-based SiO2/TiO2 hybrid microsphere composite aerogel films for water-in-oil emulsion <span class="hlt">separation</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yang, Xue; Ma, Jianjun; Ling, Jing; Li, Na; Wang, Di; Yue, Fan; Xu, Shimei</p> <p>2018-03-01</p> <p>The cellulose acetate (CA)/SiO2-TiO2 hybrid microsphere composite aerogel films were successfully fabricated via water <span class="hlt">vapor-induced</span> <span class="hlt">phase</span> inversion of CA solution and simultaneous hydrolysis/condensation of 3-aminopropyltrimethoxysilane (APTMS) and tetrabutyl titanate (TBT) at room temperature. Micro-nano hierarchical structure was constructed on the surface of the film. The film could <span class="hlt">separate</span> nano-sized surfactant-stabilized water-in-oil emulsions only under gravity. The flux of the film for the emulsion <span class="hlt">separation</span> was up to 667 L m-2 h-1, while the <span class="hlt">separation</span> efficiency was up to 99.99 wt%. Meanwhile, the film exhibited excellent stability during multiple cycles. Moreover, the film performed excellent photo-degradation performance under UV light due to the photocatalytic ability of TiO2. Facile preparation, good <span class="hlt">separation</span> and potential biodegradation maked the CA/SiO2-TiO2 hybrid microsphere composite aerogel films a candidate in oil/water <span class="hlt">separation</span> application.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19790021432','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19790021432"><span>Alternate methods of applying diffusants to silicon solar cells. [screen printing of thick-film paste materials and <span class="hlt">vapor</span> <span class="hlt">phase</span> transport from solid sources</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Brock, T. W.; Field, M. B.</p> <p>1979-01-01</p> <p>Low-melting phosphate and borate glasses were screen printed on silicon wafers and heated to form n and p junctions. Data on surface appearance, sheet resistance and junction depth are presented. Similar data are reported for <span class="hlt">vapor</span> <span class="hlt">phase</span> transport from sintered aluminum metaphosphate and boron-containing glass-ceramic solid sources. Simultaneous diffusion of an N(+) layer with screen-printed glass and a p(+) layer with screen-printed Al alloy paste was attempted. No p(+) back surface field formation was achieved. Some good cells were produced but the heating in an endless-belt furnace caused a large scatter in sheet resistance and junction depth for three <span class="hlt">separate</span> lots of wafers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22598914-phase-transition-enthalpy-measurements-organic-organometallic-compounds-sublimation-vaporization-fusion-enthalpies-from-part-sub-sub','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22598914-phase-transition-enthalpy-measurements-organic-organometallic-compounds-sublimation-vaporization-fusion-enthalpies-from-part-sub-sub"><span><span class="hlt">Phase</span> Transition Enthalpy Measurements of Organic and Organometallic Compounds. Sublimation, <span class="hlt">Vaporization</span> and Fusion Enthalpies From 1880 to 2015. Part 1. C{sub 1} − C{sub 10}</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Acree, William; Chickos, James S.</p> <p>2016-09-15</p> <p>A compendium of <span class="hlt">phase</span> change enthalpies published in 2010 is updated to include the period 1880–2015. <span class="hlt">Phase</span> change enthalpies including fusion, <span class="hlt">vaporization</span>, and sublimation enthalpies are included for organic, organometallic, and a few inorganic compounds. Part 1 of this compendium includes organic compounds from C{sub 1} to C{sub 10}. Part 2 of this compendium, to be published <span class="hlt">separately</span>, will include organic and organometallic compounds from C{sub 11} to C{sub 192}. Sufficient data are presently available to permit thermodynamic cycles to be constructed as an independent means of evaluating the reliability of the data. Temperature adjustments of <span class="hlt">phase</span> change enthalpies frommore » the temperature of measurement to the standard reference temperature, T = 298.15 K, and a protocol for doing so are briefly discussed.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1344215-kinetic-model-gaas-growth-hydride-vapor-phase-epitaxy','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1344215-kinetic-model-gaas-growth-hydride-vapor-phase-epitaxy"><span>A Kinetic Model for GaAs Growth by Hydride <span class="hlt">Vapor</span> <span class="hlt">Phase</span> Epitaxy</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Schulte, Kevin L.; Simon, John; Jain, Nikhil</p> <p>2016-11-21</p> <p>Precise control of the growth of III-V materials by hydride <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy (HVPE) is complicated by the fact that the growth rate depends on the concentrations of nearly all inputs to the reactor and also the reaction temperature. This behavior is in contrast to metalorganic <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy (MOVPE), which in common practice operates in a mass transport limited regime where growth rate and alloy composition are controlled almost exclusively by flow of the Group III precursor. In HVPE, the growth rate and alloy compositions are very sensitive to temperature and reactant concentrations, which are strong functions of themore » reactor geometry. HVPE growth, particularly the growth of large area materials and devices, will benefit from the development of a growth model that can eventually be coupled with a computational fluid dynamics (CFD) model of a specific reactor geometry. In this work, we develop a growth rate law using a Langmuir-Hinshelwood (L-H) analysis, fitting unknown parameters to growth rate data from the literature that captures the relevant kinetic and thermodynamic phenomena of the HVPE process. We compare the L-H rate law to growth rate data from our custom HVPE reactor, and develop quantitative insight into reactor performance, demonstrating the utility of the growth model.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvB..97i4415M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvB..97i4415M"><span><span class="hlt">Phase</span> <span class="hlt">separation</span> and second-order <span class="hlt">phase</span> transition in the phenomenological model for a Coulomb-frustrated two-dimensional system</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mamin, R. F.; Shaposhnikova, T. S.; Kabanov, V. V.</p> <p>2018-03-01</p> <p>We have considered the model of the <span class="hlt">phase</span> transition of the second order for the Coulomb frustrated 2D charged system. The coupling of the order parameter with the charge was considered as the local temperature. We have found that in such a system, an appearance of the <span class="hlt">phase-separated</span> state is possible. By numerical simulation, we have obtained different types ("stripes," "rings," "snakes") of <span class="hlt">phase-separated</span> states and determined the parameter ranges for these states. Thus the system undergoes a series of <span class="hlt">phase</span> transitions when the temperature decreases. First, the system moves from the homogeneous state with a zero order parameter to the <span class="hlt">phase-separated</span> state with two <span class="hlt">phases</span> in one of which the order parameter is zero and, in the other, it is nonzero (τ >0 ). Then a first-order transition occurs to another <span class="hlt">phase-separated</span> state, in which both <span class="hlt">phases</span> have different and nonzero values of the order parameter (for τ <0 ). Only a further decrease of temperature leads to a transition to a homogeneous ordered state.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015JCrGr.418..145K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015JCrGr.418..145K"><span>Effect of gamma-ray irradiation on structural properties of GaAsN films grown by metal organic <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Klangtakai, Pawinee; Sanorpim, Sakuntam; Wattanawareekul, Atiwat; Suwanyangyaun, Pattana; Srepusharawoot, Pornjuk; Onabe, Kentaro</p> <p>2015-05-01</p> <p>The effects of gamma-ray irradiation on the structural properties of GaAs1-xNx films (N concentration=1.9 and 5.1 at%) grown by metal organic <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy on GaAs (001) substrates were investigated. The GaAs1-xNx films were irradiated by gamma rays with irradiation strength of 0-2.0 MGy. Scanning electron microscopy and atomic force microscopy results showed that a gamma ray with a strength of 0, 0.5, 1.0, 1.5, and 2.0 MGy formed holes with a density of 0.0, 8.8, 9.4, 11.5, and 11.9 μm-2, respectively, on the surface of a GaAs0.981N0.019 film with low N content. On the other hand, the irradiated high-N-content GaAs0.949N0.051 film exhibited a cross-hatch pattern, which was <span class="hlt">induced</span> by partial strain relaxation at high N levels, with a line density of 0.0, 0.21, 0.37, 0.67, and 0.26 μm-1 corresponding to an irradiation strength of 0, 0.5, 1.0, 1.5, and 2.0 MGy, respectively. The high-resolution X-ray diffraction and Raman scattering results revealed an increase in N incorporation and strain relaxation after irradiation. In addition, the GaAs0.949N0.051 films exhibited <span class="hlt">phase</span> <span class="hlt">separation</span>, which took place via N out-diffusion across the interface when the irradiation strength exceeded 1.0 MGy. Based on these results, the main cause of structural change was determined to be the irradiation effects including displacement damage and gamma-ray heating.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017NatSR...746281C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017NatSR...746281C"><span>Ultrahigh Responsivity and Detectivity Graphene-Perovskite Hybrid Phototransistors by Sequential <span class="hlt">Vapor</span> Deposition</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chang, Po-Han; Liu, Shang-Yi; Lan, Yu-Bing; Tsai, Yi-Chen; You, Xue-Qian; Li, Chia-Shuo; Huang, Kuo-You; Chou, Ang-Sheng; Cheng, Tsung-Chin; Wang, Juen-Kai; Wu, Chih-I.</p> <p>2017-04-01</p> <p>In this work, graphene-methylammonium lead iodide (MAPbI3) perovskite hybrid phototransistors fabricated by sequential <span class="hlt">vapor</span> deposition are demonstrated. Ultrahigh responsivity of 1.73 × 107 A W-1 and detectivity of 2 × 1015 Jones are achieved, with extremely high effective quantum efficiencies of about 108% in the visible range (450-700 nm). This excellent performance is attributed to the ultra-flat perovskite films grown by <span class="hlt">vapor</span> deposition on the graphene sheets. The hybrid structure of graphene covered with uniform perovskite has high exciton <span class="hlt">separation</span> ability under light exposure, and thus efficiently generates photocurrents. This paper presents photoluminescence (PL) images along with statistical analysis used to study the photo-<span class="hlt">induced</span> exciton behavior. Both uniform and dramatic PL intensity quenching has been observed over entire measured regions, consistently demonstrating excellent exciton <span class="hlt">separation</span> in the devices.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_16");'>16</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li class="active"><span>18</span></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_18 --> <div id="page_19" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li class="active"><span>19</span></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="361"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhRvM...2d0401N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhRvM...2d0401N"><span>In situ ESEM imaging of the <span class="hlt">vapor</span>-pressure-dependent sublimation-<span class="hlt">induced</span> morphology of ice</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nair, Malavika; Husmann, Anke; Cameron, Ruth E.; Best, Serena M.</p> <p>2018-04-01</p> <p>Sublimation is a fundamental <span class="hlt">phase</span> transition that has a profound impact on both natural phenomena and advanced manufacturing technologies. Although great strides have been made in the study of ice growth from melt and <span class="hlt">vapor</span>, little consideration has been given to the effect of sublimation on the morphological features that develop in the ice microstructure. In this experimental study, we demonstrate the effect of <span class="hlt">vapor</span> pressure on the mesoscopic faceting observed and show that a <span class="hlt">vapor</span>-pressure-specific wavelength characterizes the periodic features that arise during sublimation. The ability to control the length scale of these features not only provides us with new insights into the mesoscopic roughness of ice crystals, but also presents the potential to exploit this effect in a plethora of applications from comet dating to ice-templated tissue engineering scaffolds.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013AGUFM.A54C..07O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013AGUFM.A54C..07O"><span>Spectro-microscopic Characterization of Physical Properties and <span class="hlt">Phase</span> <span class="hlt">Separations</span> in Individual Atmospheric Particles</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>OBrien, R. E.; Wang, B.; Neu, A.; Kelly, S. T.; Lundt, N.; Epstein, S. A.; MacMillan, A.; You, Y.; Laskin, A.; Nizkorodov, S.; Bertram, A. K.; Moffet, R.; Gilles, M.</p> <p>2013-12-01</p> <p>The <span class="hlt">phase</span> state and liquid-liquid <span class="hlt">phase</span> <span class="hlt">separations</span> of ambient and laboratory generated aerosol particles were investigated using (1) scanning transmission x-ray microscopy/near-edge x-ray absorption fine structure spectroscopy (STXM/NEXAFS) coupled to a relative humidity (RH) controlled in-situ chamber and (2) environmental scanning electron microscopy (ESEM). The <span class="hlt">phase</span> states of the particles were determined from measurements of their size and optical density. A comparison is made between the observed <span class="hlt">phase</span> states of ambient samples and of laboratory generated aerosols to determine how well laboratory samples represent the <span class="hlt">phase</span> of ambient samples. In addition, liquid-liquid <span class="hlt">phase</span> <span class="hlt">separations</span> in laboratory generated particles were investigated. Preliminary results showing that liquid-liquid <span class="hlt">phase</span> <span class="hlt">separations</span> occur at RH's between the deliquescence and efflorescence points and that the organic <span class="hlt">phase</span> surrounds the inorganic <span class="hlt">phase</span> will be presented. The STXM/NEXAFS technique provides insight into the degree of mixing at the deliquescence point and the degree of <span class="hlt">phase</span> <span class="hlt">separation</span> for particles of atmospherically relevant sizes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009PhRvL.102g7205S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009PhRvL.102g7205S"><span>Intrinsic Tunneling in <span class="hlt">Phase</span> <span class="hlt">Separated</span> Manganites</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Singh-Bhalla, G.; Selcuk, S.; Dhakal, T.; Biswas, A.; Hebard, A. F.</p> <p>2009-02-01</p> <p>We present evidence of direct electron tunneling across intrinsic insulating regions in submicrometer wide bridges of the <span class="hlt">phase-separated</span> ferromagnet (La,Pr,Ca)MnO3. Upon cooling below the Curie temperature, a predominantly ferromagnetic supercooled state persists where tunneling across the intrinsic tunnel barriers (ITBs) results in metastable, temperature-independent, high-resistance plateaus over a large range of temperatures. Upon application of a magnetic field, our data reveal that the ITBs are extinguished resulting in sharp, colossal, low-field resistance drops. Our results compare well to theoretical predictions of magnetic domain walls coinciding with the intrinsic insulating <span class="hlt">phase</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1185819-role-chalcogen-vapor-annealing-inducing-bulk-superconductivity-fe1+yte1-xsex-how-does-annealing-chalcogen-vapor-induce-superconductivity-fe1+yte-xsex','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1185819-role-chalcogen-vapor-annealing-inducing-bulk-superconductivity-fe1+yte1-xsex-how-does-annealing-chalcogen-vapor-induce-superconductivity-fe1+yte-xsex"><span>Role of chalcogen <span class="hlt">vapor</span> annealing in <span class="hlt">inducing</span> bulk superconductivity in Fe 1+yTe 1-xSe x [How does annealing in chalcogen <span class="hlt">vapor</span> <span class="hlt">induce</span> superconductivity in Fe 1+yTe -xSe x?</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Lin, Wenzhi; Ganesh, P.; Gianfrancesco, Anthony; ...</p> <p>2015-02-27</p> <p>Recent investigations have shown that Fe 1+yTe 1-xSe x can be made superconducting by annealing it in Se and O <span class="hlt">vapors</span>. The current lore is that these chalcogen <span class="hlt">vapors</span> <span class="hlt">induce</span> superconductivity by removing the magnetic excess Fe atoms. To investigate this phenomenon we performed a combination of magnetic susceptibility, specific heat and transport measurements together with scanning tunneling microscopy and spectroscopy and density functional theory calculations on Fe 1+yTe 1-xSe x treated with Te <span class="hlt">vapor</span>. We conclude that the main role of the Te <span class="hlt">vapor</span> is to quench the magnetic moments of the excess Fe atoms by forming FeTe mmore » (m ≥ 1) complexes. We show that the remaining FeTe m complexes are still damaging to the superconductivity and therefore that their removal potentially could further improve superconductive properties in these compounds.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23394745','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23394745"><span><span class="hlt">Separation</span> of piracetam derivatives on polysaccharide-based chiral stationary <span class="hlt">phases</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kažoka, H; Koliškina, O; Veinberg, G; Vorona, M</p> <p>2013-03-15</p> <p>High-performance liquid chromatography was used for the enantiomeric <span class="hlt">separation</span> of two chiral piracetam derivatives. The suitability of six commercially available polysaccharide-based chiral stationary <span class="hlt">phases</span> (CSPs) under normal <span class="hlt">phase</span> mode for direct enantioseparation has been investigated. The influence of the CSPs as well the nature and content of an alcoholic modifier in the mobile <span class="hlt">phase</span> on <span class="hlt">separation</span> and elution order was studied. It was established that CSP Lux Amylose-2 shows high chiral recognition ability towards 4-phenylsubstituted piracetam derivatives. Copyright © 2013 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20120010417','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20120010417"><span>Microgravity Passive <span class="hlt">Phase</span> <span class="hlt">Separator</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Paragano, Matthew; Indoe, William; Darmetko, Jeffrey</p> <p>2012-01-01</p> <p>A new invention disclosure discusses a structure and process for <span class="hlt">separating</span> gas from liquids in microgravity. The Microgravity Passive <span class="hlt">Phase</span> <span class="hlt">Separator</span> consists of two concentric, pleated, woven stainless- steel screens (25-micrometer nominal pore) with an axial inlet, and an annular outlet between both screens (see figure). Water enters at one end of the center screen at high velocity, eventually passing through the inner screen and out through the annular exit. As gas is introduced into the flow stream, the drag force exerted on the bubble pushes it downstream until flow stagnation or until it reaches an equilibrium point between the surface tension holding bubble to the screen and the drag force. Gas bubbles of a given size will form a front that is moved further down the length of the inner screen with increasing velocity. As more bubbles are added, the front location will remain fixed, but additional bubbles will move to the end of the unit, eventually coming to rest in the large cavity between the unit housing and the outer screen (storage area). Owing to the small size of the pores and the hydrophilic nature of the screen material, gas does not pass through the screen and is retained within the unit for emptying during ground processing. If debris is picked up on the screen, the area closest to the inlet will become clogged, so high-velocity flow will persist farther down the length of the center screen, pushing the bubble front further from the inlet of the inner screen. It is desired to keep the velocity high enough so that, for any bubble size, an area of clean screen exists between the bubbles and the debris. The primary benefits of this innovation are the lack of any need for additional power, strip gas, or location for venting the <span class="hlt">separated</span> gas. As the unit contains no membrane, the transport fluid will not be lost due to evaporation in the process of gas <span class="hlt">separation</span>. <span class="hlt">Separation</span> is performed with relatively low pressure drop based on the large surface</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24617952','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24617952"><span>New mechanism for autocatalytic decomposition of H2CO3 in the <span class="hlt">vapor</span> <span class="hlt">phase</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ghoshal, Sourav; Hazra, Montu K</p> <p>2014-04-03</p> <p>In this article, we present high level ab initio calculations investigating the energetics of a new autocatalytic decomposition mechanism for carbonic acid (H2CO3) in the <span class="hlt">vapor</span> <span class="hlt">phase</span>. The calculation have been performed at the MP2 level of theory in conjunction with aug-cc-pVDZ, aug-cc-pVTZ, and 6-311++G(3df,3pd) basis sets as well as at the CCSD(T)/aug-cc-pVTZ level. The present study suggests that this new decomposition mechanism is effectively a near-barrierless process at room temperature and makes <span class="hlt">vapor</span> <span class="hlt">phase</span> of H2CO3 unstable even in the absence of water molecules. Our calculation at the MP2/aug-cc-pVTZ level predicts that the effective barrier, defined as the difference between the zero-point vibrational energy (ZPE) corrected energy of the transition state and the total energy of the isolated starting reactants in terms of bimolecular encounters, is nearly zero for the autocatalytic decomposition mechanism. The results at the CCSD(T)/aug-cc-pVTZ level of calculations suggest that the effective barrier, as defined above, is sensitive to some extent to the levels of calculations used, nevertheless, we find that the effective barrier height predicted at the CCSD(T)/aug-cc-pVTZ level is very small or in other words the autocatalytic decomposition mechanism presented in this work is a near-barrierless process as mentioned above. Thus, we suggest that this new autocatalytic decomposition mechanism has to be considered as the primary mechanism for the decomposition of carbonic acid, especially at its source, where the <span class="hlt">vapor</span> <span class="hlt">phase</span> concentration of H2CO3 molecules reaches its highest levels.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017ApPhL.110w2102K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017ApPhL.110w2102K"><span>Metalorganic <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy of AlN on sapphire with low etch pit density</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Koleske, D. D.; Figiel, J. J.; Alliman, D. L.; Gunning, B. P.; Kempisty, J. M.; Creighton, J. R.; Mishima, A.; Ikenaga, K.</p> <p>2017-06-01</p> <p>Using metalorganic <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy, methods were developed to achieve AlN films on sapphire with low etch pit density (EPD). Key to this achievement was using the same AlN growth recipe and only varying the pre-growth conditioning of the quartz-ware. After AlN growth, the quartz-ware was removed from the growth chamber and either exposed to room air or moved into the N2 purged glove box and exposed to H2O <span class="hlt">vapor</span>. After the quartz-ware was exposed to room air or H2O, the AlN film growth was found to be more reproducible, resulting in films with (0002) and (10-12) x-ray diffraction (XRD) rocking curve linewidths of 200 and 500 arc sec, respectively, and EPDs < 100 cm-2. The EPD was found to correlate with (0002) linewidths, suggesting that the etch pits are associated with open core screw dislocations similar to GaN films. Once reproducible AlN conditions were established using the H2O pre-treatment, it was found that even small doses of trimethylaluminum (TMAl)/NH3 on the quartz-ware surfaces generated AlN films with higher EPDs. The presence of these residual TMAl/NH3-derived coatings in metalorganic <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy (MOVPE) systems and their impact on the sapphire surface during heating might explain why reproducible growth of AlN on sapphire is difficult.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23097215','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23097215"><span><span class="hlt">Vapor-phase</span> fabrication of β-iron oxide nanopyramids for lithium-ion battery anodes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Carraro, Giorgio; Barreca, Davide; Cruz-Yusta, Manuel; Gasparotto, Alberto; Maccato, Chiara; Morales, Julián; Sada, Cinzia; Sánchez, Luis</p> <p>2012-12-07</p> <p>The other polymorph: A <span class="hlt">vapor-phase</span> route for the fabrication of β-Fe(2)O(3) nanomaterials on Ti substrates at 400-500 °C is reported. For the first time, the β polymorph is tested as anode for lithium batteries, exhibiting promising performances in terms of Li storage and rate capability. Copyright © 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016JChPh.144u4502F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016JChPh.144u4502F"><span><span class="hlt">Separating</span> the effects of repulsive and attractive forces on the <span class="hlt">phase</span> diagram, interfacial, and critical properties of simple fluids</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fuentes-Herrera, M.; Moreno-Razo, J. A.; Guzmán, O.; López-Lemus, J.; Ibarra-Tandi, B.</p> <p>2016-06-01</p> <p>Molecular simulations in the canonical and isothermal-isobaric ensembles were performed to study the effect of varying the shape of the intermolecular potential on the <span class="hlt">phase</span> diagram, critical, and interfacial properties of model fluids. The molecular interactions were modeled by the Approximate Non-Conformal (ANC) theory potentials. Unlike the Lennard-Jones or Morse potentials, the ANC interactions incorporate parameters (called softnesses) that modulate the steepness of the potential in their repulsive and attractive parts independently. This feature allowed us to <span class="hlt">separate</span> unambiguously the role of each region of the potential on setting the thermophysical properties. In particular, we found positive linear correlation between all critical coordinates and the attractive and repulsive softness, except for the critical density and the attractive softness which are negatively correlated. Moreover, we found that the physical properties related to <span class="hlt">phase</span> coexistence (such as span of the liquid <span class="hlt">phase</span> between the critical and triple points, variations in the P-T <span class="hlt">vaporization</span> curve, interface width, and surface tension) are more sensitive to changes in the attractive softness than to the repulsive one. Understanding the different roles of attractive and repulsive forces on <span class="hlt">phase</span> coexistence may contribute to developing more accurate models of liquids and their mixtures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27276958','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27276958"><span><span class="hlt">Separating</span> the effects of repulsive and attractive forces on the <span class="hlt">phase</span> diagram, interfacial, and critical properties of simple fluids.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Fuentes-Herrera, M; Moreno-Razo, J A; Guzmán, O; López-Lemus, J; Ibarra-Tandi, B</p> <p>2016-06-07</p> <p>Molecular simulations in the canonical and isothermal-isobaric ensembles were performed to study the effect of varying the shape of the intermolecular potential on the <span class="hlt">phase</span> diagram, critical, and interfacial properties of model fluids. The molecular interactions were modeled by the Approximate Non-Conformal (ANC) theory potentials. Unlike the Lennard-Jones or Morse potentials, the ANC interactions incorporate parameters (called softnesses) that modulate the steepness of the potential in their repulsive and attractive parts independently. This feature allowed us to <span class="hlt">separate</span> unambiguously the role of each region of the potential on setting the thermophysical properties. In particular, we found positive linear correlation between all critical coordinates and the attractive and repulsive softness, except for the critical density and the attractive softness which are negatively correlated. Moreover, we found that the physical properties related to <span class="hlt">phase</span> coexistence (such as span of the liquid <span class="hlt">phase</span> between the critical and triple points, variations in the P-T <span class="hlt">vaporization</span> curve, interface width, and surface tension) are more sensitive to changes in the attractive softness than to the repulsive one. Understanding the different roles of attractive and repulsive forces on <span class="hlt">phase</span> coexistence may contribute to developing more accurate models of liquids and their mixtures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20722461','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20722461"><span>Film thickness dependence of <span class="hlt">phase</span> <span class="hlt">separation</span> and dewetting behaviors in PMMA/SAN blend films.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>You, Jichun; Liao, Yonggui; Men, Yongfeng; Shi, Tongfei; An, Lijia</p> <p>2010-09-21</p> <p>Film thickness dependence of complex behaviors coupled by <span class="hlt">phase</span> <span class="hlt">separation</span> and dewetting in blend [poly(methyl methacrylate) (PMMA) and poly(styrene-ran-acrylonitrile) (SAN)] films on silicon oxide substrate at 175 °C was investigated by grazing incidence ultrasmall-angle X-ray scattering (GIUSAX) and in situ atomic force microscopy (AFM). It was found that the dewetting pathway was under the control of the parameter U(q0)/E, which described the initial amplitude of the surface undulation and original thickness of film, respectively. Furthermore, our results showed that interplay between <span class="hlt">phase</span> <span class="hlt">separation</span> and dewetting depended crucially on film thickness. Three mechanisms including dewetting-<span class="hlt">phase</span> <span class="hlt">separation</span>/wetting, dewetting/wetting-<span class="hlt">phase</span> <span class="hlt">separation</span>, and <span class="hlt">phase</span> <span class="hlt">separation</span>/wetting-pseudodewetting were discussed in detail. In conclusion, it is relative rates of <span class="hlt">phase</span> <span class="hlt">separation</span> and dewetting that dominate the interplay between them.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19174874','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19174874"><span>Correlations between water-soluble organic aerosol and water <span class="hlt">vapor</span>: a synergistic effect from biogenic emissions?</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hennigan, Christopher J; Bergin, Michael H; Weber, Rodney J</p> <p>2008-12-15</p> <p>Ground-based measurements of meteorological parameters and water-soluble organic carbon in the gas(WSOCg) and particle (WSOCp) <span class="hlt">phases</span> were carried out in Atlanta, Georgia, from May to September 2007. Fourteen <span class="hlt">separate</span> events were observed throughout the summer in which WSOCp and water <span class="hlt">vapor</span> concentrations were highly correlated (average WSOCp-water <span class="hlt">vapor</span> r = 0.92); however, for the entire summer, no well-defined relationship existed between the two. The correlation events, which lasted on average 19 h, were characterized by a wide range of WSOCp and water <span class="hlt">vapor</span> concentrations. Several hypotheses for the correlation are explored, including heterogeneous liquid <span class="hlt">phase</span> SOA formation and the co-emission of biogenic VOCs and water <span class="hlt">vapor</span>. The data provide supporting evidence for contributions from both and suggest the possibility of a synergistic effect between the co-emission of water <span class="hlt">vapor</span> and VOCs from biogenic sources on SOA formation. Median WSOCp concentrations were also correlated with elemental carbon (EC), although this correlation extended over the entire summer. Despite the emission of water <span class="hlt">vapor</span> from anthropogenic mobile sources and the WSOCp-EC correlation, mobile sources were not considered a potential cause for the WSOCp-water <span class="hlt">vapor</span> correlations because of their low contribution to the water <span class="hlt">vapor</span> budget. Meteorology could perhaps have influenced the WSOCp-EC correlation, but other factors are implicated as well. Overall, the results suggest that the temperature-dependent co-emission of water <span class="hlt">vapor</span> through evapotranspiration and SOA precursor-VOCs by vegetation may be an important process contributing to SOA in some environments.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19850003853','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19850003853"><span>Sintering behavior of ultrafine silicon carbide powders obtained by <span class="hlt">vapor</span> <span class="hlt">phase</span> reaction</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Okabe, Y.; Miyachi, K.; Hojo, J.; Kato, A.</p> <p>1984-01-01</p> <p>The sintering behavior of ultrafine SiC powder with average particle size of about 0.01-0.06 microns produced by a <span class="hlt">vapor</span> <span class="hlt">phase</span> reaction of the Me4Si-H2 system was studied at the temperature range of 1400-2050 deg. It was found that the homogeneous dispersion of C on SiC particles is important to remove the surface oxide layer effectively. B and C and inhibitive effect on SiC grain growth.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018APExp..11f5502N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018APExp..11f5502N"><span>Ultrahigh-yield growth of GaN via halogen-free <span class="hlt">vapor-phase</span> epitaxy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nakamura, Daisuke; Kimura, Taishi</p> <p>2018-06-01</p> <p>The material yield of Ga during GaN growth via halogen-free <span class="hlt">vapor-phase</span> epitaxy (HF-VPE) was systematically investigated and found to be much higher than that obtained using conventional hydride VPE. This is attributed to the much lower process pressure and shorter seed-to-source distance, owing to the inherent chemical reactions and corresponding reactor design used for HF-VPE growth. Ultrahigh-yield GaN growth was demonstrated on a 4-in.-diameter sapphire seed substrate.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29152026','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29152026"><span>Particle <span class="hlt">separation</span> by <span class="hlt">phase</span> modulated surface acoustic waves.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Simon, Gergely; Andrade, Marco A B; Reboud, Julien; Marques-Hueso, Jose; Desmulliez, Marc P Y; Cooper, Jonathan M; Riehle, Mathis O; Bernassau, Anne L</p> <p>2017-09-01</p> <p>High efficiency isolation of cells or particles from a heterogeneous mixture is a critical processing step in lab-on-a-chip devices. Acoustic techniques offer contactless and label-free manipulation, preserve viability of biological cells, and provide versatility as the applied electrical signal can be adapted to various scenarios. Conventional acoustic <span class="hlt">separation</span> methods use time-of-flight and achieve <span class="hlt">separation</span> up to distances of quarter wavelength with limited <span class="hlt">separation</span> power due to slow gradients in the force. The method proposed here allows <span class="hlt">separation</span> by half of the wavelength and can be extended by repeating the modulation pattern and can ensure maximum force acting on the particles. In this work, we propose an optimised <span class="hlt">phase</span> modulation scheme for particle <span class="hlt">separation</span> in a surface acoustic wave microfluidic device. An expression for the acoustic radiation force arising from the interaction between acoustic waves in the fluid was derived. We demonstrated, for the first time, that the expression of the acoustic radiation force differs in surface acoustic wave and bulk devices, due to the presence of a geometric scaling factor. Two <span class="hlt">phase</span> modulation schemes are investigated theoretically and experimentally. Theoretical findings were experimentally validated for different mixtures of polystyrene particles confirming that the method offers high selectivity. A Monte-Carlo simulation enabled us to assess performance in real situations, including the effects of particle size variation and non-uniform acoustic field on sorting efficiency and purity, validating the ability to <span class="hlt">separate</span> particles with high purity and high resolution.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1436439','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1436439"><span>Kinetic and Mechanistic Study of <span class="hlt">Vapor-Phase</span> Free Radical Polymerization onto Liquid Surfaces</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Gupta, Malancha</p> <p></p> <p>The primary objective of this proposal was to study <span class="hlt">vapor</span> deposition of polymers onto liquid surfaces. Deposition onto liquid surfaces is a relatively new area of research because the past few decades have focused on deposition onto solid materials. We used initiated chemical <span class="hlt">vapor</span> deposition to deposit polymers onto the liquid surfaces. The process is a one-step, solventless, free-radical polymerization process in which monomer and initiator molecules are flowed into a vacuum chamber. We found that the surface tension interaction between the polymer and the liquid determines whether a film or nanoparticles are formed. We also found that we couldmore » form gels by using soluble monomers. We found that we could tune the size of the nanoparticles by varying the viscosity of the liquid and the process parameters including pressure and time. These insights allow scalable synthesis of polymer materials for a variety of <span class="hlt">separation</span> and catalysis applications.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70035590','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70035590"><span>Columnar jointing in <span class="hlt">vapor-phase</span>-altered, non-welded Cerro Galán Ignimbrite, Paycuqui, Argentina</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Wright, Heather M.; Lesti, Chiara; Cas, Ray A.F.; Porreca, Massimiliano; Viramonte, Jose G.; Folkes, Christopher B.; Giordano, Guido</p> <p>2011-01-01</p> <p>Columnar jointing is thought to occur primarily in lavas and welded pyroclastic flow deposits. However, the non-welded Cerro Galán Ignimbrite at Paycuqui, Argentina, contains well-developed columnar joints that are instead due to high-temperature <span class="hlt">vapor-phase</span> alteration of the deposit, where devitrification and <span class="hlt">vapor-phase</span> crystallization have increased the density and cohesion of the upper half of the section. Thermal remanent magnetization analyses of entrained lithic clasts indicate high emplacement temperatures, above 630°C, but the lack of welding textures indicates temperatures below the glass transition temperature. In order to remain below the glass transition at 630°C, the minimum cooling rate prior to deposition was 3.0 × 10−3–8.5 × 10−2°C/min (depending on the experimental data used for comparison). Alternatively, if the deposit was emplaced above the glass transition temperature, conductive cooling alone was insufficient to prevent welding. Crack patterns (average, 4.5 sides to each polygon) and column diameters (average, 75 cm) are consistent with relatively rapid cooling, where advective heat loss due to <span class="hlt">vapor</span> fluxing increases cooling over simple conductive heat transfer. The presence of regularly spaced, complex radiating joint patterns is consistent with fumarolic gas rise, where volatiles originated in the valley-confined drainage system below. Joint spacing is a proxy for cooling rates and is controlled by depositional thickness/valley width. We suggest that the formation of joints in high-temperature, non-welded deposits is aided by the presence of underlying external water, where <span class="hlt">vapor</span> transfer causes crystallization in pore spaces, densifies the deposit, and helps prevent welding.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MMTB..tmp...82L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MMTB..tmp...82L"><span>Rapid <span class="hlt">Separation</span> of Copper <span class="hlt">Phase</span> and Iron-Rich <span class="hlt">Phase</span> From Copper Slag at Low Temperature in a Super-Gravity Field</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lan, Xi; Gao, Jintao; Huang, Zili; Guo, Zhancheng</p> <p>2018-03-01</p> <p>A novel approach for quickly <span class="hlt">separating</span> a metal copper <span class="hlt">phase</span> and iron-rich <span class="hlt">phase</span> from copper slag at low temperature is proposed based on a super-gravity method. The morphology and mineral evolution of the copper slag with increasing temperature were studied using in situ high-temperature confocal laser scanning microscopy and ex situ scanning electron microscopy and X-ray diffraction methods. Fe3O4 particles dispersed among the copper slag were transformed into FeO by adding an appropriate amount of carbon as a reducing agent, forming the slag melt with SiO2 at low temperature and assisting <span class="hlt">separation</span> of the copper <span class="hlt">phase</span> from the slag. Consequently, in a super-gravity field, the metallic copper and copper matte were concentrated as the copper <span class="hlt">phase</span> along the super-gravity direction, whereas the iron-rich slag migrated in the opposite direction and was quickly <span class="hlt">separated</span> from the copper <span class="hlt">phase</span>. Increasing the gravity coefficient (G) significantly enhanced the <span class="hlt">separation</span> efficiency. After super-gravity <span class="hlt">separation</span> at G = 1000 and 1473 K (1200 °C) for 3 minutes, the mass fraction of Cu in the <span class="hlt">separated</span> copper <span class="hlt">phase</span> reached 86.11 wt pct, while that in the <span class="hlt">separated</span> iron-rich <span class="hlt">phase</span> was reduced to 0.105 wt pct. The recovery ratio of Cu in the copper <span class="hlt">phase</span> was as high as up to 97.47 pct.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MMTB...49.1165L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MMTB...49.1165L"><span>Rapid <span class="hlt">Separation</span> of Copper <span class="hlt">Phase</span> and Iron-Rich <span class="hlt">Phase</span> From Copper Slag at Low Temperature in a Super-Gravity Field</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lan, Xi; Gao, Jintao; Huang, Zili; Guo, Zhancheng</p> <p>2018-06-01</p> <p>A novel approach for quickly <span class="hlt">separating</span> a metal copper <span class="hlt">phase</span> and iron-rich <span class="hlt">phase</span> from copper slag at low temperature is proposed based on a super-gravity method. The morphology and mineral evolution of the copper slag with increasing temperature were studied using in situ high-temperature confocal laser scanning microscopy and ex situ scanning electron microscopy and X-ray diffraction methods. Fe3O4 particles dispersed among the copper slag were transformed into FeO by adding an appropriate amount of carbon as a reducing agent, forming the slag melt with SiO2 at low temperature and assisting <span class="hlt">separation</span> of the copper <span class="hlt">phase</span> from the slag. Consequently, in a super-gravity field, the metallic copper and copper matte were concentrated as the copper <span class="hlt">phase</span> along the super-gravity direction, whereas the iron-rich slag migrated in the opposite direction and was quickly <span class="hlt">separated</span> from the copper <span class="hlt">phase</span>. Increasing the gravity coefficient (G) significantly enhanced the <span class="hlt">separation</span> efficiency. After super-gravity <span class="hlt">separation</span> at G = 1000 and 1473 K (1200 °C) for 3 minutes, the mass fraction of Cu in the <span class="hlt">separated</span> copper <span class="hlt">phase</span> reached 86.11 wt pct, while that in the <span class="hlt">separated</span> iron-rich <span class="hlt">phase</span> was reduced to 0.105 wt pct. The recovery ratio of Cu in the copper <span class="hlt">phase</span> was as high as up to 97.47 pct.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_17");'>17</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li class="active"><span>19</span></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_19 --> <div id="page_20" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li class="active"><span>20</span></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="381"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4664338','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4664338"><span>Wetting and <span class="hlt">phase</span> <span class="hlt">separation</span> in soft adhesion</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Jensen, Katharine E.; Sarfati, Raphael; Style, Robert W.; Boltyanskiy, Rostislav; Chakrabarti, Aditi; Chaudhury, Manoj K.; Dufresne, Eric R.</p> <p>2015-01-01</p> <p>In the classic theory of solid adhesion, surface energy drives deformation to increase contact area whereas bulk elasticity opposes it. Recently, solid surface stress has been shown also to play an important role in opposing deformation of soft materials. This suggests that the contact line in soft adhesion should mimic that of a liquid droplet, with a contact angle determined by surface tensions. Consistent with this hypothesis, we observe a contact angle of a soft silicone substrate on rigid silica spheres that depends on the surface functionalization but not the sphere size. However, to satisfy this wetting condition without a divergent elastic stress, the gel <span class="hlt">phase</span> <span class="hlt">separates</span> from its solvent near the contact line. This creates a four-<span class="hlt">phase</span> contact zone with two additional contact lines hidden below the surface of the substrate. Whereas the geometries of these contact lines are independent of the size of the sphere, the volume of the <span class="hlt">phase-separated</span> region is not, but rather depends on the indentation volume. These results indicate that theories of adhesion of soft gels need to account for both the compressibility of the gel network and a nonzero surface stress between the gel and its solvent. PMID:26553989</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24148387','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24148387"><span>Development of a novel amide-silica stationary <span class="hlt">phase</span> for the reversed-<span class="hlt">phase</span> HPLC <span class="hlt">separation</span> of different classes of phytohormones.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Aral, Hayriye; Aral, Tarık; Ziyadanoğulları, Berrin; Ziyadanoğulları, Recep</p> <p>2013-11-15</p> <p>A novel amide-bonded silica stationary <span class="hlt">phase</span> was prepared starting from N-Boc-phenylalanine, cyclohexylamine and spherical silica gel (4 µm, 60 Å). The amide ligand was synthesised with high yield. The resulting amide bonded stationary <span class="hlt">phase</span> was characterised by SEM, IR and elemental analysis. The resulting selector bearing a polar amide group is used for the reversed-<span class="hlt">phase</span> chromatography <span class="hlt">separation</span> of different classes of thirteen phytohormones (plant hormones). The chromatographic behaviours of these analytes on the amide-silica stationary <span class="hlt">phase</span> were compared with those of RP-C18 column under same conditions. The effects of different <span class="hlt">separation</span> conditions, such as mobile <span class="hlt">phase</span>, pH value, flow rate and temperature, on the <span class="hlt">separation</span> and retention behaviours of the 13 phytohormones in this system were studied. The optimum <span class="hlt">separation</span> was achieved using reversed-<span class="hlt">phase</span> HPLC gradient elution with an aqueous mobile <span class="hlt">phase</span> containing pH=6.85 potassium phosphate buffer (20 mM) and acetonitrile with a 22 °C column temperature. Under these experimental conditions, the 12 phytohormones could be <span class="hlt">separated</span> and detected at 230 or 270 nm within 26 min. Copyright © 2013 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17410252','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17410252"><span>Four-wave parametric oscillation in sodium <span class="hlt">vapor</span> by electromagnetically <span class="hlt">induced</span> diffraction.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Harada, Ken-ichi; Ogata, Minoru; Mitsunaga, Masaharu</p> <p>2007-05-01</p> <p>We have observed a novel type of parametric oscillation in sodium atomic <span class="hlt">vapor</span> where four off-axis signal waves simultaneously build up under resonant and counterpropagating pump beams with elliptical beam profiles. The four waves, two of them Stokes shifted and the other two anti-Stokes shifted, have similar output powers of up to 10 mW with a conversion efficiency of 30% and are parametrically coupled by electromagnetically <span class="hlt">induced</span> diffraction.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23369269','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23369269"><span>Biodegradation of high concentrations of benzene <span class="hlt">vapors</span> in a two <span class="hlt">phase</span> partition stirred tank bioreactor.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Karimi, Ali; Golbabaei, Farideh; Neghab, Masoud; Pourmand, Mohammad Reza; Nikpey, Ahmad; Mohammad, Kazem; Mehrnia, Momammad Reza</p> <p>2013-01-15</p> <p>The present study examined the biodegradation rate of benzene <span class="hlt">vapors</span> in a two <span class="hlt">phase</span> stirred tank bioreactor by a bacterial consortium obtained from wastewater of an oil industry refinery house. Initially, the ability of the microbial consortium for degrading benzene was evaluated before running the bioreactor. The gaseous samples from inlet and outlet of bioreactor were directly injected into a gas chromatograph to determine benzene concentrations. Carbone oxide concentration at the inlet and outlet of bioreactor were also measured with a CO2 meter to determine the mineralization rate of benzene. Influence of the second non-aqueous <span class="hlt">phase</span> (silicon oil) has been emphasized, so at the first stage the removal efficiency (RE) and elimination capacity (EC) of benzene <span class="hlt">vapors</span> were evaluated without any organic <span class="hlt">phase</span> and in the second stage, 10% of silicon oil was added to bioreactor media as an organic <span class="hlt">phase</span>. Addition of silicon oil increased the biodegradation performance up to an inlet loading of 5580 mg/m3, a condition at which, the elimination capacity and removal efficiency were 181 g/m3/h and 95% respectively. The elimination rate of benzene increased by 38% in the presence of 10% of silicone oil. The finding of this study demonstrated that two <span class="hlt">phase</span> partition bioreactors (TPPBs) are potentially effective tools for the treatment of gas streams contaminated with high concentrations of poorly water soluble organic contaminant, such as benzene.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19730000383','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19730000383"><span><span class="hlt">Separation</span> of gas from liquid in a two-<span class="hlt">phase</span> flow system</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hayes, L. G.; Elliott, D. G.</p> <p>1973-01-01</p> <p><span class="hlt">Separation</span> system causes jets which leave two-<span class="hlt">phase</span> nozzles to impinge on each other, so that liquid from jets tends to coalesce in center of combined jet streams while gas <span class="hlt">phase</span> is forced to outer periphery. Thus, because liquid coalescence is achieved without resort to <span class="hlt">separation</span> with solid surfaces, cycle efficiency is improved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19930065665&hterms=Fuel+injection&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DFuel%2Binjection','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19930065665&hterms=Fuel+injection&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DFuel%2Binjection"><span>Atomization and <span class="hlt">vaporization</span> characteristics of airblast fuel injection inside a venturi tube</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Sun, H.; Chue, T.-H.; Lai, M.-C.; Tacina, R. R.</p> <p>1993-01-01</p> <p>This paper describes the experimental and numerical characterization of the capillary fuel injection, atomization, dispersion, and <span class="hlt">vaporization</span> of liquid fuel in a coflowing air stream inside a single venturi tube. The experimental techniques used are all laser-based. <span class="hlt">Phase</span> Doppler analyzer was used to characterize the atomization and <span class="hlt">vaporization</span> process. Planar laser-<span class="hlt">induced</span> fluorescence visualizations give good qualitative picture of the fuel droplet and <span class="hlt">vapor</span> distribution. Limited quantitative capabilities of the technique are also demonstrated. A modified version of the KIVA-II was used to simulate the entire spray process, including breakup and <span class="hlt">vaporization</span>. The advantage of venturi nozzle is demonstrated in terms of better atomization, more uniform F/A distribution, and less pressure drop. Multidimensional spray calculations can be used as a design tool only if care is taken for the proper breakup model, and wall impingement process.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27482542','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27482542"><span>Self-optimized superconductivity attainable by interlayer <span class="hlt">phase</span> <span class="hlt">separation</span> at cuprate interfaces.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Misawa, Takahiro; Nomura, Yusuke; Biermann, Silke; Imada, Masatoshi</p> <p>2016-07-01</p> <p>Stabilizing superconductivity at high temperatures and elucidating its mechanism have long been major challenges of materials research in condensed matter physics. Meanwhile, recent progress in nanostructuring offers unprecedented possibilities for designing novel functionalities. Above all, thin films of cuprate and iron-based high-temperature superconductors exhibit remarkably better superconducting characteristics (for example, higher critical temperatures) than in the bulk, but the underlying mechanism is still not understood. Solving microscopic models suitable for cuprates, we demonstrate that, at an interface between a Mott insulator and an overdoped nonsuperconducting metal, the superconducting amplitude is always pinned at the optimum achieved in the bulk, independently of the carrier concentration in the metal. This is in contrast to the dome-like dependence in bulk superconductors but consistent with the astonishing independence of the critical temperature from the carrier density x observed at the interfaces of La2CuO4 and La2-x Sr x CuO4. Furthermore, we identify a self-organization mechanism as responsible for the pinning at the optimum amplitude: An emergent electronic structure <span class="hlt">induced</span> by interlayer <span class="hlt">phase</span> <span class="hlt">separation</span> eludes bulk <span class="hlt">phase</span> <span class="hlt">separation</span> and inhomogeneities that would kill superconductivity in the bulk. Thus, interfaces provide an ideal tool to enhance and stabilize superconductivity. This interfacial example opens up further ways of shaping superconductivity by suppressing competing instabilities, with direct perspectives for designing devices.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4966878','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4966878"><span>Self-optimized superconductivity attainable by interlayer <span class="hlt">phase</span> <span class="hlt">separation</span> at cuprate interfaces</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Misawa, Takahiro; Nomura, Yusuke; Biermann, Silke; Imada, Masatoshi</p> <p>2016-01-01</p> <p>Stabilizing superconductivity at high temperatures and elucidating its mechanism have long been major challenges of materials research in condensed matter physics. Meanwhile, recent progress in nanostructuring offers unprecedented possibilities for designing novel functionalities. Above all, thin films of cuprate and iron-based high-temperature superconductors exhibit remarkably better superconducting characteristics (for example, higher critical temperatures) than in the bulk, but the underlying mechanism is still not understood. Solving microscopic models suitable for cuprates, we demonstrate that, at an interface between a Mott insulator and an overdoped nonsuperconducting metal, the superconducting amplitude is always pinned at the optimum achieved in the bulk, independently of the carrier concentration in the metal. This is in contrast to the dome-like dependence in bulk superconductors but consistent with the astonishing independence of the critical temperature from the carrier density x observed at the interfaces of La2CuO4 and La2−xSrxCuO4. Furthermore, we identify a self-organization mechanism as responsible for the pinning at the optimum amplitude: An emergent electronic structure <span class="hlt">induced</span> by interlayer <span class="hlt">phase</span> <span class="hlt">separation</span> eludes bulk <span class="hlt">phase</span> <span class="hlt">separation</span> and inhomogeneities that would kill superconductivity in the bulk. Thus, interfaces provide an ideal tool to enhance and stabilize superconductivity. This interfacial example opens up further ways of shaping superconductivity by suppressing competing instabilities, with direct perspectives for designing devices. PMID:27482542</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5395820','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5395820"><span>Ultrahigh Responsivity and Detectivity Graphene–Perovskite Hybrid Phototransistors by Sequential <span class="hlt">Vapor</span> Deposition</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Chang, Po-Han; Liu, Shang-Yi; Lan, Yu-Bing; Tsai, Yi-Chen; You, Xue-Qian; Li, Chia-Shuo; Huang, Kuo-You; Chou, Ang-Sheng; Cheng, Tsung-Chin; Wang, Juen-Kai; Wu, Chih-I</p> <p>2017-01-01</p> <p>In this work, graphene-methylammonium lead iodide (MAPbI3) perovskite hybrid phototransistors fabricated by sequential <span class="hlt">vapor</span> deposition are demonstrated. Ultrahigh responsivity of 1.73 × 107 A W−1 and detectivity of 2 × 1015 Jones are achieved, with extremely high effective quantum efficiencies of about 108% in the visible range (450–700 nm). This excellent performance is attributed to the ultra-flat perovskite films grown by <span class="hlt">vapor</span> deposition on the graphene sheets. The hybrid structure of graphene covered with uniform perovskite has high exciton <span class="hlt">separation</span> ability under light exposure, and thus efficiently generates photocurrents. This paper presents photoluminescence (PL) images along with statistical analysis used to study the photo-<span class="hlt">induced</span> exciton behavior. Both uniform and dramatic PL intensity quenching has been observed over entire measured regions, consistently demonstrating excellent exciton <span class="hlt">separation</span> in the devices. PMID:28422117</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18193910','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18193910"><span><span class="hlt">Phase</span> diagram of nanoscale alloy particles used for <span class="hlt">vapor</span>-liquid-solid growth of semiconductor nanowires.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sutter, Eli; Sutter, Peter</p> <p>2008-02-01</p> <p>We use transmission electron microscopy observations to establish the parts of the <span class="hlt">phase</span> diagram of nanometer sized Au-Ge alloy drops at the tips of Ge nanowires (NWs) that determine their temperature-dependent equilibrium composition and, hence, their exchange of semiconductor material with the NWs. We find that the <span class="hlt">phase</span> diagram of the nanoscale drop deviates significantly from that of the bulk alloy, which explains discrepancies between actual growth results and predictions on the basis of the bulk-<span class="hlt">phase</span> equilibria. Our findings provide the basis for tailoring <span class="hlt">vapor</span>-liquid-solid growth to achieve complex one-dimensional materials geometries.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20593796','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20593796"><span><span class="hlt">Separation</span> of multiphosphorylated peptide isomers by hydrophilic interaction chromatography on an aminopropyl <span class="hlt">phase</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Singer, David; Kuhlmann, Julia; Muschket, Matthias; Hoffmann, Ralf</p> <p>2010-08-01</p> <p>The <span class="hlt">separation</span> of isomeric phosphorylated peptides is challenging and often impossible for multiphosphorylated isomers using chromatographic and capillary electrophoretic methods. In this study we investigated the <span class="hlt">separation</span> of a set of single-, double-, and triple-phosphorylated peptides (corresponding to the human tau protein) by ion-pair reversed-<span class="hlt">phase</span> chromatography (IP-RPC) and hydrophilic interaction chromatography (HILIC). In HILIC both hydroxyl and aminopropyl stationary <span class="hlt">phases</span> were tested with aqueous acetonitrile in order to assess their <span class="hlt">separation</span> efficiency. The hydroxyl <span class="hlt">phase</span> <span class="hlt">separated</span> the phosphopeptides very well from the unphosphorylated analogue, while on the aminopropyl <span class="hlt">phase</span> even isomeric phosphopeptides attained baseline <span class="hlt">separation</span>. Thus, up to seven phosphorylated versions of a given tau domain were <span class="hlt">separated</span>. Furthermore, the low concentration of an acidic ammonium formate buffer allowed an online analysis with electrospray ionization tandem mass spectrometry (ESI-MS/MS) to be conducted, enabling peptide sequencing and identification of phosphorylation sites.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23800085','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23800085"><span><span class="hlt">Separating</span> homeologs by <span class="hlt">phasing</span> in the tetraploid wheat transcriptome.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Krasileva, Ksenia V; Buffalo, Vince; Bailey, Paul; Pearce, Stephen; Ayling, Sarah; Tabbita, Facundo; Soria, Marcelo; Wang, Shichen; Akhunov, Eduard; Uauy, Cristobal; Dubcovsky, Jorge</p> <p>2013-06-25</p> <p>The high level of identity among duplicated homoeologous genomes in tetraploid pasta wheat presents substantial challenges for de novo transcriptome assembly. To solve this problem, we develop a specialized bioinformatics workflow that optimizes transcriptome assembly and <span class="hlt">separation</span> of merged homoeologs. To evaluate our strategy, we sequence and assemble the transcriptome of one of the diploid ancestors of pasta wheat, and compare both assemblies with a benchmark set of 13,472 full-length, non-redundant bread wheat cDNAs. A total of 489 million 100 bp paired-end reads from tetraploid wheat assemble in 140,118 contigs, including 96% of the benchmark cDNAs. We used a comparative genomics approach to annotate 66,633 open reading frames. The multiple k-mer assembly strategy increases the proportion of cDNAs assembled full-length in a single contig by 22% relative to the best single k-mer size. Homoeologs are <span class="hlt">separated</span> using a post-assembly pipeline that includes polymorphism identification, <span class="hlt">phasing</span> of SNPs, read sorting, and re-assembly of <span class="hlt">phased</span> reads. Using a reference set of genes, we determine that 98.7% of SNPs analyzed are correctly <span class="hlt">separated</span> by <span class="hlt">phasing</span>. Our study shows that de novo transcriptome assembly of tetraploid wheat benefit from multiple k-mer assembly strategies more than diploid wheat. Our results also demonstrate that <span class="hlt">phasing</span> approaches originally designed for heterozygous diploid organisms can be used to <span class="hlt">separate</span> the close homoeologous genomes of tetraploid wheat. The predicted tetraploid wheat proteome and gene models provide a valuable tool for the wheat research community and for those interested in comparative genomic studies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3673304','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3673304"><span>Quantitative analysis of aqueous <span class="hlt">phase</span> composition of model dentin adhesives experiencing <span class="hlt">phase</span> <span class="hlt">separation</span></span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Ye, Qiang; Park, Jonggu; Parthasarathy, Ranganathan; Pamatmat, Francis; Misra, Anil; Laurence, Jennifer S.; Marangos, Orestes; Spencer, Paulette</p> <p>2013-01-01</p> <p>There have been reports of the sensitivity of our current dentin adhesives to excess moisture, for example, water-blisters in adhesives placed on over-wet surfaces, and <span class="hlt">phase</span> <span class="hlt">separation</span> with concomitant limited infiltration of the critical dimethacrylate component into the demineralized dentin matrix. To determine quantitatively the hydrophobic/hydrophilic components in the aqueous <span class="hlt">phase</span> when exposed to over-wet environments, model adhesives were mixed with 16, 33, and 50 wt % water to yield well-<span class="hlt">separated</span> <span class="hlt">phases</span>. Based upon high-performance liquid chromatography coupled with photodiode array detection, it was found that the amounts of hydrophobic BisGMA and hydrophobic initiators are less than 0.1 wt % in the aqueous <span class="hlt">phase</span>. The amount of these compounds decreased with an increase in the initial water content. The major components of the aqueous <span class="hlt">phase</span> were hydroxyethyl methacrylate (HEMA) and water, and the HEMA content ranged from 18.3 to 14.7 wt %. Different BisGMA homologues and the relative content of these homologues in the aqueous <span class="hlt">phase</span> have been identified; however, the amount of crosslinkable BisGMA was minimal and, thus, could not help in the formation of a crosslinked polymer network in the aqueous <span class="hlt">phase</span>. Without the protection afforded by a strong crosslinked network, the poorly photoreactive compounds of this aqueous <span class="hlt">phase</span> could be leached easily. These results suggest that adhesive formulations should be designed to include hydrophilic multimethacrylate monomers and water compatible initiators. PMID:22331596</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1351887','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1351887"><span>Low temperature <span class="hlt">vapor</span> <span class="hlt">phase</span> digestion of graphite</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Pierce, Robert A.</p> <p>2017-04-18</p> <p>A method for digestion and gasification of graphite for removal from an underlying surface is described. The method can be utilized to remove graphite remnants of a formation process from the formed metal piece in a cleaning process. The method can be particularly beneficial in cleaning castings formed with graphite molding materials. The method can utilize <span class="hlt">vaporous</span> nitric acid (HNO.sub.3) or <span class="hlt">vaporous</span> HNO.sub.3 with air/oxygen to digest the graphite at conditions that can avoid damage to the underlying surface.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008JPCM...20Q4228H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008JPCM...20Q4228H"><span>Strategies towards controlling strain-<span class="hlt">induced</span> mesoscopic <span class="hlt">phase</span> <span class="hlt">separation</span> in manganite thin films</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Habermeier, H.-U.</p> <p>2008-10-01</p> <p>Complex oxides represent a class of materials with a plethora of fascinating intrinsic physical functionalities. The intriguing interplay of charge, spin and orbital ordering in these systems superimposed by lattice effects opens a scientifically rewarding playground for both fundamental as well as application oriented research. The existence of nanoscale electronic <span class="hlt">phase</span> <span class="hlt">separation</span> in correlated complex oxides is one of the areas in this field whose impact on the current understanding of their physics and potential applications is not yet clear. In this paper this issue is treated from the point of view of complex oxide thin film technology. Commenting on aspects of complex oxide thin film growth gives an insight into the complexity of a reliable thin film technology for these materials. Exploring fundamentals of interfacial strain generation and strain accommodation paves the way to intentionally manipulate thin film properties. Furthermore, examples are given for an extrinsic continuous tuning of intrinsic electronic inhomogeneities in perovskite-type complex oxide thin films.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25534540','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25534540"><span>Hour-glass magnetic excitations <span class="hlt">induced</span> by nanoscopic <span class="hlt">phase</span> <span class="hlt">separation</span> in cobalt oxides.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Drees, Y; Li, Z W; Ricci, A; Rotter, M; Schmidt, W; Lamago, D; Sobolev, O; Rütt, U; Gutowski, O; Sprung, M; Piovano, A; Castellan, J P; Komarek, A C</p> <p>2014-12-23</p> <p>The magnetic excitations in the cuprate superconductors might be essential for an understanding of high-temperature superconductivity. In these cuprate superconductors the magnetic excitation spectrum resembles an hour-glass and certain resonant magnetic excitations within are believed to be connected to the pairing mechanism, which is corroborated by the observation of a universal linear scaling of superconducting gap and magnetic resonance energy. So far, charge stripes are widely believed to be involved in the physics of hour-glass spectra. Here we study an isostructural cobaltate that also exhibits an hour-glass magnetic spectrum. Instead of the expected charge stripe order we observe nano <span class="hlt">phase</span> <span class="hlt">separation</span> and unravel a microscopically split origin of hour-glass spectra on the nano scale pointing to a connection between the magnetic resonance peak and the spin gap originating in islands of the antiferromagnetic parent insulator. Our findings open new ways to theories of magnetic excitations and superconductivity in cuprate superconductors.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26754414','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26754414"><span><span class="hlt">Phase</span> <span class="hlt">separation</span> of comb polymer nanocomposite melts.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Xu, Qinzhi; Feng, Yancong; Chen, Lan</p> <p>2016-02-07</p> <p>In this work, the spinodal <span class="hlt">phase</span> demixing of branched comb polymer nanocomposite (PNC) melts is systematically investigated using the polymer reference interaction site model (PRISM) theory. To verify the reliability of the present method in characterizing the <span class="hlt">phase</span> behavior of comb PNCs, the intermolecular correlation functions of the system for nonzero particle volume fractions are compared with our molecular dynamics simulation data. After verifying the model and discussing the structure of the comb PNCs in the dilute nanoparticle limit, the interference among the side chain number, side chain length, nanoparticle-monomer size ratio and attractive interactions between the comb polymer and nanoparticles in spinodal demixing curves is analyzed and discussed in detail. The results predict two kinds of distinct <span class="hlt">phase</span> <span class="hlt">separation</span> behaviors. One is called classic fluid <span class="hlt">phase</span> boundary, which is mediated by the entropic depletion attraction and contact aggregation of nanoparticles at relatively low nanoparticle-monomer attraction strength. The second demixing transition occurs at relatively high attraction strength and involves the formation of an equilibrium physical network <span class="hlt">phase</span> with local bridging of nanoparticles. The <span class="hlt">phase</span> boundaries are found to be sensitive to the side chain number, side chain length, nanoparticle-monomer size ratio and attractive interactions. As the side chain length is fixed, the side chain number has a large effect on the <span class="hlt">phase</span> behavior of comb PNCs; with increasing side chain number, the miscibility window first widens and then shrinks. When the side chain number is lower than a threshold value, the <span class="hlt">phase</span> boundaries undergo a process from enlarging the miscibility window to narrowing as side chain length increases. Once the side chain number overtakes this threshold value, the <span class="hlt">phase</span> boundary shifts towards less miscibility. With increasing nanoparticle-monomer size ratio, a crossover of particle size occurs, above which the <span class="hlt">phase</span> <span class="hlt">separation</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/984340','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/984340"><span>Stratified <span class="hlt">vapor</span> generator</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Bharathan, Desikan [Lakewood, CO; Hassani, Vahab [Golden, CO</p> <p>2008-05-20</p> <p>A stratified <span class="hlt">vapor</span> generator (110) comprises a first heating section (H.sub.1) and a second heating section (H.sub.2). The first and second heating sections (H.sub.1, H.sub.2) are arranged so that the inlet of the second heating section (H.sub.2) is operatively associated with the outlet of the first heating section (H.sub.1). A moisture <span class="hlt">separator</span> (126) having a <span class="hlt">vapor</span> outlet (164) and a liquid outlet (144) is operatively associated with the outlet (124) of the second heating section (H.sub.2). A cooling section (C.sub.1) is operatively associated with the liquid outlet (144) of the moisture <span class="hlt">separator</span> (126) and includes an outlet that is operatively associated with the inlet of the second heating section (H.sub.2).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22494683-gas-phase-emitter-effect-lanthanum-within-ceramic-metal-halide-lamps-its-dependence-la-vapor-pressure-operating-frequency','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22494683-gas-phase-emitter-effect-lanthanum-within-ceramic-metal-halide-lamps-its-dependence-la-vapor-pressure-operating-frequency"><span>The gas <span class="hlt">phase</span> emitter effect of lanthanum within ceramic metal halide lamps and its dependence on the La <span class="hlt">vapor</span> pressure and operating frequency</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Ruhrmann, C.; Hoebing, T.; Bergner, A.</p> <p>2015-08-07</p> <p>The gas <span class="hlt">phase</span> emitter effect increases the lamp lifetime by lowering the work function and, with it, the temperature of the tungsten electrodes of metal halide lamps especially for lamps in ceramic vessels due to their high rare earth pressures. It is generated by a monolayer on the electrode surface of electropositive atoms of certain emitter elements, which are inserted into the lamp bulb by metal iodide salts. They are <span class="hlt">vaporized</span>, dissociated, ionized, and deposited by an emitter ion current onto the electrode surface within the cathodic <span class="hlt">phase</span> of lamp operation with a switched-dc or ac-current. The gas <span class="hlt">phase</span> emittermore » effect of La and the influence of Na on the emitter effect of La are studied by spatially and <span class="hlt">phase</span>-resolved pyrometric measurements of the electrode tip temperature, La atom, and ion densities by optical emission spectroscopy as well as optical broadband absorption spectroscopy and arc attachment images by short time photography. An addition of Na to the lamp filling increases the La <span class="hlt">vapor</span> pressure within the lamp considerably, resulting in an improved gas <span class="hlt">phase</span> emitter effect of La. Furthermore, the La <span class="hlt">vapor</span> pressure is raised by a heating of the cold spot. In this way, conditions depending on the La <span class="hlt">vapor</span> pressure and operating frequency are identified, at which the temperature of the electrodes becomes a minimum.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=MSFC-0300208&hterms=physics+experiment&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dphysics%2Bexperiment','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=MSFC-0300208&hterms=physics+experiment&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dphysics%2Bexperiment"><span>Movie of <span class="hlt">phase</span> <span class="hlt">separation</span> during physics of colloids in space experiment</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>2002-01-01</p> <p>Still photographs taken over 16 hours on Nov. 13, 2001, on the International Space Station have been condensed into a few seconds to show the de-mixing -- or <span class="hlt">phase</span> <span class="hlt">separation</span> -- process studied by the Experiment on Physics of Colloids in Space. Commanded from the ground, dozens of similar tests have been conducted since the experiment arrived on ISS in 2000. The sample is a mix of polymethylmethacrylate (PMMA or acrylic) colloids, polystyrene polymers and solvents. The circular area in the video is 2 cm (0.8 in.) in diameter. The <span class="hlt">phase</span> <span class="hlt">separation</span> process occurs spontaneously after the sample is mechanically mixed. The evolving lighter regions are rich in colloid and have the structure of a liquid. The dark regions are poor in colloids and have the structure of a gas. This behavior carnot be observed on Earth because gravity causes the particles to fall out of solution faster than the <span class="hlt">phase</span> <span class="hlt">separation</span> can occur. While similar to a gas-liquid <span class="hlt">phase</span> transition, the growth rate observed in this test is different from any atomic gas-liquid or liquid-liquid <span class="hlt">phase</span> transition ever measured experimentally. Ultimately, the sample <span class="hlt">separates</span> into colloid-poor and colloid-rich areas, just as oil and vinegar <span class="hlt">separate</span>. The fundamental science of de-mixing in this colloid-polymer sample is the same found in the annealing of metal alloys and plastic polymer blends. Improving the understanding of this process may lead to improving processing of these materials on Earth.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_18");'>18</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li class="active"><span>20</span></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_20 --> <div id="page_21" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li class="active"><span>21</span></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="401"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=MSFC-0300207&hterms=physics+experiment&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dphysics%2Bexperiment','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=MSFC-0300207&hterms=physics+experiment&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dphysics%2Bexperiment"><span><span class="hlt">Phase</span> <span class="hlt">separation</span> during the Experiment on Physics of Colloids in Space</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>2003-01-01</p> <p>Still photographs taken over 16 hours on Nov. 13, 2001, on the International Space Station have been condensed into a few seconds to show the de-mixing -- or <span class="hlt">phase</span> <span class="hlt">separation</span> -- process studied by the Experiment on Physics of Colloids in Space. Commanded from the ground, dozens of similar tests have been conducted since the experiment arrived on ISS in 2000. The sample is a mix of polymethylmethacrylate (PMMA or acrylic) colloids, polystyrene polymers and solvents. The circular area is 2 cm (0.8 in.) in diameter. The <span class="hlt">phase</span> <span class="hlt">separation</span> process occurs spontaneously after the sample is mechanically mixed. The evolving lighter regions are rich in colloid and have the structure of a liquid. The dark regions are poor in colloids and have the structure of a gas. This behavior carnot be observed on Earth because gravity causes the particles to fall out of solution faster than the <span class="hlt">phase</span> <span class="hlt">separation</span> can occur. While similar to a gas-liquid <span class="hlt">phase</span> transition, the growth rate observed in this test is different from any atomic gas-liquid or liquid-liquid <span class="hlt">phase</span> transition ever measured experimentally. Ultimately, the sample <span class="hlt">separates</span> into colloid-poor and colloid-rich areas, just as oil and vinegar <span class="hlt">separate</span>. The fundamental science of de-mixing in this colloid-polymer sample is the same found in the annealing of metal alloys and plastic polymer blends. Improving the understanding of this process may lead to improving processing of these materials on Earth.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JMMM..456..212S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JMMM..456..212S"><span>Thermal cycling effects on static and dynamic properties of a <span class="hlt">phase</span> <span class="hlt">separated</span> manganite</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sacanell, J.; Sievers, B.; Quintero, M.; Granja, L.; Ghivelder, L.; Parisi, F.</p> <p>2018-06-01</p> <p>In this work we address the interplay between two phenomena which are signatures of the out-of-equilibrium state in <span class="hlt">phase</span> <span class="hlt">separated</span> manganites: irreversibility against thermal cycling and aging/rejuvenation process. The sample investigated is La0.5Ca0.5MnO3, a prototypical manganite exhibiting <span class="hlt">phase</span> <span class="hlt">separation</span>. Two regimes for isothermal relaxation were observed according to the temperature range: for T > 100 K, aging/rejuvenation effects are observed, while for T < 100 K an irreversible aging was found. Our results show that thermal cycles act as a tool to unveil the dynamical behavior of the <span class="hlt">phase</span> <span class="hlt">separated</span> state in manganites, revealing the close interplay between static and dynamic properties of <span class="hlt">phase</span> <span class="hlt">separated</span> manganites.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4646775','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4646775"><span>Hierarchical multiscale hyperporous block copolymer membranes via tunable dual-<span class="hlt">phase</span> <span class="hlt">separation</span></span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Yoo, Seungmin; Kim, Jung-Hwan; Shin, Myoungsoo; Park, Hyungmin; Kim, Jeong-Hoon; Lee, Sang-Young; Park, Soojin</p> <p>2015-01-01</p> <p>The rational design and realization of revolutionary porous structures have been long-standing challenges in membrane science. We demonstrate a new class of amphiphilic polystyrene-block-poly(4-vinylpyridine) block copolymer (BCP)–based porous membranes featuring hierarchical multiscale hyperporous structures. The introduction of surface energy–modifying agents and the control of major <span class="hlt">phase</span> <span class="hlt">separation</span> parameters (such as nonsolvent polarity and solvent drying time) enable tunable dual-<span class="hlt">phase</span> <span class="hlt">separation</span> of BCPs, eventually leading to macro/nanoscale porous structures and chemical functionalities far beyond those accessible with conventional approaches. Application of this BCP membrane to a lithium-ion battery <span class="hlt">separator</span> affords exceptional improvement in electrochemical performance. The dual-<span class="hlt">phase</span> separation–driven macro/nanopore construction strategy, owing to its simplicity and tunability, is expected to be readily applicable to a rich variety of membrane fields including molecular <span class="hlt">separation</span>, water purification, and energy-related devices. PMID:26601212</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20090039479&hterms=kellogg&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dkellogg','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20090039479&hterms=kellogg&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dkellogg"><span>Architecture Study on Telemetry Coverage for Immediate Post-<span class="hlt">Separation</span> <span class="hlt">Phase</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Cheung, Kar-Ming; Lee, Charles H.; Kellogg, Kent H.; Stocklin, Frank J.; Zillig, David J.; Fielhauer, Karl B.</p> <p>2008-01-01</p> <p>This paper presents the preliminary results of an architecture study that provides continuous telemetry coverage for NASA missions for immediate post-<span class="hlt">separation</span> <span class="hlt">phase</span>. This study is a collaboration effort between Jet Propulsion Laboratory (JPL), Goddard Space Flight Center (GSFC), and Applied Physics Laboratory (APL). After launch when the spacecraft <span class="hlt">separated</span> from the upper stage, the spacecraft typically executes a number of mission-critical operations prior to the deployment of solar panels and the activation of the primary communication subsystem. JPL, GSFC, and APL have similar design principle statements that require continuous coverage of mission-critical telemetry during the immediate post-<span class="hlt">separation</span> <span class="hlt">phase</span>. To conform to these design principles, an architecture that consists of a <span class="hlt">separate</span> spacecraft transmitter and a robust communication network capable of tracking the spacecraft signals is needed.This paper presents the preliminary results of an architecture study that provides continuous telemetry coverage for NASA missions for immediate post-<span class="hlt">separation</span> <span class="hlt">phase</span>. This study is a collaboration effort between Jet Propulsion Laboratory (JPL), Goddard Space Flight Center (GSFC), and Applied Physics Laboratory (APL). After launch when the spacecraft <span class="hlt">separated</span> from the upper stage, the spacecraft typically executes a number of mission-critical operations prior to the deployment of solar panels and the activation of the primary communication subsystem. JPL, GSFC, and APL have similar design principle statements that require continuous coverage of mission-critical telemetry during the immediate post-<span class="hlt">separation</span> <span class="hlt">phase</span>. To conform to these design principles, an architecture that consists of a <span class="hlt">separate</span> spacecraft transmitter and a robust communication network capable of tracking the spacecraft signals is needed. The main results of this study are as follows: 1) At low altitude (< 10000 km) when most post-<span class="hlt">separation</span> critical operations are executed, Earth-based network</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..DFDG17004W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..DFDG17004W"><span>Low-frequency dynamics of pressure-<span class="hlt">induced</span> turbulent <span class="hlt">separation</span> bubbles</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Weiss, Julien; Mohammed-Taifour, Abdelouahab; Lefloch, Arnaud</p> <p>2017-11-01</p> <p>We experimentally investigate a pressure-<span class="hlt">induced</span> turbulent <span class="hlt">separation</span> bubble (TSB), which is generated on a flat test surface through a combination of adverse and favorable pressure gradients imposed on a nominally two-dimensional, incompressible, turbulent boundary layer. We probe the flow using piezo-resistive pressure transducers, MEMS shear-stress sensors, and high-speed, 2D-2C, PIV measurements. Through the use of Fourier analysis of the wall-pressure fluctuations and Proper Orthogonal Decomposition of the velocity fields, we show that this type of flow is characterized by a self-<span class="hlt">induced</span>, low-frequency contraction and expansion - called breathing - of the TSB. The dominant Strouhal number of this motion, based on the TSB length and the incoming velocity in the potential flow, is of the order of 0.01. We compare this motion to the low-frequency dynamics observed in laminar <span class="hlt">separation</span> bubbles (LSBs), geometry-<span class="hlt">induced</span> TSBs, and shock-<span class="hlt">induced</span> <span class="hlt">separated</span> flows.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26704546','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26704546"><span>Extent and mechanism of <span class="hlt">phase</span> <span class="hlt">separation</span> during the extrusion of calcium phosphate pastes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>O'Neill, Rory; McCarthy, Helen O; Cunningham, Eoin; Montufar, Edgar; Ginebra, Maria-Pau; Wilson, D Ian; Lennon, Alex; Dunne, Nicholas</p> <p>2016-02-01</p> <p>The aim of this study was to increase understanding of the mechanism and dominant drivers influencing <span class="hlt">phase</span> <span class="hlt">separation</span> during ram extrusion of calcium phosphate (CaP) paste for orthopaedic applications. The liquid content of extrudate was determined, and the flow of liquid and powder <span class="hlt">phases</span> within the syringe barrel during extrusion were observed, subject to various extrusion parameters. Increasing the initial liquid-to-powder mass ratio, LPR, (0.4-0.45), plunger rate (5-20 mm/min), and tapering the barrel exit (45°-90°) significantly reduced the extent of <span class="hlt">phase</span> <span class="hlt">separation</span>. <span class="hlt">Phase</span> <span class="hlt">separation</span> values ranged from (6.22 ± 0.69 to 18.94 ± 0.69 %). However altering needle geometry had no significant effect on <span class="hlt">phase</span> <span class="hlt">separation</span>. From powder tracing and liquid content determination, static zones of powder and a non-uniform liquid distribution was observed within the barrel. Measurements of extrudate and paste LPR within the barrel indicated that extrudate LPR remained constant during extrusion, while LPR of paste within the barrel decreased steadily. These observations indicate the mechanism of <span class="hlt">phase</span> <span class="hlt">separation</span> was located within the syringe barrel. Therefore <span class="hlt">phase</span> <span class="hlt">separation</span> can be attributed to either; (1) the liquid being forced downstream by an increase in pore pressure as a result of powder consolidation due to the pressure exerted by the plunger or (2) the liquid being drawn from paste within the barrel, due to suction, driven by dilation of the solids matrix at the barrel exit. Differentiating between these two mechanisms is difficult; however results obtained suggest that suction is the dominant <span class="hlt">phase</span> <span class="hlt">separation</span> mechanism occurring during extrusion of CaP paste.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1054363-hydration-induced-phase-separation-amphiphilic-polymer-matrices-its-influence-voclosporin-release','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1054363-hydration-induced-phase-separation-amphiphilic-polymer-matrices-its-influence-voclosporin-release"><span>Hydration-<span class="hlt">Induced</span> <span class="hlt">Phase</span> <span class="hlt">Separation</span> in Amphiphilic Polymer Matrices and its Influence on Voclosporin Release</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Khan, I. John; Murthy, N. Sanjeeva; Kohn, Joachim</p> <p>2015-10-30</p> <p>Voclosporin is a highly potent, new cyclosporine -- a derivative that is currently in <span class="hlt">Phase</span> 3 clinical trials in the USA as a potential treatment for inflammatory diseases of the eye. Voclosporin represents a number of very sparingly soluble drugs that are difficult to administer. It was selected as a model drug that is dispersed within amphiphilic polymer matrices, and investigated the changing morphology of the matrices using neutron and x-ray scattering during voclosporin release and polymer resorption. The hydrophobic segments of the amphiphilic polymer chain are comprised of desaminotyrosyl-tyrosine ethyl ester (DTE) and desaminotyrosyl-tyrosine (DT), and the hydrophilic componentmore » is poly(ethylene glycol) (PEG). Water uptake in these matrices resulted in the <span class="hlt">phase</span> <span class="hlt">separation</span> of hydrophobic and hydrophilic domains that are a few hundred Angstroms apart. These water-driven morphological changes influenced the release profile of voclosporin and facilitated a burst-free release from the polymer. No such morphological reorganization was observed in poly(lactide-co-glycolide) (PLGA), which exhibits an extended lag period, followed by a burst-like release of voclosporin when the polymer was degraded. An understanding of the effect of polymer composition on the hydration behavior is central to understanding and controlling the <span class="hlt">phase</span> behavior and resorption characteristics of the matrix for achieving long-term controlled release of hydrophobic drugs such as voclosporin.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003JPCRD..32.1387M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003JPCRD..32.1387M"><span>Correlations for <span class="hlt">Vapor</span> Nucleating Critical Embryo Parameters</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Magnusson, Lars-Erik; Koropchak, John A.; Anisimov, Michael P.; Poznjakovskiy, Valeriy M.; de la Mora, Juan Fernandez</p> <p>2003-12-01</p> <p>Condensation nucleation light scattering detection in principle works by converting the effluent of the chromatographic <span class="hlt">separation</span> into an aerosol and then selectively evaporating the mobile <span class="hlt">phase</span>, leaving less volatile analytes and nonvolatile impurities as dry aerosol particles. The dry particles produced are then exposed to an environment that is saturated with the <span class="hlt">vapors</span> of an organic solvent (commonly n-butanol). The blend of aerosol particles and organic <span class="hlt">vapor</span> is then cooled so that conditions of <span class="hlt">vapor</span> supersaturation are achieved. In principle, the <span class="hlt">vapor</span> then condenses onto the dry particles, growing each particle (ideally) from as small as a few nanometers in diameter into a droplet with a diameter up to about 10 μm. The grown droplets are then passed through a beam of light, and the light scattered by the droplets is detected and used as the detector response. This growth and detection step is generally carried out using commercial continuous-flow condensation nucleus counters. In the present research, the possibility of using other fluids than the commonly used n-butanol is investigated. The Kelvin equation and the Nucleation theorem [Anisimov et al. (1978)] are used to evaluate a range of fluids for efficacy of growing small particles by condensation nucleation. Using the available experimental data on <span class="hlt">vapor</span> nucleation, the correlations of Kelvin diameters (the critical embryo sizes) and the bulk surface tension with dielectric constants of working liquids are found. A simple method for choosing the most efficient fluid, within a class of fluids, for growth of small particles is suggested.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20000032224&hterms=jayakumar&qs=N%3D0%26Ntk%3DAuthor-Name%26Ntx%3Dmode%2Bmatchall%26Ntt%3Djayakumar','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20000032224&hterms=jayakumar&qs=N%3D0%26Ntk%3DAuthor-Name%26Ntx%3Dmode%2Bmatchall%26Ntt%3Djayakumar"><span>Analysis of <span class="hlt">Phase</span> <span class="hlt">Separation</span> in Czochralski Grown Single Crystal Ilmenite</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Wilkins, R.; Powell, Kirk St. A.; Loregnard, Kieron R.; Lin, Sy-Chyi; Muthusami, Jayakumar; Zhou, Feng; Pandey, R. K.; Brown, Geoff; Hawley, M. E.</p> <p>1998-01-01</p> <p>Ilmenite (FeTiOs) is a wide bandgap semiconductor with an energy gap of 2.58 eV. Ilmenite has properties suited for radiation tolerant applications, as well as a variety of other electronic applications. Single crystal ilmenite has been grown from the melt using the Czochralski method. Growth conditions have a profound effect on the microstructure of the samples. Here we present data from a variety of analytical techniques which indicate that some grown crystals exhibit distinct <span class="hlt">phase</span> <span class="hlt">separation</span> during growth. This <span class="hlt">phase</span> <span class="hlt">separation</span> is apparent for both post-growth annealed and unannealed samples. Under optical microscopy, there appear two distinct areas forming a matrix with an array of dots on order of 5 pm diameter. While appearing bright in the optical micrograph, atomic force microscope (AFM) shows the dots to be shallow pits on the surface. Magnetic force microscope (MFM) shows the dots to be magnetic. <span class="hlt">Phase</span> identification via electron microprobe analysis (EMPA) indicates two major <span class="hlt">phases</span> in the unannealed samples and four in the annealed samples, where the dots appear to be almost pure iron. This is consistent with micrographs taken with a scanning probe microscope used in the magnetic force mode. Samples that do not exhibit the <span class="hlt">phase</span> <span class="hlt">separation</span> have little or no discernible magnetic structure detectable by the MFM.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015NatSR...516335W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015NatSR...516335W"><span>Liquid-liquid <span class="hlt">phase</span> <span class="hlt">separation</span> of freely falling undercooled ternary Fe-Cu-Sn alloy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, W. L.; Wu, Y. H.; Li, L. H.; Zhai, W.; Zhang, X. M.; Wei, B.</p> <p>2015-11-01</p> <p>The active modulation and control of the liquid <span class="hlt">phase</span> <span class="hlt">separation</span> for high-temperature metallic systems are still challenging the development of advanced immiscible alloys. Here we present an attempt to manipulate the dynamic process of liquid-liquid <span class="hlt">phase</span> <span class="hlt">separation</span> for ternary Fe47.5Cu47.5Sn5 alloy. It was firstly dispersed into numerous droplets with 66 ~ 810 μm diameters and then highly undercooled and rapidly solidified under the containerless microgravity condition inside drop tube. 3-D <span class="hlt">phase</span> field simulation was performed to explore the kinetic evolution of liquid <span class="hlt">phase</span> <span class="hlt">separation</span>. Through regulating the combined effects of undercooling level, <span class="hlt">phase</span> <span class="hlt">separation</span> time and Marangoni migration, three types of <span class="hlt">separation</span> patterns were yielded: monotectic cell, core shell and dispersive structures. The two-layer core-shell morphology proved to be the most stable <span class="hlt">separation</span> configuration owing to its lowest chemical potential. Whereas the monotectic cell and dispersive microstructures were both thermodynamically metastable transition states because of their highly active energy. The Sn solute partition profiles of Fe-rich core and Cu-rich shell in core-shell structures varied only slightly with cooling rate.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/4343403','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/biblio/4343403"><span><span class="hlt">VAPOR</span> SHIELD FOR INDUCTION FURNACE</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Reese, S.L.; Samoriga, S.A.</p> <p>1958-03-11</p> <p>This patent relates to a water-cooled <span class="hlt">vapor</span> shield for an inductlon furnace that will condense metallic <span class="hlt">vapors</span> arising from the crucible and thus prevent their condensation on or near the induction coils, thereby eliminating possible corrosion or shorting out of the coils. This is accomplished by placing, about the top, of the crucible a disk, apron, and cooling jacket that <span class="hlt">separates</span> the area of the coils from the interior of the cruclbIe and provides a cooled surface upon whlch the <span class="hlt">vapors</span> may condense.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4053977','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4053977"><span><span class="hlt">Separating</span> homeologs by <span class="hlt">phasing</span> in the tetraploid wheat transcriptome</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p></p> <p>2013-01-01</p> <p>Background The high level of identity among duplicated homoeologous genomes in tetraploid pasta wheat presents substantial challenges for de novo transcriptome assembly. To solve this problem, we develop a specialized bioinformatics workflow that optimizes transcriptome assembly and <span class="hlt">separation</span> of merged homoeologs. To evaluate our strategy, we sequence and assemble the transcriptome of one of the diploid ancestors of pasta wheat, and compare both assemblies with a benchmark set of 13,472 full-length, non-redundant bread wheat cDNAs. Results A total of 489 million 100 bp paired-end reads from tetraploid wheat assemble in 140,118 contigs, including 96% of the benchmark cDNAs. We used a comparative genomics approach to annotate 66,633 open reading frames. The multiple k-mer assembly strategy increases the proportion of cDNAs assembled full-length in a single contig by 22% relative to the best single k-mer size. Homoeologs are <span class="hlt">separated</span> using a post-assembly pipeline that includes polymorphism identification, <span class="hlt">phasing</span> of SNPs, read sorting, and re-assembly of <span class="hlt">phased</span> reads. Using a reference set of genes, we determine that 98.7% of SNPs analyzed are correctly <span class="hlt">separated</span> by <span class="hlt">phasing</span>. Conclusions Our study shows that de novo transcriptome assembly of tetraploid wheat benefit from multiple k-mer assembly strategies more than diploid wheat. Our results also demonstrate that <span class="hlt">phasing</span> approaches originally designed for heterozygous diploid organisms can be used to <span class="hlt">separate</span> the close homoeologous genomes of tetraploid wheat. The predicted tetraploid wheat proteome and gene models provide a valuable tool for the wheat research community and for those interested in comparative genomic studies. PMID:23800085</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19880041353&hterms=Liquid+refractive+index&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DLiquid%2Brefractive%2Bindex','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19880041353&hterms=Liquid+refractive+index&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DLiquid%2Brefractive%2Bindex"><span><span class="hlt">Phase</span> <span class="hlt">separation</span> kinetics in immiscible liquids</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Ng, Lee H.; Sadoway, Donald R.</p> <p>1987-01-01</p> <p>The kinetics of <span class="hlt">phase</span> <span class="hlt">separation</span> in the succinonitrile-water system are being investigated. Experiments involve initial physical mixing of the two immiscible liquids at a temperature above the consolute, decreasing the temperature into the miscibility gap, followed by iamging of the resultant microstructure as it evolves with time. Refractive index differences allow documentation of the changing microstructures by noninvasive optical techniques without the need to quench the liquid structures for analysis.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19890004302','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19890004302"><span><span class="hlt">Phase</span> <span class="hlt">separation</span> kinetics in immiscible liquids</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Sadoway, D. R.</p> <p>1986-01-01</p> <p>The kinetics of <span class="hlt">phase</span> <span class="hlt">separation</span> in the succinonitrile-water system are being investigated. Experiments involve initial physical mixing of the two immiscible liquids at a temperature above the consolute, decreasing the temperature into the miscibility gap, followed by imaging of the resultant microstructure as it evolves with time. Refractive index differences allow documentation of the changing microstructures by noninvasive optical techniques without the need to quench the liquid structures for analysis.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20382068','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20382068"><span>Infrared spectroscopic studies of the conformation in ethyl alpha-haloacetates in the <span class="hlt">vapor</span>, liquid and solid <span class="hlt">phases</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jassem, Naserallah A; El-Bermani, Muhsin F</p> <p>2010-07-01</p> <p>Infrared spectra of ethyl alpha-fluoroacetate, ethyl alpha-chloroacetate, ethyl alpha-bromoacetate and ethyl alpha-iodoacetate have been measured in the solid, liquid and <span class="hlt">vapor</span> <span class="hlt">phases</span> in the region 4000-200 cm(-1). Vibrational frequency assignment of the observed bands to the appropriate modes of vibration was made. Calculations at DFT B3LYP/6-311+G** level, Job: conformer distribution, using Spartan program '08, release 132 was made to determine which conformers exist in which molecule. The results indicated that the first compound exists as an equilibrium mixture of cis and trans conformers and the other three compounds exist as equilibrium mixtures of cis and gauche conformers. Enthalpy differences between the conformers have been determined experimentally for each compound and for every <span class="hlt">phase</span>. The values indicated that the trans of the first compound is more stable in the <span class="hlt">vapor</span> <span class="hlt">phase</span>, while the cis is the more stable in both the liquid and solid <span class="hlt">phases</span>. In the other three compounds the gauche is more stable in the <span class="hlt">vapor</span> and liquid <span class="hlt">phases</span>, while the cis conformer is the more stable in the solid <span class="hlt">phase</span> for each of the second and third compound, except for ethyl alpha-iodoacetate, the gauche conformer is the more stable over the three <span class="hlt">phases</span>. Molar energy of activation Ea and the pseudo-thermodynamic parameters of activation DeltaH(double dagger), DeltaS(double dagger) and DeltaG(double dagger) were determined in the solid <span class="hlt">phase</span> by applying Arrhenius equation; using bands arising from single conformers. The respective E(a) values of these compounds are 5.1+/-0.4, 6.7+/-0.1, 7.5+/-1.3 and 12.0+/-0.6 kJ mol(-1). Potential energy surface calculations were made at two levels; for ethyl alpha-fluoroacetate and ethyl alpha-chloroacetate; the calculations were established at DFT B3LYP/6-311+G** level and for ethyl alpha-bromoacetate and ethyl alpha-iodoacetate at DFT B3LYP/6-311G* level. The results showed no potential energy minimum exists for the gauche conformer in</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/864487','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/864487"><span>Electrolyte <span class="hlt">vapor</span> condenser</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Sederquist, Richard A.; Szydlowski, Donald F.; Sawyer, Richard D.</p> <p>1983-01-01</p> <p>A system is disclosed for removing electrolyte from a fuel cell gas stream. The gas stream containing electrolyte <span class="hlt">vapor</span> is supercooled utilizing conventional heat exchangers and the thus supercooled gas stream is passed over high surface area passive condensers. The condensed electrolyte is then drained from the condenser and the remainder of the gas stream passed on. The system is particularly useful for electrolytes such as phosphoric acid and molten carbonate, but can be used for other electrolyte cells and simple <span class="hlt">vapor</span> <span class="hlt">separation</span> as well.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/7152985','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/biblio/7152985"><span>Electrolyte <span class="hlt">vapor</span> condenser</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Sederquist, R.A.; Szydlowski, D.F.; Sawyer, R.D.</p> <p>1983-02-08</p> <p>A system is disclosed for removing electrolyte from a fuel cell gas stream. The gas stream containing electrolyte <span class="hlt">vapor</span> is supercooled utilizing conventional heat exchangers and the thus supercooled gas stream is passed over high surface area passive condensers. The condensed electrolyte is then drained from the condenser and the remainder of the gas stream passed on. The system is particularly useful for electrolytes such as phosphoric acid and molten carbonate, but can be used for other electrolyte cells and simple <span class="hlt">vapor</span> <span class="hlt">separation</span> as well. 3 figs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhyA..489...65N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhyA..489...65N"><span>Numerical investigation of the pseudopotential lattice Boltzmann modeling of liquid-<span class="hlt">vapor</span> for multi-<span class="hlt">phase</span> flows</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nemati, Maedeh; Shateri Najaf Abady, Ali Reza; Toghraie, Davood; Karimipour, Arash</p> <p>2018-01-01</p> <p>The incorporation of different equations of state into single-component multiphase lattice Boltzmann model is considered in this paper. The original pseudopotential model is first detailed, and several cubic equations of state, the Redlich-Kwong, Redlich-Kwong-Soave, and Peng-Robinson are then incorporated into the lattice Boltzmann model. A comparison of the numerical simulation achievements on the basis of density ratios and spurious currents is used for presentation of the details of <span class="hlt">phase</span> <span class="hlt">separation</span> in these non-ideal single-component systems. The paper demonstrates that the scheme for the inter-particle interaction force term as well as the force term incorporation method matters to achieve more accurate and stable results. The velocity shifting method is demonstrated as the force term incorporation method, among many, with accuracy and stability results. Kupershtokh scheme also makes it possible to achieve large density ratio (up to 104) and to reproduce the coexistence curve with high accuracy. Significant reduction of the spurious currents at <span class="hlt">vapor</span>-liquid interface is another observation. High-density ratio and spurious current reduction resulted from the Redlich-Kwong-Soave and Peng-Robinson EOSs, in higher accordance with the Maxwell construction results.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27775180','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27775180"><span>Robust water fat <span class="hlt">separated</span> dual-echo MRI by <span class="hlt">phase</span>-sensitive reconstruction.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Romu, Thobias; Dahlström, Nils; Leinhard, Olof Dahlqvist; Borga, Magnus</p> <p>2017-09-01</p> <p>The purpose of this work was to develop and evaluate a robust water-fat <span class="hlt">separation</span> method for T1-weighted symmetric two-point Dixon data. A method for water-fat <span class="hlt">separation</span> by <span class="hlt">phase</span> unwrapping of the opposite-<span class="hlt">phase</span> images by <span class="hlt">phase</span>-sensitive reconstruction (PSR) is introduced. PSR consists of three steps; (1), identification of clusters of tissue voxels; (2), unwrapping of the <span class="hlt">phase</span> in each cluster by solving Poisson's equation; and (3), finding the correct sign of each unwrapped opposite-<span class="hlt">phase</span> cluster, so that the water-fat images are assigned the correct identities. Robustness was evaluated by counting the number of water-fat swap artifacts in a total of 733 image volumes. The method was also compared to commercial software. In the water-fat <span class="hlt">separated</span> image volumes, the PSR method failed to unwrap the <span class="hlt">phase</span> of one cluster and misclassified 10. One swap was observed in areas affected by motion and was constricted to the affected area. Twenty swaps were observed surrounding susceptibility artifacts, none of which spread outside the artifact affected regions. The PSR method had fewer swaps when compared to commercial software. The PSR method can robustly produce water-fat <span class="hlt">separated</span> whole-body images based on symmetric two-echo spoiled gradient echo images, under both ideal conditions and in the presence of common artifacts. Magn Reson Med 78:1208-1216, 2017. © 2016 International Society for Magnetic Resonance in Medicine. © 2016 International Society for Magnetic Resonance in Medicine.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/9192111','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/9192111"><span>Cyclohexylamine additives for enhanced peptide <span class="hlt">separations</span> in reversed <span class="hlt">phase</span> liquid chromatography.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cole, S R; Dorsey, J G</p> <p>1997-01-01</p> <p>While the choice of stationary <span class="hlt">phase</span>, organic modifier, and gradient strength can have significant effects on biomolecule <span class="hlt">separations</span>, mobile <span class="hlt">phase</span> additives can also have a significant effect on the chromatographic selectivity, recovery, efficiency and resolution. Given the importance of stationary <span class="hlt">phase</span> coverage, the beneficial, silanol-masking properties of amines, and the potential for selectivity modification through ion-pair interactions, cyclohexylamine was examined as a mobile <span class="hlt">phase</span> additive and compared with triethylamine and trifluoroacetic acid. Greatly improved <span class="hlt">separation</span> was possible when cyclohexylamine was used as compared with phosphate buffer, and cyclohexylamine did not require purification before use, while triethylamine required distillation before 'clean' chromatograms were obtained.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_19");'>19</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li class="active"><span>21</span></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_21 --> <div id="page_22" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li class="active"><span>22</span></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="421"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/12075966','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12075966"><span>In vivo droplet <span class="hlt">vaporization</span> for occlusion therapy and <span class="hlt">phase</span> aberration correction.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kripfgans, Oliver D; Fowlkes, J Brian; Woydt, Michael; Eldevik, Odd P; Carson, Paul L</p> <p>2002-06-01</p> <p>The objective was to determine whether a transpulmonary droplet emulsion (90%, <6 microm diameter) could be used to form large gas bubbles (>30 microm) temporarily in vivo. Such bubbles could occlude a targeted capillary bed when used in a large number density. Alternatively, for a very sparse population of droplets, the resulting gas bubbles could serve as point beacons for <span class="hlt">phase</span> aberration corrections in ultrasonic imaging. Gas bubbles can be made in vivo by acoustic droplet <span class="hlt">vaporization</span> (ADV) of injected, superheated, dodecafluoropentane droplets. Droplets <span class="hlt">vaporize</span> in an acoustic field whose peak rarefactional pressure exceeds a well-defined threshold. In this new work, it has been found that intraarterial and intravenous injections can be used to introduce the emulsion into the blood stream for subsequent ADV (B- and M-mode on a clinical scanner) in situ. Intravenous administration results in a lower gas bubble yield, possibly because of filtering in the lung, dilution in the blood volume, or other circulatory effects. Results show that for occlusion purposes, a reduction in regional blood flow of 34% can be achieved. Individual point beacons with a +24 dB backscatter amplitude relative to white matter were created by intravenous injection and ADV.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4609299','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4609299"><span>Formation and Maturation of <span class="hlt">Phase</span> <span class="hlt">Separated</span> Liquid Droplets by RNA Binding Proteins</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Lin, Yuan; Protter, David S. W.; Rosen, Michael K.; Parker, Roy</p> <p>2015-01-01</p> <p>Eukaryotic cells possess numerous dynamic membrane-less organelles, RNP granules, enriched in RNA and RNA binding proteins containing disordered regions. We demonstrate that the disordered regions of key RNP granule components, and the full-length granule protein hnRNPA1, can <span class="hlt">phase</span> <span class="hlt">separate</span> in vitro, producing dynamic liquid droplets. <span class="hlt">Phase</span> <span class="hlt">separation</span> is promoted by low salt concentrations or RNA. Over time, the droplets mature to more stable states, as assessed by slowed fluorescence recovery after photobleaching and resistance to salt. Maturation often coincides with formation of fibrous structures. Different disordered domains can co-assemble into <span class="hlt">phase-separated</span> droplets. These biophysical properties demonstrate a plausible mechanism by which interactions between disordered regions, coupled with RNA binding, could contribute to RNP granule assembly in vivo through promoting <span class="hlt">phase</span> <span class="hlt">separation</span>. Progression from dynamic liquids to stable fibers may be regulated to produce cellular structures with diverse physiochemical properties and functions. Misregulation could contribute to diseases involving aberrant RNA granules. PMID:26412307</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1356883-upright-inverted-single-junction-gaas-solar-cells-grown-hydride-vapor-phase-epitaxy','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1356883-upright-inverted-single-junction-gaas-solar-cells-grown-hydride-vapor-phase-epitaxy"><span>Upright and Inverted Single-Junction GaAs Solar Cells Grown by Hydride <span class="hlt">Vapor</span> <span class="hlt">Phase</span> Epitaxy</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Simon, John; Schulte, Kevin L.; Jain, Nikhil; ...</p> <p>2016-10-19</p> <p>Hydride <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy (HVPE) is a low-cost alternative to conventional metal-organic <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy (MOVPE) growth of III-V solar cells. In this work, we show continued improvement of the performance of HVPE-grown single-junction GaAs solar cells. We show over an order of magnitude improvement in the interface recombination velocity between GaAs and GaInP layers through the elimination of growth interrupts, leading to increased short-circuit current density and open-circuit voltage compared with cells with interrupts. One-sun conversion efficiencies as high as 20.6% were achieved with this improved growth process. Solar cells grown in an inverted configuration that were removed frommore » the substrate showed nearly identical performance to on-wafer cells, demonstrating the viability of HVPE to be used together with conventional wafer reuse techniques for further cost reduction. As a result, these devices utilized multiple heterointerfaces, showing the potential of HVPE for the growth of complex and high-quality III-V devices.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/380985-vapor-phase-particulate-associated-pesticides-pcb-concentrations-eastern-north-dakota-air-samples','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/380985-vapor-phase-particulate-associated-pesticides-pcb-concentrations-eastern-north-dakota-air-samples"><span><span class="hlt">Vapor-phase</span> and particulate-associated pesticides and PCB concentrations in eastern North Dakota air samples</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Hawthorne, S.B.; Miller, D.J.; Louie, P.K.K.</p> <p>1996-05-01</p> <p><span class="hlt">Vapor-phase</span> and suspended particulate (<50 {mu}m) samples were collected on polyurethane foam (PUF) and quartz fiber filters in rural North Dakota to determine the air concentrations of pesticides in an area where agriculture is a primary source of semivolatile pollutants. Samples were collected at two sites from 1992 to 1994 that were at least 0.4 km from the nearest farmed fields and known application of pesticides, and analyzed for 22 different organochlorine, triazine, and acid herbicide pesticides. Fourteen pesticides were found above the detection limits (typically <1 pg/m{sup 3}). Concentrations of polychlorinated biphenyl (PCB) congeners were much lower (<50 pg/m{supmore » 3} in all cases) than many of the pesticides. These results demonstrate that pesticides are among the most prevalent chlorinated semivolatile pollutants present in rural North Dakota, that significant transport of pesticides occurs both in the <span class="hlt">vapor-phase</span> and on suspended particulate matter, and that blown soil may be a significant mechanism for introducing pesticides into surface and ground waters. 32 refs., 2 figs., 4 tabs.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1121200-local-bias-induced-phase-transitions','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1121200-local-bias-induced-phase-transitions"><span>Local bias-<span class="hlt">induced</span> <span class="hlt">phase</span> transitions</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Seal, Katyayani; Baddorf, Arthur P.; Jesse, Stephen; ...</p> <p>2008-11-27</p> <p>Electrical bias-<span class="hlt">induced</span> <span class="hlt">phase</span> transitions underpin a wide range of applications from data storage to energy generation and conversion. The mechanisms behind these transitions are often quite complex and in many cases are extremely sensitive to local defects that act as centers for local transformations or pinning. Furthermore, using ferroelectrics as an example, we review methods for probing bias-<span class="hlt">induced</span> <span class="hlt">phase</span> transitions and discuss the current limitations and challenges for extending the methods to field-<span class="hlt">induced</span> <span class="hlt">phase</span> transitions and electrochemical reactions in energy storage, biological and molecular systems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/210483-biofiltration-innovative-approach-vapor-phase-treatment-silvex-hazardous-waste-site-florida','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/210483-biofiltration-innovative-approach-vapor-phase-treatment-silvex-hazardous-waste-site-florida"><span>Biofiltration - an innovative approach to <span class="hlt">vapor</span> <span class="hlt">phase</span> treatment at the Silvex hazardous waste site in Florida</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Hartsfield, B.</p> <p>1995-12-31</p> <p>Biofiltration is an emerging technology that is being used for <span class="hlt">vapor</span> <span class="hlt">phase</span> treatment at the Silvex hazardous waste site. Biofiltration works by directing the off-gas from the groundwater treatment system through a bed of soil, compost or other medium that supports the growth of bacteria. Contaminants are absorbed into the water present in the medium, and are subsequently degraded by the microorganisms. The biofiltration system at the Silvex hazardous waste site has been effective in removing contaminants from the off-gas. The biofiltration system has also been effective in minimizing the odor problem resulting from mercaptans in the off-gas. Biofiltration hasmore » been used for many years at wastewater and industrial plants to control odor and remove organic contaminants. This technology has only recently been used for hazardous waste site cleanups. The hazardous waste literature is now listing biofiltration as a <span class="hlt">vapor</span> <span class="hlt">phase</span> treatment technology, along with carbon, thermal oxidation and others.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017BVol...79...74I','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017BVol...79...74I"><span><span class="hlt">Vapor-phase</span> cristobalite as a durable indicator of magmatic pore structure and halogen degassing: an example from White Island volcano (New Zealand)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ian Schipper, C.; Mandon, Céline; Maksimenko, Anton; Castro, Jonathan M.; Conway, Chris E.; Hauer, Peter; Kirilova, Martina; Kilgour, Geoff</p> <p>2017-10-01</p> <p>Vesicles in volcanic rocks are physical records of magmatic degassing; however, the interpretation of their textures is complicated by resorption, coalescence, and collapse. We discuss the textural significance of vesicle-hosted <span class="hlt">vapor-phase</span> cristobalite (high-T, low-P SiO2 polymorph), and its utility as a complement to textural assessments of magmatic degassing, using a representative dacite bomb erupted from White Island volcano (New Zealand) in 1999. Imaging in 2D (SEM) and 3D (CT) shows the bomb to have 56% bulk porosity, almost all of which is connected ( 99%) and devoid of SiO2 <span class="hlt">phases</span>. The remaining ( 1%) of porosity is in isolated, sub-spherical vesicles that have corroded walls and contain small (< 30 μm across) prismatic <span class="hlt">vapor-phase</span> cristobalite crystals (98.4 ± 0.4 wt.% SiO2 with diagnostic laser Raman spectra). Halogen degassing models show <span class="hlt">vapor-phase</span> cristobalite to be indicative of closed-system chlorine and fluorine partitioning into H2O-rich fluid in isolated pores. At White Island, this occurred during shallow (< 100s of meters) ascent and extensive ( 50%) groundmass crystallization associated with slow cooling in a volcanic plug. Pristine textures in this White Island bomb demonstrate the link between pore isolation and <span class="hlt">vapor-phase</span> cristobalite deposition. We suggest that because these crystals have higher preservation potential than the bubbles in which they form, they can serve as durable, qualitative textural indicators of halogen degassing and pre-quench bubble morphologies in slowly cooled volcanic rocks (e.g., lava flows and domes), even where emplacement mechanisms have overprinted original bubble textures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/12541946','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/12541946"><span>[Influence of mobile <span class="hlt">phase</span> composition on chiral <span class="hlt">separation</span> of organic selenium racemates].</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Han, Xiao-qian; Qi, Bang-feng; Dun, Hui-juan; Zhu, Xin-yi; Na, Peng-jun; Jiang, Sheng-xiang; Chen, Li-ren</p> <p>2002-05-01</p> <p>The chiral <span class="hlt">separation</span> of some chiral compounds with similar structure on the cellulose tris (3,5-dimethylphenylcarbamate) chiral stationary <span class="hlt">phase</span> prepared by us was obtained. Ternary mobile <span class="hlt">phases</span> influencing chiral recognition were investigated. A mode of interaction between the structural character of samples and chiral stationary <span class="hlt">phase</span> is discussed. The results indicated that the retention and chiral <span class="hlt">separation</span> of the analytes had a bigger change with minute addition of alcohols or acetonitrile as modifier in n-hexane/2-propanol (80/20, volume ratio) binary mobile <span class="hlt">phase</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005SPIE.5864..161B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005SPIE.5864..161B"><span><span class="hlt">Vapor-phase</span> infrared spectroscopy on solid organic compounds with a pulsed resonant photoacoustic detection scheme</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bartlome, Richard; Fischer, Cornelia; Sigrist, Markus W.</p> <p>2005-08-01</p> <p>There is a great need for a low cost and sensitive method to measure infrared spectra of solid organic compounds in the gas <span class="hlt">phase</span>. To record such spectra, we propose an optical parametric generator-based photoacoustic spectrometer, which emits in the mid-infrared fingerprint region between 3 and 4 microns. In this system, the sample is heated in a vessel before entering a home built photoacoustic cell, where the gaseous molecules are excited by a tunable laser source with a frequency repetition rate that matches the first longitudinal resonance frequency of the photocaoustic cell. In a first <span class="hlt">phase</span>, we have focused on low-melting point stimulants such as Nikethamide, Mephentermine sulfate, Methylephedrine, Ephedrine and Pseudoephedrine. The <span class="hlt">vapor-phase</span> spectra of these doping substances were measured between 2800 and 3100 cm-1, where fundamental C-H stretching vibrations take place. Our spectra show notable differences with commercially available condensed <span class="hlt">phase</span> spectra. Our scheme enables to measure very low <span class="hlt">vapor</span> pressures of low-melting point (<160 °C) solid organic compounds. Furthermore, the optical resolution of 8 cm-1 is good enough to distinguish closely related chemical structures such as the Ephedra alkaloids Ephedrine and Methylephedrine, but doesn't allow to differentiate diastereoisomeric pairs such as Ephedrine and Pseudoephedrine, two important neurotransmitters which reveal different biological activities. Therefore, higher resolution and a system capable of measuring organic compounds with higher melting points are required.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17949216','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17949216"><span>Integral equation theory study on the <span class="hlt">phase</span> <span class="hlt">separation</span> in star polymer nanocomposite melts.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhao, Lei; Li, Yi-Gui; Zhong, Chongli</p> <p>2007-10-21</p> <p>The polymer reference interaction site model theory is used to investigate <span class="hlt">phase</span> <span class="hlt">separation</span> in star polymer nanocomposite melts. Two kinds of spinodal curves were obtained: classic fluid <span class="hlt">phase</span> boundary for relatively low nanoparticle-monomer attraction strength and network <span class="hlt">phase</span> boundary for relatively high nanoparticle-monomer attraction strength. The network <span class="hlt">phase</span> boundaries are much more sensitive with nanoparticle-monomer attraction strength than the fluid <span class="hlt">phase</span> boundaries. The interference among the arm number, arm length, and nanoparticle-monomer attraction strength was systematically investigated. When the arm lengths are short, the network <span class="hlt">phase</span> boundary shows a marked shift toward less miscibility with increasing arm number. When the arm lengths are long enough, the network <span class="hlt">phase</span> boundaries show opposite trends. There exists a crossover arm number value for star polymer nanocomposite melts, below which the network <span class="hlt">phase</span> <span class="hlt">separation</span> is consistent with that of chain polymer nanocomposite melts. However, the network <span class="hlt">phase</span> <span class="hlt">separation</span> shows qualitatively different behaviors when the arm number is larger than this value.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29582482','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29582482"><span>Metal-Organic Frameworks for <span class="hlt">Separation</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhao, Xiang; Wang, Yanxiang; Li, Dong-Sheng; Bu, Xianhui; Feng, Pingyun</p> <p>2018-03-27</p> <p><span class="hlt">Separation</span> is an important industrial step with critical roles in the chemical, petrochemical, pharmaceutical, and nuclear industries, as well as in many other fields. Although much progress has been made, the development of better <span class="hlt">separation</span> technologies, especially through the discovery of high-performance <span class="hlt">separation</span> materials, continues to attract increasing interest due to concerns over factors such as efficiency, health and environmental impacts, and the cost of existing methods. Metal-organic frameworks (MOFs), a rapidly expanding family of crystalline porous materials, have shown great promise to address various <span class="hlt">separation</span> challenges due to their well-defined pore size and unprecedented tunability in both composition and pore geometry. In the past decade, extensive research is performed on applications of MOF materials, including <span class="hlt">separation</span> and capture of many gases and <span class="hlt">vapors</span>, and liquid-<span class="hlt">phase</span> <span class="hlt">separation</span> involving both liquid mixtures and solutions. MOFs also bring new opportunities in enantioselective <span class="hlt">separation</span> and are amenable to morphological control such as fabrication of membranes for enhanced <span class="hlt">separation</span> outcomes. Here, some of the latest progress in the applications of MOFs for several key <span class="hlt">separation</span> issues, with emphasis on newly synthesized MOF materials and the impact of their compositional and structural features on <span class="hlt">separation</span> properties, are reviewed and highlighted. © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1068840','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/1068840"><span><span class="hlt">Phase-separated</span>, epitaxial composite cap layers for electronic device applications and method of making the same</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Aytug, Tolga [Knoxville, TN; Paranthaman, Mariappan Parans [Knoxville, TN; Polat, Ozgur [Knoxville, TN</p> <p>2012-07-17</p> <p>An electronic component that includes a substrate and a <span class="hlt">phase-separated</span> layer supported on the substrate and a method of forming the same are disclosed. The <span class="hlt">phase-separated</span> layer includes a first <span class="hlt">phase</span> comprising lanthanum manganate (LMO) and a second <span class="hlt">phase</span> selected from a metal oxide (MO), metal nitride (MN), a metal (Me), and combinations thereof. The <span class="hlt">phase-separated</span> material can be an epitaxial layer and an upper surface of the <span class="hlt">phase-separated</span> layer can include interfaces between the first <span class="hlt">phase</span> and the second <span class="hlt">phase</span>. The <span class="hlt">phase-separated</span> layer can be supported on a buffer layer comprising a composition selected from the group consisting of IBAD MgO, LMO/IBAD-MgO, homoepi-IBAD MgO and LMO/homoepi-MgO. The electronic component can also include an electronically active layer supported on the <span class="hlt">phase-separated</span> layer. The electronically active layer can be a superconducting material, a ferroelectric material, a multiferroic material, a magnetic material, a photovoltaic material, an electrical storage material, and a semiconductor material.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eric.ed.gov/?q=Separation+AND+pigments&id=EJ541774','ERIC'); return false;" href="https://eric.ed.gov/?q=Separation+AND+pigments&id=EJ541774"><span><span class="hlt">Separation</span> of Chloroplast Pigments Using Reverse <span class="hlt">Phase</span> Chromatography.</span></a></p> <p><a target="_blank" href="http://www.eric.ed.gov/ERICWebPortal/search/extended.jsp?_pageLabel=advanced">ERIC Educational Resources Information Center</a></p> <p>Reese, R. Neil</p> <p>1997-01-01</p> <p>Presents a protocol that uses reverse <span class="hlt">phase</span> chromatography for the <span class="hlt">separation</span> of chloroplast pigments. Provides a simple and relatively safe procedure for use in teaching laboratories. Discusses pigment extraction, chromatography, results, and advantages of the process. (JRH)</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19920065606&hterms=refraction&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Drefraction','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19920065606&hterms=refraction&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Drefraction"><span>The slant path atmospheric refraction calibrator - An instrument to measure the microwave propagation delays <span class="hlt">induced</span> by atmospheric water <span class="hlt">vapor</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Walter, Steven J.; Bender, Peter L.</p> <p>1992-01-01</p> <p>The water <span class="hlt">vapor-induced</span> propagation delay experienced by a radio signal traversing the atmosphere is characterized by the Slant Path Atmospheric Refraction Calibrator (SPARC), which measures the difference in the travel times between an optical and a microwave signal propagating along the same atmospheric path with an accuracy of 15 picosec or better. Attention is given to the theoretical and experimental issues involved in measuring the delay <span class="hlt">induced</span> by water <span class="hlt">vapor</span>; SPARC measurements conducted along a 13.35-km ground-based path are presented, illustrating the instrument's stability, precision, and accuracy.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23669785','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23669785"><span>Effects of water <span class="hlt">vapor</span> on flue gas conditioning in the electric fields with corona discharge.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Liqiang, Qi; Yajuan, Zhang</p> <p>2013-07-15</p> <p>Sulfur dioxide (SO2) removal via pulsed discharge nonthermal plasma in the absence of ammonia was investigated to determine how electrostatic precipitators (ESPs) can effectively collect particulate matter less than 2.5μm in diameter from flue gas. SO2 removal increased as water <span class="hlt">vapor</span> concentration increased. In a wet-type plasma reactor, directing a gas-<span class="hlt">phase</span> discharge plasma toward the water film surface significantly enhanced the liquid-<span class="hlt">phase</span> oxidation of HSO3(-) to SO4(2-). Comparisons of various absorbents revealed that the hydroxyl radical is a key factor in plasma-<span class="hlt">induced</span> liquid-<span class="hlt">phase</span> reactions. The resistivity, size distribution, and cohesive force of fly ash at different water <span class="hlt">vapor</span> contents were measured using a Bahco centrifuge, which is a dust electrical resistivity test instrument, as well as a cohesive force test apparatus developed by the researchers. When water <span class="hlt">vapor</span> content increased by 5%, fly ash resistivity in flue gas decreased by approximately two orders of magnitude, adhesive force and size increased, and specific surface area decreased. Therefore, ESP efficiency increased. Copyright © 2013 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.dtic.mil/docs/citations/ADA163430','DTIC-ST'); return false;" href="http://www.dtic.mil/docs/citations/ADA163430"><span>The Inhibition of <span class="hlt">Vapor-Phase</span> Corrosion. A Review</span></a></p> <p><a target="_blank" href="http://www.dtic.mil/">DTIC Science & Technology</a></p> <p></p> <p>1985-10-01</p> <p><span class="hlt">vaporization</span> of the inhibitor in a nondissociated molecular form, followed by hydrolysis on the surface of the metal. The products of hydrolysis may...Patent No. 600328) was assigned to Shell in 1945 . Some time ago, camphor was used to protect military materials made of ferrous metals. Naphthalene <span class="hlt">vapor</span>...reduce moisture, they also "reduce corrosion. More importantly, they decompose as they absorb water, and the decomposition products (as illustrated by</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5884133','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5884133"><span>Development of a passive <span class="hlt">phase</span> <span class="hlt">separator</span> for space and earth applications</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Wu, Xiongjun; Loraine, Greg; Hsiao, Chao-Tsung; Chahine, Georges L.</p> <p>2018-01-01</p> <p>The limited amount of liquids and gases that can be carried to space makes it imperative to recycle and reuse these fluids for extended human operations. During recycling processes gas and liquid <span class="hlt">phases</span> are often intermixed. In the absence of gravity, <span class="hlt">separating</span> gases from liquids is challenging due to the absence of buoyancy. This paper describes development of a passive <span class="hlt">phase</span> <span class="hlt">separator</span> that is capable of efficiently and reliably <span class="hlt">separating</span> gas–liquid mixtures of both high and low void fractions in a wide range of flow rates that is applicable to for both space and earth applications. PMID:29628785</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013APS..MARW34005L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013APS..MARW34005L"><span>A Laterally-Mobile Mixed Polymer/Polyelectrolyte Brush Undergoes a Macroscopic <span class="hlt">Phase</span> <span class="hlt">Separation</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lee, Hoyoung; Park, Hae-Woong; Tsouris, Vasilios; Choi, Je; Mustafa, Rafid; Lim, Yunho; Meron, Mati; Lin, Binhua; Won, You-Yeon</p> <p>2013-03-01</p> <p>We studied mixed PEO and PDMAEMA brushes. The question we attempted to answer was: When the chain grafting points are laterally mobile, how will this lateral mobility influence the structure and <span class="hlt">phase</span> behavior of the mixed brush? Two different model mixed PEO/PDMAEMA brush systems were prepared: a mobile mixed brush by spreading a mixture of two diblock copolymers, PEO-PnBA and PDMAEMA-PnBA, onto the air-water interface, and an inseparable mixed brush using a PEO-PnBA-PDMAEMA triblock copolymer having respective brush molecular weights matched to those of the diblock copolymers. These two systems were investigated by surface pressure-area isotherm, X-ray reflectivity and AFM imaging measurements. The results suggest that the mobile mixed brush undergoes a lateral macroscopic <span class="hlt">phase</span> <span class="hlt">separation</span> at high chain grafting densities, whereas the inseparable system is only microscopically <span class="hlt">phase</span> <span class="hlt">separated</span> under comparable brush density conditions. We also conducted an SCF analysis of the <span class="hlt">phase</span> behavior of the mixed brush system. This analysis further supported the experimental findings. The macroscopic <span class="hlt">phase</span> <span class="hlt">separation</span> observed in the mobile system is in contrast to the microphase <span class="hlt">separation</span> behavior commonly observed in two-dimensional laterally-mobile small molecule mixtures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..MARF17011K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..MARF17011K"><span>Hybrid films with <span class="hlt">phase-separated</span> domains: A new class of functional materials</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kang, Minjee; Leal, Cecilia</p> <p></p> <p>The cell membrane is highly compartmentalized over micro-and nano scale. The compartmentalized domains play an important role in regulating the diffusion and distribution of species within and across the membrane. In this work, we introduced nanoscale heterogeneities into lipid films for the purpose of developing nature-mimicking <span class="hlt">phase-separated</span> materials. The mixtures of phospholipids and amphiphilic block copolymers self-assemble into supported 1D multi-bilayers. We observed that in each lamella, mixtures of lipid and polymer <span class="hlt">phase-separate</span> into domains that differ in their composition akin to sub-<span class="hlt">phases</span> in cholesterol-containing lipid bilayers. Interestingly, we found evidence that like-domains are in registry across multilayers, making <span class="hlt">phase</span> <span class="hlt">separation</span> three-dimensional. To exploit such distinctive domain structure for surface-mediated drug delivery, we incorporated pharmaceutical molecules into the films. The drug release study revealed that the presence of domains in hybrid films modifies the diffusion pathways of drugs that become confined within <span class="hlt">phase-separated</span> domains. A comprehensive domain structure coupled with drug diffusion pathways in films will be presented, offering new perspectives in designing a thin-film matrix system for controlled drug delivery. This work was supported by the National Science Foundation under Grant No. DMR-1554435.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27384744','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27384744"><span>Stability and Oil Migration of Oil-in-Water Emulsions Emulsified by <span class="hlt">Phase-Separating</span> Biopolymer Mixtures.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yang, Nan; Mao, Peng; Lv, Ruihe; Zhang, Ke; Fang, Yapeng; Nishinari, Katsuyoshi; Phillips, Glyn O</p> <p>2016-08-01</p> <p>Oil-in-water (O/W) emulsions with varying concentration of oil <span class="hlt">phase</span>, medium-chain triglyceride (MCT), were prepared using <span class="hlt">phase-separating</span> gum arabic (GA)/sugar beet pectin (SBP) mixture as an emulsifier. Stability of the emulsions including emulsion <span class="hlt">phase</span> <span class="hlt">separation</span>, droplet size change, and oil migration were investigated by means of visual observation, droplet size analysis, oil partition analysis, backscattering of light, and interfacial tension measurement. It was found that in the emulsions prepared with 4.0% GA/1.0% SBP, when the concentration of MCT was greater than 2.0%, emulsion <span class="hlt">phase</span> <span class="hlt">separation</span> was not observed and the emulsions were stable with droplet size unchanged during storage. This result proves the emulsification ability of <span class="hlt">phase-separating</span> biopolymer mixtures and their potential usage as emulsifiers to prepare O/W emulsion. However, when the concentration of MCT was equal or less than 2.0%, emulsion <span class="hlt">phase</span> <span class="hlt">separation</span> occurred after preparation resulting in an upper SBP-rich <span class="hlt">phase</span> and a lower GA-rich <span class="hlt">phase</span>. The droplet size increased in the upper <span class="hlt">phase</span> whereas decreased slightly in the lower <span class="hlt">phase</span> with time, compared to the freshly prepared emulsions. During storage, the oil droplets exhibited a complex migration process: first moving to the SBP-rich <span class="hlt">phase</span>, then to the GA-rich <span class="hlt">phase</span> and finally gathering at the interface between the two <span class="hlt">phases</span>. The mechanisms of the emulsion stability and oil migration in the <span class="hlt">phase-separated</span> emulsions were discussed. © 2016 Institute of Food Technologists®</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_20");'>20</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li class="active"><span>22</span></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_22 --> <div id="page_23" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li class="active"><span>23</span></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="441"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27723115','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27723115"><span><span class="hlt">Phase-Separated</span> Polyaniline/Graphene Composite Electrodes for High-Rate Electrochemical Supercapacitors.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wu, Jifeng; Zhang, Qin'e; Zhou, An'an; Huang, Zhifeng; Bai, Hua; Li, Lei</p> <p>2016-12-01</p> <p>Polyaniline/graphene hydrogel composites with a macroscopically <span class="hlt">phase-separated</span> structure are prepared. The composites show high specific capacitance and excellent rate performance. Further investigation demonstrates that polyaniline inside the graphene hydrogel has low rate performance, thus a <span class="hlt">phase-separated</span> structure, in which polyaniline is mainly outside the graphene hydrogel matrix, can enhance the rate performance of the composites. © 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27966862','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27966862"><span>Controlling Microstructure-Transport Interplay in Highly <span class="hlt">Phase-Separated</span> Perfluorosulfonated Aromatic Multiblock Ionomers via Molecular Architecture Design.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Nguyen, Huu-Dat; Assumma, Luca; Judeinstein, Patrick; Mercier, Regis; Porcar, Lionel; Jestin, Jacques; Iojoiu, Cristina; Lyonnard, Sandrine</p> <p>2017-01-18</p> <p>Proton-conducting multiblock polysulfones bearing perfluorosulfonic acid side chains were designed to encode nanoscale <span class="hlt">phase-separation</span>, well-defined hydrophilic/hydrophobic interfaces, and optimized transport properties. Herein, we show that the superacid side chains yield highly ordered morphologies that can be tailored by best compromising ion-exchange capacity and block lengths. The obtained microstructures were extensively characterized by small-angle neutron scattering (SANS) over an extended range of hydration. Peculiar swelling behaviors were evidenced at two different scales and attributed to the dilution of locally flat polymer particles. We evidence the direct correlation between the quality of interfaces, the topology and connectivity of ionic nanodomains, the block superstructure long-range organization, and the transport properties. In particular, we found that the proton conductivity linearly depends on the microscopic expansion of both ionic and block domains. These findings indicate that neat nanoscale <span class="hlt">phase-separation</span> and block-<span class="hlt">induced</span> long-range connectivity can be optimized by designing aromatic ionomers with controlled architectures to improve the performances of polymer electrolyte membranes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1999ApJ...525..482M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1999ApJ...525..482M"><span>Evolutionary Calculations of <span class="hlt">Phase</span> <span class="hlt">Separation</span> in Crystallizing White Dwarf Stars</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Montgomery, M. H.; Klumpe, E. W.; Winget, D. E.; Wood, M. A.</p> <p>1999-11-01</p> <p>We present an exploration of the significance of carbon/oxygen <span class="hlt">phase</span> <span class="hlt">separation</span> in white dwarf stars in the context of self-consistent evolutionary calculations. Because <span class="hlt">phase</span> <span class="hlt">separation</span> can potentially increase the calculated ages of the oldest white dwarfs, it can affect the age of the Galactic disk as derived from the downturn in the white dwarf luminosity function. We find that the largest possible increase in ages due to <span class="hlt">phase</span> <span class="hlt">separation</span> is ~1.5 Gyr, with a most likely value of approximately 0.6 Gyr, depending on the parameters of our white dwarf models. The most important factors influencing the size of this delay are the total stellar mass, the initial composition profile, and the <span class="hlt">phase</span> diagram assumed for crystallization. We find a maximum age delay in models with masses of ~0.6 Msolar, which is near the peak in the observed white dwarf mass distribution. In addition, we note that the prescription that we have adopted for the mixing during crystallization provides an upper bound for the efficiency of this process, and hence a maximum for the age delays. More realistic treatments of the mixing process may reduce the size of this effect. We find that varying the opacities (via the metallicity) has little effect on the calculated age delays. In the context of Galactic evolution, age estimates for the oldest Galactic globular clusters range from 11.5 to 16 Gyr and depend on a variety of parameters. In addition, a 4-6 Gyr delay is expected between the formation of the globular clusters and the formation of the Galactic thin disk, while the observed white dwarf luminosity function gives an age estimate for the thin disk of 9.5+1.1-0.8 Gyr, without including the effect of <span class="hlt">phase</span> <span class="hlt">separation</span>. Using the above numbers, we see that <span class="hlt">phase</span> <span class="hlt">separation</span> could add between 0 and 3 Gyr to the white dwarf ages and still be consistent with the overall picture of Galaxy formation. Our calculated maximum value of <~1.5 Gyr fits within these bounds, as does our best-guess value of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21175693-substrate-misorientation-induced-strong-increase-hole-concentration-mg-doped-gan-grown-metalorganic-vapor-phase-epitaxy','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21175693-substrate-misorientation-induced-strong-increase-hole-concentration-mg-doped-gan-grown-metalorganic-vapor-phase-epitaxy"><span>Substrate misorientation <span class="hlt">induced</span> strong increase in the hole concentration in Mg doped GaN grown by metalorganic <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Suski, T.; Litwin-Staszewska, E.; Piotrzkowski, R.</p> <p></p> <p>We demonstrate that relatively small GaN substrate misorientation can strongly change hole carrier concentration in Mg doped GaN layers grown by metalorganic <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy. In this work intentionally misoriented GaN substrates (up to 2 deg. with respect to ideal <0001> plane) were employed. An increase in the hole carrier concentration to the level above 10{sup 18} cm{sup -3} and a decrease in GaN:Mg resistivity below 1 {omega} cm were achieved. Using secondary ion mass spectroscopy we found that Mg incorporation does not change with varying misorientation angle. This finding suggests that the compensation rate, i.e., a decrease in unintentionalmore » donor density, is responsible for the observed increase in the hole concentration. Analysis of the temperature dependence of electrical transport confirms this interpretation.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=63846&keyword=moisture+AND+removal&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50','EPA-EIMS'); return false;" href="https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=63846&keyword=moisture+AND+removal&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50"><span>THE EFFECT OF WATER (<span class="hlt">VAPOR-PHASE</span>) AND CARBON ON ELEMENTAL MERCURY REMOVAL IN A FLOW REACTOR</span></a></p> <p><a target="_blank" href="http://oaspub.epa.gov/eims/query.page">EPA Science Inventory</a></p> <p></p> <p></p> <p>The paper gives results of studying the effect of <span class="hlt">vapor-phase</span> moisture on elemental mercury (Hgo) removal by activated carbon (AC) in a flow reactor. tests involved injecting AC into both a dry and a 4% moisture nitrogen (N2) /Hgo gas stream. A bituminous-coal-based AC (Calgon WP...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1983JaJAP..22..240N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1983JaJAP..22..240N"><span>Growth of NH4Cl Single Crystal from <span class="hlt">Vapor</span> <span class="hlt">Phase</span> in Vertical Furnace</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nigara, Yutaka; Yoshizawa, Masahito; Fujimura, Tadao</p> <p>1983-02-01</p> <p>A pure and internally stress-free single crystal of NH4Cl was grown successfully from the <span class="hlt">vapor</span> <span class="hlt">phase</span>. The crystal measured 1.6 cmφ× 2 cm and had the disordered CsCl structure, which was stable below 184°C. The crystal was grown in an ampoule in a vertical furnace, in which the <span class="hlt">vapor</span> was efficiently transported both by diffusion and convection. In line with the growth mechanism of a single crystal, the temperature fluctuation (°C/min) on the growth interface was kept smaller than the product of the temperature gradient (°C/cm) and the growth rate (cm/min). The specific heat of the crystal was measured around -31°C (242 K) during cooling and heating cycles by AC calorimetry. The thermal hysteresis (0.4 K) obtained here was smaller than that (0.89 K) of an NH4Cl crystal grown from its aqueous solution with urea added as a habit modifier.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/5898779','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/biblio/5898779"><span>Drying of pulverized material with heated condensible <span class="hlt">vapor</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Carlson, L.W.</p> <p>1984-08-16</p> <p>Apparatus for drying pulverized material utilizes a high enthalpy condensable <span class="hlt">vapor</span> such as steam for removing moisture from the individual particles of the pulverized material. The initially wet particulate material is tangentially delivered by a carrier <span class="hlt">vapor</span> flow to an upper portion of a generally vertical cylindrical <span class="hlt">separation</span> drum. The lateral wall of the <span class="hlt">separation</span> drum is provided with a plurality of flow guides for directing the <span class="hlt">vapor</span> tangentially therein in the direction of particulate material flow. Positioned concentrically within the <span class="hlt">separation</span> drum and along the longitudinal axis thereof is a water-cooled condensation cylinder which is provided with a plurality of collection plates, or fines, on the outer lateral surface thereof. The cooled collection fines are aligned counter to the flow of the pulverized material and high enthalpy <span class="hlt">vapor</span> mixture to maximize water <span class="hlt">vapor</span> condensation thereon. The condensed liquid which includes moisture removed from the pulverized materials then flows downward along the outer surface of the coolant cylinder and is collected and removed. The particles travel in a shallow helix due to respective centrifugal and vertical acceleration forces applied thereto. The individual particles of the pulverized material are directed outwardly by the vortex flow where they contact the inner cylindrical surface of the <span class="hlt">separation</span> drum and are then deposited at the bottom thereof for easy collection and removal. The pulverized material drying apparatus is particularly adapted for drying coal fines and facilitates the recovery of the pulverized coal. 2 figs.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/865878','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/865878"><span>Drying of pulverized material with heated condensible <span class="hlt">vapor</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Carlson, Larry W.</p> <p>1986-01-01</p> <p>Apparatus for drying pulverized material utilizes a high enthalpy condensable <span class="hlt">vapor</span> such as steam for removing moisture from the individual particles of the pulverized material. The initially wet particulate material is tangentially delivered by a carrier <span class="hlt">vapor</span> flow to an upper portion of a generally vertical cylindrical <span class="hlt">separation</span> drum. The lateral wall of the <span class="hlt">separation</span> drum is provided with a plurality of flow guides for directing the <span class="hlt">vapor</span> tangentially therein in the direction of particulate material flow. Positioned concentrically within the <span class="hlt">separation</span> drum and along the longitudinal axis thereof is a water-cooled condensation cylinder which is provided with a plurality of collection plates, or fins, on the outer lateral surface thereof. The cooled collection fins are aligned counter to the flow of the pulverized material and high enthalpy <span class="hlt">vapor</span> mixture to maximize water <span class="hlt">vapor</span> condensation thereon. The condensed liquid which includes moisture removed from the pulverized material then flows downward along the outer surface of the coolant cylinder and is collected and removed. The particles travel in a shallow helix due to respective centrifugal and vertical acceleration forces applied thereto. The individual particles of the pulverized material are directed outwardly by the vortex flow where they contact the inner cylindrical surface of the <span class="hlt">separation</span> drum and are then deposited at the bottom thereof for easy collection and removal. The pulverized material drying apparatus is particularly adapted for drying coal fines and facilitates the recovery of the pulverized coal.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015PhRvB..91f0513L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015PhRvB..91f0513L"><span>Role of chalcogen <span class="hlt">vapor</span> annealing in <span class="hlt">inducing</span> bulk superconductivity in Fe1 +yTe1 -xSex</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lin, Wenzhi; Ganesh, P.; Gianfrancesco, Anthony; Wang, Jun; Berlijn, Tom; Maier, Thomas A.; Kalinin, Sergei V.; Sales, Brian C.; Pan, Minghu</p> <p>2015-02-01</p> <p>Recent investigations have shown that Fe1 +yTe1 -xSex can be made superconducting by annealing it in Se and O <span class="hlt">vapors</span>. The current lore is that these chalcogen <span class="hlt">vapors</span> <span class="hlt">induce</span> superconductivity by removing the magnetic excess Fe atoms. To investigate this phenomenon, we performed a combination of magnetic susceptibility, specific heat, and transport measurements together with scanning tunneling microscopy and spectroscopy and density functional theory calculations on Fe1 +yTe1 -xSex treated with Te <span class="hlt">vapor</span>. We conclude that the main role of the Te <span class="hlt">vapor</span> is to quench the magnetic moments of the excess Fe atoms by forming FeTem (m ≥1 ) complexes. We show that the remaining FeTem complexes are still damaging to the superconductivity and therefore that their removal potentially could further improve superconductive properties in these compounds.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009APS..MARW35004A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009APS..MARW35004A"><span><span class="hlt">Phase</span> <span class="hlt">separation</span> of electrons strongly coupled with phonons in cuprates and manganites</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Alexandrov, Sasha</p> <p>2009-03-01</p> <p>Recent advanced Monte Carlo simulations have not found superconductivity and <span class="hlt">phase</span> <span class="hlt">separation</span> in the Hubbard model with on-site repulsive electron-electron correlations. I argue that microscopic <span class="hlt">phase</span> <span class="hlt">separations</span> in cuprate superconductors and colossal magnetoresistance (CMR) manganites originate from a strong electron-phonon interaction (EPI) combined with unavoidable disorder. Attractive electron correlations, caused by an almost unretarded EPI, are sufficient to overcome the direct inter-site Coulomb repulsion in these charge-transfer Mott-Hubbard insulators, so that low energy physics is that of small polarons and small bipolarons. They form clusters localized by disorder below the mobility edge, but propagate as the Bloch states above the mobility edge. I identify the Froehlich EPI as the most essential for pairing and <span class="hlt">phase</span> <span class="hlt">separation</span> in superconducting layered cuprates. The pairing of oxygen holes into heavy bipolarons in the paramagnetic <span class="hlt">phase</span> (current-carrier density collapse (CCDC)) explains also CMR and high and low-resistance <span class="hlt">phase</span> coexistence near the ferromagnetic transition of doped manganites.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26845079','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26845079"><span>Gas-<span class="hlt">Phase</span> Dopant-<span class="hlt">Induced</span> Conformational Changes Monitored with Transversal Modulation Ion Mobility Spectrometry.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Meyer, Nicole Andrea; Root, Katharina; Zenobi, Renato; Vidal-de-Miguel, Guillermo</p> <p>2016-02-16</p> <p>The potential of a Transversal Modulation Ion Mobility Spectrometry (TMIMS) instrument for protein analysis applications has been evaluated. The Collision Cross Section (CCS) of cytochrome c measured with the TMIMS is in agreement with values reported in the literature. Additionally, it enables tandem IMS-IMS prefiltration in dry gas and in <span class="hlt">vapor</span> doped gas. The chemical specificity of the different dopants enables interesting studies on the structure of proteins as CCS changed strongly depending on the specific dopant. Hexane produced an unexpectedly high CCS shift, which can be utilized to evaluate the exposure of hydrophobic parts of the protein. Alcohols produced higher shifts with a dual behavior: an increase in CCS due to <span class="hlt">vapor</span> uptake at specific absorption sites, followed by a linear shift typical for unspecific and unstable <span class="hlt">vapor</span> uptake. The molten globule +8 shows a very specific transition. Initially, its CCS follows the trend of the compact folded states, and then it rapidly increases to the levels of the unfolded states. This strong variation suggests that the +8 charge state undergoes a dopant-<span class="hlt">induced</span> conformational change. Interestingly, more sterically demanding alcohols seem to unfold the protein more effectively also in the gas <span class="hlt">phase</span>. This study shows the capabilities of the TMIMS device for protein analysis and how tandem IMS-IMS with dopants could provide better understanding of the conformational changes of proteins.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/953670','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/953670"><span>Method and Apparatus for Concentrating <span class="hlt">Vapors</span> for Analysis</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Grate, Jay W.; Baldwin, David L.; Anheier, Jr., Norman C.</p> <p>2008-10-07</p> <p>An apparatus and method are disclosed for pre-concentrating gaseous <span class="hlt">vapors</span> for analysis. The invention finds application in conjunction with, e.g., analytical instruments where low detection limits for gaseous <span class="hlt">vapors</span> are desirable. <span class="hlt">Vapors</span> sorbed and concentrated within the bed of the apparatus can be thermally desorbed achieving at least partial <span class="hlt">separation</span> of <span class="hlt">vapor</span> mixtures. The apparatus is suitable, e.g., for preconcentration and sample injection, and provides greater resolution of peaks for <span class="hlt">vapors</span> within <span class="hlt">vapor</span> mixtures, yielding detection levels that are 10-10,000 times better than for direct sampling and analysis systems. Features are particularly useful for continuous unattended monitoring applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006PhRvA..74f3831Q','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006PhRvA..74f3831Q"><span>Dimension-sensitive optical responses of electromagnetically <span class="hlt">induced</span> transparency <span class="hlt">vapor</span> in a waveguide</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Qi Shen, Jian; He, Sailing</p> <p>2006-12-01</p> <p>A three-level EIT (electromagnetically <span class="hlt">induced</span> transparency) <span class="hlt">vapor</span> is used to manipulate the transparency and absorption properties of the probe light in a waveguide. The most remarkable feature of the present scheme is such that the optical responses resulting from both electromagnetically <span class="hlt">induced</span> transparency and large spontaneous emission enhancement are very sensitive to the frequency detunings of the probe light as well as to the small changes of the waveguide dimension. The potential applications of the dimension- and dispersion-sensitive EIT responses are discussed, and the sensitivity limits of some waveguide-based sensors, including electric absorption modulator, optical switch, wavelength sensor, and sensitive magnetometer, are analyzed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=335740&Lab=NERL&keyword=forensics&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50','EPA-EIMS'); return false;" href="https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=335740&Lab=NERL&keyword=forensics&actType=&TIMSType=+&TIMSSubTypeID=&DEID=&epaNumber=&ntisID=&archiveStatus=Both&ombCat=Any&dateBeginCreated=&dateEndCreated=&dateBeginPublishedPresented=&dateEndPublishedPresented=&dateBeginUpdated=&dateEndUpdated=&dateBeginCompleted=&dateEndCompleted=&personID=&role=Any&journalID=&publisherID=&sortBy=revisionDate&count=50"><span>On the Implications of aerosol liquid water and <span class="hlt">phase</span> <span class="hlt">separation</span> for modeled organic aerosol mass</span></a></p> <p><a target="_blank" href="http://oaspub.epa.gov/eims/query.page">EPA Science Inventory</a></p> <p></p> <p></p> <p>Current chemical transport models assume that organic aerosol (OA)-forming compounds partition mostly to a water-poor, organic-rich <span class="hlt">phase</span> in accordance with their <span class="hlt">vapor</span> pressures. However, in the southeast United States, a significant fraction of ambient organic compounds are wat...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014AGUFM.A23Q..03M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014AGUFM.A23Q..03M"><span><span class="hlt">Vapor</span> Wall Deposition in Chambers: Theoretical Considerations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>McVay, R.; Cappa, C. D.; Seinfeld, J.</p> <p>2014-12-01</p> <p>In order to constrain the effects of <span class="hlt">vapor</span> wall deposition on measured secondary organic aerosol (SOA) yields in laboratory chambers, Zhang et al. (2014) varied the seed aerosol surface area in toluene oxidation and observed a clear increase in the SOA yield with increasing seed surface area. Using a coupled <span class="hlt">vapor</span>-particle dynamics model, we examine the extent to which this increase is the result of <span class="hlt">vapor</span> wall deposition versus kinetic limitations arising from imperfect accommodation of organic species into the particle <span class="hlt">phase</span>. We show that a seed surface area dependence of the SOA yield is present only when condensation of <span class="hlt">vapors</span> onto particles is kinetically limited. The existence of kinetic limitation can be predicted by comparing the characteristic timescales of gas-<span class="hlt">phase</span> reaction, <span class="hlt">vapor</span> wall deposition, and gas-particle equilibration. The gas-particle equilibration timescale depends on the gas-particle accommodation coefficient αp. Regardless of the extent of kinetic limitation, <span class="hlt">vapor</span> wall deposition depresses the SOA yield from that in its absence since <span class="hlt">vapor</span> molecules that might otherwise condense on particles deposit on the walls. To accurately extrapolate chamber-derived yields to atmospheric conditions, both <span class="hlt">vapor</span> wall deposition and kinetic limitations must be taken into account.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25084590','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25084590"><span>Pollution level and distribution of PCDD/PCDF congeners between <span class="hlt">vapor</span> <span class="hlt">phase</span> and particulate <span class="hlt">phase</span> in winter air of Dalian, China.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Wei; Qin, Songtao; Song, Yu; Xu, Qian; Ni, Yuwen; Chen, Jiping; Zhang, Xueping; Mu, Jim; Zhu, Xiuhua</p> <p>2011-06-01</p> <p>In December 2009, ambient air was sampled with active high-volume air samplers at two sites: on the roof of the No. l building of Dalian Jiaotong University and on the roof of the building of Dalian Meteorological Observatory. The concentrations and the congeners between <span class="hlt">vapor</span> <span class="hlt">phase</span> and particulate <span class="hlt">phase</span> of polychlorinated dibenzo-p-dioxins and dibenzofurans (PCDD/Fs) in the air were measured. Sample analysis results showed that the concentrations of PCDD/Fs in particulate <span class="hlt">phase</span> was higher than that in gaseous <span class="hlt">phase</span>. The ratio of PCDD to PCDF in gaseous <span class="hlt">phase</span> and particulate <span class="hlt">phase</span> was lower than 0.4 in all samples. The total I-TEQ value in gaseous <span class="hlt">phase</span> and particulate <span class="hlt">phase</span> was 5.5 and 453.8 fg/m(3) at Dalian Jiaotong University, 16.6 and 462.1 fg/m(3) at Dalian Meteorological Observatory, respectively. The I-TEQ value of Dalian atmosphere was 5.5-462.1 fg/m(3) which was lower than international standard, the atmospheric quality in Dalian is better. Copyright © 2011 The Research Centre for Eco-Environmental Sciences, Chinese Academy of Sciences. Published by Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22299732-existence-vapor-liquid-phase-transition-dusty-plasmas','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22299732-existence-vapor-liquid-phase-transition-dusty-plasmas"><span>On the existence of <span class="hlt">vapor</span>-liquid <span class="hlt">phase</span> transition in dusty plasmas</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Kundu, M.; Sen, A.; Ganesh, R.</p> <p>2014-10-15</p> <p>The phenomenon of <span class="hlt">phase</span> transition in a dusty-plasma system (DPS) has attracted some attention in the past. Earlier Farouki and Hamaguchi [J. Chem. Phys. 101, 9876 (1994)] have demonstrated the existence of a liquid to solid transition in DPS where the dust particles interact through a Yukawa potential. However, the question of the existence of a <span class="hlt">vapor</span>-liquid (VL) transition in such a system remains unanswered and relatively unexplored so far. We have investigated this problem by performing extensive molecular dynamics simulations which show that the VL transition does not have a critical curve in the pressure versus volume diagram formore » a large range of the Yukawa screening parameter κ and the Coulomb coupling parameter Γ. Thus, the VL <span class="hlt">phase</span> transition is found to be super-critical, meaning that this transition is continuous in the dusty plasma model given by Farouki and Hamaguchi. We provide an approximate analytic explanation of this finding by means of a simple model calculation.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/907978','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/907978"><span>Device for two-dimensional gas-<span class="hlt">phase</span> <span class="hlt">separation</span> and characterization of ion mixtures</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Tang, Keqi [Richland, WA; Shvartsburg, Alexandre A [Richland, WA; Smith, Richard D [Richland, WA</p> <p>2006-12-12</p> <p>The present invention relates to a device for <span class="hlt">separation</span> and characterization of gas-<span class="hlt">phase</span> ions. The device incorporates an ion source, a field asymmetric waveform ion mobility spectrometry (FAIMS) analyzer, an ion mobility spectrometry (IMS) drift tube, and an ion detector. In one aspect of the invention, FAIMS operating voltages are electrically floated on top of the IMS drift voltage. In the other aspect, the FAIMS/IMS interface is implemented employing an electrodynamic ion funnel, including in particular an hourglass ion funnel. The present invention improves the efficiency (peak capacity) and sensitivity of gas-<span class="hlt">phase</span> <span class="hlt">separations</span>; the online FAIMS/IMS coupling creates a fundamentally novel two-dimensional gas-<span class="hlt">phase</span> <span class="hlt">separation</span> technology with high peak capacity, specificity, and exceptional throughput.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016APS..MAR.M1021I','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016APS..MAR.M1021I"><span><span class="hlt">Phase</span> <span class="hlt">separated</span> microstructure and dynamics of polyurethane elastomers under strain</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Iacob, Ciprian; Padsalgikar, Ajay; Runt, James</p> <p></p> <p>The molecular mobility of polyurethane elastomers is of the utmost importance in establishing physical properties for uses ranging from automotive tires and shoe soles to more sophisticated aerospace and biomedical applications. In many of these applications, chain dynamics as well as mechanical properties under external stresses/strains are critical for determining ultimate performance. In order to develop a more complete understanding of their mechanical response, we explored the effect of uniaxial strain on the <span class="hlt">phase</span> <span class="hlt">separated</span> microstructure and molecular dynamics of the elastomers. We utilize X-ray scattering to investigate soft segment and hard domain orientation, and broadband dielectric spectroscopy for interrogation of the dynamics. Uniaxial deformation is found to significantly perturb the <span class="hlt">phase-separated</span> microstructure and chain orientation, and results in a considerable slowing down of the dynamics of the elastomers. Attenuated total reflectance Fourier transform infrared spectroscopy measurements of the polyurethanes under uniaxial deformation are also employed and the results are quantitatively correlated with mechanical tensile tests and the degree of <span class="hlt">phase</span> <span class="hlt">separation</span> from small-angle X-ray scattering measurements.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20030060500','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20030060500"><span>Real-Time Optical Monitoring and Simulations of Gas <span class="hlt">Phase</span> Kinetics in InN <span class="hlt">Vapor</span> <span class="hlt">Phase</span> Epitaxy at High Pressure</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Dietz, Nikolaus; Woods, Vincent; McCall, Sonya D.; Bachmann, Klaus J.</p> <p>2003-01-01</p> <p>Understanding the kinetics of nucleation and coalescence of heteroepitaxial thin films is a crucial step in controlling a chemical <span class="hlt">vapor</span> deposition process, since it defines the perfection of the heteroepitaxial film both in terms of extended defect formation and chemical integrity of the interface. The initial nucleation process also defines the film quality during the later stages of film growth. The growth of emerging new materials heterostructures such as InN or In-rich Ga(x)In(1-x)N require deposition methods operating at higher <span class="hlt">vapor</span> densities due to the high thermal decomposition pressure in these materials. High nitrogen pressure has been demonstrated to suppress thermal decomposition of InN, but has not been applied yet in chemical <span class="hlt">vapor</span> deposition or etching experiments. Because of the difficulty with maintaining stochiometry at elevated temperature, current knowledge regarding thermodynamic data for InN, e.g., its melting point, temperature-dependent heat capacity, heat and entropy of formation are known with far less accuracy than for InP, InAs and InSb. Also, no information exists regarding the partial pressures of nitrogen and phosphorus along the liquidus surfaces of mixed-anion alloys of InN, of which the InN(x)P(1-x) system is the most interesting option. A miscibility gap is expected for InN(x)P(1-x) pseudobinary solidus compositions, but its extent is not established at this point by experimental studies under near equilibrium conditions. The extension of chemical <span class="hlt">vapor</span> deposition to elevated pressure is also necessary for retaining stoichiometric single <span class="hlt">phase</span> surface composition for materials that are characterized by large thermal decomposition pressures at optimum processing temperatures.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li class="active"><span>23</span></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_23 --> <div id="page_24" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li class="active"><span>24</span></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="461"> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20020086971&hterms=Shrink&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3DShrink','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20020086971&hterms=Shrink&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D80%26Ntt%3DShrink"><span>The Collapse of <span class="hlt">Vapor</span> Bubbles in a Spatially Non-Uniform Flow</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hao, Y.; Prosperetti, A.</p> <p>2000-01-01</p> <p>Pressure gradients act differently on liquid particles and suspended bubbles and are, therefore, capable of <span class="hlt">inducing</span> a relative motion between the <span class="hlt">phases</span> even when no relative velocity initially exists. As a consequence of the enhanced heat transfer in the presence of convection, this fact may have a major impact on the evolution of a <span class="hlt">vapor</span> bubble. The effect is particularly strong in the case of a collapsing bubble for which, due to the conservation of the system's impulse, the <span class="hlt">induced</span> relative velocity tends to be magnified when the bubble volume shrinks. A practical application could be, for instance, the enhancement of the condensation rate of bubbles downstream of a heated region, thereby reducing the quality of a flowing liquid-<span class="hlt">vapor</span> mixture. A simple model of the process, in which the bubble is assumed to be spherical and the flow potential, is developed in the paper.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/4309536','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/biblio/4309536"><span>METHOD OF <span class="hlt">SEPARATING</span> ISOTOPES OF URANIUM IN A CALUTRON</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Jenkins, F.A.</p> <p>1958-05-01</p> <p>Mass <span class="hlt">separation</span> devices of the calutron type and the use of uranium hexachloride as a charge material in the calutron ion source are described. The method for using this material in a mass <span class="hlt">separator</span> includes heating the uranium hexachloride to a temperature in the range of 60 to 100 d C in a vacuum and thereby forming a <span class="hlt">vapor</span> of the material. The <span class="hlt">vaporized</span> uranium hexachloride is then ionized in a <span class="hlt">vapor</span> ionizing device for subsequent mass <span class="hlt">separation</span> processing.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015APS..DFDM22010T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015APS..DFDM22010T"><span>Suppression of turbulent energy cascade due to <span class="hlt">phase</span> <span class="hlt">separation</span> in homogenous binary mixture fluid</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Takagi, Youhei; Okamoto, Sachiya</p> <p>2015-11-01</p> <p>When a multi-component fluid mixture becomes themophysically unstable state by quenching from well-melting condition, <span class="hlt">phase</span> <span class="hlt">separation</span> due to spinodal decomposition occurs, and a self-organized structure is formed. During <span class="hlt">phase</span> <span class="hlt">separation</span>, free energy is consumed for the structure formation. In our previous report, the <span class="hlt">phase</span> <span class="hlt">separation</span> in homogenous turbulence was numerically simulated and the coarsening process of <span class="hlt">phase</span> <span class="hlt">separation</span> was discussed. In this study, we extended our numerical model to a high Schmidt number fluid corresponding to actual polymer solution. The governing equations were continuity, Navier-Stokes, and Chan-Hiliard equations as same as our previous report. The flow filed was an isotropic homogenous turbulence, and the dimensionless parameters in the Chan-Hilliard equation were estimated based on the thermophysical condition of binary mixture. From the numerical results, it was found that turbulent energy cascade was drastically suppressed in the inertial subrange by <span class="hlt">phase</span> <span class="hlt">separation</span> for the high Schmidt number flow. By using the identification of turbulent and <span class="hlt">phase</span> <span class="hlt">separation</span> structure, we discussed the relation between total energy balance and the structures formation processes. This study is financially supported by the Grand-in-Aid for Young Scientists (B) (No. T26820045) from the Ministry of Education, Cul-ture, Sports, Science and Technology of Japan.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1999PhyA..268...50C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1999PhyA..268...50C"><span><span class="hlt">Phase</span> <span class="hlt">separation</span> in living micellar networks</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cristobal, G.; Rouch, J.; Curély, J.; Panizza, P.</p> <p></p> <p>We present a lattice model based on two n→0 spin vectors, capable of treating the thermodynamics of living networks in micellar solutions at any surfactant concentration. We establish an isomorphism between the coupling constants in the two spin vector Hamiltonian and the surfactant energies involved in the micellar situation. Solving this Hamiltonian in the mean-field approximation allows one to calculate osmotic pressure, aggregation number, free end and cross-link densities at any surfactant concentration. We derive a <span class="hlt">phase</span> diagram, including changes in topology such as the transition between spheres and rods and between saturated and unsaturated networks. A <span class="hlt">phase</span> <span class="hlt">separation</span> can be found between a saturated network and a dilute solution composed of long flexible micelles or a saturated network and a solution of spherical micelles.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006AGUFMMR31A..06L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006AGUFMMR31A..06L"><span>Determination of Methane Hydrate Solubility in the Absence of <span class="hlt">Vapor</span> <span class="hlt">Phase</span> by in-situ Raman Spectroscopy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lu, W.; Chou, I.; Burruss, R.</p> <p>2006-12-01</p> <p>Prediction of the occurrence, distribution, and evolution of methane hydrate in porous marine sediments requires information on solubilities of methane hydrate in water. Solubilities of methane hydrate in the presence of a <span class="hlt">vapor</span> <span class="hlt">phase</span> are well established, but those in the absence of a <span class="hlt">vapor</span> <span class="hlt">phase</span> are not well defined with differences up to 30%. We have measured methane concentrations in pure water in equilibrium with sI methane hydrate, in the absence of <span class="hlt">vapor</span> <span class="hlt">phase</span>, by in-situ Raman spectroscopy at temperatures (T) from 2 to 20 (± 0.3) °C and pressures (P) at 10, 20, 30, and 40 (± 0.4%) MPa. Methane hydrate was synthesized in a high-pressure capillary optical cell (Chou et al., 2005; Advances in High-Pressure Technology for Geophysical Applications. Ed. J. Chen et al., Chapter 24, p. 475, Elsevier). A small quantity of methane was first loaded in an evacuated cell and then pressurized by water. Hydrate crystals were formed near the liquid-<span class="hlt">vapor</span> interface near the enclosed end of the optical tube at room T, and were then placed at the center of a USGS-type heating-cooling stage. By adjusting sample P and T, the crystals went through dissolution-formation cycles three to four times in three days until the <span class="hlt">vapor</span> <span class="hlt">phase</span> was completely consumed and several crystals (typically 40 x 40 x 10 μm) were formed. These crystals were located at about 200 μm from the enclosed end and were about 20 to 40 μm from each other. Raman spectra were collected for the liquid <span class="hlt">phase</span> adjacent to hydrate crystals near the enclosed end of the tube. A volumetric decrease in crystal size was observed away from the sampling spot; however, no such volumetric decrease was observed in or near the sampling spot. Therefore, equilibrium was likely established locally within the sampling area. The results are represented by the following linear isobaric equations: 10 MPa: ln [X(CH4)] = 0.06175 T - 6.79507; r2 = 0.9991 (n = 6) 20 MPa: ln [X(CH4)] = 0.06170 T - 6.82816; r2 = 0.9985 (n = 6) 30 MPa</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29390781','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29390781"><span>A model for acoustic <span class="hlt">vaporization</span> dynamics of a bubble/droplet system encapsulated within a hyperelastic shell.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lacour, Thomas; Guédra, Matthieu; Valier-Brasier, Tony; Coulouvrat, François</p> <p>2018-01-01</p> <p>Nanodroplets have great, promising medical applications such as contrast imaging, embolotherapy, or targeted drug delivery. Their functions can be mechanically activated by means of focused ultrasound <span class="hlt">inducing</span> a <span class="hlt">phase</span> change of the inner liquid known as the acoustic droplet <span class="hlt">vaporization</span> (ADV) process. In this context, a four-<span class="hlt">phases</span> (<span class="hlt">vapor</span> + liquid + shell + surrounding environment) model of ADV is proposed. Attention is especially devoted to the mechanical properties of the encapsulating shell, incorporating the well-known strain-softening behavior of Mooney-Rivlin material adapted to very large deformations of soft, nearly incompressible materials. Various responses to ultrasound excitation are illustrated, depending on linear and nonlinear mechanical shell properties and acoustical excitation parameters. Different classes of ADV outcomes are exhibited, and a relevant threshold ensuring complete <span class="hlt">vaporization</span> of the inner liquid layer is defined. The dependence of this threshold with acoustical, geometrical, and mechanical parameters is also provided.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19880039338&hterms=gallium+vapor&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dgallium%2Bvapor','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19880039338&hterms=gallium+vapor&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dgallium%2Bvapor"><span>Reaction mechanisms in the organometallic <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxial growth of GaAs</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Larsen, C. A.; Buchan, N. I.; Stringfellow, G. B.</p> <p>1988-01-01</p> <p>The decomposition mechanisms of AsH3, trimethylgallium (TMGa), and mixtures of the two have been studied in an atmospheric-pressure flow system with the use of D2 to label the reaction products which are analyzed in a time-of-flight mass spectrometer. AsH3 decomposes entirely heterogeneously to give H2. TMGa decomposes by a series of gas-<span class="hlt">phase</span> steps, involving methyl radicals and D atoms to produce CH3D, CH4, C2H6, and HD. TMGa decomposition is accelerated by the presence of AsH3. When the two are mixed, as in the organometallic <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxial growth of GaAs, both compounds decompose in concert to produce only CH4. A likely model is that of a Lewis acid-base adduct that forms and subsequently eliminates CH4.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1988ApPhL..52..480L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1988ApPhL..52..480L"><span>Reaction mechanisms in the organometallic <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxial growth of GaAs</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Larsen, C. A.; Buchan, N. I.; Stringfellow, G. B.</p> <p>1988-02-01</p> <p>The decomposition mechanisms of AsH3, trimethylgallium (TMGa), and mixtures of the two have been studied in an atmospheric-pressure flow system with the use of D2 to label the reaction products which are analyzed in a time-of-flight mass spectrometer. AsH3 decomposes entirely heterogeneously to give H2. TMGa decomposes by a series of gas-<span class="hlt">phase</span> steps, involving methyl radicals and D atoms to produce CH3D, CH4, C2H6, and HD. TMGa decomposition is accelerated by the presence of AsH3. When the two are mixed, as in the organometallic <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxial growth of GaAs, both compounds decompose in concert to produce only CH4. A likely model is that of a Lewis acid-base adduct that forms and subsequently eliminates CH4.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JOM...tmp..155K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JOM...tmp..155K"><span>Microstructure Evolution and Related Magnetic Properties of Cu-Zr-Al-Gd <span class="hlt">Phase-Separating</span> Metallic Glasses</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kim, Sang Jun; Kim, Jinwoo; Park, Eun Soo</p> <p>2018-04-01</p> <p>We carefully investigated the correlation between microstructures and magnetic properties of Cu-Zr-Al-Gd <span class="hlt">phase-separating</span> metallic glasses (PSMGs). The saturation magnetizations of the PSMGs were determined by total Gd contents of the alloys, while their coercivity exhibits a large deviation by the occurrence of <span class="hlt">phase</span> <span class="hlt">separation</span> due to the boundary pinning effect of hierarchically <span class="hlt">separated</span> amorphous <span class="hlt">phases</span>. Especially, the PSMGs containing Gd-rich amorphous nanoparticles show the highest coercivity which can be attributed to the size effect of the ferromagnetic amorphous <span class="hlt">phase</span>. Furthermore, the selective crystallization of ferromagnetic amorphous <span class="hlt">phases</span> can affect the magnetization behavior of the PSMGs. Our results could provide a novel strategy for tailoring unique soft magnetic properties of metallic glasses by introducing hierarchically <span class="hlt">separated</span> amorphous <span class="hlt">phases</span> and controlling their crystallinity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JOM....70f.988K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JOM....70f.988K"><span>Microstructure Evolution and Related Magnetic Properties of Cu-Zr-Al-Gd <span class="hlt">Phase-Separating</span> Metallic Glasses</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kim, Sang Jun; Kim, Jinwoo; Park, Eun Soo</p> <p>2018-06-01</p> <p>We carefully investigated the correlation between microstructures and magnetic properties of Cu-Zr-Al-Gd <span class="hlt">phase-separating</span> metallic glasses (PSMGs). The saturation magnetizations of the PSMGs were determined by total Gd contents of the alloys, while their coercivity exhibits a large deviation by the occurrence of <span class="hlt">phase</span> <span class="hlt">separation</span> due to the boundary pinning effect of hierarchically <span class="hlt">separated</span> amorphous <span class="hlt">phases</span>. Especially, the PSMGs containing Gd-rich amorphous nanoparticles show the highest coercivity which can be attributed to the size effect of the ferromagnetic amorphous <span class="hlt">phase</span>. Furthermore, the selective crystallization of ferromagnetic amorphous <span class="hlt">phases</span> can affect the magnetization behavior of the PSMGs. Our results could provide a novel strategy for tailoring unique soft magnetic properties of metallic glasses by introducing hierarchically <span class="hlt">separated</span> amorphous <span class="hlt">phases</span> and controlling their crystallinity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19920030261&hterms=oil+disposition&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Doil%2Bdisposition','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19920030261&hterms=oil+disposition&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3Doil%2Bdisposition"><span>Cell <span class="hlt">separations</span> and the demixing of aqueous two <span class="hlt">phase</span> polymer solutions in microgravity</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Brooks, Donald E.; Bamberger, Stephan; Harris, J. M.; Van Alstine, James M.</p> <p>1991-01-01</p> <p>Partition in <span class="hlt">phase</span> <span class="hlt">separated</span> aqueous polymer solutions is a cell <span class="hlt">separation</span> procedure thought to be adversely influenced by gravity. In preparation for performing cell partitioning experiments in space, and to provide general information concerning the demixing of immiscible liquids in low gravity, a series of <span class="hlt">phase</span> <span class="hlt">separated</span> aqueous polymer solutions have been flown on two shuttle flights. Fluorocarbon oil and water emulsions were also flown on the second flight. The aqueous polymer emulsions, which in one g demix largely by sedimentation and convection due to the density differences between the <span class="hlt">phases</span>, demixed more slowly than on the ground and the final disposition of the <span class="hlt">phases</span> was determined by the wetting of the container wall by the <span class="hlt">phases</span>. The demixing behavior and kinetics were influenced by the <span class="hlt">phase</span> volume ratio, physical properties of the systems and chamber wall interaction. The average domain size increased linearly with time as the systems demixed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25375491','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25375491"><span><span class="hlt">Phase</span> <span class="hlt">separation</span> and emergent structures in an active nematic fluid.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Putzig, Elias; Baskaran, Aparna</p> <p>2014-10-01</p> <p>We consider a phenomenological continuum theory for an active nematic fluid and show that there exists a universal, model-independent instability which renders the homogeneous nematic state unstable to order fluctuations. Using numerical and analytic tools we show that, in the vicinity of a critical point, this instability leads to a <span class="hlt">phase-separated</span> state in which the ordered regions form bands in which the direction of nematic order is perpendicular to the direction of the density gradient. We argue that the underlying mechanism that leads to this <span class="hlt">phase</span> <span class="hlt">separation</span> is a universal feature of active fluids of different symmetries.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JLTP..191..153W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JLTP..191..153W"><span>Morphological Simulation of <span class="hlt">Phase</span> <span class="hlt">Separation</span> Coupled Oscillation Shear and Varying Temperature Fields</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, Heping; Li, Xiaoguang; Lin, Kejun; Geng, Xingguo</p> <p>2018-05-01</p> <p>This paper explores the effect of the shear frequency and Prandtl number ( Pr) on the procedure and pattern formation of <span class="hlt">phase</span> <span class="hlt">separation</span> in symmetric and asymmetric systems. For the symmetric system, the periodic shear significantly prolongs the spinodal decomposition stage and enlarges the <span class="hlt">separated</span> domain in domain growth stage. By adjusting the Pr and shear frequency, the number and orientation of <span class="hlt">separated</span> steady layer structures can be controlled during domain stretch stage. The numerical results indicate that the increase in Pr and decrease in the shear frequency can significantly increase in the layer number of the lamellar structure, which relates to the decrease in domain size. Furthermore, the lamellar orientation parallel to the shear direction is altered into that perpendicular to the shear direction by further increasing the shear frequency, and also similar results for larger systems. For asymmetric system, the quantitative analysis shows that the decrease in the shear frequency enlarges the size of <span class="hlt">separated</span> minority <span class="hlt">phases</span>. These numerical results provide guidance for setting the optimum condition for the <span class="hlt">phase</span> <span class="hlt">separation</span> under periodic shear and slow cooling.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..DFDF15003A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..DFDF15003A"><span>Dynamical <span class="hlt">phase</span> <span class="hlt">separation</span> using a microfluidic device: experiments and modeling</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Aymard, Benjamin; Vaes, Urbain; Radhakrishnan, Anand; Pradas, Marc; Gavriilidis, Asterios; Kalliadasis, Serafim; Complex Multiscale Systems Team</p> <p>2017-11-01</p> <p>We study the dynamical <span class="hlt">phase</span> <span class="hlt">separation</span> of a binary fluid by a microfluidic device both from the experimental and from the modeling points of view. The experimental device consists of a main channel (600 μm wide) leading into an array of 276 trapezoidal capillaries of 5 μm width arranged on both sides and <span class="hlt">separating</span> the lateral channels from the main channel. Due to geometrical effects as well as wetting properties of the substrate, and under well chosen pressure boundary conditions, a multiphase flow introduced into the main channel gets <span class="hlt">separated</span> at the capillaries. Understanding this dynamics via modeling and numerical simulation is a crucial step in designing future efficient micro-<span class="hlt">separators</span>. We propose a diffuse-interface model, based on the classical Cahn-Hilliard-Navier-Stokes system, with a new nonlinear mobility and new wetting boundary conditions. We also propose a novel numerical method using a finite-element approach, together with an adaptive mesh refinement strategy. The complex geometry is captured using the same computer-aided design files as the ones adopted in the fabrication of the actual device. Numerical simulations reveal a very good qualitative agreement between model and experiments, demonstrating also a clear <span class="hlt">separation</span> of <span class="hlt">phases</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016APS..MARS37015R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016APS..MARS37015R"><span><span class="hlt">Phase</span> <span class="hlt">separation</span> in artificial vesicles driven by light and curvature</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rinaldin, Melissa; Pomp, Wim; Schmidt, Thomas; Giomi, Luca; Kraft, Daniela; Physics of Life Processes Team; Soft; Bio Mechanics Collaboration; Self-Assembly in Soft Matter Systems Collaboration</p> <p></p> <p>The role of <span class="hlt">phase</span>-demixing in living cells, leading to the lipid-raft hypothesis, has been extensively studied. Lipid domains of higher lipid chain order are proposed to regulate protein spatial organization. Giant Unilamellar Vesicles provide an artificial model to study <span class="hlt">phase</span> <span class="hlt">separation</span>. So far temperature was used to initiate the process. Here we introduce a new methodology based on the induction of <span class="hlt">phase</span> <span class="hlt">separation</span> by light. To this aim, the composition of the lipid membrane is varied by photo-oxidation of lipids. The control of the process gained by using light allowed us to observe vesicle shape fluctuations during <span class="hlt">phase</span>-demixing. The presence of fluctuations near the critical mixing point resembles features of a critical process. We quantitatively analyze these fluctuations using a 2d elastic model, from which we can estimate the material parameters such as bending rigidity and surface tension, demonstrating the non-equilibrium critical behaviour. Finally, I will describe recent attempts toward tuning the membrane composition by controlling the vesicle curvature.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..DFD.G6001B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..DFD.G6001B"><span>Acoustically-Enhanced Direct Contact <span class="hlt">Vapor</span> Bubble Condensation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Boziuk, Thomas; Smith, Marc; Glezer, Ari</p> <p>2017-11-01</p> <p>Rate-limited, direct contact <span class="hlt">vapor</span> condensation of <span class="hlt">vapor</span> bubbles that are formed by direct steam injection through a nozzle in a quiescent subcooled liquid bath is accelerated using ultrasonic (MHz-range) actuation. A submerged, low power actuator produces an acoustic beam whose radiation pressure deforms the liquid-<span class="hlt">vapor</span> interface, leading to the formation of a liquid spear that penetrates the <span class="hlt">vapor</span> bubble to form a <span class="hlt">vapor</span> torus with a significantly larger surface area and condensation rate. Ultrasonic focusing along the spear leads to the ejection of small, subcooled droplets through the <span class="hlt">vapor</span> volume that impact the <span class="hlt">vapor</span>-liquid interface and further enhance the condensation. High-speed Schlieren imaging of the formation and collapse of the <span class="hlt">vapor</span> bubbles in the absence and presence of actuation shows that the impulse associated with the collapse of the toroidal volume leads to the formation of a turbulent vortex ring in the liquid <span class="hlt">phase</span>. Liquid motions near the condensing <span class="hlt">vapor</span> volume are investigated in the absence and presence of acoustic actuation using high-magnification PIV and show the evolution of a liquid jet through the center of the condensing toroidal volume and the formation and advection of vortex ring structures whose impulse appear to increase with temperature difference between the liquid and <span class="hlt">vapor</span> <span class="hlt">phases</span>. High-speed image processing is used to assess the effect of the actuation on the temporal and spatial variations in the characteristic scales and condensation rates of the <span class="hlt">vapor</span> bubbles.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4585945','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4585945"><span>Ternary <span class="hlt">Phase-Separation</span> Investigation of Sol-Gel Derived Silica from Ethyl Silicate 40</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Wang, Shengnan; Wang, David K.; Smart, Simon; Diniz da Costa, João C.</p> <p>2015-01-01</p> <p>A ternary <span class="hlt">phase-separation</span> investigation of the ethyl silicate 40 (ES40) sol-gel process was conducted using ethanol and water as the solvent and hydrolysing agent, respectively. This oligomeric silica precursor underwent various degrees of <span class="hlt">phase</span> <span class="hlt">separation</span> behaviour in solution during the sol-gel reactions as a function of temperature and H2O/Si ratios. The solution composition within the immiscible region of the ES40 <span class="hlt">phase-separated</span> system shows that the hydrolysis and condensation reactions decreased with decreasing reaction temperature. A mesoporous structure was obtained at low temperature due to weak drying forces from slow solvent evaporation on one hand and formation of unreacted ES40 cages in the other, which reduced network shrinkage and produced larger pores. This was attributed to the concentration of the reactive sites around the <span class="hlt">phase-separated</span> interface, which enhanced the condensation and crosslinking. Contrary to dense silica structures obtained from sol-gel reactions in the miscible region, higher microporosity was produced via a <span class="hlt">phase-separated</span> sol-gel system by using high H2O/Si ratios. This tailoring process facilitated further condensation reactions and crosslinking of silica chains, which coupled with stiffening of the network, made it more resistant to compression and densification. PMID:26411484</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29096360','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29096360"><span>Detection of <span class="hlt">vapor-phase</span> organophosphate threats using wearable conformable integrated epidermal and textile wireless biosensor systems.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Mishra, Rupesh K; Martín, Aida; Nakagawa, Tatsuo; Barfidokht, Abbas; Lu, Xialong; Sempionatto, Juliane R; Lyu, Kay Mengjia; Karajic, Aleksandar; Musameh, Mustafa M; Kyratzis, Ilias L; Wang, Joseph</p> <p>2018-03-15</p> <p>Flexible epidermal tattoo and textile-based electrochemical biosensors have been developed for <span class="hlt">vapor-phase</span> detection of organophosphorus (OP) nerve agents. These new wearable sensors, based on stretchable organophosphorus hydrolase (OPH) enzyme electrodes, are coupled with a fully integrated conformal flexible electronic interface that offers rapid and selective square-wave voltammetric detection of OP <span class="hlt">vapor</span> threats and wireless data transmission to a mobile device. The epidermal tattoo and textile sensors display a good reproducibility (with RSD of 2.5% and 4.2%, respectively), along with good discrimination against potential interferences and linearity over the 90-300mg/L range, with a sensitivity of 10.7µA∙cm 3 ∙mg -1 (R 2 = 0.983) and detection limit of 12mg/L in terms of OP air density. Stress-enduring inks, used for printing the electrode transducers, ensure resilience against mechanical deformations associated with textile and skin-based on-body sensing operations. Theoretical simulations are used to estimate the OP air density over the sensor surface. These fully integrated wearable wireless tattoo and textile-based nerve-agent <span class="hlt">vapor</span> biosensor systems offer considerable promise for rapid warning regarding personal exposure to OP nerve-agent <span class="hlt">vapors</span> in variety of decentralized security applications. Copyright © 2017 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20050239577&hterms=ammonia&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dammonia','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20050239577&hterms=ammonia&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dammonia"><span>Performance Testing of the <span class="hlt">Vapor</span> <span class="hlt">Phase</span> Catalytic Ammonia Removal Engineering Development Unit</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Flynn, Michael; Tleimat, Maher; Nalette, Tim; Quinn, Gregory</p> <p>2005-01-01</p> <p>This paper describes the results of performance testing of the <span class="hlt">Vapor</span> <span class="hlt">Phase</span> Catalytic Ammonia Removal (VPCAR) technology. The VPCAR technology is currently being developed by NASA as a Mars transit vehicle water recycling system. NASA has recently completed-a grant-to develop a next generation VPCAR system. This grant concluded with the shipment of the final deliverable to NASA on 8/31/03. This paper presents the results of mass, power, volume, and acoustic measurements for the delivered system. Product water purity analysis for a Mars transit mission and a simulated planetary base wastewater ersatz are also provided.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1050875','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/1050875"><span>Method and apparatus for concentrating <span class="hlt">vapors</span> for analysis</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Grate, Jay W [West Richland, WA; Baldwin, David L [Kennewick, WA; Anheier, Jr., Norman C.</p> <p>2012-06-05</p> <p>A pre-concentration device and a method are disclosed for concentrating gaseous <span class="hlt">vapors</span> for analysis. <span class="hlt">Vapors</span> sorbed and concentrated within the bed of the pre-concentration device are thermally desorbed, achieving at least partial <span class="hlt">separation</span> of the <span class="hlt">vapor</span> mixtures. The pre-concentration device is suitable, e.g., for pre-concentration and sample injection, and provides greater resolution of peaks for <span class="hlt">vapors</span> within <span class="hlt">vapor</span> mixtures, yielding detection levels that are 10-10,000 times better than direct sampling and analysis systems. Features are particularly useful for continuous unattended monitoring applications. The invention finds application in conjunction with, e.g., analytical instruments where low detection limits for gaseous <span class="hlt">vapors</span> are desirable.</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li class="active"><span>24</span></li> <li><a href="#" onclick='return showDiv("page_25");'>25</a></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_24 --> <div id="page_25" class="hiddenDiv"> <div class="row"> <div class="col-sm-12"> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li class="active"><span>25</span></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div> </div> <div class="row"> <div class="col-sm-12"> <ol class="result-class" start="481"> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015APS..MARL45002K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015APS..MARL45002K"><span>Formation of ion clusters in the <span class="hlt">phase</span> <span class="hlt">separated</span> structures of neutral-charged polymer blends</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kwon, Ha-Kyung; Olvera de La Cruz, Monica</p> <p>2015-03-01</p> <p>Polyelectrolyte blends, consisting of at least one charged species, are promising candidate materials for fuel cell membranes, for their mechanical stability and high selectivity for proton conduction. The <span class="hlt">phase</span> behavior of the blends is important to understand, as this can significantly affect the performance of the device. The <span class="hlt">phase</span> behavior is controlled by χN, the Flory-Huggins parameter multiplied by the number of mers, as well as the electrostatic interactions between the charged backbone and the counterions. It has recently been shown that local ionic correlations, incorporated via liquid state (LS) theory, enhance <span class="hlt">phase</span> <span class="hlt">separation</span> of the blend, even in the absence of polymer interactions. In this study, we show <span class="hlt">phase</span> diagrams of neutral-charged polymer blends including ionic correlations via LS theory. In addition to enhanced <span class="hlt">phase</span> <span class="hlt">separation</span> at low χN, the blends show liquid-liquid <span class="hlt">phase</span> <span class="hlt">separation</span> at high electrostatic interaction strengths. Above the critical strength, the charged polymer <span class="hlt">phase</span> <span class="hlt">separates</span> into ion-rich and ion-poor regions, resulting in the formation of ion clusters within the charged polymer <span class="hlt">phase</span>. This can be shown by the appearance of multiple spinodal and critical points, indicating the coexistence of several charge <span class="hlt">separated</span> <span class="hlt">phases</span>. This work was performed under the following financial assistance award 70NANB14H012 from U.S. Department of Commerce, National Institute of Standards and Technology as part of the Center for Hierarchical Materials Design (CHiMaD).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015PhaTr..88..396S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015PhaTr..88..396S"><span>On the synthesis of AlPO4-21 molecular sieve by <span class="hlt">vapor</span> <span class="hlt">phase</span> transport method and its <span class="hlt">phase</span> transformation to AlPO4-15 molecular sieve</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shao, Hui; Chen, Jingjing; Chen, Xia; Leng, Yixin; Zhong, Jing</p> <p>2015-04-01</p> <p>An experimental design was applied to the synthesis of AlPO4-21 molecular sieve (AWO structure) by <span class="hlt">vapor</span> <span class="hlt">phase</span> transport (VPT) method, using tetramethylguanidine (TMG) as the template. In this study, the effects of crystallization time, crystallization temperature, phosphor content, template content and water content in the synthesis gel were investigated. The materials obtained were characterized by X-ray diffraction, scanning electron microscopy and fourier transform infrared spectroscopy (FT-IR). Microstructural analysis of the crystal growth in <span class="hlt">vapor</span> synthetic conditions revealed a revised crystal growth route from zeolite AlPO4-21 to AlPO4-15 in the presence of the TMG. Homogenous hexagonal prism AlPO4-21 crystals with size of 7 × 3 μm were synthesized at a lower temperature (120 °C), which were completely different from the typical tabular parallelogram crystallization microstructure of AlPO4-21 <span class="hlt">phase</span>. The crystals were transformed into AlPO4-21 <span class="hlt">phase</span> with higher crystallization temperature, longer crystallization time, higher P2O5/Al2O3 ratio and higher TMG/Al2O3 ratio.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26471552','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26471552"><span>Integrated method of thermosensitive triblock copolymer-salt aqueous two <span class="hlt">phase</span> extraction and dialysis membrane <span class="hlt">separation</span> for purification of lycium barbarum polysaccharide.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Yun; Hu, Xiaowei; Han, Juan; Ni, Liang; Tang, Xu; Hu, Yutao; Chen, Tong</p> <p>2016-03-01</p> <p>A polymer-salt aqueous two-<span class="hlt">phase</span> system (ATPS) consisting of thermosensitive copolymer ethylene-oxide-b-propylene-oxide-b-ethylene-oxide (EOPOEO) and NaH2PO4 was employed in deproteinization for lycium barbarum polysaccharide (LBP). The effects of salt type and concentration, EOPOEO concentration, amount of crude LBP solution and temperature were studied. In the primary extraction process, LBP was preferentially partitioned to the bottom (salt-rich) <span class="hlt">phase</span> with high recovery ratio of 96.3%, while 94.4% of impurity protein was removed to the top (EOPOEO-rich) <span class="hlt">phase</span>. Moreover, the majority of pigments could be discarded to top <span class="hlt">phase</span>. After <span class="hlt">phase-separation</span>, the LBP in the bottom <span class="hlt">phase</span> was further purified by dialysis membrane to remove salt and other small molecular impurities. The purity of LBP was enhanced to 64%. Additionally, the FT-IR spectrum was used to identify LBP. EOPOEO was recovered by a temperature-<span class="hlt">induced</span> <span class="hlt">separation</span>, and reused in a new ATPS. An ideal extraction and recycle result were achieved. Copyright © 2015 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009APS..MARX40006H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009APS..MARX40006H"><span>The effect of protein on <span class="hlt">phase</span> <span class="hlt">separation</span> in giant unilamellar lipid vesicles.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hutchison, J. B.; Weis, R. M.; Dinsmore, A. D.</p> <p>2009-03-01</p> <p>We explore the coarsening and out of plane curvature (budding) of domains in lipid bilayer vesicles composed of DOPC (unsaturated), PSM (saturated), and cholesterol. Green fluorescent protein (GFP) was added to the membrane in controlled amounts by binding to the Ni-chelating lipid, Ni-DOGS. Vesicles with diameters between 10 and 50 microns were prepared via a standard electroformation procedure. As a sample is lowered through temperature Tmix, a previously homogeneous vesicle <span class="hlt">phase</span> <span class="hlt">separates</span> into two fluid <span class="hlt">phases</span> with distinct compositions. <span class="hlt">Phase-separated</span> domains have a line tension (energy/length) at the boundary with the major <span class="hlt">phase</span> which competes with bending energy and lateral tension to determine the overall configuration of the vesicle. Domain budding and coarsening were observed and recorded using both bright field and fluorescence microscopy during temperature scans and with varying concentrations of GFP. The addition of a model protein into our system allows for a broader understanding of the effect of protein, which are ubiquitous in cell membranes, on <span class="hlt">phase</span> <span class="hlt">separation</span>, budding, and coarsening.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JCrGr.464..190G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JCrGr.464..190G"><span><span class="hlt">Phase</span> degradation in BxGa1-xN films grown at low temperature by metalorganic <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gunning, Brendan P.; Moseley, Michael W.; Koleske, Daniel D.; Allerman, Andrew A.; Lee, Stephen R.</p> <p>2017-04-01</p> <p>Using metalorganic <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy, a comprehensive study of BxGa1-xN growth on GaN and AlN templates is described. BGaN growth at high-temperature and high-pressure results in rough surfaces and poor boron incorporation efficiency, while growth at low-temperature and low-pressure (750-900 °C and 20 Torr) using nitrogen carrier gas results in improved surface morphology and boron incorporation up to 7.4% as determined by nuclear reaction analysis. However, further structural analysis by transmission electron microscopy and x-ray pole figures points to severe degradation of the high boron composition films, into a twinned cubic structure with a high density of stacking faults and little or no room temperature photoluminescence emission. Films with <1% triethylboron (TEB) flow show more intense, narrower x-ray diffraction peaks, near-band-edge photoluminescence emission at 362 nm, and primarily wurtzite-<span class="hlt">phase</span> structure in the x-ray pole figures. For films with >1% TEB flow, the crystal structure becomes dominated by the cubic <span class="hlt">phase</span>. Only when the TEB flow is zero (pure GaN), does the cubic <span class="hlt">phase</span> entirely disappear from the x-ray pole figure, suggesting that under these growth conditions even very low boron compositions lead to mixed crystalline <span class="hlt">phases</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22494734-enhanced-thermoelectric-properties-phase-separating-bismuth-selenium-telluride-thin-films-via-two-step-method','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22494734-enhanced-thermoelectric-properties-phase-separating-bismuth-selenium-telluride-thin-films-via-two-step-method"><span>Enhanced thermoelectric properties of <span class="hlt">phase-separating</span> bismuth selenium telluride thin films via a two-step method</span></a></p> <p><a target="_blank" href="http://www.osti.gov/search">DOE Office of Scientific and Technical Information (OSTI.GOV)</a></p> <p>Takashiri, Masayuki, E-mail: takashiri@tokai-u.jp; Kurita, Kensuke; Hagino, Harutoshi</p> <p>2015-08-14</p> <p>A two-step method that combines homogeneous electron beam (EB) irradiation and thermal annealing has been developed to enhance the thermoelectric properties of nanocrystalline bismuth selenium telluride thin films. The thin films, prepared using a flash evaporation method, were treated with EB irradiation in a N{sub 2} atmosphere at room temperature and an acceleration voltage of 0.17 MeV. Thermal annealing was performed under Ar/H{sub 2} (5%) at 300 °C for 60 min. X-ray diffraction was used to determine that compositional <span class="hlt">phase</span> <span class="hlt">separation</span> between bismuth telluride and bismuth selenium telluride developed in the thin films exposed to higher EB doses and thermal annealing. We proposemore » that the <span class="hlt">phase</span> <span class="hlt">separation</span> was <span class="hlt">induced</span> by fluctuations in the distribution of selenium atoms after EB irradiation, followed by the migration of selenium atoms to more stable sites during thermal annealing. As a result, thin film crystallinity improved and mobility was significantly enhanced. This indicates that the <span class="hlt">phase</span> <span class="hlt">separation</span> resulting from the two-step method enhanced, rather than disturbed, the electron transport. Both the electrical conductivity and the Seebeck coefficient were improved following the two-step method. Consequently, the power factor of thin films that underwent the two-step method was enhanced to 20 times (from 0.96 to 21.0 μW/(cm K{sup 2}) that of the thin films treated with EB irradiation alone.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70017627','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70017627"><span>Magmatic <span class="hlt">vapor</span> source for sulfur dioxide released during volcanic eruptions: Evidence from Mount Pinatubo</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Wallace, P.J.; Gerlach, T.M.</p> <p>1994-01-01</p> <p>Sulfur dioxide (SO2) released by the explosive eruption of Mount Pinatubo on 15 June 1991 had an impact on climate and stratospheric ozone. The total mass of SO2 released was much greater than the amount dissolved in the magma before the eruption, and thus an additional source for the excess SO2 is required. Infrared spectroscopic analyses of dissolved water and carbon dioxide in glass inclusions from quartz phenocrysts demonstrate that before eruption the magma contained a <span class="hlt">separate</span>, SO2-bearing <span class="hlt">vapor</span> <span class="hlt">phase</span>. Data for gas emissions from other volcanoes in subduction-related arcs suggest that preeruptive magmatic <span class="hlt">vapor</span> is a major source of the SO2 that is released during many volcanic eruptions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016Nanot..27y4004G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016Nanot..27y4004G"><span>An atom probe perspective on <span class="hlt">phase</span> <span class="hlt">separation</span> and precipitation in duplex stainless steels</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Guo, Wei; Garfinkel, David A.; Tucker, Julie D.; Haley, Daniel; Young, George A.; Poplawsky, Jonathan D.</p> <p>2016-06-01</p> <p>Three-dimensional chemical imaging of Fe-Cr alloys showing Fe-rich (α)/Cr-rich (α‧) <span class="hlt">phase</span> <span class="hlt">separation</span> is reported using atom probe tomography techniques. The extent of <span class="hlt">phase</span> <span class="hlt">separation</span>, i.e., amplitude and wavelength, has been quantitatively assessed using the Langer-Bar-on-Miller, proximity histogram, and autocorrelation function methods for two <span class="hlt">separate</span> Fe-Cr alloys, designated 2101 and 2205. Although the 2101 alloy possesses a larger wavelength and amplitude after annealing at 427 °C for 100-10 000 h, it exhibits a lower hardness than the 2205 alloy. In addition to this <span class="hlt">phase</span> <span class="hlt">separation</span>, ultra-fine Ni-Mn-Si-Cu-rich G-<span class="hlt">phase</span> precipitates form at the α/α‧ interfaces in both alloys. For the 2101 alloy, Cu clusters act to form a nucleus, around which a Ni-Mn-Si shell develops during the precipitation process. For the 2205 alloy, the Ni and Cu atoms enrich simultaneously and no core-shell chemical distribution was found. This segregation phenomenon may arise from the exact Ni/Cu ratio inside the ferrite. After annealing for 10 000 h, the number density of the G-<span class="hlt">phase</span> within the 2205 alloy was found to be roughly one order of magnitude higher than in the 2101 alloy. The G-<span class="hlt">phase</span> precipitates have an additional deleterious effect on the thermal embrittlement, as evaluated by the Ashby-Orowan equation, which explains the discrepancy between the hardness and the rate of <span class="hlt">phase</span> <span class="hlt">separation</span> with respect to annealing time (Gladman T 1999 Mater. Sci. Tech. Ser. 15 30-36). ).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1255672-atom-probe-perspective-phase-separation-precipitation-duplex-stainless-steels','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1255672-atom-probe-perspective-phase-separation-precipitation-duplex-stainless-steels"><span>An atom probe perspective on <span class="hlt">phase</span> <span class="hlt">separation</span> and precipitation in duplex stainless steels</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Garfinkel, David A.; Tucker, Julie D.; Haley, Daniel A.; ...</p> <p>2016-05-16</p> <p>Here, three-dimensional chemical imaging of Fe–Cr alloys showing Fe-rich (α)/Cr-rich (α') <span class="hlt">phase</span> <span class="hlt">separation</span> is reported using atom probe tomography techniques. The extent of <span class="hlt">phase</span> <span class="hlt">separation</span>, i.e., amplitude and wavelength, has been quantitatively assessed using the Langer-Bar-on-Miller, proximity histogram, and autocorrelation function methods for two <span class="hlt">separate</span> Fe–Cr alloys, designated 2101 and 2205. Although the 2101 alloy possesses a larger wavelength and amplitude after annealing at 427 °C for 100–10 000 h, it exhibits a lower hardness than the 2205 alloy. In addition to this <span class="hlt">phase</span> <span class="hlt">separation</span>, ultra-fine Ni–Mn–Si–Cu-rich G-<span class="hlt">phase</span> precipitates form at the α/α' interfaces in both alloys. For the 2101more » alloy, Cu clusters act to form a nucleus, around which a Ni–Mn–Si shell develops during the precipitation process. For the 2205 alloy, the Ni and Cu atoms enrich simultaneously and no core–shell chemical distribution was found. This segregation phenomenon may arise from the exact Ni/Cu ratio inside the ferrite. After annealing for 10 000 h, the number density of the G-<span class="hlt">phase</span> within the 2205 alloy was found to be roughly one order of magnitude higher than in the 2101 alloy. The G-<span class="hlt">phase</span> precipitates have an additional deleterious effect on the thermal embrittlement, as evaluated by the Ashby–Orowan equation, which explains the discrepancy between the hardness and the rate of <span class="hlt">phase</span> <span class="hlt">separation</span> with respect to annealing time (Gladman T 1999 Mater. Sci. Tech. Ser. 15 30–36).« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24835016','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24835016"><span>Lo/Ld <span class="hlt">phase</span> coexistence modulation <span class="hlt">induced</span> by GM1.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Puff, Nicolas; Watanabe, Chiho; Seigneuret, Michel; Angelova, Miglena I; Staneva, Galya</p> <p>2014-08-01</p> <p>Lipid rafts are assumed to undergo biologically important size-modulations from nanorafts to microrafts. Due to the complexity of cellular membranes, model systems become important tools, especially for the investigation of the factors affecting "raft-like" Lo domain size and the search for Lo nanodomains as precursors in Lo microdomain formation. Because lipid compositional change is the primary mechanism by which a cell can alter membrane <span class="hlt">phase</span> behavior, we studied the effect of the ganglioside GM1 concentration on the Lo/Ld lateral <span class="hlt">phase</span> <span class="hlt">separation</span> in PC/SM/Chol/GM1 bilayers. GM1 above 1mol % abolishes the formation of the micrometer-scale Lo domains observed in GUVs. However, the apparently homogeneous <span class="hlt">phase</span> observed in optical microscopy corresponds in fact, within a certain temperature range, to a Lo/Ld lateral <span class="hlt">phase</span> <span class="hlt">separation</span> taking place below the optical resolution. This nanoscale <span class="hlt">phase</span> <span class="hlt">separation</span> is revealed by fluorescence spectroscopy, including C12NBD-PC self-quenching and Laurdan GP measurements, and is supported by Gaussian spectral decomposition analysis. The temperature of formation of nanoscale Lo <span class="hlt">phase</span> domains over an Ld <span class="hlt">phase</span> is determined, and is shifted to higher values when the GM1 content increases. A "morphological" <span class="hlt">phase</span> diagram could be made, and it displays three regions corresponding respectively to Lo/Ld micrometric <span class="hlt">phase</span> <span class="hlt">separation</span>, Lo/Ld nanometric <span class="hlt">phase</span> <span class="hlt">separation</span>, and a homogeneous Ld <span class="hlt">phase</span>. We therefore show that a lipid only-based mechanism is able to control the existence and the sizes of <span class="hlt">phase-separated</span> membrane domains. GM1 could act on the line tension, "arresting" domain growth and thereby stabilizing Lo nanodomains. Copyright © 2014 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15600621','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15600621"><span>Coarsening mechanism of <span class="hlt">phase</span> <span class="hlt">separation</span> caused by a double temperature quench in an off-symmetric binary mixture.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sigehuzi, Tomoo; Tanaka, Hajime</p> <p>2004-11-01</p> <p>We study <span class="hlt">phase-separation</span> behavior of an off-symmetric fluid mixture <span class="hlt">induced</span> by a "double temperature quench." We first quench a system into the unstable region. After a large <span class="hlt">phase-separated</span> structure is formed, we again quench the system more deeply and follow the pattern-evolution process. The second quench makes the domains formed by the first quench unstable and leads to double <span class="hlt">phase</span> <span class="hlt">separation</span>; that is, small droplets are formed inside the large domains created by the first quench. The complex coarsening behavior of this hierarchic structure having two characteristic length scales is studied in detail by using the digital image analysis. We find three distinct time regimes in the time evolution of the structure factor of the system. In the first regime, small droplets coarsen with time inside large domains. There a large domain containing small droplets in it can be regarded as an isolated system. Later, however, the coarsening of small droplets stops when they start to interact via diffusion with the large domain containing them. Finally, small droplets disappear due to the Lifshitz-Slyozov mechanism. Thus the observed behavior can be explained by the crossover of the nature of a large domain from the isolated to the open system; this is a direct consequence of the existence of the two characteristic length scales.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APLM....5l0701D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APLM....5l0701D"><span>Research Update: A minimal region of squid reflectin for <span class="hlt">vapor-induced</span> light scattering</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dennis, Patrick B.; Singh, Kristi M.; Vasudev, Milana C.; Naik, Rajesh R.; Crookes-Goodson, Wendy J.</p> <p>2017-12-01</p> <p>Reflectins are a family of proteins found in the light manipulating cells of cephalopods. These proteins are made up of a series of conserved repeats that contain highly represented amino acids thought to be important for function. Previous studies demonstrated that recombinant reflectins cast into thin films produced structural colors that could be dynamically modulated via changing environmental conditions. In this study, we demonstrate light scattering from reflectin films following exposure to a series of water <span class="hlt">vapor</span> pulses. Analysis of film surface topography shows that the induction of light scatter is accompanied by self-assembly of reflectins into micro- and nanoscale features. Using a reductionist strategy, we determine which reflectin repeats and sub-repeats are necessary for these events following water <span class="hlt">vapor</span> pulsing. With this approach, we identify a singly represented, 23-amino acid region in reflectins as being sufficient to recapitulate the light scattering properties observed in thin films of the full-length protein. Finally, the aqueous stability of reflectin films is leveraged to show that pre-exposure to buffers of varying pH can modulate the ability of water <span class="hlt">vapor</span> pulses to <span class="hlt">induce</span> light scatter and protein self-assembly.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19930032221&hterms=chloride&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dchloride','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19930032221&hterms=chloride&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dchloride"><span>A semi-empirical model for the complete orientation dependence of the growth rate for <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy - Chloride VPE of GaAs</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Seidel-Salinas, L. K.; Jones, S. H.; Duva, J. M.</p> <p>1992-01-01</p> <p>A semi-empirical model has been developed to determine the complete crystallographic orientation dependence of the growth rate for <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy (VPE). Previous researchers have been able to determine this dependence for a limited range of orientations; however, our model yields relative growth rate information for any orientation. This model for diamond and zincblende structure materials is based on experimental growth rate data, gas <span class="hlt">phase</span> diffusion, and surface reactions. Data for GaAs chloride VPE is used to illustrate the model. The resulting growth rate polar diagrams are used in conjunction with Wulff constructions to simulate epitaxial layer shapes as grown on patterned substrates. In general, this model can be applied to a variety of materials and <span class="hlt">vapor</span> <span class="hlt">phase</span> epitaxy systems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=34051','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=34051"><span>Macroscopic <span class="hlt">phase</span> <span class="hlt">separation</span> in high-temperature superconductors</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Wen, Hai-Hu</p> <p>2000-01-01</p> <p>High-temperature superconductivity is recovered by introducing extra holes to the Cu-O planes, which initially are insulating with antiferromagnetism. In this paper I present data to show the macroscopic electronic <span class="hlt">phase</span> <span class="hlt">separation</span> that is caused by either mobile doping or electronic instability in the overdoped region. My results clearly demonstrate that the electronic inhomogeneity is probably a general feature of high-temperature superconductors. PMID:11027323</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19900035313&hterms=exchange+theory&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dexchange%2Btheory','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19900035313&hterms=exchange+theory&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dexchange%2Btheory"><span><span class="hlt">Phase</span> <span class="hlt">separation</span> in the t-J model. [in theory of high-temperature superconductors</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Emery, V. J.; Lin, H. Q.; Kivelson, S. A.</p> <p>1990-01-01</p> <p>A detailed understanding of the motion of 'holes' in an antiferromagnet is of fundamental importance for the theory of high-temperature superconductors. It is shown here that, for the t-J model, dilute holes in an antiferromagnet are unstable against <span class="hlt">phase</span> <span class="hlt">separation</span> into a hole-rich and a no-hole <span class="hlt">phase</span>. When the spin-exchange interaction J exceeds a critical value Jc, the hole-rich <span class="hlt">phase</span> has no electrons. It is proposed that, for J slightly less than Jc, the hole-rich <span class="hlt">phase</span> is a low-density superfluid of electron pairs. <span class="hlt">Phase</span> <span class="hlt">separation</span> in related models is briefly discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011PhDT........13O','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011PhDT........13O"><span>The rheology and <span class="hlt">phase</span> <span class="hlt">separation</span> kinetics of mixed-matrix membrane dopes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Olanrewaju, Kayode Olaseni</p> <p></p> <p>Mixed-matrix hollow fiber membranes are being developed to offer more efficient gas <span class="hlt">separations</span> applications than what the current technologies allow. Mixed-matrix membranes (MMMs) are membranes in which molecular sieves incorporated in a polymer matrix enhance <span class="hlt">separation</span> of gas mixtures based on the molecular size difference and/or adsorption properties of the component gases in the molecular sieve. The major challenges encountered in the efficient development of MMMs are associated with some of the paradigm shifts involved in their processing, as compared to pure polymer membranes. For instance, mixed-matrix hollow fiber membranes are prepared by a dry-wet jet spinning method. Efficient large scale processing of hollow fibers by this method requires knowledge of two key process variables: the rheology and kinetics of <span class="hlt">phase</span> <span class="hlt">separation</span> of the MMM dopes. Predicting the rheological properties of MMM dopes is not trivial; the presence of particles significantly affects neat polymer membrane dopes. Therefore, the need exists to characterize and develop predictive capabilities for the rheology of MMM dopes. Furthermore, the kinetics of <span class="hlt">phase</span> <span class="hlt">separation</span> of polymer solutions is not well understood. In the case of MMM dopes, the kinetics of <span class="hlt">phase</span> <span class="hlt">separation</span> are further complicated by the presence of porous particles in a polymer solution. Thus, studies on the <span class="hlt">phase</span> <span class="hlt">separation</span> kinetics of polymer solutions and suspensions of zeolite particles in polymer solutions are essential. Therefore, this research thesis aims to study the rheology and <span class="hlt">phase</span> <span class="hlt">separation</span> kinetics of mixed-matrix membrane dopes. In our research efforts to develop predictive models for the shear rheology of suspensions of zeolite particles in polymer solutions, it was found that MFI zeolite suspensions have relative viscosities that dramatically exceed the Krieger-Dougherty predictions for hard sphere suspensions. Our investigations showed that the major origin of this discrepancy is the selective</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23207470','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23207470"><span>Rationale for the development and the mechanism of action of endoscopic thermal <span class="hlt">vapor</span> ablation (Inter<span class="hlt">Vapor</span>) for the treatment of emphysema.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kesten, Steven; Anderson, Joseph C; Tuck, Stephanie A</p> <p>2012-07-01</p> <p>Emphysema remains a disabling disease despite current treatment. Novel approaches to the underlying physiological abnormalities responsible for symptom generation are warranted. A review of current hypotheses and preclinical and clinical data on the utility of endoscopic thermal <span class="hlt">vapor</span> ablation (Inter<span class="hlt">Vapor</span>) in the treatment of emphysema. In animal studies, thermal energy in the form of heated water <span class="hlt">vapor</span> both in healthy and in papain-<span class="hlt">induced</span> emphysema in dogs and sheep leads to an inflammatory response followed by healing with airway and parenchymal fibrosis. The fibrosis and associated distal atelectasis result in volume reduction. The amount of thermal energy delivered has been based on the amount of target tissue mass determined from a high-resolution computed tomogram. Early human studies indicated the feasibility of Inter<span class="hlt">Vapor</span> with 5 cal/g tissue; however, the dose appeared insufficient to <span class="hlt">induce</span> lobar volume reduction. A study using 10 cal/g to 1 upper lobe (n=44) <span class="hlt">induced</span> a mean of 46% lobar volume reduction at 12 months along with significant improvements in the physiology and health outcomes. Inter<span class="hlt">Vapor</span> <span class="hlt">induces</span> lung volume reduction in patients with emphysema. The mechanism of action is through a thermally <span class="hlt">induced</span> inflammatory response followed by healing with subsequent remodeling of tissue (fibrosis and distal atelectasis).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2003SPIE.5089.1088S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2003SPIE.5089.1088S"><span>Overview: MURI Center on spectroscopic and time domain detection of trace explosives in condensed and <span class="hlt">vapor</span> <span class="hlt">phases</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Spicer, James B.; Dagdigian, Paul; Osiander, Robert; Miragliotta, Joseph A.; Zhang, Xi-Cheng; Kersting, Roland; Crosley, David R.; Hanson, Ronald K.; Jeffries, Jay</p> <p>2003-09-01</p> <p>The research center established by Army Research Office under the Multidisciplinary University Research Initiative program pursues a multidisciplinary approach to investigate and advance the use of complementary analytical techniques for sensing of explosives and/or explosive-related compounds as they occur in the environment. The techniques being investigated include Terahertz (THz) imaging and spectroscopy, Laser-<span class="hlt">Induced</span> Breakdown Spectroscopy (LIBS), Cavity Ring Down Spectroscopy (CRDS) and Resonance Enhanced Multiphoton Ionization (REMPI). This suite of techniques encompasses a diversity of sensing approaches that can be applied to detection of explosives in condensed <span class="hlt">phases</span> such as adsorbed species in soil or can be used for <span class="hlt">vapor</span> <span class="hlt">phase</span> detection above the source. Some techniques allow for remote detection while others have highly specific and sensitive analysis capabilities. This program is addressing a range of fundamental, technical issues associated with trace detection of explosive related compounds using these techniques. For example, while both LIBS and THz can be used to carry-out remote analysis of condensed <span class="hlt">phase</span> analyte from a distance in excess several meters, the sensitivities of these techniques to surface adsorbed explosive-related compounds are not currently known. In current implementations, both CRDS and REMPI require sample collection techniques that have not been optimized for environmental applications. Early program elements will pursue the fundamental advances required for these techniques including signature identification for explosive-related compounds/interferents and trace analyte extraction. Later program tasks will explore simultaneous application of two or more techniques to assess the benefits of sensor fusion.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26849155','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26849155"><span>Therapeutic Antibody Engineering To Improve Viscosity and <span class="hlt">Phase</span> <span class="hlt">Separation</span> Guided by Crystal Structure.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Chow, Chi-Kin; Allan, Barrett W; Chai, Qing; Atwell, Shane; Lu, Jirong</p> <p>2016-03-07</p> <p>Antibodies at high concentrations often reveal unanticipated biophysical properties suboptimal for therapeutic development. The purpose of this work was to explore the use of point mutations based on crystal structure information to improve antibody physical properties such as viscosity and <span class="hlt">phase</span> <span class="hlt">separation</span> (LLPS) at high concentrations. An IgG4 monoclonal antibody (Mab4) that exhibited high viscosity and <span class="hlt">phase</span> <span class="hlt">separation</span> at high concentration was used as a model system. Guided by the crystal structure, four CDR point mutants were made to evaluate the role of hydrophobic and charge interactions on solution behavior. Surprisingly and unpredictably, two of the charge mutants, R33G and N35E, showed a reduction in viscosity and a lower propensity to form LLPS at high concentration compared to the wild-type (WT), while a third charge mutant S28K showed an increased propensity to form LLPS compared to the WT. A fourth mutant, F102H, had reduced hydrophobicity, but unchanged viscosity and <span class="hlt">phase</span> <span class="hlt">separation</span> behavior. We further evaluated the correlation of various biophysical measurements including second virial coefficient (A2), interaction parameter (kD), weight-average molecular weight (WAMW), and hydrodynamic diameters (DH), at relatively low protein concentration (4 to 15 mg/mL) to physical properties, such as viscosity and liquid-liquid <span class="hlt">phase</span> <span class="hlt">separation</span> (LLPS), at high concentration. Surprisingly, kD measured using dynamic light scattering (DLS) at low antibody concentration correlated better with viscosity and <span class="hlt">phase</span> <span class="hlt">separation</span> than did A2 for Mab4. Our results suggest that the high viscosity and <span class="hlt">phase</span> <span class="hlt">separation</span> observed at high concentration for Mab4 are mainly driven by charge and not hydrophobicity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4547336','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4547336"><span>Two-Point Microrheology of <span class="hlt">Phase-Separated</span> Domains in Lipid Bilayers</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Hormel, Tristan T.; Reyer, Matthew A.; Parthasarathy, Raghuveer</p> <p>2015-01-01</p> <p>Though the importance of membrane fluidity for cellular function has been well established for decades, methods for measuring lipid bilayer viscosity remain challenging to devise and implement. Recently, approaches based on characterizing the Brownian dynamics of individual tracers such as colloidal particles or lipid domains have provided insights into bilayer viscosity. For fluids in general, however, methods based on single-particle trajectories provide a limited view of hydrodynamic response. The technique of two-point microrheology, in which correlations between the Brownian dynamics of pairs of tracers report on the properties of the intervening medium, characterizes viscosity at length-scales that are larger than that of individual tracers and has less sensitivity to tracer-<span class="hlt">induced</span> distortions, but has never been applied to lipid membranes. We present, to our knowledge, the first two-point microrheological study of lipid bilayers, examining the correlated motion of domains in <span class="hlt">phase-separated</span> lipid vesicles and comparing one- and two-point results. We measure two-point correlation functions in excellent agreement with the forms predicted by two-dimensional hydrodynamic models, analysis of which reveals a viscosity intermediate between those of the two lipid <span class="hlt">phases</span>, indicative of global fluid properties rather than the viscosity of the local neighborhood of the tracer. PMID:26287625</p> </li> </ol> <div class="pull-right"> <ul class="pagination"> <li><a href="#" onclick='return showDiv("page_1");'>«</a></li> <li><a href="#" onclick='return showDiv("page_21");'>21</a></li> <li><a href="#" onclick='return showDiv("page_22");'>22</a></li> <li><a href="#" onclick='return showDiv("page_23");'>23</a></li> <li><a href="#" onclick='return showDiv("page_24");'>24</a></li> <li class="active"><span>25</span></li> <li><a href="#" onclick='return showDiv("page_25");'>»</a></li> </ul> </div> </div><!-- col-sm-12 --> </div><!-- row --> </div><!-- page_25 --> <div class="footer-extlink text-muted" style="margin-bottom:1rem; text-align:center;">Some links on this page may take you to non-federal websites. Their policies may differ from this site.</div> </div><!-- container --> <footer><a id="backToTop" href="#top"> </a><nav><a id="backToTop" href="#top"> </a><ul class="links"><a id="backToTop" href="#top"> </a><li><a id="backToTop" href="#top"></a><a href="/sitemap.html">Site Map</a></li> <li><a href="/members/index.html">Members Only</a></li> <li><a href="/website-policies.html">Website Policies</a></li> <li><a href="https://doe.responsibledisclosure.com/hc/en-us" target="_blank">Vulnerability Disclosure Program</a></li> <li><a href="/contact.html">Contact Us</a></li> </ul> <div class="small">Science.gov is maintained by the U.S. Department of Energy's <a href="https://www.osti.gov/" target="_blank">Office of Scientific and Technical Information</a>, in partnership with <a href="https://www.cendi.gov/" target="_blank">CENDI</a>.</div> </nav> </footer> <script type="text/javascript"><!-- // var lastDiv = ""; function showDiv(divName) { // hide last div if (lastDiv) { document.getElementById(lastDiv).className = "hiddenDiv"; } //if value of the box is not nothing and an object with that name exists, then change the class if (divName && document.getElementById(divName)) { document.getElementById(divName).className = "visibleDiv"; lastDiv = divName; } } //--> </script> <script> /** * Function that tracks a click on an outbound link in Google Analytics. * This function takes a valid URL string as an argument, and uses that URL string * as the event label. */ var trackOutboundLink = function(url,collectionCode) { try { h = window.open(url); setTimeout(function() { ga('send', 'event', 'topic-page-click-through', collectionCode, url); }, 1000); } catch(err){} }; </script> <!-- Google Analytics --> <script> (function(i,s,o,g,r,a,m){i['GoogleAnalyticsObject']=r;i[r]=i[r]||function(){ (i[r].q=i[r].q||[]).push(arguments)},i[r].l=1*new Date();a=s.createElement(o), m=s.getElementsByTagName(o)[0];a.async=1;a.src=g;m.parentNode.insertBefore(a,m) })(window,document,'script','//www.google-analytics.com/analytics.js','ga'); ga('create', 'UA-1122789-34', 'auto'); ga('send', 'pageview'); </script> <!-- End Google Analytics --> <script> showDiv('page_1') </script> </body> </html>