Sample records for charged solvent droplets

  1. Reactions of Microsolvated Organic Compounds at Ambient Surfaces: Droplet Velocity, Charge State, and Solvent Effects

    NASA Astrophysics Data System (ADS)

    Badu-Tawiah, Abraham K.; Campbell, Dahlia I.; Cooks, R. Graham

    2012-06-01

    The exposure of charged microdroplets containing organic ions to solid-phase reagents at ambient surfaces results in heterogeneous ion/surface reactions. The electrosprayed droplets were driven pneumatically in ambient air and then electrically directed onto a surface coated with reagent. Using this reactive soft landing approach, acid-catalyzed Girard condensation was achieved at an ambient surface by directing droplets containing Girard T ions onto a dry keto-steroid. The charged droplet/surface reaction was much more efficient than the corresponding bulk solution-phase reaction performed on the same scale. The increase in product yield is ascribed to solvent evaporation, which causes moderate pH values in the starting droplet to reach extreme values and increases reagent concentrations. Comparisons are made with an experiment in which the droplets were pneumatically accelerated onto the ambient surface (reactive desorption electrospray ionization, DESI). The same reaction products were observed but differences in spatial distribution were seen associated with the "splash" of the high velocity DESI droplets. In a third type of experiment, the reactions of charged droplets with vapor phase reagents were examined by allowing electrosprayed droplets containing a reagent to intercept the headspace vapor of an analyte. Deposition onto a collector surface and mass analysis showed that samples in the vapor phase were captured by the electrospray droplets, and that instantaneous derivatization of the captured sample is possible in the open air. The systems examined under this condition included the derivatization of cortisone vapor with Girard T and that of 4-phenylpyridine N-oxide and 2-phenylacetophenone vapors with ethanolamine.

  2. Release of Native-like Gaseous Proteins from Electrospray Droplets via the Charged Residue Mechanism: Insights from Molecular Dynamics Simulations.

    PubMed

    McAllister, Robert G; Metwally, Haidy; Sun, Yu; Konermann, Lars

    2015-10-07

    The mechanism whereby gaseous protein ions are released from charged solvent droplets during electrospray ionization (ESI) remains a matter of debate. Also, it is unclear to what extent electrosprayed proteins retain their solution structure. Molecular dynamics (MD) simulations offer insights into the temporal evolution of protein systems. Surprisingly, there have been no all-atom simulations of the protein ESI process to date. The current work closes this gap by investigating the behavior of protein-containing aqueous nanodroplets that carry excess positive charge. We focus on "native ESI", where proteins initially adopt their biologically active solution structures. ESI proceeds while the protein remains entrapped within the droplet. Protein release into the gas phase occurs upon solvent evaporation to dryness. Droplet shrinkage is accompanied by ejection of charge carriers (Na(+) for the conditions chosen here), keeping the droplet at ∼85% of the Rayleigh limit throughout its life cycle. Any remaining charge carriers bind to the protein as the final solvent molecules evaporate. The outcome of these events is largely independent of the initial protein charge and the mode of charge carrier binding. ESI charge states and collision cross sections of the MD structures agree with experimental data. Our results confirm the Rayleigh/charged residue model (CRM). Field emission of excess Na(+) plays an ancillary role by governing the net charge of the shrinking droplet. Models that envision protein ejection from the droplet are not supported. Most nascent CRM ions retain native-like conformations. For unfolded proteins ESI likely proceeds along routes that are different from the native state mechanism explored here.

  3. Electrostatic field and charge distribution in small charged dielectric droplets

    NASA Astrophysics Data System (ADS)

    Storozhev, V. B.

    2004-08-01

    The charge distribution in small dielectric droplets is calculated on the basis of continuum medium approximation. There are considered charged liquid spherical droplets of methanol in the range of nanometer sizes. The problem is solved by the following way. We find the free energy of some ion in dielectric droplet, which is a function of distribution of other ions in the droplet. The probability of location of the ion in some element of volume in the droplet is a function of its free energy in this element of volume. The same approach can be applied to other ions in the droplet. The obtained charge distribution differs considerably from the surface distribution. The curve of the charge distribution in the droplet as a function of radius has maximum near the surface. Relative concentration of charges in the vicinity of the center of the droplet does not equal to zero, and it is the higher, the less is the total charge of the droplet. According to the estimates the model is applicable if the droplet radius is larger than 10 nm.

  4. Interaction between electrically charged droplets in microgravity

    NASA Astrophysics Data System (ADS)

    Brandenbourger, Martin; Caps, Herve; Hardouin, Jerome; Vitry, Youen; Boigelot, Bernard; Dorbolo, Stephane; Grasp Team; Beams Collaboration

    2015-11-01

    The past ten years, electrically charged droplets have been studied tremendously for their applications in industry (electrospray, electrowetting,...). However, charged droplets are also present in nature. Indeed, it has been shown that the droplets falling from thunderclouds possess an excess of electric charges. Moreover, some research groups try to use the electrical interaction between drops in order to control the coalescence between cloud droplets and control rain generation. The common way to study this kind of system is to make hypothesis on the interaction between two charged drops. Then, these hypothesis are extended to a system of thousands of charged droplets. Thanks to microgravity conditions, we were able to study the interaction between two electrically charged droplets. In practice, the charged droplets were propelled one in front of the other at low speed (less than 1 m/s). The droplets trajectory is studied for various charges and volumes. The repulsion between two charged drops is correctly fitted by a simple Coulomb repulsion law. In the case of attractive interactions, we discuss the collisions observed as a function of the droplets speed, volume and electric charges. Thanks to FNRS for financial support.

  5. Ion-induced nucleation in solution: promotion of solute nucleation in charged levitated droplets.

    PubMed

    Draper, Neil D; Bakhoum, Samuel F; Haddrell, Allen E; Agnes, George R

    2007-09-19

    We have investigated the nucleation and growth of sodium chloride in both single quiescent charged droplets and charged droplet populations that were levitated in an electrodynamic levitation trap (EDLT). In both cases, the magnitude of a droplet's net excess charge (ions(DNEC)) influenced NaCl nucleation and growth, albeit in different capacities. We have termed the phenomenon ion-induced nucleation in solution. For single quiescent levitated droplets, an increase in ions(DNEC) resulted in a significant promotion of NaCl nucleation, as determined by the number of crystals observed. For levitated droplet populations, a change in NaCl crystal habit, from regular cubic shapes to dome-shaped dendrites, was observed once a surface charge density threshold of -9 x 10(-4) e.nm(-2) was surpassed. Although promotion of NaCl nucleation was observed for droplet population experiments, this can be attributed in part to the increased rate of solvent evaporation observed for levitated droplet populations having a high net charge. Promotion of nucleation was also observed for two organic acids, 2,4,6-trihydroxyacetophenone monohydrate (THAP) and alpha-cyano-4-hydroxycinnamic acid (CHCA). These results are of direct relevance to processes that occur in both soft-ionization techniques for mass spectrometry and to a variety of industrial processes. To this end, we have demonstrated the use of ion-induced nucleation in solution to form ammonium nitrate particles from levitated droplets to be used in in vitro toxicology studies of ambient particle types.

  6. Geoengineering with Charged Droplets

    NASA Astrophysics Data System (ADS)

    Gokturk, H.

    2011-12-01

    Water molecules in a droplet are held together by intermolecular forces generated by hydrogen bonding which has a bonding energy of only about 0.2 eV. One can create a more rugged droplet by using an ion as a condensation nucleus. In that case, water molecules are held together by the interaction between the ion and the dipole moments of the water molecules surrounding the ion, in addition to any hydrogen bonding. In this research, properties of such charged droplets were investigated using first principle quantum mechanical calculations. A molecule which exhibits positive electron affinity is a good candidate to serve as the ionic condensation nucleus, because addition of an electron to such a molecule creates an energetically more stable state than the neutral molecule. A good example is the oxygen molecule (O2) where energy of O2 negative (O2-) ion is lower than that of the neutral O2 by about 0.5 eV. Examples of other molecules which have positive electron affinity include ozone (O3), nitrogen dioxide (NO2) and sulfur oxides (SOx, x=1-3). Atomic models used in the calculations consisted of a negative ion of one of the molecules mentioned above surrounded by water molecules. Calculations were performed using the DFT method with B3LYP hybrid functional and Pople type basis sets with polarization and diffuse functions. Energy of interaction between O2- ion and the water molecule was found to be ~0.7 eV. This energy is an order of magnitude greater than the thermal energy of even the highest temperatures encountered in the atmosphere. Once created, charged rugged droplets can survive in hot and dry climates where they can be utilized to create humidity and precipitation. The ion which serves as the nucleus of the droplet can attract not only water molecules but also other dipolar gases in the atmosphere. Such dipolar gases include industrial pollutants, for example nitrogen dioxide (NO2) or sulfur dioxide (SO2). Energy of interaction between O2- ion and pollutant

  7. Self-arraying of charged levitating droplets.

    PubMed

    Kauffmann, Paul; Nussbaumer, Jérémie; Masse, Alain; Jeandey, Christian; Grateau, Henri; Pham, Pascale; Reyne, Gilbert; Haguet, Vincent

    2011-06-01

    Diamagnetic levitation of water droplets in air is a promising phenomenon to achieve contactless manipulation of chemical or biochemical samples. This noncontact handling technique prevents contaminations of samples as well as provides measurements of interaction forces between levitating reactors. Under a nonuniform magnetic field, diamagnetic bodies such as water droplets experience a repulsive force which may lead to diamagnetic levitation of a single or few micro-objects. The levitation of several repulsively charged picoliter droplets was successfully performed in a ~1 mm(2) adjustable flat magnetic well provided by a centimeter-sized cylindrical permanent magnet structure. Each droplet position results from the balance between the centripetal diamagnetic force and the repulsive Coulombian forces. Levitating water droplets self-organize into satellite patterns or thin clouds, according to their charge and size. Small triangular lattices of identical droplets reproduce magneto-Wigner crystals. Repulsive forces and inner charges can be measured in the piconewton and the femtocoulomb ranges, respectively. Evolution of interaction forces is accurately followed up over time during droplet evaporation.

  8. Freezing, fragmentation, and charge separation in sonic sprayed water droplets

    NASA Astrophysics Data System (ADS)

    Zilch, Lloyd W.; Maze, Joshua T.; Smith, John W.; Jarrold, Martin F.

    2009-06-01

    Water droplets are generated by sonic spray, transferred into vacuum through a capillary interface, and then passed through two image charge detectors separated by a drift region. The image charge detectors measure the charge and velocity of each droplet. For around 1% of the droplets, the charge changes significantly between the detectors. In some cases it increases, in others it decreases, and for some droplets the charge changes polarity. We attribute the charge changing behavior to fragmentation caused by freezing. Simulations indicate that the time required for a droplet to cool and freeze in vacuum depends on its size, and that droplets with radii of 1-2 [mu]m have the right size to freeze between the two detectors. These sizes correspond to the smaller end of the distribution present in the experiment. When the charge on a droplet increases or changes polarity, fragmentation must be accompanied by charge separation where fragments carry away opposite charges. In some cases, two fission fragments were observed in the second charge detector. We show examples where the droplet breaks apart to give fragments of the same charge and opposite charges. The fragmentation and charge changing behavior found here is consistent with what has been found in the freezing of larger suspended and supported droplets.

  9. New mechanisms of macroion-induced disintegration of charged droplets

    NASA Astrophysics Data System (ADS)

    Consta, Styliani; Oh, Myong In; Malevanets, Anatoly

    2016-10-01

    Molecular modeling has revealed that the presence of charged macromolecules (macroions) in liquid droplets dramatically changes the pathways of droplet fission. These mechanisms are not captured by the traditional theories such as ion-evaporation and charge-residue models. We review the general mechanisms by which macroions emerge from droplets and the factors that determine the droplet fission. These mechanisms include counter-intuitive ;star; droplet formations and extrusion of linear macroions from droplets. These findings may play a direct role in determining macromolecule charge states in electrospray mass spectrometry experiments.

  10. Charging and Release Mechanisms of Flexible Macromolecules in Droplets

    NASA Astrophysics Data System (ADS)

    Oh, Myong In; Consta, Styliani

    2017-08-01

    We study systematically the charging and release mechanisms of a flexible macromolecule, modeled by poly(ethylene glycol) (PEG), in a droplet by using molecular dynamics simulations. We compare how PEG is solvated and charged by sodium Na+ ions in a droplet of water (H2O), acetonitrile (MeCN), and their mixtures. Initially, we examine the location and the conformation of the macromolecule in a droplet bearing no net charge. It is revealed that the presence of charge carriers do not affect the location of PEG in aqueous and MeCN droplets compared with that in the neutral droplets, but the location of the macromolecule and the droplet size do affect the PEG conformation. PEG is charged on the surface of a sodiated aqueous droplet that is found close to the Rayleigh limit. Its charging is coupled to the extrusion mechanism, where PEG segments leave the droplet once they coordinate a Na+ ion or in a correlated motion with Na+ ions. In contrast, as PEG resides in the interior of a MeCN droplet, it is sodiated inside the droplet. The compact macro-ion transitions through partially unwound states to an extended conformation, a process occurring during the final stage of desolvation and in the presence of only a handful of MeCN molecules. For charged H2O/MeCN droplets, the sodiation of PEG is determined by the H2O component, reflecting its slower evaporation and preference over MeCN for solvating Na+ ions. We use the simulation data to construct an analytical model that suggests that the droplet surface electric field may play a role in the macro-ion-droplet interactions that lead to the extrusion of the macro-ion. This study provides the first evidence of the effect of the surface electric field by using atomistic simulations. [Figure not available: see fulltext.

  11. Droplet Charging Effects in the Space Environment

    DTIC Science & Technology

    2010-06-16

    in GEO during periods of high geomagnetic or solar activity. An experiment was conducted to assess the charging of silcon- oil droplets due to...experiment was conducted to assess the charging of silcon- oil droplets due to photoemission. The photoemission yield in the 120-200nm wavelength range was...For the application of interest in this study, a liquid droplet stream of low- vapor-pressure, silicon-based oil is being proposed as a potential

  12. Charged Slurry Droplet Research

    DTIC Science & Technology

    1989-02-20

    IEEE/IAS annual meeting, Denver, CO, Sept. 28 - Oct. 3, 1986, p.1434. Accepted for publication IEEE Transactions on Industry Applications. 6. Lord...34Analysis of the Description of Evaporating Charged Droplets, IEEE Transactions on Industry Applications, IA-19, 771, 1983. 9. H.M.A. Elghazaly, G.S.P. Castle...34Analysis of the Instability of Evaporating Charged Liquid Drops", IEEE Transactions on Industry Applications, IA-22, 892, 1986. 10. H.M.A

  13. Surface area generation and droplet size control in solvent extraction systems utilizing high intensity electric fields

    DOEpatents

    Scott, Timothy C.; Wham, Robert M.

    1988-01-01

    A method and system for solvent extraction where droplets are shattered by a high intensity electric field. These shattered droplets form a plurality of smaller droplets which have a greater combined surface area than the original droplet. Dispersion, coalescence and phase separation are accomplished in one vessel through the use of the single pulsing high intensity electric field. Electric field conditions are chosen so that simultaneous dispersion and coalescence are taking place in the emulsion formed in the electric field. The electric field creates a large amount of interfacial surface area for solvent extraction when the droplet is disintegrated and is capable of controlling droplet size and thus droplet stability. These operations take place in the presence of a counter current flow of the continuous phase.

  14. Non-coalescence of oppositely charged droplets in pH-sensitive emulsions

    PubMed Central

    Liu, Tingting; Seiffert, Sebastian; Thiele, Julian; Abate, Adam R.; Weitz, David A.; Richtering, Walter

    2012-01-01

    Like charges stabilize emulsions, whereas opposite charges break emulsions. This is the fundamental principle for many industrial and practical processes. Using micrometer-sized pH-sensitive polymeric hydrogel particles as emulsion stabilizers, we prepare emulsions that consist of oppositely charged droplets, which do not coalesce. We observe noncoalescence of oppositely charged droplets in bulk emulsification as well as in microfluidic devices, where oppositely charged droplets are forced to collide within channel junctions. The results demonstrate that electrostatic interactions between droplets do not determine their stability and reveal the unique pH-dependent properties of emulsions stabilized by soft microgel particles. The noncoalescence can be switched to coalescence by neutralizing the microgels, and the emulsion can be broken on demand. This unusual feature of the microgel-stabilized emulsions offers fascinating opportunities for future applications of these systems. PMID:22203968

  15. Droplet charging regimes for ultrasonic atomization of a liquid electrolyte in an external electric field.

    PubMed

    Forbes, Thomas P; Degertekin, F Levent; Fedorov, Andrei G

    2011-01-01

    Distinct regimes of droplet charging, determined by the dominant charge transport process, are identified for an ultrasonic droplet ejector using electrohydrodynamic computational simulations, a fundamental scale analysis, and experimental measurements. The regimes of droplet charging are determined by the relative magnitudes of the dimensionless Strouhal and electric Reynolds numbers, which are a function of the process (pressure forcing), advection, and charge relaxation time scales for charge transport. Optimal (net maximum) droplet charging has been identified to exist for conditions in which the electric Reynolds number is of the order of the inverse Strouhal number, i.e., the charge relaxation time is on the order of the pressure forcing (droplet formation) time scale. The conditions necessary for optimal droplet charging have been identified as a function of the dimensionless Debye number (i.e., liquid conductivity), external electric field (magnitude and duration), and atomization drive signal (frequency and amplitude). The specific regime of droplet charging also determines the functional relationship between droplet charge and charging electric field strength. The commonly expected linear relationship between droplet charge and external electric field strength is only found when either the inverse of the Strouhal number is less than the electric Reynolds number, i.e., the charge relaxation is slower than both the advection and external pressure forcing, or in the electrostatic limit, i.e., when charge relaxation is much faster than all other processes. The analysis provides a basic understanding of the dominant physics of droplet charging with implications to many important applications, such as electrospray mass spectrometry, ink jet printing, and drop-on-demand manufacturing.

  16. Droplet charging regimes for ultrasonic atomization of a liquid electrolyte in an external electric field

    PubMed Central

    Forbes, Thomas P.; Degertekin, F. Levent; Fedorov, Andrei G.

    2011-01-01

    Distinct regimes of droplet charging, determined by the dominant charge transport process, are identified for an ultrasonic droplet ejector using electrohydrodynamic computational simulations, a fundamental scale analysis, and experimental measurements. The regimes of droplet charging are determined by the relative magnitudes of the dimensionless Strouhal and electric Reynolds numbers, which are a function of the process (pressure forcing), advection, and charge relaxation time scales for charge transport. Optimal (net maximum) droplet charging has been identified to exist for conditions in which the electric Reynolds number is of the order of the inverse Strouhal number, i.e., the charge relaxation time is on the order of the pressure forcing (droplet formation) time scale. The conditions necessary for optimal droplet charging have been identified as a function of the dimensionless Debye number (i.e., liquid conductivity), external electric field (magnitude and duration), and atomization drive signal (frequency and amplitude). The specific regime of droplet charging also determines the functional relationship between droplet charge and charging electric field strength. The commonly expected linear relationship between droplet charge and external electric field strength is only found when either the inverse of the Strouhal number is less than the electric Reynolds number, i.e., the charge relaxation is slower than both the advection and external pressure forcing, or in the electrostatic limit, i.e., when charge relaxation is much faster than all other processes. The analysis provides a basic understanding of the dominant physics of droplet charging with implications to many important applications, such as electrospray mass spectrometry, ink jet printing, and drop-on-demand manufacturing. PMID:21301636

  17. Investigation of the charging characteristics of micrometer sized droplets based on parallel plate capacitor model.

    PubMed

    Zhang, Yanzhen; Liu, Yonghong; Wang, Xiaolong; Shen, Yang; Ji, Renjie; Cai, Baoping

    2013-02-05

    The charging characteristics of micrometer sized aqueous droplets have attracted more and more attentions due to the development of the microfluidics technology since the electrophoretic motion of a charged droplet can be used as the droplet actuation method. This work proposed a novel method of investigating the charging characteristics of micrometer sized aqueous droplets based on parallel plate capacitor model. With this method, the effects of the electric field strength, electrolyte concentration, and ion species on the charging characteristics of the aqueous droplets was investigated. Experimental results showed that the charging characteristics of micrometer sized droplets can be investigated by this method.

  18. The Orbit of Water Droplets around Charged Rod

    ERIC Educational Resources Information Center

    Ferstl, Andrew; Burns, Andrew

    2013-01-01

    The motion of charges around a centrally charged object is often compared to gravitational orbits (such as satellites around planets). Recently, a video taken by astronaut Don Pettit onboard the International Space Station shows water droplets orbiting a charged knitting needle. Here we attempt to model this motion and estimate the charges on the…

  19. Tuning aggregation of microemulsion droplets and silica nanoparticles using solvent mixtures.

    PubMed

    Salabat, Alireza; Eastoe, Julian; Mutch, Kevin J; Tabor, Rico F

    2008-02-15

    The effect of solvent on stability of water-in-oil microemulsions has been studied with AOT (sodium bis(2-ethylhexyl)sulfosuccinate) and different solvent mixtures of n-heptane, toluene and dodecane. Dynamic light scattering DLS was used to monitor the apparent diffusion coefficient D(A) and effective microemulsion droplet diameter on changing composition of the solvent. Interdroplet attractive interactions, as indicated by variations in D(A), can be tuned by formulation of appropriate solvent mixtures using heptane, toluene, and dodecane. In extreme cases, solvent mixtures can be used to induce phase transitions in the microemulsions. Aggregation and stability of model AOT-stabilized silica nanoparticles in different solvents were also investigated to explore further these solvent effects. For both systems the state of aggregation can be correlated with the effective molecular volume of the solvent V(mol)(eff) mixture.

  20. Charge Assisted Laser Desorption/Ionization Mass Spectrometry of Droplets

    PubMed Central

    Jorabchi, Kaveh; Westphall, Michael S.; Smith, Lloyd M.

    2008-01-01

    We propose and evaluate a new mechanism to account for analyte ion signal enhancement in ultraviolet-laser desorption mass spectrometry of droplets in the presence of corona ions. Our new insights are based on timing control of corona ion production, laser desorption, and peptide ion extraction achieved by a novel pulsed corona apparatus. We demonstrate that droplet charging rather than gas-phase ion-neutral reactions is the major contributor to analyte ion generation from an electrically isolated droplet. Implications of the new mechanism, termed charge assisted laser desorption/ionization (CALDI), are discussed and contrasted to those of the laser desorption atmospheric pressure chemical ionization method (LD-APCI). It is also demonstrated that analyte ion generation in CALDI occurs with external electric fields about one order of magnitude lower than those needed for atmospheric pressure matrix assisted laser desorption/ionization or electrospray ionization of droplets. PMID:18387311

  1. Electrostatic Model Applied to ISS Charged Water Droplet Experiment

    NASA Technical Reports Server (NTRS)

    Stevenson, Daan; Schaub, Hanspeter; Pettit, Donald R.

    2015-01-01

    The electrostatic force can be used to create novel relative motion between charged bodies if it can be isolated from the stronger gravitational and dissipative forces. Recently, Coulomb orbital motion was demonstrated on the International Space Station by releasing charged water droplets in the vicinity of a charged knitting needle. In this investigation, the Multi-Sphere Method, an electrostatic model developed to study active spacecraft position control by Coulomb charging, is used to simulate the complex orbital motion of the droplets. When atmospheric drag is introduced, the simulated motion closely mimics that seen in the video footage of the experiment. The electrostatic force's inverse dependency on separation distance near the center of the needle lends itself to analytic predictions of the radial motion.

  2. Charge Effects on the Efflorescence in Single Levitated Droplets.

    PubMed

    Hermann, Gunter; Zhang, Yan; Wassermann, Bernhard; Fischer, Henry; Quennet, Marcel; Rühl, Eckart

    2017-09-14

    The influence of electrical excess charges on the crystallization from supersaturated aqueous sodium chloride solutions is reported. This is accomplished by efflorescence studies on single levitated microdroplets using optical and electrodynamic levitation. Specifically, a strong increase in efflorescence humidity is observed as a function of the droplet's negative excess charge, ranging up to -2.1 pC, with a distinct threshold behavior, increasing the relative efflorescence humidity, at which spontaneous nucleation occurs, from 44% for the neutral microparticle to 60%. These findings are interpreted by using molecular dynamics simulations for determining plausible structural patterns located near the particle surface that could serve as suitable precursors for the formation of critical clusters overcoming the nucleation barrier. These results, facilitating heterogeneous nucleation in the case of negatively charged microparticles, are compared to recent work on charge-induced nucleation of neat supercooled water, where a distinctly different nucleation behavior as a function of droplet charge has been observed.

  3. Visualization of the evolution of charged droplet formation and jet transition in electrostatic atomization

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Huo, Yuanping, E-mail: huoyuanping@gmail.com; Wang, Junfeng, E-mail: wangjunfeng@ujs.edu.cn; Zuo, Ziwen

    2015-11-15

    A detailed experimental study on the evolution of charged droplet formation and jet transition from a capillary is reported. By means of high-speed microscopy, special attention has been paid to the dynamics of the liquid thread and satellite droplets in the dripping mode, and a method for calculating the surface charge on the satellite droplet is proposed. Jet transition behavior based on the electric Bond number has been visualized, droplet sizes and velocities are measured to obtain the ejection characteristic of the spray plume, and the charge and hydrodynamic relaxation are linked to give explanations for ejection dynamics with differentmore » properties. The results show that the relative length is very sensitive to the hydrodynamic relaxation time. The magnitude of the electric field strength dominates the behavior of coalescence and noncoalescence, with the charge relationship between the satellite droplet and the main droplet being clear for every noncoalescence movement. Ejection mode transitions mainly depend on the magnitude of the electric Bond number, and the meniscus dynamics is determined by the ratio of the charge relaxation time to the hydrodynamic relaxation time.« less

  4. Numerical Investigation of Two-Phase Flows With Charged Droplets in Electrostatic Field

    NASA Technical Reports Server (NTRS)

    Kim, Sang-Wook

    1996-01-01

    A numerical method to solve two-phase turbulent flows with charged droplets in an electrostatic field is presented. The ensemble-averaged Navier-Stokes equations and the electrostatic potential equation are solved using a finite volume method. The transitional turbulence field is described using multiple-time-scale turbulence equations. The equations of motion of droplets are solved using a Lagrangian particle tracking scheme, and the inter-phase momentum exchange is described by the Particle-In-Cell scheme. The electrostatic force caused by an applied electrical potential is calculated using the electrostatic field obtained by solving a Laplacian equation and the force exerted by charged droplets is calculated using the Coulombic force equation. The method is applied to solve electro-hydrodynamic sprays. The calculated droplet velocity distributions for droplet dispersions occurring in a stagnant surrounding are in good agreement with the measured data. For droplet dispersions occurring in a two-phase flow, the droplet trajectories are influenced by aerodynamic forces, the Coulombic force, and the applied electrostatic potential field.

  5. Microfluidic production of single micrometer-sized hydrogel beads utilizing droplet dissolution in a polar solvent

    PubMed Central

    Sugaya, Sari; Yamada, Masumi; Hori, Ayaka; Seki, Minoru

    2013-01-01

    In this study, a microfluidic process is proposed for preparing monodisperse micrometer-sized hydrogel beads. This process utilizes non-equilibrium aqueous droplets formed in a polar organic solvent. The water-in-oil droplets of the hydrogel precursor rapidly shrunk owing to the dissolution of water molecules into the continuous phase. The shrunken and condensed droplets were then gelled, resulting in the formation of hydrogel microbeads with sizes significantly smaller than the initial droplet size. This study employed methyl acetate as the polar organic solvent, which can dissolve water at 8%. Two types of monodisperse hydrogel beads—Ca-alginate and chitosan—with sizes of 6–10 μm (coefficient of variation < 6%) were successfully produced. In addition, we obtained hydrogel beads with non-spherical morphologies by controlling the degree of droplet shrinkage at the time of gelation and by adjusting the concentration of the gelation agent. Furthermore, the encapsulation and concentration of DNA molecules within the hydrogel beads were demonstrated. The process presented in this study has great potential to produce small and highly concentrated hydrogel beads that are difficult to obtain by using conventional microfluidic processes. PMID:24396529

  6. Microfluidic production of single micrometer-sized hydrogel beads utilizing droplet dissolution in a polar solvent.

    PubMed

    Sugaya, Sari; Yamada, Masumi; Hori, Ayaka; Seki, Minoru

    2013-01-01

    In this study, a microfluidic process is proposed for preparing monodisperse micrometer-sized hydrogel beads. This process utilizes non-equilibrium aqueous droplets formed in a polar organic solvent. The water-in-oil droplets of the hydrogel precursor rapidly shrunk owing to the dissolution of water molecules into the continuous phase. The shrunken and condensed droplets were then gelled, resulting in the formation of hydrogel microbeads with sizes significantly smaller than the initial droplet size. This study employed methyl acetate as the polar organic solvent, which can dissolve water at 8%. Two types of monodisperse hydrogel beads-Ca-alginate and chitosan-with sizes of 6-10 μm (coefficient of variation < 6%) were successfully produced. In addition, we obtained hydrogel beads with non-spherical morphologies by controlling the degree of droplet shrinkage at the time of gelation and by adjusting the concentration of the gelation agent. Furthermore, the encapsulation and concentration of DNA molecules within the hydrogel beads were demonstrated. The process presented in this study has great potential to produce small and highly concentrated hydrogel beads that are difficult to obtain by using conventional microfluidic processes.

  7. Enhancing physicochemical properties of emulsions by heteroaggregation of oppositely charged lactoferrin coated lutein droplets and whey protein isolate coated DHA droplets.

    PubMed

    Li, Xin; Wang, Xu; Xu, Duoxia; Cao, Yanping; Wang, Shaojia; Wang, Bei; Sun, Baoguo; Yuan, Fang; Gao, Yanxiang

    2018-01-15

    The formation and physicochemical stability of mixed functional components (lutein & DHA) emulsions through heteroaggregation were studied. It was formed by controlled heteroaggregation of oppositely charged lutein and DHA droplets coated by cationic lactoferrin (LF) and anionic whey protein isolate (WPI), respectively. Heteroaggregation was induced by mixing the oppositely charged LF-lutein and WPI-DHA emulsions together at pH 6.0. Droplet size, zeta-potential, transmission-physical stability, microrheological behavior and microstructure of the heteroaggregates formed were measured as a function of LF-lutein to WPI-DHA droplet ratio. Lutein degradation and DHA oxidation by measurement of lipid hydroperoxides and thiobarbituric acid reactive substances were determined. Upon mixing the two types of bioactive compounds droplets together, it was found that the largest aggregates and highest physical stability occurred at a droplet ratio of 40% LF-lutein droplets to 60% WPI-DHA droplets. Heteroaggregates formation altered the microrheological properties of the mixed emulsions mainly by the special network structure of the droplets. When LF-coated lutein droplets ratios were more than 30% and less than 60%, the mixed emulsions exhibited distinct decreases in the Mean Square Displacement, which indicated that their limited scope of Brownian motion and stable structure. Mixed emulsions with LF-lutein/WPI-DHA droplets ratio of 4:6 exhibited Macroscopic Viscosity Index with 13 times and Elasticity Index with 3 times of magnitudes higher than the individual emulsions from which they were prepared. Compared with the WPI-DHA emulsion or LF-lutein emulsion, the oxidative stability of the heteroaggregate of LF-lutein/WPI-DHA emulsions was improved. Heteroaggregates formed by oppositely charged bioactive compounds droplets may be useful for creating specific food structures that lead to desirable physicochemical properties, such as microrheological property, physical and chemical

  8. Solvent Dependence of Lateral Charge Transfer in a Porphyrin Monolayer

    DOE PAGES

    Brennan, Bradley J.; Regan, Kevin P.; Durrell, Alec C.; ...

    2016-12-19

    Lateral charge transport in a redox)active monolayer can be utilized for solar energy harvesting. We chose the porphyrin system to study the influence of the solvent on lateral hole hopping, which plays a crucial role in the charge)transfer kinetics. We also examined the influence of water, acetonitrile, and propylene carbonate as solvents. Hole)hopping lifetimes varied by nearly three orders of magnitude among solvents, ranging from 3 ns in water to 2800 ns in propylene carbonate, and increased nonlinearly as a function of added acetonitrile in aqueous solvent mixtures. Our results elucidate the important roles of solvation, molecular packing dynamics, andmore » lateral charge)transfer mechanisms that have implications for all dye)sensitized photoelectrochemical device designs.« less

  9. Charged Water Droplets can Melt Metallic Electrodes

    NASA Astrophysics Data System (ADS)

    Elton, Eric; Rosenberg, Ethan; Ristenpart, William

    2016-11-01

    A water drop, when immersed in an insulating fluid, acquires charge when it contacts an energized electrode. Provided the electric field is strong enough, the drop will move away to the opposite electrode, acquire the opposite charge, and repeat the process, effectively 'bouncing' back and forth between the electrodes. A key implicit assumption, dating back to Maxwell, has been that the electrode remains unaltered by the charging process. Here we demonstrate that the electrode is physically deformed during each charge transfer event with an individual water droplet or other conducting object. We used optical, electron, and atomic force microscopy to characterize a variety of different metallic electrodes before and after drops were electrically bounced on them. Although the electrodes appear unchanged to the naked eye, the microscopy reveals that each charge transfer event yielded a crater approximately 1 micron wide and 50 nm deep, with the exact dimensions proportional to the applied field strength. We present evidence that the craters are formed by localized melting of the electrodes via Joule heating in the metal and concurrent dielectric breakdown of the surrounding fluid, suggesting that the electrode locally achieves temperatures exceeding 3400°C. Present address: Dept. Materials Sci. Engineering, MIT.

  10. Fluorescence and Nonlinear Optical Properties of Alizarin Red S in Solvents and Droplet.

    PubMed

    Sangsefedi, Seyed Ahmad; Sharifi, Soheil; Rezaion, Hadi Rastegar Moghaddam; Azarpour, Afshin

    2018-05-28

    The enhancement of the nonlinear properties of materials is an interesting topic since it has many applications in optical devices and medicines. The Z-scan technique was used to study the values of the two-photon absorption (β), second-order molecular hyperpolarizability (γ R ), third-order susceptibility (χ R ), and nonlinear refractive index (n 2 ) of Alizarin Red S in different media using a continuous-wave diode-pump laser radiation at 532 nm. For Alizarin Red S in a droplet, the β, n 2 , χ R, and γ R were estimated at the order of 10 -7  cm 2 /W and 10 -12  cm/W, 10 -3  m 3  W -1  s -1 and 10 -24  m 6  W -1  s -1 , respectively. The results indicated that the values of β and n 2 reduced, whereas the values of χ R and γ R were enhanced when the solvent was changed from droplet to water, DMF, and dimethyl sulfoxide due to the change in the solvent's dielectric constant (ε). Moreover, the values of β were enhanced by an increase in the concentration of the surfactant in the aqueous solution. The absorption spectra of Alizarin Red S in the aqueous solution was observed at 428 nm, and a few red shifts in the absorption spectra were observed with a reduction in the dielectric constant of the medium. The same effect was observed in the absorption spectra of Alizarin Red S in the droplet when the bulk dielectric constant reduced. The dielectric constant can affect the fluorescence spectra of Alizarin Red S when the solution is changed from water to dimethyl sulfoxide. The dipole moments of Alizarin Red S in the different media were studied using the quantum perturbation theory.

  11. Microfluidic solvent extraction of poly(vinyl alcohol) droplets: effect of polymer structure on particle and capsule formation.

    PubMed

    Sharratt, W N; Brooker, A; Robles, E S J; Cabral, J T

    2018-04-26

    We investigate the formation of poly(vinyl alcohol) microparticles by the selective extraction of aqueous polymer solution droplets, templated by microfluidics and subsequently immersed in a non-solvent bath. The role of polymer molecular mass (18-105 kg mol-1), degree of hydrolysis (88-99%) and thus solubility, and initial solution concentration (0.01-10% w/w) are quantified. Monodisperse droplets with radii ranging from 50 to 500 μm were produced at a flow-focusing junction with carrier phase hexadecane and extracted into ethyl acetate. Solvent exchange and extraction result in droplet shrinkage, demixing, coarsening and phase-inversion, yielding polymer microparticles with well-defined dimensions and internal microstructure. Polymer concentration, varied from below the overlap concentration c* to above the concentrated crossover c**, as estimated by viscosity measurements, was found to have the largest impact on the final particle size and extraction timescale, while polymer mass and hydrolysis played a secondary role. These results are consistent with the observation that the average polymer concentration upon solidification greatly exceeds c**, and that the internal microparticle porosity is largely unchanged. However, reducing the initial polymer concentration to well below c* (approximately 100×) and increasing droplet size yields thin-walled (100's of nm) capsules which controllably crumple upon extraction. The symmetry of the process can be readily broken by imposing extraction conditions at an impermeable surface, yielding large, buckled, cavity morphologies. Based on these results, we establish robust design criteria for polymer capsules and particles, demonstrated here for poly(vinyl alcohol), with well-defined shape, dimensions and internal microstructure.

  12. Influence of calcium-induced droplet heteroaggregation on the physicochemical properties of oppositely charged lactoferrin coated lutein droplets and whey protein isolate-coated DHA droplets.

    PubMed

    Li, Xin; Wang, Xu; Xu, Duoxia; Cao, Yanping; Wang, Shaojia; Wang, Bei; Wang, Chengtao; Sun, Baoguo

    2017-08-01

    The influence of calcium-induced droplet heteroaggregation on the formation and physicochemical stability of mixed lutein and DHA emulsions was studied. Heteroaggregation was induced by mixing oppositely charged lactoferrin (LF)-coated lutein and whey protein isolate (WPI)-coated DHA emulsions with different CaCl 2 concentrations at pH 6.0. The droplet size, zeta-potential, transmission-physical stability and microstructure behavior (CLSM and Cryo-SEM) of single-protein emulsions and mixed emulsions were measured as a function of different CaCl 2 concentrations. Lutein degradation and DHA oxidation by measurement of lipid hydroperoxides and thiobarbituric acid reactive substances were determined during storage. The physical stability of the mixed emulsions could be modulated by controlling CaCl 2 concentrations. Microstructure behavior indicated that a mixed emulsion with 30 mM CaCl 2 promoted more droplets to form a special three-dimensional network and microcluster structures. The chemical stability of the mixed lutein and DHA emulsions was obviously enhanced by the addition of 30 mM CaCl 2 . The decreased surface areas of the DHA and lutein droplets and the physical barrier of the network of heteroaggregates against transition metals and free radicals could mainly explain the improvement in chemical stability. Calcium-induced droplet aggregation may be useful for creating specific food structures that lead to desirable physicochemical properties of multiple functional components.

  13. Flexible particle flow-focusing in microchannel driven by droplet-directed induced-charge electroosmosis.

    PubMed

    Ren, Yukun; Liu, Xianyu; Liu, Weiyu; Tao, Ye; Jia, Yankai; Hou, Likai; Li, Wenying; Jiang, Hongyuan

    2018-02-01

    We report herein a novel microfluidic particle concentrator that utilizes constriction microchannels to enhance the flow-focusing performance of induced-charge electroosmosis (ICEO), where viscous hemi-spherical oil droplets are embedded within the mainchannel to form deformable converging-diverging constriction structures. The constriction region between symmetric oil droplets partially coated on the electrode strips can improve the focusing performance by inducing a granular wake flow area at the diverging channel, which makes almost all of the scattered sample particles trapped within a narrow stream on the floating electrode. Another asymmetric droplet pair arranged near the outlets can further direct the trajectory of focused particle stream to one specified outlet port depending on the symmetry breaking in the shape of opposing phase interfaces. By fully exploiting rectification properties of induced-charge electrokinetic phenomena at immiscible water/oil interfaces of tunable geometry, the expected function of continuous and switchable flow-focusing is demonstrated by preconcentrating both inorganic silica particles and biological yeast cells. Physical mechanisms responsible for particle focusing and locus deflection in the droplet-assisted concentrentor are analyzed in detail, and simulation results are in good accordance with experimental observations. Our work provides new routes to construct flexible electrokinetic framework for preprocessing on-chip biological samples before performing subsequent analysis. © 2017 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.

  14. WITHDRAWN: Fragmentation of charged aqueous nanodroplets

    NASA Astrophysics Data System (ADS)

    Ichiki, Kengo

    2005-11-01

    The whole evaporating process of charged aqueous nanodroplets is studied by systematic molecular dynamics simulations until most of the solvent molecules are evaporated. % The solvent evaporation makes the droplet smaller and smaller, and at a certain point the repulsive force among ions causes an instability, where typically single ion and 10 to 20 water molecules are disintegrated from the main droplet. % This ion fragmentation occurs around 70 to 80% of the charge predicted by the Rayleigh theory [Lord Rayleigh, Phil. Mag. 14, 184 (1882)]. % The numerical results are summarized in the function R(z) which is the fragmentation radius at the charge z. From the fitting by the power law Rz^β, we find that at lower temperature T=350 and 370 K the result is close to the Rayleigh theory β= 2/3, while at higher temperature T=400 and 450 K it is like β= 1/2. % Another fitting on R(z) by the extended ion evaporation mechanism [M. Gamero-Castaño and J. Fern'andez de la Mora, Anal. Chim. Acta 406, 67 (2000)] works well for both cases. % The final state of the evaporation process is typically a single ion with several water molecules. If we put an alanine dipeptide in zwitterionic form at the beginning, two charges remain in some cases.

  15. Structure and Dynamics of Solvent Landscapes in Charge-Transfer Reactions

    NASA Astrophysics Data System (ADS)

    Leite, Vitor B. Pereira

    The dynamics of solvent polarization plays a major role in the control of charge transfer reactions. The success of Marcus theory describing the solvent influence via a single collective quadratic polarization coordinate has been remarkable. Onuchic and Wolynes have recently proposed (J. Chem Phys 98 (3) 2218, 1993) a simple model demonstrating how a many-dimensional-complex model composed by several dipole moments (representing solvent molecules or polar groups in proteins) can be reduced under the appropriate limits into the Marcus Model. This work presents a dynamical study of the same model, which is characterized by two parameters, an average dipole-dipole interaction as a term associated with the potential energy landscape roughness. It is shown why the effective potential, obtained using a thermodynamic approach, is appropriate for the dynamics of the system. At high temperatures, the system exhibits effective diffusive one-dimensional dynamics, where the Born-Marcus limit is recovered. At low temperatures, a glassy phase appears with a slow non-self-averaging dynamics. At intermediate temperatures, the concept of equivalent diffusion paths and polarization dependence effects are discussed. This approach is extended to treat more realistic solvent models. Real solvents are discussed in terms of simple parameters described above, and an analysis of how different regimes affect the rate of charge transfer is presented. Finally, these ideas are correlated to analogous problems in other areas.

  16. Solvent Additive-Assisted Anisotropic Assembly and Enhanced Charge Transport of π-Conjugated Polymer Thin Films.

    PubMed

    Jeong, Jae Won; Jo, Gyounglyul; Choi, Solip; Kim, Yoong Ahm; Yoon, Hyeonseok; Ryu, Sang-Wan; Jung, Jaehan; Chang, Mincheol

    2018-05-30

    Charge transport in π-conjugated polymer films involves π-π interactions within or between polymer chains. Here, we demonstrate a facile solution processing strategy that provides enhanced intra- and interchain π-π interactions of the resultant polymer films using a good solvent additive with low volatility. These increased interactions result in enhanced charge transport properties. The effect of the good solvent additive on the intra- and intermolecular interactions, morphologies, and charge transport properties of poly(3-hexylthiophene) (P3HT) films is systematically investigated. We found that the good solvent additive facilitates the self-assembly of P3HT chains into crystalline fibrillar nanostructures by extending the solvent drying time during thin-film formation. As compared to the prior approach using a nonsolvent additive with low volatility, the solvent blend system containing a good solvent additive results in enhanced charge transport in P3HT organic field-effect transistor (OFET) devices [from ca. 1.7 × 10 -2 to ca. 8.2 × 10 -2 cm 2 V -1 s -1 for dichlorobenzene (DCB) versus 4.4 × 10 -2 cm 2 V -1 s -1 for acetonitrile]. The mobility appears to be maximized over a broad spectrum of additive concentrations (1-7 vol %), indicative of a wide processing window. Detailed analysis results regarding the charge injection and transport characteristics of the OFET devices reveal that a high-boiling-point solvent additive decreases both the contact resistance ( R c ) and channel resistance ( R ch ), contributing to the mobility enhancement of the devices. Finally, the platform presented here is proven to be applicable to alternative good solvent additives with low volatility, such as chlorobenzene (CB) and trichlorobenzene (TCB). Specifically, the mobility enhancement of the resultant P3HT films increases in the order CB (bp 131 °C) < DCB (bp 180 °C) < TCB (bp 214 °C), suggesting that solvent additives with higher boiling points provide resultant

  17. The effect of solvent relaxation time constants on free energy gap law for ultrafast charge recombination following photoinduced charge separation.

    PubMed

    Mikhailova, Valentina A; Malykhin, Roman E; Ivanov, Anatoly I

    2018-05-16

    To elucidate the regularities inherent in the kinetics of ultrafast charge recombination following photoinduced charge separation in donor-acceptor dyads in solutions, the simulations of the kinetics have been performed within the stochastic multichannel point-transition model. Increasing the solvent relaxation time scales has been shown to strongly vary the dependence of the charge recombination rate constant on the free energy gap. In slow relaxing solvents the non-equilibrium charge recombination occurring in parallel with solvent relaxation is very effective so that the charge recombination terminates at the non-equilibrium stage. This results in a crucial difference between the free energy gap laws for the ultrafast charge recombination and the thermal charge transfer. For the thermal reactions the well-known Marcus bell-shaped dependence of the rate constant on the free energy gap is realized while for the ultrafast charge recombination only a descending branch is predicted in the whole area of the free energy gap exceeding 0.2 eV. From the available experimental data on the population kinetics of the second and first excited states for a series of Zn-porphyrin-imide dyads in toluene and tetrahydrofuran solutions, an effective rate constant of the charge recombination into the first excited state has been calculated. The obtained rate constant being very high is nearly invariable in the area of the charge recombination free energy gap from 0.2 to 0.6 eV that supports the theoretical prediction.

  18. "Inverted" Solvent Effect on Charge Transfer in the Excited State.

    PubMed

    Nau; Pischel

    1999-10-04

    Faster in cyclohexane than in acetonitrile is the fluorescence quenching of the azoalkane 2,3-diazabicyclo[2.2.2]oct-2-ene (DBO) by amines and sulfides. Although this photoreaction is induced by charge transfer (CT; see picture) and exciplexes are formed, the increase in the dipole moment of the exciplex is not large enough to offset the solvent stabilization of the excited reactants, and an "inverted" solvent effect results.

  19. Probing chemical transformation in picolitre volume aerosol droplets

    NASA Astrophysics Data System (ADS)

    Miloserdov, Anatolij; Day, Calum P. F.; Rosario, Gabriela L.; Horrocks, Benjamin R.; Carruthers, Antonia E.

    2017-08-01

    We have demonstrated chemical transformation in single microscopic-sized aerosol droplets localised in optical tweezers. Droplets in situ are measured during chemical transformation processes of solvent exchange and solute transformation through an ion exchange reaction. Solvent exchange between deionised water and heavy water in aerosol droplets is monitored through observation of the OH and OD Raman stretches. A change in solute chemistry of aerosol is achieved through droplet coalescence events between calcium chloride and sodium carbonate to promote ion exchange. The transformation forming meta-stable and stable states of CaCO3 is observed and analysed using Gaussian peak decomposition to reveal polymorphs.

  20. Hidden topological constellations and polyvalent charges in chiral nematic droplets

    PubMed Central

    Posnjak, Gregor; Čopar, Simon; Muševič, Igor

    2017-01-01

    Topology has an increasingly important role in the physics of condensed matter, quantum systems, material science, photonics and biology, with spectacular realizations of topological concepts in liquid crystals. Here we report on long-lived hidden topological states in thermally quenched, chiral nematic droplets, formed from string-like, triangular and polyhedral constellations of monovalent and polyvalent singular point defects. These topological defects are regularly packed into a spherical liquid volume and stabilized by the elastic energy barrier due to the helical structure and confinement of the liquid crystal in the micro-sphere. We observe, for the first time, topological three-dimensional point defects of the quantized hedgehog charge q=−2, −3. These higher-charge defects act as ideal polyvalent artificial atoms, binding the defects into polyhedral constellations representing topological molecules. PMID:28220770

  1. Hidden topological constellations and polyvalent charges in chiral nematic droplets

    NASA Astrophysics Data System (ADS)

    Posnjak, Gregor; Čopar, Simon; Muševič, Igor

    2017-02-01

    Topology has an increasingly important role in the physics of condensed matter, quantum systems, material science, photonics and biology, with spectacular realizations of topological concepts in liquid crystals. Here we report on long-lived hidden topological states in thermally quenched, chiral nematic droplets, formed from string-like, triangular and polyhedral constellations of monovalent and polyvalent singular point defects. These topological defects are regularly packed into a spherical liquid volume and stabilized by the elastic energy barrier due to the helical structure and confinement of the liquid crystal in the micro-sphere. We observe, for the first time, topological three-dimensional point defects of the quantized hedgehog charge q=-2, -3. These higher-charge defects act as ideal polyvalent artificial atoms, binding the defects into polyhedral constellations representing topological molecules.

  2. Mass spectrometry of acoustically levitated droplets.

    PubMed

    Westphall, Michael S; Jorabchi, Kaveh; Smith, Lloyd M

    2008-08-01

    Containerless sample handling techniques such as acoustic levitation offer potential advantages for mass spectrometry, by eliminating surfaces where undesired adsorption/desorption processes can occur. In addition, they provide a unique opportunity to study fundamental aspects of the ionization process as well as phenomena occurring at the air-droplet interface. Realizing these advantages is contingent, however, upon being able to effectively interface levitated droplets with a mass spectrometer, a challenging task that is addressed in this report. We have employed a newly developed charge and matrix-assisted laser desorption/ionization (CALDI) technique to obtain mass spectra from a 5-microL acoustically levitated droplet containing peptides and an ionic matrix. A four-ring electrostatic lens is used in conjunction with a corona needle to produce bursts of corona ions and to direct those ions toward the droplet, resulting in droplet charging. Analyte ions are produced from the droplet by a 337-nm laser pulse and detected by an atmospheric sampling mass spectrometer. The ion generation and extraction cycle is repeated at 20 Hz, the maximum operating frequency of the laser employed. It is shown in delayed ion extraction experiments that both positive and negative ions are produced, behavior similar to that observed for atmospheric pressure matrix-assisted laser absorption/ionization. No ion signal is observed in the absence of droplet charging. It is likely, although not yet proven, that the role of the droplet charging is to increase the strength of the electric field at the surface of the droplet, reducing charge recombination after ion desorption.

  3. Charge-Induced Saffman-Taylor Instabilities in Toroidal Droplets

    NASA Astrophysics Data System (ADS)

    Fragkopoulos, A. A.; Aizenman, A.; Fernández-Nieves, A.

    2017-06-01

    We show that charged toroidal droplets can develop fingerlike structures as they expand due to Saffman-Taylor instabilities. While these are commonly observed in quasi-two-dimensional geometries when a fluid displaces another fluid of higher viscosity, we show that the toroidal confinement breaks the symmetry of the problem, effectively making it quasi-two-dimensional and enabling the instability to develop in this three-dimensional situation. We control the expansion speed of the torus with the imposed electric stress and show that fingers are observed provided the characteristic time scale associated with this instability is smaller than the characteristic time scale associated with Rayleigh-Plateau break-up. We confirm our interpretation of the results by showing that the number of fingers is consistent with expectations from linear stability analysis in radial Hele-Shaw cells.

  4. Mass Spectrometry of Acoustically Levitated Droplets

    PubMed Central

    Westphall, Michael S.; Jorabchi, Kaveh; Smith, Lloyd M.

    2008-01-01

    Containerless sample handling techniques such as acoustic levitation offer potential advantages for mass spectrometry, by eliminating surfaces where undesired adsorption/desorption processes can occur. In addition, they provide a unique opportunity to study fundamental aspects of the ionization process as well as phenomena occurring at the air–droplet interface. Realizing these advantages is contingent, however, upon being able to effectively interface levitated droplets with a mass spectrometer, a challenging task that is addressed in this report. We have employed a newly developed charge and matrix-assisted laser desorption/ionization (CALDI) technique to obtain mass spectra from a 5-μL acoustically levitated droplet containing peptides and an ionic matrix. A four-ring electrostatic lens is used in conjunction with a corona needle to produce bursts of corona ions and to direct those ions toward the droplet, resulting in droplet charging. Analyte ions are produced from the droplet by a 337-nm laser pulse and detected by an atmospheric sampling mass spectrometer. The ion generation and extraction cycle is repeated at 20 Hz, the maximum operating frequency of the laser employed. It is shown in delayed ion extraction experiments that both positive and negative ions are produced, behavior similar to that observed for atmospheric pressure matrix-assisted laser absorption/ionization. No ion signal is observed in the absence of droplet charging. It is likely, although not yet proven, that the role of the droplet charging is to increase the strength of the electric field at the surface of the droplet, reducing chargere combination after ion desorption. PMID:18582090

  5. Influence of solvent species on the charge-discharge characteristics of a natural graphite electrode

    NASA Astrophysics Data System (ADS)

    Fujimoto, Masahisa; Shoji, Yoshihiro; Kida, Yoshinori; Ohshita, Ryuji; Nohma, Toshiyuki; Nishio, Koji

    The charge-discharge characteristics of a natural graphite electrode are examined in a mixed solvent composed of ethylene carbonate (EC) and propylene carbonate (PC). The characteristics are influenced largely by the solvent species. Natural graphite electrode displays good charge-discharge characteristics in an electrolyte containing EC with a high volume fraction. In an electrolyte containing PC, however, the electrode cannot be charged and the solvent is decomposed. X-ray photoelectron spectroscopy is used to obtain information about the surface of natural graphite. A thin LiF layer, the decomposition product of lithium hexafluorophosphate (LiPF 6), is formed on the surface of the natural graphite charged to 0.5 V (vs. Li/Li +) in an electrolyte containing a high volume fraction of EC. On the other hand, LiF and a carbonate compound are formed in the bulk and on the surface of natural graphite when the volume fraction of PC is high. These results suggest that the thin LiF layer, which is produced at a potential higher than 0.5 V (vs. Li/Li +) on the surface of natural graphite, enables the lithium ions to intercalate into the natural graphite without further decomposition of the electrolyte.

  6. Electric field makes Leidenfrost droplets take a leap.

    PubMed

    Wildeman, Sander; Sun, Chao

    2016-12-06

    Leidenfrost droplets, i.e. droplets whose mobility is ensured by a thin vapor film between the droplet and a hot plate, are exposed to an external electric field. We find that in a strong vertical electric field the droplet can start to bounce progressively higher, defying gravitational attraction. From the droplet's trajectory we infer the temporal evolution of the amount of charge on the droplet. This reveals that the charge starts high and then decreases in steps as the droplet slowly evaporates. After each discharge event the charge is in a fixed proportion to the droplet's surface area. We show that this behavior can be accurately modeled by treating the droplet as a conducting sphere that occasionally makes electrical contact with the hot plate, at intervals dictated by an electro-capillary instability in the vapor film. An analysis of the kinetic and potential energies of the bouncing droplet reveals that, while the overall motion is damped, the droplet occasionally experiences a sudden boost, keeping its energy close to the value for which the free fall trajectory and droplet oscillation are in sync. This helps the droplet to escape from the hot surface when finally the electrical surface forces overtake gravity.

  7. Droplet-Based Production of Liposomes

    NASA Technical Reports Server (NTRS)

    Ackley, Donald E.; Forster, Anita

    2009-01-01

    A process for making monodisperse liposomes having lipid bilayer membranes involves fewer, simpler process steps than do related prior methods. First, a microfluidic, cross junction droplet generator is used to produce vesicles comprising aqueous solution droplets contained in single layer lipid membranes. The vesicles are collected in a lipid-solvent mix that is at most partially soluble in water and is less dense than is water. A layer of water is dispensed on top of the solvent. By virtue of the difference in densities, the water sinks to the bottom and the solvent floats to the top. The vesicles, which have almost the same density as that of water, become exchanged into the water instead of floating to the top. As there are excess lipids in the solvent solution, in order for the vesicles to remain in the water, the addition of a second lipid layer to each vesicle is energetically favored. The resulting lipid bilayers present the hydrophilic ends of the lipid molecules to both the inner and outer membrane surfaces. If lipids of a second kind are dissolved in the solvent in sufficient excess before use, then asymmetric liposomes may be formed.

  8. Electro-coalescence of particle-coated droplets

    NASA Astrophysics Data System (ADS)

    Shum, Anderson Ho Cheung

    Droplets in air or in an immiscible liquid phase are used widely in applications ranging from personal hygiene products to drug delivery. The stability of the droplets are highly linked to their utility, and thus have been systematically studied. To enhance the stability of the droplets, particles are often added to the droplets. In this talk, I will discuss how the particle layer at droplet interfaces responds to electrical charging of the droplets. The electrical forces can distort the droplet shape, which is opposed by the layer of particles adsorbed. A balance of the electrical and interfacial effects provides a quantitative indicator of the droplet instability. The coalescence of droplets in both air and liquid induced by electrically charging, which we call ``electro-coalescence'', will be introduced, with its potential application in devising a digital millifluidic platform. We thank the Research Grants Council of Hong Kong (No. HKU 719813E, 17304514 and 17306315 and C6004-14G) from the and National Natural Science Foundation of China (No. 21476189/B060201 and 91434202).

  9. Scaling laws for first and second generation electrospray droplets

    NASA Astrophysics Data System (ADS)

    Basaran, Osman; Sambath, Krishnaraj; Anthony, Christopher; Collins, Robert; Wagoner, Brayden; Harris, Michael

    2017-11-01

    When uncharged liquid interfaces of pendant and free drops (hereafter referred to as parent drops) or liquid films are subject to a sufficiently strong electric field, they can emit thin fluid jets from conical tip structures that form at their surfaces. The disintegration of such jets into a spray consisting of charged droplets (hereafter referred to as daughter droplets) is common to electrospray ionization mass spectrometry, printing and coating processes, and raindrops in thunderclouds. We use simulation to determine the sizes and charges of these first-generation daughter droplets which are shown to be Coulombically stable and charged below the Rayleigh limit of stability. Once these daughter droplets shrink in size due to evaporation, they in turn reach their respective Rayleigh limits and explode by emitting yet even smaller second-generation daughter droplets from their conical tips. Once again, we use simulation and theory to deduce scaling laws for the sizes and charges of these second-generation droplets. A comparison is also provided for scaling laws pertaining to different generations of daughter droplets.

  10. Buckling in armored droplets.

    PubMed

    Sicard, François; Striolo, Alberto

    2017-06-29

    The buckling mechanism in droplets stabilized by solid particles (armored droplets) is tackled at a mesoscopic level using dissipative particle dynamics simulations. We consider one spherical water droplet in a decane solvent coated with nanoparticle monolayers of two different types: Janus (particles whose surface shows two regions with different wetting properties) and homogeneous. The chosen particles yield comparable initial three-phase contact angles, selected to maximize the adsorption energy at the interface. We study the interplay between the evolution of droplet shape, layering of the particles, and their distribution at the interface when the volume of the droplets is reduced. We show that Janus particles affect strongly the shape of the droplet with the formation of a crater-like depression. This evolution is actively controlled by a close-packed particle monolayer at the curved interface. In contrast, homogeneous particles follow passively the volume reduction of the droplet, whose shape does not deviate too much from spherical, even when a nanoparticle monolayer/bilayer transition is detected at the interface. We discuss how these buckled armored droplets might be of relevance in various applications including potential drug delivery systems and biomimetic design of functional surfaces.

  11. Coalescence of repelling colloidal droplets: a route to monodisperse populations.

    PubMed

    Roger, Kevin; Botet, Robert; Cabane, Bernard

    2013-05-14

    Populations of droplets or particles dispersed in a liquid may evolve through Brownian collisions, aggregation, and coalescence. We have found a set of conditions under which these populations evolve spontaneously toward a narrow size distribution. The experimental system consists of poly(methyl methacrylate) (PMMA) nanodroplets dispersed in a solvent (acetone) + nonsolvent (water) mixture. These droplets carry electrical charges, located on the ionic end groups of the macromolecules. We used time-resolved small angle X-ray scattering to determine their size distribution. We find that the droplets grow through coalescence events: the average radius (R) increases logarithmically with elapsed time while the relative width σR/(R) of the distribution decreases as the inverse square root of (R). We interpret this evolution as resulting from coalescence events that are hindered by ionic repulsions between droplets. We generalize this evolution through a simulation of the Smoluchowski kinetic equation, with a kernel that takes into account the interactions between droplets. In the case of vanishing or attractive interactions, all droplet encounters lead to coalescence. The corresponding kernel leads to the well-known "self-preserving" particle distribution of the coalescence process, where σR/(R) increases to a plateau value. However, for droplets that interact through long-range ionic repulsions, "large + small" droplet encounters are more successful at coalescence than "large + large" encounters. We show that the corresponding kernel leads to a particular scaling of the droplet-size distribution-known as the "second-scaling law" in the theory of critical phenomena, where σR/(R) decreases as 1/√(R) and becomes independent of the initial distribution. We argue that this scaling explains the narrow size distributions of colloidal dispersions that have been synthesized through aggregation processes.

  12. Crater Formation on Electrodes during Charge Transfer with Aqueous Droplets or Solid Particles

    NASA Astrophysics Data System (ADS)

    Elton, Eric S.; Rosenberg, Ethan R.; Ristenpart, William D.

    2017-11-01

    We report that metallic electrodes are physically pitted during charge transfer events with water droplets or other conductive objects moving in strong electric fields (>1 kV/cm). Post situ microscopic inspection of the electrode shows that an individual charge transfer event yields a crater approximately 1 to 3 microns wide, often with features similar to splash coronae. We interpret the crater formation in terms of localized melting of the electrode via resistive heating concurrent with dielectric breakdown through the surrounding insulating fluid. A scaling analysis indicates that the crater diameter scales as the inverse cube root of the melting point temperature Tm of the metal, in accord with measurements on several metals (660°C <=Tm <= 3414°C). The process of crater formation provides a possible explanation for the longstanding difficulty in quantitatively corroborating Maxwell's prediction for the amount of charge acquired by spheres contacting a planar electrode.

  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. Systems and methods for producing metal clusters; functionalized surfaces; and droplets including solvated metal ions

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Cooks, Robert Graham; Li, Anyin; Luo, Qingjie

    The invention generally relates to systems and methods for producing metal clusters; functionalized surfaces; and droplets including solvated metal ions. In certain aspects, the invention provides methods that involve providing a metal and a solvent. The methods additionally involve applying voltage to the solvated metal to thereby produce solvent droplets including ions of the metal containing compound, and directing the solvent droplets including the metal ions to a target. In certain embodiments, once at the target, the metal ions can react directly or catalyze reactions.

  15. Systems and methods for producing metal clusters; functionalized surfaces; and droplets including solvated metal ions

    DOEpatents

    Cooks, Robert Graham; Li, Anyin; Luo, Qingjie

    2017-01-24

    The invention generally relates to systems and methods for producing metal clusters; functionalized surfaces; and droplets including solvated metal ions. In certain aspects, the invention provides methods that involve providing a metal and a solvent. The methods additionally involve applying voltage to the solvated metal to thereby produce solvent droplets including ions of the metal containing compound, and directing the solvent droplets including the metal ions to a target. In certain embodiments, once at the target, the metal ions can react directly or catalyze reactions.

  16. Electrospray Charging of Minerals: Surface Chemistry and Applications to High-Velocity Microparticle Impacts

    NASA Astrophysics Data System (ADS)

    Daly, T.; Call, S.; Austin, D. E.

    2010-12-01

    Electrospray is a soft ionization technique commonly used to charge large biomolecules; it has, however, also been applied to inorganic compounds. We are extending this technique to mineral microparticles. Electrospray-charged mineral microparticles are interesting in the context of surface science because surface chemistry dictates where and how charge carriers can bond to mineral surfaces. In addition, using electrospray to charge mineral particles allows these particles to be electrostatically accelerated as projectiles in high- and hyper-velocity impacts. Since current techniques for producing high- and hyper-velocity microparticle impacts are largely limited to metal or metal-coated projectiles, using minerals as projectiles is a significant innovation. Electrospray involves three steps: creation of charged droplets containing solute/particles, evaporation and bifurcation of droplets, and desolvation of the solute/particles. An acidified solution is slowly pumped through a needle in a strong DC field, which causes the solution to break into tiny, charged droplets laden with protons. Solvent evaporates from the electrosprayed droplets as they move through the electric field toward a grounded plate, causing the charge on the droplet to increase relative to its mass. When the electrosprayed droplet’s charge becomes such that the droplet is no longer stable, it bifurcates, and each of the resulting droplets carries some of the original droplet’s charge. Evaporation and bifurcation continues until the solute particle is completely desolvated. The result is a protonated solute molecule or particle. We built an instrument that electrosprays particles into vacuum and measures them using an image charge detector. Mineral microparticles were prepared by grinding natural mineral samples to ~2 µm diameter. These microparticles are then added to a 4:1 methanol:water solution to create a 0.005% w/v suspension. The suspension is electrosprayed into vacuum, where the

  17. Controlled Gelation of Particle Suspensions Using Controlled Solvent Removal in Picoliter Droplets

    NASA Astrophysics Data System (ADS)

    Vuong, Sharon; Walker, Lynn; Anna, Shelley

    2013-11-01

    Droplets in microfluidic devices have proven useful as uniform picoliter reactors for nanoparticle synthesis and as components in tunable emulsions. However, there can be significant transport between the component phases depending on solubility and other factors. In the present talk, we show that water droplets trapped within a microfluidic device for tens of hours slowly dehydrate, concentrating the contents encapsulated within. We use this slow dehydration along with control of the initial droplet composition to monitor gelation of aqueous suspensions of spherical silica particles (Ludox) and disk-shaped clay particles (Laponite). Droplets are generated in a microfluidic device containing small wells that trap the droplets. We monitor the concentration process through size and shape changes of these droplets as a function of time in tens of droplets and use the large number of individual reactors to generate statistics regarding the gelation process. We also examine changes in suspension viscosity through fluorescent particle tracking as a function of dehydration rate, initial suspension concentration and initial droplet volume, and added salt, and compare the results with the Krieger-Dougherty model in which viscosity increases dramatically with particle volume fraction.

  18. A simple model of solvent-induced symmetry-breaking charge transfer in excited quadrupolar molecules

    NASA Astrophysics Data System (ADS)

    Ivanov, Anatoly I.; Dereka, Bogdan; Vauthey, Eric

    2017-04-01

    A simple model has been developed to describe the symmetry-breaking of the electronic distribution of AL-D-AR type molecules in the excited state, where D is an electron donor and AL and AR are identical acceptors. The origin of this process is usually associated with the interaction between the molecule and the solvent polarization that stabilizes an asymmetric and dipolar state, with a larger charge transfer on one side than on the other. An additional symmetry-breaking mechanism involving the direct Coulomb interaction of the charges on the acceptors is proposed. At the same time, the electronic coupling between the two degenerate states, which correspond to the transferred charge being localised either on AL or AR, favours a quadrupolar excited state with equal amount of charge-transfer on both sides. Because of these counteracting effects, symmetry breaking is only feasible when the electronic coupling remains below a threshold value, which depends on the solvation energy and the Coulomb repulsion energy between the charges located on AL and AR. This model allows reproducing the solvent polarity dependence of the symmetry-breaking reported recently using time-resolved infrared spectroscopy.

  19. Preventing droplet deformation during dielectrophoretic centering of a compound emulsion droplet

    NASA Astrophysics Data System (ADS)

    Randall, Greg; Blue, Brent

    2012-11-01

    Compound droplets, or droplets-within-droplets, are traditionally key components in applications ranging from drug delivery to the food industry. Presently, millimeter-sized compound droplets are precursors for shell targets in inertial fusion energy work. However, a key constraint in target fabrication is a uniform shell wall thickness, which in turn requires a centered core droplet in the compound droplet precursor. Previously, Bei et al. (2009, 2010) have shown that compound droplets could be centered in a static fluid using an electric field of 0.7 kV/cm at 20 MHz. Randall et al. (2012) developed a process to center the core of a moving compound droplet, though the ~kV/cm field induced small (< 5%) but undesirable droplet stretching. This work shows that by using macromolecular emulsifiers to strengthen the droplet's interfaces, (proteins, tunable peptides, or biotinylated streptavidin) droplet stretching can be greatly inhibited. Proof-of-principle experiments are performed in either a stagnant density-matched aquarium or a vertical channel of buoyancy-driven droplets in a ~kV/cm electric field. A scaling analysis is given from a fluid mechanics and interfacial rheology perspective and we discuss the effective interfacial charge from an emulsifier and its impact on centering. Work funded by General Atomics Internal R&D.

  20. Crater Formation on Electrodes during Charge Transfer with Aqueous Droplets or Solid Particles

    NASA Astrophysics Data System (ADS)

    Elton, E. S.; Rosenberg, E. R.; Ristenpart, W. D.

    2017-09-01

    We report that metallic electrodes are physically pitted during charge transfer events with water droplets or other conductive objects moving in strong electric fields (>1 kV /cm ). Post situ microscopic inspection of the electrode shows that an individual charge transfer event yields a crater approximately 1-3 μ m wide, often with features similar to a splash corona. We interpret the crater formation in terms of localized melting of the electrode via resistive heating concurrent with dielectric breakdown through the surrounding insulating fluid. A scaling analysis indicates that the crater diameter scales as the inverse cube root of the melting point temperature Tm of the metal, in accord with measurements on several metals (660 °C ≤Tm≤3414 °C ). The process of crater formation provides a possible explanation for the longstanding difficulty in quantitatively corroborating Maxwell's prediction for the amount of charge acquired by spheres contacting a planar electrode.

  1. Solvent jet desorption capillary photoionization-mass spectrometry.

    PubMed

    Haapala, Markus; Teppo, Jaakko; Ollikainen, Elisa; Kiiski, Iiro; Vaikkinen, Anu; Kauppila, Tiina J; Kostiainen, Risto

    2015-03-17

    A new ambient mass spectrometry method, solvent jet desorption capillary photoionization (DCPI), is described. The method uses a solvent jet generated by a coaxial nebulizer operated at ambient conditions with nitrogen as nebulizer gas. The solvent jet is directed onto a sample surface, from which analytes are extracted into the solvent and ejected from the surface in secondary droplets formed in collisions between the jet and the sample surface. The secondary droplets are directed into the heated capillary photoionization (CPI) device, where the droplets are vaporized and the gaseous analytes are ionized by 10 eV photons generated by a vacuum ultraviolet (VUV) krypton discharge lamp. As the CPI device is directly connected to the extended capillary inlet of the MS, high ion transfer efficiency to the vacuum of MS is achieved. The solvent jet DCPI provides several advantages: high sensitivity for nonpolar and polar compounds with limit of detection down to low fmol levels, capability of analyzing small and large molecules, and good spatial resolution (250 μm). Two ionization mechanisms are involved in DCPI: atmospheric pressure photoionization, capable of ionizing polar and nonpolar compounds, and solvent assisted inlet ionization capable of ionizing larger molecules like peptides. The feasibility of DCPI was successfully tested in the analysis of polar and nonpolar compounds in sage leaves and chili pepper.

  2. The Evolution of Electrospray Generated Droplets is Not Affected by Ionization Mode

    NASA Astrophysics Data System (ADS)

    Liigand, Piia; Heering (Suu), Agnes; Kaupmees, Karl; Leito, Ivo; Girod, Marion; Antoine, Rodolphe; Kruve, Anneli

    2017-10-01

    Ionization efficiency and mechanism in ESI is strongly affected by the properties of mobile phase. The use of mobile-phase properties to accurately describe droplets in ESI source is convenient but may be inadequate as the composition of the droplets is changing in the plume due to electrochemical reactions occurring in the needle tip as well as continuous drying and fission of droplets. Presently, there is paucity of research on the effect of the polarity of the ESI mode on mobile phase composition in the droplets. In this paper, the change in the organic solvent content, pH, and droplet size are studied in the ESI plume in both ESI+ and ESI- ionization mode. We introduce a rigorous way - the absolute pH (pHabs H 2 O) - to describe pH change in the plume that takes into account organic solvent content in the mobile phase. pHabs H 2 O enables comparing acidities of ESI droplets with different organic solvent contents. The results are surprisingly similar for both ionization modes, indicating that the dynamics of the change of mobile-phase properties is independent from the ESI mode used. This allows us to conclude that the evolution of ESI droplets first of all proceeds via the evaporation of the organic modifier and to a lesser extent via fission of smaller droplets from parent droplets. Secondly, our study shows that qualitative findings related to the ESI process obtained on the ESI+ mode can almost directly be applied also in the ESI- mode. [Figure not available: see fulltext.

  3. Substituent and Solvent Effects on Excited State Charge Transfer Behavior of Highly Fluorescent Dyes Containing Thiophenylimidazole-Based Aldehydes

    NASA Technical Reports Server (NTRS)

    Santos, Javier; Bu, Xiu R.; Mintz, Eric A.

    2001-01-01

    The excited state charge transfer for a series of highly fluorescent dyes containing thiophenylimidazole moiety was investigated. These systems follow the Twisted Intramolecular Charge Transfer (TICT) model. Dual fluorescence was observed for each substituted dye. X-ray structures analysis reveals a twisted ground state geometry for the donor substituted aryl on the 4 and 5 position at the imidazole ring. The excited state charge transfer was modeled by a linear solvation energy relationship using Taft's pi and Dimroth's E(sub T)(30) as solvent parameters. There is linear relation between the energy of the fluorescence transition and solvent polarity. The degree of stabilization of the excited state charge transfer was found to be consistent with the intramolecular molecular charge transfer. Excited dipole moment was studied by utilizing the solvatochromic shift method.

  4. Regimes of electrostatic collapse of a highly charged polyelectrolyte in a poor solvent.

    PubMed

    Tom, Anvy Moly; Vemparala, Satyavani; Rajesh, R; Brilliantov, Nikolai V

    2017-03-01

    We perform extensive molecular dynamics simulations of a highly charged, collapsed, flexible polyelectrolyte chain in a poor solvent for the case when the electrostatic interactions, characterized by the reduced Bjerrum length l B , are strong. We find the existence of several sub-regimes in the dependence of the gyration radius of the chain R g on l B characterized by R g ∼ l. In contrast to a good solvent, the exponent γ for a poor solvent crucially depends on the size and valency of the counterions. To explain the different sub-regimes, we generalize the existing counterion fluctuation theory by including a more complete account of all possible volume interactions in the free energy of the polyelectrolyte chain. We also show that the presence of condensed counterions modifies the effective attraction among the chain monomers and modulates the sign of the second virial coefficient under poor solvent conditions.

  5. Digitally controlled droplet microfluidic system based on electrophoretic actuation

    NASA Astrophysics Data System (ADS)

    Im, Do Jin; Yoo, Byeong Sun; Ahn, Myung Mo; Moon, Dustin; Kang, In Seok

    2012-11-01

    Most researches on direct charging and the subsequent manipulation of a charged droplet were focused on an on-demand sorting in microchannel where carrier fluid transports droplets. Only recently, an individual actuation of a droplet without microchannel and carrier fluid was tried. However, in the previous work, the system size was too large and the actuation voltage was too high (1.5 kV), which limits the applicability of the technology to mobile use. Therefore, in the current research, we have developed a miniaturized digital microfluidic system based on the electrophoresis of a charged droplet (ECD). By using a pin header socket for an array of electrodes, much smaller microfluidic system can be made from simple fabrication process with low cost. A full two dimensional manipulation (0.4 cm/s) of a droplet (300 nL) suspended in silicone oil (6 cSt) and multiple droplet actuation have been performed with reasonable actuation voltage (300 V). By multiple droplet actuation and coalescence, a practical biochemical application also has been demonstrated. We hope the current droplet manipulation method (ECD) can be a good alternative or complimentary technology to the conventional ones and therefore contributes to the development of droplet microfluidics. This work has been supported by BK21 program of the Ministry of Education, Science and Technology (MEST) of Korea.

  6. Liquid crystal Janus emulsion droplets: preparation, tumbling, and swimming.

    PubMed

    Jeong, Joonwoo; Gross, Adam; Wei, Wei-Shao; Tu, Fuquan; Lee, Daeyeon; Collings, Peter J; Yodh, A G

    2015-09-14

    This study introduces liquid crystal (LC) Janus droplets. We describe a process for the preparation of these droplets, which consist of nematic LC and polymer compartments. The process employs solvent-induced phase separation in emulsion droplets generated by microfluidics. The droplet morphology was systematically investigated and demonstrated to be sensitive to the surfactant concentration in the background phase, the compartment volume ratio, and the possible coalescence of multiple Janus droplets. Interestingly, the combination of a polymer and an anisotropic LC introduces new functionalities into Janus droplets, and these properties lead to unusual dynamical behaviors. The different densities and solubilities of the two compartments produce gravity-induced alignment, tumbling, and directional self-propelled motion of Janus droplets. LC Janus droplets with remarkable optical properties and dynamical behaviors thus offer new avenues for applications of Janus colloids and active soft matter.

  7. Single photon emission of a charge-tunable GaAs/Al{sub 0.25}Ga{sub 0.75}As droplet quantum dot device

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Langer, Fabian, E-mail: fabian.langer@physik.uni-wuerzburg.de; Plischke, David; Kamp, Martin

    2014-08-25

    In this work, we report the fabrication of a charge-tunable GaAs/Al{sub 0.25}Ga{sub 0.75}As quantum dot (QD) device containing QDs deposited by modified droplet epitaxy producing almost strain and composition gradient free QDs. We obtained a QD density in the low 10{sup 9 }cm{sup −2} range that enables us to perform spectroscopy on single droplet QDs showing linewidths as narrow as 40 μeV. The integration of the QDs into a Schottky diode allows us to controllably charge a single QD with up to four electrons, while non-classical photoluminescence is proven by photon auto-correlation measurements showing photon-antibunching (g{sup (2)}(0) = 0.05).

  8. Emulsion droplet interactions: a front-tracking treatment

    NASA Astrophysics Data System (ADS)

    Mason, Lachlan; Juric, Damir; Chergui, Jalel; Shin, Seungwon; Craster, Richard V.; Matar, Omar K.

    2017-11-01

    Emulsion coalescence influences a multitude of industrial applications including solvent extraction, oil recovery and the manufacture of fast-moving consumer goods. Droplet interaction models are vital for the design and scale-up of processing systems, however predictive modelling at the droplet-scale remains a research challenge. This study simulates industrially relevant moderate-inertia collisions for which a high degree of droplet deformation occurs. A hybrid front-tracking/level-set approach is used to automatically account for interface merging without the need for `bookkeeping' of interface connectivity. The model is implemented in Code BLUE using a parallel multi-grid solver, allowing both film and droplet-scale dynamics to be resolved efficiently. Droplet interaction simulations are validated using experimental sequences from the literature in the presence and absence of background turbulence. The framework is readily extensible for modelling the influence of surfactants and non-Newtonian fluids on droplet interaction processes. EPSRC, UK, MEMPHIS program Grant (EP/K003976/1), RAEng Research Chair (OKM), PETRONAS.

  9. Universal fluid droplet ejector

    DOEpatents

    Lee, E.R.; Perl, M.L.

    1999-08-24

    A droplet generator comprises a fluid reservoir having a side wall made of glass or quartz, and an end cap made from a silicon plate. The end cap contains a micromachined aperture through which the fluid is ejected. The side wall is thermally fused to the end cap, and no adhesive is necessary. This means that the fluid only comes into contact with the side wall and the end cap, both of which are chemically inert. Amplitudes of drive pulses received by reservoir determine the horizontal displacements of droplets relative to the ejection aperture. The drive pulses are varied such that the dropper generates a two-dimensional array of vertically-falling droplets. Vertical and horizontal inter-droplet spacings may be varied in real time. Applications include droplet analysis experiments such as Millikan fractional charge searches and aerosol characterization, as well as material deposition applications. 8 figs.

  10. Solvent electronic polarization effects on a charge transfer excitation studied by the mean-field QM/MM method

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Nakano, Hiroshi; Elements Strategy Initiative for Catalysts and Batteries, Kyoto University, Kyoto 615-8245

    2015-12-31

    Electronic polarization effects of a medium can have a significant impact on a chemical reaction in condensed phases. We discuss the effects on the charge transfer excitation of a chromophore, N,N-dimethyl-4-nitroaniline, in various solvents using the mean-field QM/MM method with a polarizable force field. The results show that the explicit consideration of the solvent electronic polarization effects is important especially for a solvent with a low dielectric constant when we study the solvatochromism of the chromophore.

  11. On-demand Droplet Manipulation via Triboelectrification

    NASA Astrophysics Data System (ADS)

    Wang, Wei; Vahabi, Hamed; Cackovic, Matthew; Jiang, Rui; Kota, Arun

    2017-11-01

    Controlled manipulation of liquid droplets has attracted tremendous interest across different scientific fields over the past two decades. To date, a variety of external stimuli-mediated methods such as magnetic field, electric field, and light have been developed for manipulating droplets on surfaces. However, these methods usually have drawbacks such as complex fabrication of manipulation platform, low droplet motility, expensive actuation system and lack of precise control. In this work, we demonstrate the controlled manipulation of liquid droplet with both high (e.g., water) and low (e.g., n-hexadecane) dielectric strengths on a smooth, slippery surface via triboelectric effect. Our highly simple, facile and portable methodology enables on-demand, precise manipulation of droplets using solely the electrostatic attraction or repulsion force, which is exerted on the droplet by a simple charged actuator (e.g., Teflon film). We envision that our triboelectric effect enabled droplet manipulation methodology will open a new avenue for droplet based lab-on-a-chip systems, energy harvesting devices and biomedical applications.

  12. Theoretical evidence of charge transfer interaction between SO₂ and deep eutectic solvents formed by choline chloride and glycerol.

    PubMed

    Li, Hongping; Chang, Yonghui; Zhu, Wenshuai; Wang, Changwei; Wang, Chao; Yin, Sheng; Zhang, Ming; Li, Huaming

    2015-11-21

    The nature of the interaction between deep eutectic solvents (DESs), formed by ChCl and glycerol, and SO2 has been systematically investigated using the M06-2X density functional combined with cluster models. Block-localized wave function energy decomposition (BLW-ED) analysis shows that the interaction between SO2 and DESs is dominated by a charge transfer interaction. After this interaction, the SO2 molecule becomes negatively charged, whereas the ChCl-glycerol molecule is positively charged, which is the result of Lewis acid-base interaction. The current result affords a theoretical proof that it is highly useful and efficient to manipulate the Lewis acidity of absorbents for SO2 capture. Moreover, hydrogen bonding as well as electrostatic interactions may also contribute to the stability of the complex. Structure analysis shows that solvent molecules will adjust their geometries to interact with SO2. In addition, the structure of SO2 is barely changed after interaction. The interaction energy between different cluster models and SO2 ranges from -6.8 to -14.4 kcal mol(-1). It is found that the interaction energy is very sensitive to the solvent structure. The moderate interaction between ChCl-glycerol and SO2 is consistent with the concept that highly efficient solvents for SO2 absorption should not only be solvable but also regenerable.

  13. Droplet microfluidics with a nanoemulsion continuous phase.

    PubMed

    Gu, Tonghan; Yeap, Eunice W Q; Somasundar, Ambika; Chen, Ran; Hatton, T Alan; Khan, Saif A

    2016-07-05

    We present the first study of a novel, generalizable method that uses a water-in-oil nanoemulsion as the continuous phase to generate uniform aqueous micro-droplets in a capillary-based microfluidic system. We first study the droplet generation mechanism in this system and compare it to the more conventional case where a simple oil/solvent (with surfactant) is used as the continuous phase. Next, we present two versatile methods - adding demulsifying chemicals and heat treatment - to allow active online chemical interaction between the continuous and dispersed phases. These methods allow each generated micro-droplet to act as a well-mixed micro-reactor with walls that are 'permeable' to the nanoemulsion droplets and their contents. Finally, we demonstrate an application of this system in the fabrication of uniform hydrogel (alginate) micro-beads with control over particle properties such as size and swelling. Our work expands the toolbox of droplet-based microfluidics, enabling new opportunities and applications involving active colloidal continuous phases carrying chemical payloads, both in advanced materials synthesis and droplet-based screening and diagnostic methods.

  14. Universal fluid droplet ejector

    DOEpatents

    Lee, Eric R.; Perl, Martin L.

    1999-08-24

    A droplet generator comprises a fluid reservoir having a side wall made of glass or quartz, and an end cap made from a silicon plate. The end cap contains a micromachined aperture through which the fluid is ejected. The side wall is thermally fused to the end cap, and no adhesive is necessary. This means that the fluid only comes into contact with the side wall and the end cap, both of which are chemically inert. Amplitudes of drive pulses received by reservoir determine the horizontal displacements of droplets relative to the ejection aperture. The drive pulses are varied such that the dropper generates a two-dimensional array of vertically-falling droplets. Vertical and horizontal interdroplet spacings may be varied in real time. Applications include droplet analysis experiments such as Millikan fractional charge searches and aerosol characterization, as well as material deposition applications.

  15. Treatment of charge singularities in implicit solvent models.

    PubMed

    Geng, Weihua; Yu, Sining; Wei, Guowei

    2007-09-21

    This paper presents a novel method for solving the Poisson-Boltzmann (PB) equation based on a rigorous treatment of geometric singularities of the dielectric interface and a Green's function formulation of charge singularities. Geometric singularities, such as cusps and self-intersecting surfaces, in the dielectric interfaces are bottleneck in developing highly accurate PB solvers. Based on an advanced mathematical technique, the matched interface and boundary (MIB) method, we have recently developed a PB solver by rigorously enforcing the flux continuity conditions at the solvent-molecule interface where geometric singularities may occur. The resulting PB solver, denoted as MIBPB-II, is able to deliver second order accuracy for the molecular surfaces of proteins. However, when the mesh size approaches half of the van der Waals radius, the MIBPB-II cannot maintain its accuracy because the grid points that carry the interface information overlap with those that carry distributed singular charges. In the present Green's function formalism, the charge singularities are transformed into interface flux jump conditions, which are treated on an equal footing as the geometric singularities in our MIB framework. The resulting method, denoted as MIBPB-III, is able to provide highly accurate electrostatic potentials at a mesh as coarse as 1.2 A for proteins. Consequently, at a given level of accuracy, the MIBPB-III is about three times faster than the APBS, a recent multigrid PB solver. The MIBPB-III has been extensively validated by using analytically solvable problems, molecular surfaces of polyatomic systems, and 24 proteins. It provides reliable benchmark numerical solutions for the PB equation.

  16. Treatment of charge singularities in implicit solvent models

    NASA Astrophysics Data System (ADS)

    Geng, Weihua; Yu, Sining; Wei, Guowei

    2007-09-01

    This paper presents a novel method for solving the Poisson-Boltzmann (PB) equation based on a rigorous treatment of geometric singularities of the dielectric interface and a Green's function formulation of charge singularities. Geometric singularities, such as cusps and self-intersecting surfaces, in the dielectric interfaces are bottleneck in developing highly accurate PB solvers. Based on an advanced mathematical technique, the matched interface and boundary (MIB) method, we have recently developed a PB solver by rigorously enforcing the flux continuity conditions at the solvent-molecule interface where geometric singularities may occur. The resulting PB solver, denoted as MIBPB-II, is able to deliver second order accuracy for the molecular surfaces of proteins. However, when the mesh size approaches half of the van der Waals radius, the MIBPB-II cannot maintain its accuracy because the grid points that carry the interface information overlap with those that carry distributed singular charges. In the present Green's function formalism, the charge singularities are transformed into interface flux jump conditions, which are treated on an equal footing as the geometric singularities in our MIB framework. The resulting method, denoted as MIBPB-III, is able to provide highly accurate electrostatic potentials at a mesh as coarse as 1.2Å for proteins. Consequently, at a given level of accuracy, the MIBPB-III is about three times faster than the APBS, a recent multigrid PB solver. The MIBPB-III has been extensively validated by using analytically solvable problems, molecular surfaces of polyatomic systems, and 24 proteins. It provides reliable benchmark numerical solutions for the PB equation.

  17. Understanding the influence of solvent field and fluctuations on the stability of photo-induced charge-separated state in molecular triad

    NASA Astrophysics Data System (ADS)

    Balamurugan, D.; Aquino, Adelia; Lischka, Hans; Dios, Francis; Flores, Lionel; Cheung, Margaret

    2013-03-01

    Molecular triad composed of fullerene, porphyrin, and carotene is an artificial analogue of natural photosynthetic system and is considered for applications in solar energy conversion because of its ability to produce long-lived photo-induced charge separated state. The goal of the present multiscale simulation is to understand how the stability of photo-induced charge-separated state in molecular triad is influenced by a polar organic solvent, namely tetrahydrofuran (THF). The multiscale approach is based on combined quantum, classical molecular dynamics, and statistical physics calculations. The quantum chemical calculations were performed on the triad using the second order algebraic diagrammatic perturbation and time-dependent density functional theory. Molecular dynamics simulations were performed on triad in a box of THF solvent with the replica exchange method. The two methods on different length and time scales are bridged through an important sampling technique. We have analyzed the free energy landscape, structural fluctuations, and the long- range electrostatic interactions between triad and solvent molecules. The results suggest that the polarity and re-organization of the solvent is critical in stabilization of charge-separated state in triad. Supported by DOE (DE-FG02-10ER16175)

  18. Surface tension driven aggregation of organic nanowires via lab in a droplet.

    PubMed

    Gu, Jianmin; Yin, Baipeng; Fu, Shaoyan; Feng, Man; Zhang, Ziming; Dong, Haiyun; Gao, Faming; Zhao, Yong Sheng

    2018-06-05

    Directing the architecture of complex organic nanostructures is desirable and still remains a challenge in areas of materials science due to their structure-dependent collective optoelectronic properties. Herein, we demonstrate a simple and versatile solution strategy that allows surface tension to drive low-dimensional nanostructures to aggregate into complex structures via a lab in a droplet technique. By selecting a suitable combination of a solvent and an anti-solvent with controllable surface tension difference, the droplets can be automatically cracked into micro-droplets, which provides an aggregation force directed toward the centre of the droplet to drive the low-dimensional building blocks to form the special aggregations during the self-assembly process. This synthetic strategy has been shown to be universal for organic materials, which is beneficial for further optimizing the optoelectronic properties. These results contribute to gaining an insightful understanding on the detailed growth mechanism of complex organic nanostructures and greatly promoting the development of organic nanophotonics.

  19. Discrete and continuum modeling of solvent effects in a twisted intramolecular charge transfer system: The 4-N,N-dimethylaminobenzonitrile (DMABN) molecule.

    PubMed

    Modesto-Costa, Lucas; Borges, Itamar

    2018-08-05

    The 4-N,N-dimethylaminobenzonitrile (DMABN) molecule is a prototypical system displaying twisted intramolecular (TICT) charge transfer effects. The ground and the first four electronic excited states (S 1 -S 4 ) in gas phase and upon solvation were studied. Charge transfer values as function of the torsion angle between the donor group (dimethylamine) and the acceptor moiety (benzonitrile) were explicitly computed. Potential energy curves were also obtained. The algebraic diagrammatic construction method at the second-order [ADC(2)] ab initio wave function was employed. Three solvents of increased polarities (benzene, DMSO and water) were investigated using discrete (average solvent electrostatic configuration - ASEC) and continuum (conductor-like screening model - COSMO) models. The results for the S 3 and S 4 excited states and the S 1 -S 4 charge transfer curves were not previously available in the literature. Electronic gas phase and solvent vertical spectra are in good agreement with previous theoretical and experimental results. In the twisted (90°) geometry the optical oscillator strengths have negligible values even for the S 2 bright state. Potential energy curves show two distinct pairs of curves intersecting at decreasing angles or not crossing in the more polar solvents. Charge transfer and electric dipole values allowed the rationalization of these results. The former effects are mostly independent of the solvent model and polarity. Although COSMO and ASEC solvent models mostly lead to similar results, there is an important difference: some crossings of the excitation energy curves appear only in the ASEC solvation model, which has important implications to the photochemistry of DMABN. Copyright © 2018 Elsevier B.V. All rights reserved.

  20. Electrophoretic manipulation of multiple-emulsion droplets

    NASA Astrophysics Data System (ADS)

    Schoeler, Andreas M.; Josephides, Dimitris N.; Chaurasia, Ankur S.; Sajjadi, Shahriar; Mesquida, Patrick

    2014-02-01

    Electrophoretic manipulation of multiple-emulsion oil-in-water-in-oil (O/W)/O and water-in-oil-in-water-in-oil (W/O/W)/O core-shell droplets is shown. It was found that the electrophoretic mobility of the droplets is determined solely by the outer water shell, regardless of size or composition of the inner droplets. It was observed that the surface charge of the outer water shell can be changed and the polarity can be reversed through contact with a biased electrode in a similar way as with simple W/O droplets. Furthermore, addition of the anionic surfactant, sodium dodecyl sulfate to the outer water shell reverses the initial polarity and hence, electrophoretic mobility of the core-shell droplets before contact with an electrode. The results have practical implications for the manipulation of oil droplets in a continuous oil phase.

  1. Toward Femtosecond Time-Resolved Studies of Solvent-Solute Energy Transfer in Doped Helium Nanodroplets

    NASA Astrophysics Data System (ADS)

    Bacellar, C.; Ziemkiewicz, M. P.; Leone, S. R.; Neumark, D. M.; Gessner, O.

    2015-05-01

    Superfluid helium nanodroplets provide a unique cryogenic matrix for high resolution spectroscopy and ultracold chemistry applications. With increasing photon energy and, in particular, in the increasingly important Extreme Ultraviolet (EUV) regime, the droplets become optically dense and, therefore, participate in the EUV-induced dynamics. Energy- and charge-transfer mechanisms between the host droplets and dopant atoms, however, are poorly understood. Static energy domain measurements of helium droplets doped with noble gas atoms (Xe, Kr) indicate that Penning ionization due to energy transfer from the excited droplet to dopant atoms may be a significant relaxation channel. We have set up a femtosecond time-resolved photoelectron imaging experiment to probe these dynamics directly in the time-domain. Droplets containing 104 to 106 helium atoms and a small percentage (<10-4) of dopant atoms (Xe, Kr, Ne) are excited to the 1s2p Rydberg band by 21.6 eV photons produced by high harmonic generation (HHG). Transiently populated states are probed by 1.6 eV photons, generating time-dependent photoelectron kinetic energy distributions, which are monitored by velocity map imaging (VMI). The results will provide new information about the dynamic timescales and the different relaxation channels, giving access to a more complete physical picture of solvent-solute interactions in the superfluid environment. Prospects and challenges of the novel experiment as well as preliminary experimental results will be discussed.

  2. Spectroscopy and optical imaging of coalescing droplets

    NASA Astrophysics Data System (ADS)

    Ivanov, Maksym; Viderström, Michel; Chang, Kelken; Ramírez Contreras, Claudia; Mehlig, Bernhard; Hanstorp, Dag

    2016-09-01

    We report on experimental investigations of the dynamics of colliding liquid droplets by combining optical trapping, spectroscopy and high-speed color imaging. Two droplets with diameters between 5 and 50 microns are suspended in quiescent air by optical traps. The traps allows us to control the initial positions, and hence the impact parameter and the relative velocity of the colliding droplets. Movies of the droplet dynamics are recorded using high-speed digital movie cameras at a frame rate of up to 63000 frames per second. A fluorescent dye is added to one of the colliding droplets. We investigate the temporal evolution of the scattered and fluorescence light from the colliding droplets with concurrent spectroscopy and color imaging. This technique can be used to detect the exchange of molecules between a pair of neutral or charged droplets.

  3. Scaling Atomic Partial Charges of Carbonate Solvents for Lithium Ion Solvation and Diffusion

    DOE PAGES

    Chaudhari, Mangesh I.; Nair, Jijeesh R.; Pratt, Lawrence R.; ...

    2016-10-21

    Lithium-ion solvation and diffusion properties in ethylene carbonate (EC) and propylene carbonate (PC) were studied by molecular simulation, experiments, and electronic structure calculations. Studies carried out in water provide a reference for interpretation. Classical molecular dynamics simulation results are compared to ab initio molecular dynamics to assess nonpolarizable force field parameters for solvation structure of the carbonate solvents. Quasi-chemical theory (QCT) was adapted to take advantage of fourfold occupancy of the near-neighbor solvation structure observed in simulations and used to calculate solvation free energies. The computed free energy for transfer of Li + to PC from water, based on electronicmore » structure calculations with cluster-QCT, agrees with the experimental value. The simulation-based direct-QCT results with scaled partial charges agree with the electronic structure-based QCT values. The computed Li +/PF 6 - transference numbers of 0.35/0.65 (EC) and 0.31/0.69 (PC) agree well with NMR experimental values of 0.31/0.69 (EC) and 0.34/0.66 (PC) and similar values obtained here with impedance spectroscopy. These combined results demonstrate that solvent partial charges can be scaled in systems dominated by strong electrostatic interactions to achieve trends in ion solvation and transport properties that are comparable to ab initio and experimental results. Thus, the results support the use of scaled partial charges in simple, nonpolarizable force fields in future studies of these electrolyte solutions.« less

  4. Modeling Evaporation and Particle Assembly in Colloidal Droplets.

    PubMed

    Zhao, Mingfei; Yong, Xin

    2017-06-13

    Evaporation-induced assembly of nanoparticles in a drying droplet is of great importance in many engineering applications, including printing, coating, and thin film processing. The investigation of particle dynamics in evaporating droplets can provide fundamental hydrodynamic insight for revealing the processing-structure relationship in the particle self-organization induced by solvent evaporation. We develop a free-energy-based multiphase lattice Boltzmann method coupled with Brownian dynamics to simulate evaporating colloidal droplets on solid substrates with specified wetting properties. The influence of interface-bound nanoparticles on the surface tension and evaporation of a flat liquid-vapor interface is first quantified. The results indicate that the particles at the interface reduce surface tension and enhance evaporation flux. For evaporating particle-covered droplets on substrates with different wetting properties, we characterize the increase of evaporate rate via measuring droplet volume. We find that droplet evaporation is determined by the number density and circumferential distribution of interfacial particles. We further correlate particle dynamics and assembly to the evaporation-induced convection in the bulk and on the surface of droplet. Finally, we observe distinct final deposits from evaporating colloidal droplets with bulk-dispersed and interface-bound particles. In addition, the deposit pattern is also influenced by the equilibrium contact angle of droplet.

  5. Water Entry by a Train of Droplets

    NASA Astrophysics Data System (ADS)

    Ohl, Claus-Dieter; Huang, Xin; Chan, Chon U.; Frommhold, Philipp Erhard; Lippert, Alexander

    2014-11-01

    The impact of single droplets on a deep pool is a well-studied phenomenon which reveals reach fluid mechanics. Lesser studied is the impact of a train of droplet and the accompanied formation of largely elongated cavities, in particular for well controlled droplets. The droplets with diameters of 20-40 μm and velocities of approx. 20 m/s are generated with a piezo-actuated nozzle at rates of 200-300 kHz. Individual droplets are selected by electric charging and deflection and the impact is visualized with stroboscopic photography and high-speed videos. We study in particular the formation and shape of the cavity as by varying the number of droplets from one to 64. The cavities reach centimetres in length with lateral diameters of the order of 100 of micrometres.

  6. Effect of solvent hydrogen bonding on the photophysical properties of intramolecular charge transfer probe trans-ethyl p-(dimethylamino) cinamate and its derivative

    NASA Astrophysics Data System (ADS)

    Singh, T. Sanjoy; Moyon, N. S.; Mitra, Sivaprasad

    2009-08-01

    Intramolecular charge transfer (ICT) behavior of trans-ethyl p-(dimethylamino) cinamate (EDAC) and 4-(dimethylamino) cinnamic acid (DMACA) were studied by steady state absorption and emission, picosecond time-resolved fluorescence experiments in various pure and mixed solvent systems. The large fluorescence spectral shift in more polar solvents indicates an efficient charge transfer from the donor site to the acceptor moiety in the excited state compared to the ground state. The energy for 0,0 transition ( ν0,0) for EDAC shows very good linear correlation with static solvent dielectric property; however, fluorescence emission maximum, stokes shift and fluorescence quantum yield show significant deviation from linearity in polar protic solvents, indicating a large contribution of solvent hydrogen bonding on the excited state relaxation mechanism. A quantitative estimation of contribution from different solvatochromic parameters was made using linear free energy relationship based on Kamlet-Taft equation.

  7. Solvent Reaction Field Potential inside an Uncharged Globular Protein: A Bridge between Implicit and Explicit Solvent Models?

    PubMed Central

    Baker, Nathan A.; McCammon, J. Andrew

    2008-01-01

    The solvent reaction field potential of an uncharged protein immersed in Simple Point Charge/Extended (SPC/E) explicit solvent was computed over a series of molecular dynamics trajectories, intotal 1560 ns of simulation time. A finite, positive potential of 13 to 24 kbTec−1 (where T = 300K), dependent on the geometry of the solvent-accessible surface, was observed inside the biomolecule. The primary contribution to this potential arose from a layer of positive charge density 1.0 Å from the solute surface, on average 0.008 ec/Å3, which we found to be the product of a highly ordered first solvation shell. Significant second solvation shell effects, including additional layers of charge density and a slight decrease in the short-range solvent-solvent interaction strength, were also observed. The impact of these findings on implicit solvent models was assessed by running similar explicit-solvent simulations on the fully charged protein system. When the energy due to the solvent reaction field in the uncharged system is accounted for, correlation between per-atom electrostatic energies for the explicit solvent model and a simple implicit (Poisson) calculation is 0.97, and correlation between per-atom energies for the explicit solvent model and a previously published, optimized Poisson model is 0.99. PMID:17949217

  8. Solvent reaction field potential inside an uncharged globular protein: A bridge between implicit and explicit solvent models?

    NASA Astrophysics Data System (ADS)

    Cerutti, David S.; Baker, Nathan A.; McCammon, J. Andrew

    2007-10-01

    The solvent reaction field potential of an uncharged protein immersed in simple point charge/extended explicit solvent was computed over a series of molecular dynamics trajectories, in total 1560ns of simulation time. A finite, positive potential of 13-24 kbTec-1 (where T =300K), dependent on the geometry of the solvent-accessible surface, was observed inside the biomolecule. The primary contribution to this potential arose from a layer of positive charge density 1.0Å from the solute surface, on average 0.008ec/Å3, which we found to be the product of a highly ordered first solvation shell. Significant second solvation shell effects, including additional layers of charge density and a slight decrease in the short-range solvent-solvent interaction strength, were also observed. The impact of these findings on implicit solvent models was assessed by running similar explicit solvent simulations on the fully charged protein system. When the energy due to the solvent reaction field in the uncharged system is accounted for, correlation between per-atom electrostatic energies for the explicit solvent model and a simple implicit (Poisson) calculation is 0.97, and correlation between per-atom energies for the explicit solvent model and a previously published, optimized Poisson model is 0.99.

  9. Spontaneous Droplet Jump with Electro-Bouncing

    NASA Astrophysics Data System (ADS)

    Schmidt, Erin; Weislogel, Mark

    2016-11-01

    We investigate the dynamics of water droplet jumps from superhydrophobic surfaces in the presence of an electric field during a step reduction in gravity level. In the brief free-fall environment of a drop tower, when a strong non-homogeneous electric field (with a measured strength between 0 . 39 and 2 . 36 kV/cm) is imposed, body forces acting on the jumped droplets are primarily supplied by polarization stress and Coulombic attraction instead of gravity. The droplet charge, measured to be on the order of 2 . 3 . (10-11) C, originates by electro-osmosis of charged species at the (PTFE coated) hydrophobic surface interface. This electric body force leads to a droplet bouncing behavior similar to well-known phenomena in 1-g, though occurring for larger drops 0.1 mL for a given range of impact Weber numbers, We < 20 . In 1-g, for We > 0 . 4 , impact recoil behavior on a super-hydrophobic surface is normally dominated by damping from contact line hysteresis and by air-layer interactions. However, in the strong electric field, the droplet bounce dynamics additionally include electrohydrodynamic effects on wettability and Cassie-Wenzel transition. This is qualitatively discussed in terms of coefficients of restitution and trends in contact time. This work was supported primarily by NASA Cooperative Agreement NNX12A047A.

  10. The roles of the solute and solvent cavities in charge-transfer-to-solvent dynamics: Ultrafast studies of potasside and sodide in diethyl ether

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Cavanagh, Molly C.; Young, Ryan M.; Schwartz, Benjamin J.

    2008-10-07

    Although electron transfer reactions are among the most fundamental in chemistry, it is still not clear how to isolate the roles of the solute and solvent in moving charge between reactants in solution. In this paper, we address this question by comparing the ultrafast charge-transfer-to-solvent (CTTS) dynamics of potasside (K{sup -}) in diethyl ether (DEE) to those of sodide (Na{sup -}) in both DEE and tetrahydrofuran (THF). We find that for sodide in both DEE and THF, CTTS excitation leads to delayed ejection of a solvated electron that appears with its equilibrium absorption spectrum. This indicates that the ejected electronsmore » are localized in pre-existing solvent traps, suggesting that the structure of liquid DEE is characterized by cavities that are favorably polarized to localize an excess electron, as has been previously shown is the case for liquid THF. We also find that the geminate recombination dynamics following CTTS excitation of sodide in THF and DEE are similar, suggesting that the nature of the CTTS excited states and their coupling to the electronic states supported by the naturally occurring solvent cavities are similar in the two solvents. In contrast, the geminate recombination dynamics of potasside and sodide in DEE are different, with red-edge excitation of the K{sup -} CTTS band producing a greater number of long-lived electrons than is seen following the corresponding red-edge excitation of the Na{sup -} CTTS band. This indicates that the CTTS excited states of K{sup -} are better able to couple to the electronic states supported by the naturally occurring solvent cavities, allowing us to compare the energetic positions of the potasside and sodide ground and CTTS excited states on a common absolute scale. Finally, we also observe a strong transient absorption following the CTTS excitation of potasside in DEE that correlates well with the 766 nm position of the gas-phase potassium D-line. The data indicate that CTTS excitation of

  11. An Instrument Employing a Coronal Discharge for the Determination of Droplet-Size Distribution in Clouds

    NASA Technical Reports Server (NTRS)

    Brun, Rinaldo J.; Levine, Joseph; Kleinknecht, Kenneth S.

    1951-01-01

    A flight instrument that uses electric means for measuring the droplet-size distribution in above-freezing clouds has been devised and given preliminary evaluation in flight. An electric charge is placed on the droplets and they are separated aerodynamically according to their mass. Because the charge placed on the droplets is a. function of the droplet size, the size spectrum can 'be determined by measurement of the charge deposited on cylinders of several different sizes placed to intercept the charged droplets. An expression for the rate of charge acquisition by a water droplet in a field of coronal discharge is derived. The results obtained in flight with an instrument based on the method described indicate that continuous records of droplet-size spectrum variations in clouds can be obtained. The experimental instrument was used to evaluate the method and was not refined to the extent necessary for obtaining conclusive meteorological data. The desirable features of an instrument based on the method described are (i) The instrument can be used in clouds with temperatures above freezing; (2) the size and the shape of the cylinders do not change during the exposure time; (3) the readings are instantaneous and continuous; (4) the available sensitivity permits the study of variations in cloud structures of less than 200 feet in extent.

  12. Effect of Viscosity and Polar Properties of Solvent on Dynamics of Photoinduced Charge Transfer in BTA-1 Cation — Derivative of Thioflavin T

    NASA Astrophysics Data System (ADS)

    Gogoleva, S. D.; Stsiapura, V. I.

    2018-05-01

    It was found that the spectral and fluorescent properties of BTA-1C cation in protic and aprotic solvents differ. It was shown that for solutions in long-chain alcohols viscosity is the main factor that determines the dynamics of intramolecular charge transfer in the excited state of the BTA-1C molecule. In the case of aprotic solvents a correlation was found between the rate constant of twisted intramolecular charge transfer (TICT) during rotation of fragments of the molecule in relation to each other in the excited state and the solvent relaxation rate: k TICT 1/τ S .

  13. Solidification of floating organic droplet in dispersive liquid-liquid microextraction as a green analytical tool.

    PubMed

    Mansour, Fotouh R; Danielson, Neil D

    2017-08-01

    Dispersive liquid-liquid microextraction (DLLME) is a special type of microextraction in which a mixture of two solvents (an extracting solvent and a disperser) is injected into the sample. The extraction solvent is then dispersed as fine droplets in the cloudy sample through manual or mechanical agitation. Hence, the sample is centrifuged to break the formed emulsion and the extracting solvent is manually separated. The organic solvents commonly used in DLLME are halogenated hydrocarbons that are highly toxic. These solvents are heavier than water, so they sink to the bottom of the centrifugation tube which makes the separation step difficult. By using solvents of low density, the organic extractant floats on the sample surface. If the selected solvent such as undecanol has a freezing point in the range 10-25°C, the floating droplet can be solidified using a simple ice-bath, and then transferred out of the sample matrix; this step is known as solidification of floating organic droplet (SFOD). Coupling DLLME to SFOD combines the advantages of both approaches together. The DLLME-SFOD process is controlled by the same variables of conventional liquid-liquid extraction. The organic solvents used as extractants in DLLME-SFOD must be immiscible with water, of lower density, low volatility, high partition coefficient and low melting and freezing points. The extraction efficiency of DLLME-SFOD is affected by types and volumes of organic extractant and disperser, salt addition, pH, temperature, stirring rate and extraction time. This review discusses the principle, optimization variables, advantages and disadvantages and some selected applications of DLLME-SFOD in water, food and biomedical analysis. Copyright © 2017 Elsevier B.V. All rights reserved.

  14. Charged Nanowire-Directed Growth of Amorphous Calcium Carbonate Nanosheets in a Mixed Solvent for Biomimetic Composite Films.

    PubMed

    Liu, Yang-Yi; Liu, Lei; Chen, Si-Ming; Chang, Fu-Jia; Mao, Li-Bo; Gao, Huai-Ling; Ma, Tao; Yu, Shu-Hong

    2018-05-22

    Bio-inspired mineralization is an effective way for fabricating complex inorganic materials, which inspires us to develop new methods to synthesize materials with fascinating properties. In this article, we report that the charged tellurium nanowires (TeNWs) can be used as biomacromolecule analogues to direct the formation of amorphous calcium carbonate (ACC) nanosheets (ACCNs) in a mixed solvent. The effects of surface charges and the concentration of the TeNWs on the formation of ACCNs have been investigated. Particularly, the produced ACCNs can be functionalized by Fe 3 O 4 nanoparticles to produce magnetic ACC/Fe 3 O 4 hybrid nanosheets that can be used to construct ACC/Fe 3 O 4 composite films through a self-evaporation process. Moreover, sodium alginate-ACC nanocomposite films with remarkable toughness and good transmittance can also be fabricated by using such ACCNs as nanoscale building blocks. This mineralization approach in a mixed solvent using charged TeNWs as biomacromolecule analogues provides a new way for the synthesis of ACCNs, which can be used as nanoscale building blocks for the fabrication of biomimetic composite films.

  15. Solvent control of charge transfer excited state relaxation pathways in [Fe(2,2'-bipyridine)(CN) 4] 2-

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Kjær, Kasper S.; Kunnus, Kristjan; Harlang, Tobias C. B.

    The excited state dynamics of solvated [Fe(bpy)(CN) 4] 2-, where bpy = 2,2'-bipyridine, show significant sensitivity to the solvent Lewis acidity. Using a combination of optical absorption and X-ray emission transient spectroscopies, we have previously shown that the metal to ligand charge transfer (MLCT) excited state of [Fe(bpy)(CN) 4] 2- has a 19 picosecond lifetime and no discernable contribution from metal centered (MC) states in weak Lewis acid solvents, such as dimethyl sulfoxide and acetonitrile. Here, in the present work, we use the same combination of spectroscopic techniques to measure the MLCT excited state relaxation dynamics of [Fe(bpy)(CN) 4] 2-more » in water, a strong Lewis acid solvent. The charge-transfer excited state is now found to decay in less than 100 femtoseconds, forming a quasi-stable metal centered excited state with a 13 picosecond lifetime. We find that this MC excited state has triplet ( 3MC) character, unlike other reported six-coordinate Fe(II)-centered coordination compounds, which form MC quintet ( 5MC) states. The solvent dependent changes in excited state non-radiative relaxation for [Fe(bpy)(CN) 4] 2- allows us to infer the influence of the solvent on the electronic structure of the complex. Lastly, the robust characterization of the dynamics and optical spectral signatures of the isolated 3MC intermediate provides a strong foundation for identifying 3MC intermediates in the electronic excited state relaxation mechanisms of similar Fe-centered systems being developed for solar applications.« less

  16. Solvent control of charge transfer excited state relaxation pathways in [Fe(2,2'-bipyridine)(CN) 4] 2-

    DOE PAGES

    Kjær, Kasper S.; Kunnus, Kristjan; Harlang, Tobias C. B.; ...

    2018-01-19

    The excited state dynamics of solvated [Fe(bpy)(CN) 4] 2-, where bpy = 2,2'-bipyridine, show significant sensitivity to the solvent Lewis acidity. Using a combination of optical absorption and X-ray emission transient spectroscopies, we have previously shown that the metal to ligand charge transfer (MLCT) excited state of [Fe(bpy)(CN) 4] 2- has a 19 picosecond lifetime and no discernable contribution from metal centered (MC) states in weak Lewis acid solvents, such as dimethyl sulfoxide and acetonitrile. Here, in the present work, we use the same combination of spectroscopic techniques to measure the MLCT excited state relaxation dynamics of [Fe(bpy)(CN) 4] 2-more » in water, a strong Lewis acid solvent. The charge-transfer excited state is now found to decay in less than 100 femtoseconds, forming a quasi-stable metal centered excited state with a 13 picosecond lifetime. We find that this MC excited state has triplet ( 3MC) character, unlike other reported six-coordinate Fe(II)-centered coordination compounds, which form MC quintet ( 5MC) states. The solvent dependent changes in excited state non-radiative relaxation for [Fe(bpy)(CN) 4] 2- allows us to infer the influence of the solvent on the electronic structure of the complex. Lastly, the robust characterization of the dynamics and optical spectral signatures of the isolated 3MC intermediate provides a strong foundation for identifying 3MC intermediates in the electronic excited state relaxation mechanisms of similar Fe-centered systems being developed for solar applications.« less

  17. Colloidal transport phenomena of milk components during convective droplet drying.

    PubMed

    Fu, Nan; Woo, Meng Wai; Chen, Xiao Dong

    2011-10-15

    Material segregation has been reported for industrial spray-dried milk powders, which indicates potential material migration during drying process. The relevant colloidal transport phenomenon and the underlying mechanism are still under debate. This study extended the glass-filament single droplet drying technique to observe not only the drying behaviour but also the dissolution behaviour of the correspondingly dried single particle. At progressively longer drying stage, a solvent droplet (water or ethanol) was attached to the semi-dried milk particle and the interaction between the solvent and the particle was video-recorded. Based on the different dissolution and wetting behaviours observed, material migration during milk drying was studied. Fresh skim milk and fresh whole milk were investigated using water and ethanol as solvents. Fat started to accumulate on the surface as soon as drying was started. At the initial stage of drying, the fat layer remained thin and the solubility of the semi-dried milk particle was much affected by lactose and protein present underneath the fat layer. Fat kept accumulating at the surface as drying progressed and the accumulation was completed by the middle stage of drying. The results from drying of model milk materials (pure sodium caseinate solution and lactose/sodium caseinate mixed solution) supported the colloidal transport phenomena observed for the milk drying. When mixed with lactose, sodium caseinate did not form an apparent solvent-resistant protein shell during drying. The extended technique of glass-filament single droplet approach provides a powerful tool in examining the solubility of individual particle after drying. Copyright © 2011 Elsevier B.V. All rights reserved.

  18. Raman study of bulk-heterojunction morphology in photoactive layers treated with solvent-vapor annealing

    NASA Astrophysics Data System (ADS)

    Onojima, Norio; Ishima, Yasuhisa; Izumi, Daisuke; Takahashi, Kazuyuki

    2018-03-01

    The effect of solvent-vapor annealing (SVA) on bulk-heterojunction morphology in photoactive layers composed of poly(3-hexylthiophene-2,5-diyl) (P3HT) and [6,6]-phenyl-C61-butyric acid methyl ester (PCBM) was analyzed using Raman spectroscopy. We prepared the photoactive layers by electrostatic spray deposition (ESD) and fabricated organic photovoltaic devices with a conventional cell structure. Although postdeposition annealing can be omitted when the photoactive layer is deposited using ESD under dry condition, the surface is relatively rough owing to the existence of a number of droplet traces. The SVA treatment can eliminate such droplet traces, while excessive SVA resulted in a significant decrease in open-circuit voltage. The Raman study of the bulk-heterojunction morphology demonstrated the accumulation of P3HT molecules on the surface during SVA, which induced the recombination of photogenerated charges at the interface of the cathode/photoactive layer and thereby decreased the open-circuit voltage.

  19. Liquid droplet radiator development status

    NASA Technical Reports Server (NTRS)

    White, K. Alan, III

    1987-01-01

    Development of the Liquid Droplet Radiator (LDR) is described. Significant published results of previous investigators are presented, and work currently in progress is discussed. Several proposed LDR configurations are described, and the rectangular and triangular configurations currently of most interest are examined. Development of the droplet generator, collector, and auxiliary components are discussed. Radiative performance of a droplet sheet is considered, and experimental results are seen to be in very good agreement with analytical predictions. The collision of droplets in the droplet sheet, the charging of droplets by the space plasma, and the effect of atmospheric drag on the droplet sheet are shown to be of little consequence, or can be minimized by proper design. The LDR is seen to be less susceptible than conventional technology to the effects of micrometeoroids or hostile threats. The identification of working fluids which are stable in the orbital environments of interest is also made. Methods for reducing spacecraft contamination from an LDR to an acceptable level are discussed. Preliminary results of microgravity testing of the droplet generator are presented. Possible future NASA and Air Force missions enhanced or enabled by a LDR are also discussed. System studies indicate that the LDR is potentially less massive than heat pipe radiators. Planned microgravity testing aboard the Shuttle or space station is seen to be a logical next step in LDR development.

  20. Electrically Controllable Microparticle Synthesis and Digital Microfluidic Manipulation by Electric-Field-Induced Droplet Dispensing into Immiscible Fluids

    PubMed Central

    Um, Taewoong; Hong, Jiwoo; Im, Do Jin; Lee, Sang Joon; Kang, In Seok

    2016-01-01

    The dispensing of tiny droplets is a basic and crucial process in a myriad of applications, such as DNA/protein microarray, cell cultures, chemical synthesis of microparticles, and digital microfluidics. This work systematically demonstrates droplet dispensing into immiscible fluids through electric charge concentration (ECC) method. It exhibits three main modes (i.e., attaching, uniform, and bursting modes) as a function of flow rates, applied voltages, and gap distances between the nozzle and the oil surface. Through a conventional nozzle with diameter of a few millimeters, charged droplets with volumes ranging from a few μL to a few tens of nL can be uniformly dispensed into the oil chamber without reduction in nozzle size. Based on the features of the proposed method (e.g., formation of droplets with controllable polarity and amount of electric charge in water and oil system), a simple and straightforward method is developed for microparticle synthesis, including preparation of colloidosomes and fabrication of Janus microparticles with anisotropic internal structures. Finally, a combined system consisting of ECC-induced droplet dispensing and electrophoresis of charged droplet (ECD)-driven manipulation systems is constructed. This integrated platform will provide increased utility and flexibility in microfluidic applications because a charged droplet can be delivered toward the intended position by programmable electric control. PMID:27534580

  1. Cyclic variations of fuel-droplet distribution during the early intake stroke of a lean-burn stratified-charge spark-ignition engine

    NASA Astrophysics Data System (ADS)

    Aleiferis, P. G.; Hardalupas, Y.; Taylor, A. M. K. P.; Ishii, K.; Urata, Y.

    2005-11-01

    Lean-burn spark-ignition engines exhibit higher efficiency and lower specific emissions in comparison with stoichiometrically charged engines. However, as the air-to-fuel (A/F) ratio of the mixture is made leaner than stoichiometric, cycle-by-cycle variations in the early stages of in-cylinder combustion, and subsequent indicated mean effective pressure (IMEP), become more pronounced and limit the range of lean-burn operation. Viable lean-burn engines promote charge stratification, the mixture near the spark plug being richer than the cylinder volume averaged value. Recent work has shown that cycle-by-cycle variations in the early stages of combustion in a stratified-charge engine can be associated with variations in both the local value of A/F ratio near the spark plug around ignition timing, as well as in the volume averaged value of the A/F ratio. The objective of the current work was to identify possible sources of such variability in A/F ratio by studying the in-cylinder field of fuel-droplet distribution during the early intake stroke. This field was visualised in an optical single-cylinder 4-valve pentroof-type spark-ignition engine by means of laser-sheet illumination in planes parallel to the cylinder head gasket 6 and 10 mm below the spark plug. The engine was run with port-injected isooctane at 1500 rpm with 30% volumetric efficiency and air-to-fuel ratio corresponding to both stoichiometric firing (A/F=15, Φ =1.0) and mixture strength close to the lean limit of stable operation (A/F=22, Φ =0.68). Images of Mie intensity scattered by the cloud of fuel droplets were acquired on a cycle-by-cycle basis. These were studied in order to establish possible correlations between the cyclic variations in size, location and scattered-light intensity of the cloud of droplets with the respective variations in IMEP. Because of the low level of Mie intensity scattered by the droplets and because of problems related to elastic scattering on the walls of the combustion

  2. Droplet breakup dynamics of weakly viscoelastic fluids

    NASA Astrophysics Data System (ADS)

    Marshall, Kristin; Walker, Travis

    2016-11-01

    The addition of macromolecules to solvent, even in dilute quantities, can alter a fluid's response in an extensional flow. For low-viscosity fluids, the presence of elasticity may not be apparent when measured using a standard rotational rheometer, yet it may still alter the response of a fluid when undergoing an extensional deformation, especially at small length scales where elastic effects are enhanced. Applications such as microfluidics necessitate investigating the dynamics of fluids with elastic properties that are not pronounced at large length scales. In the present work, a microfluidic cross-slot configuration is used to study the effects of elasticity on droplet breakup. Droplet breakup and the subsequent iterated-stretching - where beads form along a filament connecting two primary droplets - were observed for a variety of material and flow conditions. We present a relationship on the modes of bead formation and how and when these modes will form based on key parameters such as the properties of the outer continuous-phase fluid. The results are vital not only for simulating the droplet breakup of weakly viscoelastic fluids but also for understanding how the droplet breakup event can be used for characterizing the extensional properties of weakly-viscoelastic fluids.

  3. Quantitative Raman microspectroscopy for water permeability parameters at a droplet interface bilayer.

    PubMed

    Braziel, S; Sullivan, K; Lee, S

    2018-01-29

    Using confocal Raman microspectroscopy, we derive parameters for bilayer water transport across an isolated nanoliter aqueous droplet pair. For a bilayer formed with two osmotically imbalanced and adherent nanoliter aqueous droplets in a surrounding oil solvent, a droplet interface bilayer (DIB), the water permeability coefficient across the lipid bilayer was determined from monitoring the Raman scattering from the C[triple bond, length as m-dash]N stretching mode of K 3 Fe(CN) 6 as a measure of water uptake into the swelling droplet of a DIB pair. We also derive passive diffusional permeability coefficient for D 2 O transport across a droplet bilayer using O-D Raman signal. This method provides a significant methodological advance in determining water permeability coefficients in a convenient and reliable way.

  4. Exploring Chemistry in Microcompartments Using Guided Droplet Collisions in a Branched Quadrupole Trap Coupled to a Single Droplet, Paper Spray Mass Spectrometer

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Jacobs, Michael I.; Davies, James F.; Lee, Lance

    Recent studies suggest that reactions in aqueous microcompartments can occur at significantly different rates than those in the bulk. Most studies have used electrospray to generate a polydisperse source of highly charged microdroplets, leading to multiple confounding factors potentially influencing reaction rates (e.g., evaporation, charge, and size). Thus, the underlying mechanism for the observed enhancement remains unclear. We present a new type of electrodynamic balance - the branched quadrupole trap (BQT) - which can be used to study reactions in microdroplets in a controlled environment. The BQT allows for condensed phase chemical reactions to be initiated by colliding droplets withmore » different reactants and levitating the merged droplet indefinitely. The performance of the BQT is characterized in several ways. Sub-millisecond mixing times as fast as ~400 μs are measured for low velocity (~0.1 m/s) collisions of droplets with <40 μm diameters. The reaction of o-phthalaldehyde (OPA) with alanine in the presence of dithiolthreitol is measured using both fluorescence spectroscopy and single droplet paper spray mass spectrometry. The bimolecular rate constant for reaction of alanine with OPA is found to be 84 ± 10 and 67 ± 6 M -1s -1 in a 30 μm radius droplet and bulk solution, respectively, which demonstrates that bimolecular reaction rate coefficients can be quantified using merged microdroplets and that merged droplets can be used to study rate enhancements due to compartmentalization. Products of the reaction of OPA with alanine are detected in single droplets using paper spray mass spectrometry. Finally, we demonstrate that single droplets with <100 pg of analyte can easily be studied using single droplet mass spectrometry.« less

  5. Exploring Chemistry in Microcompartments Using Guided Droplet Collisions in a Branched Quadrupole Trap Coupled to a Single Droplet, Paper Spray Mass Spectrometer

    DOE PAGES

    Jacobs, Michael I.; Davies, James F.; Lee, Lance; ...

    2017-10-19

    Recent studies suggest that reactions in aqueous microcompartments can occur at significantly different rates than those in the bulk. Most studies have used electrospray to generate a polydisperse source of highly charged microdroplets, leading to multiple confounding factors potentially influencing reaction rates (e.g., evaporation, charge, and size). Thus, the underlying mechanism for the observed enhancement remains unclear. We present a new type of electrodynamic balance - the branched quadrupole trap (BQT) - which can be used to study reactions in microdroplets in a controlled environment. The BQT allows for condensed phase chemical reactions to be initiated by colliding droplets withmore » different reactants and levitating the merged droplet indefinitely. The performance of the BQT is characterized in several ways. Sub-millisecond mixing times as fast as ~400 μs are measured for low velocity (~0.1 m/s) collisions of droplets with <40 μm diameters. The reaction of o-phthalaldehyde (OPA) with alanine in the presence of dithiolthreitol is measured using both fluorescence spectroscopy and single droplet paper spray mass spectrometry. The bimolecular rate constant for reaction of alanine with OPA is found to be 84 ± 10 and 67 ± 6 M -1s -1 in a 30 μm radius droplet and bulk solution, respectively, which demonstrates that bimolecular reaction rate coefficients can be quantified using merged microdroplets and that merged droplets can be used to study rate enhancements due to compartmentalization. Products of the reaction of OPA with alanine are detected in single droplets using paper spray mass spectrometry. Finally, we demonstrate that single droplets with <100 pg of analyte can easily be studied using single droplet mass spectrometry.« less

  6. Charge-induced secondary atomization in diffusion flames of electrostatic sprays

    NASA Technical Reports Server (NTRS)

    Gomez, Alessandro; Chen, Gung

    1994-01-01

    The combustion of electrostatic sprays of heptane in laminar counterflow diffusion flames was experimentally studied by measuring droplet size and velocity distributions, as well as the gas-phase temperature. A detailed examination of the evolution of droplet size distribution as droplets approach the flame shows that, if substantial evaporation occurs before droplets interact with the flame, an initially monodisperse size distribution becomes bimodal. A secondary sharp peak in the size histogram develops in correspondence of diameters about one order of magnitude smaller than the mean. No evaporation mechanism can account for the development of such bimodality, that can be explained only in terms of a disintegration of droplets into finer fragments of size much smaller than that of the parent. Other evidence in support of this interpretation is offered by the measurements of droplet size-velocity correlation and velocity component distributions, showing that, as a consequence of the ejection process, the droplets responsible for the secondary peak have velocities uncorrelated with the mean flow. The fission is induced by the electric charge. When a droplet evaporates, in fact, the electric charge density on the droplet surface increases while the droplet shrinks, until the so-called Rayleigh limit is reached at which point the repulsion of electric charges overcomes the surface tension cohesive force, ultimately leading to a disintegraton into finer fragments. We report on the first observation of such fissions in combustion environments. If, on the other hand, insufficient evaporation has occurred before droplets enter the high temperature region, there appears to be no significant evidence of bimodality in their size distribution. In this case, in fact, the concentration of flame chemi-ions or, in the case of positively charged droplets, electrons may be sufficient for them to neutralize the charge on the droplets and to prevent disruption.

  7. Directional Electrostatic Accretion Process Employing Acoustic Droplet Formation

    NASA Technical Reports Server (NTRS)

    Oeftering, Richard (Inventor)

    1998-01-01

    The present invention is directed to an apparatus for manufacturing a free standing solid metal part. In the present invention, metal droplets are ejected in a nozzleless fashion from a free surface pool of molten metal by applying focused acoustic radiation pressure. The acoustic radiation pressure is produced by high intensity acoustic tone bursts emitted from an acoustic source positioned at the bottom of the pool which directs the acoustic energy at the pool surface. The metal droplets are electrostatically charged so their trajectory can be controlled by electric fields that guide the droplets to predetermined points on a target. The droplets impinge upon the target and solidify with the target material. The accretion of the electrostatically directed solidified droplets forms the free standing metal part.

  8. Increasing Protein Charge State When Using Laser Electrospray Mass Spectrometry

    NASA Astrophysics Data System (ADS)

    Karki, Santosh; Flanigan, Paul M.; Perez, Johnny J.; Archer, Jieutonne J.; Levis, Robert J.

    2015-05-01

    Femtosecond (fs) laser vaporization is used to transfer cytochrome c, myoglobin, lysozyme, and ubiquitin from the condensed phase into an electrospray (ES) plume consisting of a mixture of a supercharging reagent, m-nitrobenzyl alcohol ( m-NBA), and trifluoroacetic acid (TFA), acetic acid (AA), or formic acid (FA). Interaction of acid-sensitive proteins like cytochrome c and myoglobin with the highly charged ES droplets resulted in a shift to higher charge states in comparison with acid-stable proteins like lysozyme and ubiquitin. Laser electrospray mass spectrometry (LEMS) measurements showed an increase in both the average charge states (Zavg) and the charge state with maximum intensity (Zmode) for acid-sensitive proteins compared with conventional electrospray ionization mass spectrometry (ESI-MS) under equivalent solvent conditions. A marked increase in ion abundance of higher charge states was observed for LEMS in comparison with conventional electrospray for cytochrome c (ranging from 19+ to 21+ versus 13+ to 16+) and myoglobin (ranging from 19+ to 26+ versus 18+ to 21+) using an ES solution containing m-NBA and TFA. LEMS measurements as a function of electrospray flow rate yielded increasing charge states with decreasing flow rates for cytochrome c and myoglobin.

  9. Control of aqueous droplets using magnetic and electrostatic forces.

    PubMed

    Ohashi, Tetsuo; Kuyama, Hiroki; Suzuki, Koichi; Nakamura, Shin

    2008-04-07

    Basic control operations were successfully performed on an aqueous droplet using both magnetic and electrostatic forces. In our droplet-based microfluidics, magnetic beads were incorporated in an aqueous droplet as a force mediator. This report describes droplet anchoring and separation of the beads from the droplet using a combination of magnetic and electrostatic forces. When an aqueous droplet is placed in an oil-filled reservoir, the droplet sinks to the bottom, under which an electrode had been placed. The droplet was adsorbed (or anchored) to the bottom surface on the electrode when a DC voltage was applied to the electrode. The magnetic beads were removed with magnetic force after the droplet had been anchored. Surfactant addition into droplet solution was very effective for the elimination of electric charge, which resulted in the stable adsorption of a droplet to hydrophobic substrate under an applied voltage of DC 0.5-3 kV. In a sequential process, small volume of aqueous liquid was successfully transferred using both magnetic and electrostatic forces.

  10. Probing Ion Transfer across Liquid-Liquid Interfaces by Monitoring Collisions of Single Femtoliter Oil Droplets on Ultramicroelectrodes.

    PubMed

    Deng, Haiqiang; Dick, Jeffrey E; Kummer, Sina; Kragl, Udo; Strauss, Steven H; Bard, Allen J

    2016-08-02

    We describe a method of observing collisions of single femtoliter (fL) oil (i.e., toluene) droplets that are dispersed in water on an ultramicroelectrode (UME) to probe the ion transfer across the oil/water interface. The oil-in-water emulsion was stabilized by an ionic liquid, in which the oil droplet trapped a highly hydrophobic redox probe, rubrene. The ionic liquid also functions as the supporting electrolyte in toluene. When the potential of the UME was biased such that rubrene oxidation would be possible when a droplet collided with the electrode, no current spikes were observed. This implies that the rubrene radical cation is not hydrophilic enough to transfer into the aqueous phase. We show that current spikes are observed when tetrabutylammonium trifluoromethanesulfonate or tetrahexylammonium hexafluorophosphate are introduced into the toluene phase and when tetrabutylammonium perchlorate is introduced into the water phase, implying that the ion transfer facilitates electron transfer in the droplet collisions. The current (i)-time (t) behavior was evaluated quantitatively, which indicated the ion transfer is fast and reversible. Furthermore, the size of these emulsion droplets can also be calculated from the electrochemical collision. We further investigated the potential dependence on the electrochemical collision response in the presence of tetrabutylammonium trifluoromethanesulfonate in toluene to obtain the formal ion transfer potential of tetrabutylammonium across the toluene/water interface, which was determined to be 0.754 V in the inner potential scale. The results yield new physical insights into the charge balance mechanism in emulsion droplet collisions and indicate that the electrochemical collision technique can be used to probe formal ion transfer potentials between water and solvents with very low (ε < 5) dielectric constants.

  11. Particle formation in the emulsion-solvent evaporation process.

    PubMed

    Staff, Roland H; Schaeffel, David; Turshatov, Andrey; Donadio, Davide; Butt, Hans-Jürgen; Landfester, Katharina; Koynov, Kaloian; Crespy, Daniel

    2013-10-25

    The mechanism of particle formation from submicrometer emulsion droplets by solvent evaporation is revisited. A combination of dynamic light scattering, fluorescence resonance energy transfer, zeta potential measurements, and fluorescence cross-correlation spectroscopy is used to analyze the colloids during the evaporation process. It is shown that a combination of different methods yields reliable and quantitative data for describing the fate of the droplets during the process. The results indicate that coalescence plays a minor role during the process; the relatively large size distribution of the obtained polymer colloids can be explained by the droplet distribution after their formation. Copyright © 2013 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.

  12. Modified two-step emulsion solvent evaporation technique for fabricating biodegradable rod-shaped particles in the submicron size range.

    PubMed

    Safari, Hanieh; Adili, Reheman; Holinstat, Michael; Eniola-Adefeso, Omolola

    2018-05-15

    Though the emulsion solvent evaporation (ESE) technique has been previously modified to produce rod-shaped particles, it cannot generate small-sized rods for drug delivery applications due to the inherent coupling and contradicting requirements for the formation versus stretching of droplets. The separation of the droplet formation from the stretching step should enable the creation of submicron droplets that are then stretched in the second stage by manipulation of the system viscosity along with the surface-active molecule and oil-phase solvent. A two-step ESE protocol is evaluated where oil droplets are formed at low viscosity followed by a step increase in the aqueous phase viscosity to stretch droplets. Different surface-active molecules and oil phase solvents were evaluated to optimize the yield of biodegradable PLGA rods. Rods were assessed for drug loading via an imaging agent and vascular-targeted delivery application via blood flow adhesion assays. The two-step ESE method generated PLGA rods with major and minor axis down to 3.2 µm and 700 nm, respectively. Chloroform and sodium metaphosphate was the optimal solvent and surface-active molecule, respectively, for submicron rod fabrication. Rods demonstrated faster release of Nile Red compared to spheres and successfully targeted an inflamed endothelium under shear flow in vitro and in vivo. Copyright © 2018 Elsevier Inc. All rights reserved.

  13. Monodisperse hydrogel microspheres by forced droplet formation in aqueous two-phase systems.

    PubMed

    Ziemecka, Iwona; van Steijn, Volkert; Koper, Ger J M; Rosso, Michel; Brizard, Aurelie M; van Esch, Jan H; Kreutzer, Michiel T

    2011-02-21

    This paper presents a method to form micron-sized droplets in an aqueous two-phase system (ATPS) and to subsequently polymerize the droplets to produce hydrogel beads. Owing to the low interfacial tension in ATPS, droplets do not easily form spontaneously. We enforce the formation of drops by perturbing an otherwise stable jet that forms at the junction where the two aqueous streams meet. This is done by actuating a piezo-electric bending disc integrated in our device. The influence of forcing amplitude and frequency on jet breakup is described and related to the size of monodisperse droplets with a diameter in the range between 30 and 60 μm. Rapid on-chip polymerization of derivatized dextran inside the droplets created monodisperse hydrogel particles. This work shows how droplet-based microfluidics can be used in all-aqueous, surfactant-free, organic-solvent-free biocompatible two-phase environment.

  14. An evaporation model of colloidal suspension droplets

    NASA Astrophysics Data System (ADS)

    Sartori, Silvana; Li\\ Nán, Amable; Lasheras, Juan C.

    2009-11-01

    Colloidal suspensions of polymers in water or other solvents are widely used in the pharmaceutical industry to coat tablets with different agents. These allow controlling the rate at which the drug is delivered, taste or physical appearance. The coating is performed by simultaneously spraying and drying the tablets with the colloidal suspension at moderately high temperatures. The spreading of the coating on the pills surface depends on the droplet Webber and Reynolds numbers, angle of impact, but more importantly on the rheological properties of the drop. We present a model for the evaporation of a colloidal suspension droplet in a hot air environment with temperatures substantially lower than the boiling temperature of the carrier fluid. As the liquid vaporizes from the surface, a compacting front advances into the droplet faster than the liquid surface regresses, forming a shell of a porous medium where the particles reach their maximum packing density. While the surface regresses, the evaporation rate is determined by both the rate at which heat is transported to the droplet surface and the rate at which liquid vapor is diffused away from it. This regime continues until the compacting front reaches the center of the droplet, at which point the evaporation rate is drastically reduced.

  15. Enhancing droplet deposition through in-situ precipitation

    NASA Astrophysics Data System (ADS)

    Damak, Maher; Mahmoudi, Seyed Reza; Hyder, Md Nasim; Varanasi, Kripa K.

    2016-08-01

    Retention of agricultural sprays on plant surfaces is an important challenge. Bouncing of sprayed pesticide droplets from leaves is a major source of soil and groundwater pollution and pesticide overuse. Here we report a method to increase droplet deposition through in-situ formation of hydrophilic surface defects that can arrest droplets during impact. Defects are created by simultaneously spraying oppositely charged polyelectrolytes that induce surface precipitation when two droplets come into contact. Using high-speed imaging, we study the coupled dynamics of drop impact and surface precipitate formation. We develop a physical model to estimate the energy dissipation by the defects and predict the transition from bouncing to sticking. We demonstrate macroscopic enhancements in spray retention and surface coverage for natural and synthetic non-wetting surfaces and provide insights into designing effective agricultural sprays.

  16. Two-dimensional fluid droplet arrays generated using a single nozzle

    DOEpatents

    Lee, Eric R.; Perl, Martin L.

    1999-11-02

    Amplitudes of drive pulses received by a horizontally-placed dropper determine the horizontal displacements of droplets relative to an ejection aperture of the dropper. The drive pulses are varied such that the dropper generates a two-dimensional array of vertically-falling droplets. Vertical and horizontal interdroplet spacings may be varied in real time. Applications include droplet analysis experiments such as Millikan fractional charge searches and aerosol characterization, as well as material deposition applications.

  17. Absorption of charged particulate surfactants in microfluidics

    NASA Astrophysics Data System (ADS)

    Kong, Tiantian; Liu, Zhou; Yao, Xiaoxue; Liu, Yaming

    2017-11-01

    We use microfluidics to uncouple the generation of Pickering emulsion droplets and stability analysis against coalescence. By designing the microchannels, we control the packing time for charged particles arriving at the droplet interfaces, and subsequently test the droplet stability in a coalescence chamber. The critical particle coverage on interfaces that prevents coalescence are estimated by an adsorption model. We further investigate the dependence of the critical particle coverage on its properties such as particle sizes, surface charge densities, and bulk concentrations. Our studies are potentially beneficial to the applications involving particle-stabilized droplets including cosmetics, food products, and oil recovery. NSFC 11504238,JCYJ20160308092144035,2016A050503048.

  18. Solvent-based and solvent-free characterization of low solubility and low molecular weight polyamides by mass spectrometry: a complementary approach.

    PubMed

    Barrère, Caroline; Hubert-Roux, Marie; Lange, Catherine M; Rejaibi, Majed; Kebir, Nasreddine; Désilles, Nicolas; Lecamp, Laurence; Burel, Fabrice; Loutelier-Bourhis, Corinne

    2012-06-15

    Polyamides (PA) belong to the most used classes of polymers because of their attractive chemical and mechanical properties. In order to monitor original PA design, it is essential to develop analytical methods for the characterization of these compounds that are mostly insoluble in usual solvents. A low molecular weight polyamide (PA11), synthesized with a chain limiter, has been used as a model compound and characterized by matrix-assisted laser desorption/ionization time-of-flight mass spectrometry (MALDI-TOF-MS). In the solvent-based approach, specific solvents for PA, i.e. trifluoroacetic acid (TFA) and hexafluoroisopropanol (HFIP), were tested. Solvent-based sample preparation methods, dried-droplet and thin layer, were optimized through the choice of matrix and salt. Solvent-based (thin layer) and solvent-free methods were then compared for this low solubility polymer. Ultra-high-performance liquid chromatography/electrospray ionization (UHPLC/ESI)-TOF-MS analyses were then used to confirm elemental compositions through accurate mass measurement. Sodium iodide (NaI) and 2,5-dihydroxybenzoic acid (2,5-DHB) are, respectively, the best cationizing agent and matrix. The dried-droplet sample preparation method led to inhomogeneous deposits, but the thin-layer method could overcome this problem. Moreover, the solvent-free approach was the easiest and safest sample preparation method giving equivalent results to solvent-based methods. Linear as well as cyclic oligomers were observed. Although the PA molecular weights obtained by MALDI-TOF-MS were lower than those obtained by (1)H NMR and acido-basic titration, this technique allowed us to determine the presence of cyclic and linear species, not differentiated by the other techniques. TFA was shown to induce modification of linear oligomers that permitted cyclic and linear oligomers to be clearly highlighted in spectra. Optimal sample preparation conditions were determined for the MALDI-TOF-MS analysis of PA11, a

  19. Interdroplet attractive forces in AOT water-in-oil microemulsions formed in subcritical and supercritical solvents

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Tingey, J.M.; Fulton, J.L.; Smith, R.D.

    1990-03-08

    The van der Waals attractive interactions between aqueous droplets in water-in-oil type microemulsions have been investigated for a range of continuous-phase solvents including the alkanes from methane to isooctane and the noble gases, krypton and xenon. Hamaker constants for water droplets with surfactant shells of the sodium bis(2-ethylhexyl) sulfosuccinate (AOT) in subcritical and supercritical solvents were calculated by using Lifshitz theory and the resulting interaction potential calculations qualitatively account for many features of the phase behavior of these systems.

  20. Material forming apparatus using a directed droplet stream

    DOEpatents

    Holcomb, David E.; Viswanathan, Srinath; Blue, Craig A.; Wilgen, John B.

    2000-01-01

    Systems and methods are described for rapidly forming precision metallic and intermetallic alloy net shape parts directly from liquid metal droplets. A directed droplet deposition apparatus includes a crucible with an orifice for producing a jet of material, a jet destabilizer, a charging structure, a deflector system, and an impact zone. The systems and methods provide advantages in that fully dense, microstructurally controlled parts can be fabricated at moderate cost.

  1. Enhancing droplet deposition through in-situ precipitation

    PubMed Central

    Damak, Maher; Mahmoudi, Seyed Reza; Hyder, Md Nasim; Varanasi, Kripa K.

    2016-01-01

    Retention of agricultural sprays on plant surfaces is an important challenge. Bouncing of sprayed pesticide droplets from leaves is a major source of soil and groundwater pollution and pesticide overuse. Here we report a method to increase droplet deposition through in-situ formation of hydrophilic surface defects that can arrest droplets during impact. Defects are created by simultaneously spraying oppositely charged polyelectrolytes that induce surface precipitation when two droplets come into contact. Using high-speed imaging, we study the coupled dynamics of drop impact and surface precipitate formation. We develop a physical model to estimate the energy dissipation by the defects and predict the transition from bouncing to sticking. We demonstrate macroscopic enhancements in spray retention and surface coverage for natural and synthetic non-wetting surfaces and provide insights into designing effective agricultural sprays. PMID:27572948

  2. Explicit Solvent Simulations of Friction between Brush Layers of Charged and Neutral Bottle-Brush Macromolecules

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Carrillo, Jan-Michael; Brown, W Michael; Dobrynin, Andrey

    2012-01-01

    We study friction between charged and neutral brush layers of bottle-brush macromolecules using molecular dynamics simulations. In our simulations the solvent molecules were treated explicitly. The deformation of the bottle-brush macromolecules under the shear were studied as a function of the substrate separation and shear stress. For charged bottle-brush layers we study effect of the added salt on the brush lubricating properties to elucidate factors responsible for energy dissipation in charged and neutral brush systems. Our simulations have shown that for both charged and neutral brush systems the main deformation mode of the bottle-brush macromolecule is associated with the backbonemore » deformation. This deformation mode manifests itself in the backbone deformation ratio, , and shear viscosity, , to be universal functions of the Weissenberg number W. The value of the friction coefficient, , and viscosity, , are larger for the charged bottle-brush coatings in comparison with those for neutral brushes at the same separation distance, D, between substrates. The additional energy dissipation generated by brush sliding in charged bottle-brush systems is due to electrostatic coupling between bottle-brush and counterion motion. This coupling weakens as salt concentration, cs, increases resulting in values of the viscosity, , and friction coefficient, , approaching corresponding values obtained for neutral brush systems.« less

  3. High-throughput controllable generation of droplet arrays with low consumption

    NASA Astrophysics Data System (ADS)

    Lin, Yinyin; Wu, Zhongsheng; Gao, Yibo; Wu, Jinbo; Wen, Weijia

    2018-06-01

    We describe a controllable sliding method for fabricating millions of isolated femto- to nanoliter-sized droplets with defined volume, geometry and position and a speed of up to 375 kHz. In this work, without using a superhydrophobic or superoleophobic surface, arrays of droplets are instantly formed on the patterned substrate by sliding a strip of liquid, including water, low-surface-tension organic solvents and solution, along the substrate. To precisely control the volume of the droplets, we systemically investigate the effects of the size of the wettable pattern, the viscosity of the liquid and sliding speed, which were found to vary independently to tune the height and volume of the droplets. Through this method, we successfully fabricated an oriented single metal-organic framework crystal array with control over their XY positioning on the surface, as characterized by microscopy and X-ray diffraction (XRD) techniques.

  4. Spectrophotometric and spectroscopic studies of charge transfer complex of 1-Naphthylamine as an electron donor with picric acid as an electron acceptor in different polar solvents

    NASA Astrophysics Data System (ADS)

    Singh, Neeti; Ahmad, Afaq

    2010-08-01

    The charge transfer complex of 1-Naphthylamine as a donor with π-acceptor picric acid has been studied spectrophotometrically in different solvents at room temperature. The results indicate that the formation of charge transfer complex is high in less polar solvent. The stoichiometry of the complex was found to be 1:1 by straight line method. The data are analysed in terms of formation constant ( KCT), molar extinction coefficient ( ɛCT), standard free energy (Δ G o), oscillator strength ( ƒ), transition dipole moment ( μ EN), resonance energy ( R N) and ionization potential ( I D). It is concluded that the formation constant ( KCT) of the complex is found to be depends upon the nature of both electron acceptor and donor and also on the polarity of solvents. Further the charge transfer molecular complex between picric acid and 1-Naphthylamine is stabilized by hydrogen bonding.

  5. Energy gap law of electron transfer in nonpolar solvents.

    PubMed

    Tachiya, M; Seki, Kazuhiko

    2007-09-27

    We investigate the energy gap law of electron transfer in nonpolar solvents for charge separation and charge recombination reactions. In polar solvents, the reaction coordinate is given in terms of the electrostatic potentials from solvent permanent dipoles at solutes. In nonpolar solvents, the energy fluctuation due to solvent polarization is absent, but the energy of the ion pair state changes significantly with the distance between the ions as a result of the unscreened strong Coulomb potential. The electron transfer occurs when the final state energy coincides with the initial state energy. For charge separation reactions, the initial state is a neutral pair state, and its energy changes little with the distance between the reactants, whereas the final state is an ion pair state and its energy changes significantly with the mutual distance; for charge recombination reactions, vice versa. We show that the energy gap law of electron-transfer rates in nonpolar solvents significantly depends on the type of electron transfer.

  6. Time dependent charging of layer clouds in the global electric circuit

    NASA Astrophysics Data System (ADS)

    Zhou, Limin; Tinsley, Brian A.

    2012-09-01

    There is much observational data consistent with the hypothesis that the ionosphere-earth current density (Jz) in the global electric circuit, which is modulated by both solar activity and thunderstorm activity, affects atmospheric dynamics and cloud cover. One candidate mechanism involves Jz causing the accumulation of space charge on droplets and aerosol particles, that affects the rate of scavenging of the latter, notably those of Cloud Condensation Nuclei (CCN) and Ice Forming Nuclei (IFN) (Tinsley, 2008, 2010). Space charge is the difference, per unit volume, between total positive and total negative electrical charge that is on droplets, aerosol particles (including the CCN and IFN) and air ions. The cumulative effects of the scavenging in stratiform clouds and aerosol layers in an air mass over the lifetime of the aerosol particles of 1-10 days affects the concentration and size distribution of the CCN, so that in subsequent episodes of cloud formation (including deep convective clouds) there can be effects on droplet size distribution, coagulation, precipitation processes, and even storm dynamics.Because the time scales for charging for some clouds can be long compared to cloud lifetimes, the amount of charge at a given time, and its effect on scavenging, depend more on the charging rate than on the equilibrium charge that would eventually be attained. To evaluate this, a new time-dependent charging model has been developed. The results show that for typical altostratus clouds with typical droplet radii 10 μm and aerosol particles of radius of 0.04 μm, the time constant for charging in response to a change in Jz is about 800 s, which is comparable to cloud formation and dissipation timescales for some cloud situations. The charging timescale is found to be strong functions of altitude and aerosol concentration, with the time constant for droplet charging at 2 km in air with a high concentration of aerosols being about an hour, and for clouds at 10 km in

  7. The role of interfacial water layer in atmospherically relevant charge separation

    NASA Astrophysics Data System (ADS)

    Bhattacharyya, Indrani

    Charge separation at interfaces is important in various atmospheric processes, such as thunderstorms, lightning, and sand storms. It also plays a key role in several industrial processes, including ink-jet printing and electrostatic separation. Surprisingly, little is known about the underlying physics of these charging phenomena. Since thin films of water are ubiquitous, they may play a role in these charge separation processes. This talk will focus on the experimental investigation of the role of a water adlayer in interfacial charging, with relevance to meteorologically important phenomena, such as atmospheric charging due to wave actions on oceans and sand storms. An ocean wave generates thousands of bubbles, which upon bursting produce numerous large jet droplets and small film droplets that are charged. In the 1960s, Blanchard showed that the jet droplets are positively charged. However, the charge on the film droplets was not known. We designed an experiment to exclusively measure the charge on film droplets generated by bubble bursting on pure water and aqueous salt solution surfaces. We measured their charge to be negative and proposed a model where a slight excess of hydroxide ions in the interfacial water layer is responsible for generating these negatively charged droplets. The findings from this research led to a better understanding of the ionic disposition at the air-water interface. Sand particles in a wind-blown sand layer, or 'saltation' layer, become charged due to collisions, so much so, that it can cause lightning. Silica, being hydrophilic, is coated with a water layer even under low-humidity conditions. To investigate the importance of this water adlayer in charging the silica surfaces, we performed experiments to measure the charge on silica surfaces due to contact and collision processes. In case of contact charging, the maximum charge separation occurred at an optimum relative humidity. On the contrary, in collisional charging process, no

  8. Comparison of RESP and IPolQ-Mod Partial Charges for Solvation Free Energy Calculations of Various Solute/Solvent Pairs.

    PubMed

    Mecklenfeld, Andreas; Raabe, Gabriele

    2017-12-12

    The calculation of solvation free energies ΔG solv by molecular simulations is of great interest as they are linked to other physical properties such as relative solubility, partition coefficient, and activity coefficient. However, shortcomings in molecular models can lead to ΔG solv deviations from experimental data. Various studies have demonstrated the impact of partial charges on free energy results. Consequently, calculation methods for partial charges aimed at more accurate ΔG solv predictions are the subject of various studies in the literature. Here we compare two methods to derive partial charges for the general AMBER force field (GAFF), i.e. the default RESP as well as the physically motivated IPolQ-Mod method that implicitly accounts for polarization costs. We study 29 solutes which include characteristic functional groups of drug-like molecules, and 12 diverse solvents were examined. In total, we consider 107 solute/solvent pairs including two water models TIP3P and TIP4P/2005. Comparison with experimental results yields better agreement for TIP3P, regardless of the partial charge scheme. The overall performance of GAFF/RESP and GAFF/IPolQ-Mod is similar, though specific shortcomings in the description of ΔG solv for both RESP and IPolQ-Mod can be identified. However, the high correlation between free energies obtained with GAFF/RESP and GAFF/IPolQ-Mod demonstrates the compatibility between the modified charges and remaining GAFF parameters.

  9. Characterization and application of droplet spray ionization for real-time reaction monitoring.

    PubMed

    Zhang, Hong; Li, Na; Li, Xiao-di; Jiang, Jie; Zhao, Dan-Dan; You, Hong

    2016-08-01

    The ionization source for real-time reaction monitoring has attracted tremendous interest in recent years. We have previously reported a reliable approach in which droplet spray ionization (DSI) was used for monitoring chemical reactions in real-time. Herein, we systematically investigated the characterization and application of DSI for real-time reaction monitoring. Analyte ions are generated by loading a sample solution onto a corner of a microscope cover glass positioned in front of the MS inlet and applying a high voltage to the sample. The tolerance to positioning, solvent effect, spray angle and spray time were investigated. Extension to real-time monitoring of macromolecule reactions was also demonstrated by the charge state change of cytochrome c in the presence of acetic acid. The corner could be positioned within an area of approximately 10 × 6 × 5 mm (x, y, z) in front of the MS inlet. The broad polarities of solvents from methanol to PhF were suitable for DSI. It featured monitoring real-time changes in reactions on the time scale of seconds to minutes. A real-time charge state change of cytochrome c was captured. DSI-MS features ease of use, durability of the spray platform and reusability of the ion source. Eliminating the need for a sample transport capillary, DSI opens a new avenue for the in situ analysis and real-time monitoring of short-lived key reaction intermediates even at subsecond dead times. Copyright © 2016 John Wiley & Sons, Ltd. Copyright © 2016 John Wiley & Sons, Ltd.

  10. Insight into nanoparticle charging mechanism in nonpolar solvents to control the formation of Pt nanoparticle monolayers by electrophoretic deposition

    DOE PAGES

    Cernohorsky, Ondrej; Grym, Jan; Yatskiv, Roman; ...

    2016-08-13

    We report on the formation of Pt nanoparticle monolayers by electrophoretic deposition from nonpolar solvents. First, the growth kinetics of Pt nanoparticles prepared by the reverse micelle technique are described in detail. Second, a model of nanoparticle charging in nonpolar media is discussed and methods to control the nanoparticle charging are proposed. Lastly, essential parameters of the electrophoretic deposition process to control the deposition of nanoparticle monolayers are discussed and mechanisms of their formation are analyzed.

  11. Insight into nanoparticle charging mechanism in nonpolar solvents to control the formation of Pt nanoparticle monolayers by electrophoretic deposition

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Cernohorsky, Ondrej; Grym, Jan; Yatskiv, Roman

    We report on the formation of Pt nanoparticle monolayers by electrophoretic deposition from nonpolar solvents. First, the growth kinetics of Pt nanoparticles prepared by the reverse micelle technique are described in detail. Second, a model of nanoparticle charging in nonpolar media is discussed and methods to control the nanoparticle charging are proposed. Lastly, essential parameters of the electrophoretic deposition process to control the deposition of nanoparticle monolayers are discussed and mechanisms of their formation are analyzed.

  12. Bioreactor droplets from liposome-stabilized all-aqueous emulsions

    NASA Astrophysics Data System (ADS)

    Dewey, Daniel C.; Strulson, Christopher A.; Cacace, David N.; Bevilacqua, Philip C.; Keating, Christine D.

    2014-08-01

    Artificial bioreactors are desirable for in vitro biochemical studies and as protocells. A key challenge is maintaining a favourable internal environment while allowing substrate entry and product departure. We show that semipermeable, size-controlled bioreactors with aqueous, macromolecularly crowded interiors can be assembled by liposome stabilization of an all-aqueous emulsion. Dextran-rich aqueous droplets are dispersed in a continuous polyethylene glycol (PEG)-rich aqueous phase, with coalescence inhibited by adsorbed ~130-nm diameter liposomes. Fluorescence recovery after photobleaching and dynamic light scattering data indicate that the liposomes, which are PEGylated and negatively charged, remain intact at the interface for extended time. Inter-droplet repulsion provides electrostatic stabilization of the emulsion, with droplet coalescence prevented even for submonolayer interfacial coatings. RNA and DNA can enter and exit aqueous droplets by diffusion, with final concentrations dictated by partitioning. The capacity to serve as microscale bioreactors is established by demonstrating a ribozyme cleavage reaction within the liposome-coated droplets.

  13. Bioreactor droplets from liposome-stabilized all-aqueous emulsions.

    PubMed

    Dewey, Daniel C; Strulson, Christopher A; Cacace, David N; Bevilacqua, Philip C; Keating, Christine D

    2014-08-20

    Artificial bioreactors are desirable for in vitro biochemical studies and as protocells. A key challenge is maintaining a favourable internal environment while allowing substrate entry and product departure. We show that semipermeable, size-controlled bioreactors with aqueous, macromolecularly crowded interiors can be assembled by liposome stabilization of an all-aqueous emulsion. Dextran-rich aqueous droplets are dispersed in a continuous polyethylene glycol (PEG)-rich aqueous phase, with coalescence inhibited by adsorbed ~130-nm diameter liposomes. Fluorescence recovery after photobleaching and dynamic light scattering data indicate that the liposomes, which are PEGylated and negatively charged, remain intact at the interface for extended time. Inter-droplet repulsion provides electrostatic stabilization of the emulsion, with droplet coalescence prevented even for submonolayer interfacial coatings. RNA and DNA can enter and exit aqueous droplets by diffusion, with final concentrations dictated by partitioning. The capacity to serve as microscale bioreactors is established by demonstrating a ribozyme cleavage reaction within the liposome-coated droplets.

  14. Electrically Controllable Microparticle Synthesis and Digital Microfluidic Manipulation by Electric-Field-Induced Droplet Dispensing into Immiscible Fluids

    NASA Astrophysics Data System (ADS)

    Um, Taewoong; Hong, Jiwoo; Kang, In Seok

    2016-11-01

    The dispensing of tiny droplets is a basic and crucial process in a myriad of applications, such as DNA/protein microarray, cell cultures, chemical synthesis of microparticles, and digital microfluidics. This work demonstrates the droplet dispensing into immiscible fluids through electric charge concentration (ECC) method. Three main modes (i.e., attaching, uniform and bursting modes) are exhibited as a function of flow rates, applied voltage and gap distance between the nozzle and the oil surface. Through a conventional nozzle with diameter of a few millimeters, charged droplets with volumes ranging from a few μL to a few tens of nL can be uniformly dispensed into the oil chamber without reduction in nozzle size. Based on the features of the proposed method (e.g., formation of droplets with controllable polarity and amount of electric charge in water and oil system), a simple and straightforward method is developed for microparticle synthesis, including preparation for colloidosomes and fabrication of Janus microparticles with anisotropic internal structures. Finally, a combined system consisting of ECC-induced droplet dispensing and electrophoresis of charged droplet (ECD)-driven manipulation systems is constructed. This work was supported by the BK21Plus Program for advanced education of creative chemical engineers of the National Research Foundation of Korea (NRF) Grant funded by the Korea government (MSIP).

  15. Compressed air-assisted solvent extraction (CASX) for metal removal.

    PubMed

    Li, Chi-Wang; Chen, Yi-Ming; Hsiao, Shin-Tien

    2008-03-01

    A novel process, compressed air-assisted solvent extraction (CASX), was developed to generate micro-sized solvent-coated air bubbles (MSAB) for metal extraction. Through pressurization of solvent with compressed air followed by releasing air-oversaturated solvent into metal-containing wastewater, MSAB were generated instantaneously. The enormous surface area of MSAB makes extraction process extremely fast and achieves very high aqueous/solvent weight ratio (A/S ratio). CASX process completely removed Cr(VI) from acidic electroplating wastewater under A/S ratio of 115 and extraction time of less than 10s. When synthetic wastewater containing Cd(II) of 50mgl(-1) was treated, A/S ratios of higher than 714 and 1190 could be achieved using solvent with extractant/diluent weight ratio of 1:1 and 5:1, respectively. Also, MSAB have very different physical properties, such as size and density, compared to the emulsified solvent droplets, making separation and recovery of solvent from treated effluent very easy.

  16. A hemispherical Langmuir probe array detector for angular resolved measurements on droplet-based laser-produced plasmas

    DOE Office of Scientific and Technical Information (OSTI.GOV)

    Gambino, Nadia, E-mail: gambinon@ethz.ch; Brandstätter, Markus; Rollinger, Bob

    2014-09-15

    In this work, a new diagnostic tool for laser-produced plasmas (LPPs) is presented. The detector is based on a multiple array of six motorized Langmuir probes. It allows to measure the dynamics of a LPP in terms of charged particles detection with particular attention to droplet-based LPP sources for EUV lithography. The system design permits to temporally resolve the angular and radial plasma charge distribution and to obtain a hemispherical mapping of the ions and electrons around the droplet plasma. The understanding of these dynamics is fundamental to improve the debris mitigation techniques for droplet-based LPP sources. The device hasmore » been developed, built, and employed at the Laboratory for Energy Conversion, ETH Zürich. The experimental results have been obtained on the droplet-based LPP source ALPS II. For the first time, 2D mappings of the ion kinetic energy distribution around the droplet plasma have been obtained with an array of multiple Langmuir probes. These measurements show an anisotropic expansion of the ions in terms of kinetic energy and amount of ion charge around the droplet target. First estimations of the plasma density and electron temperature were also obtained from the analysis of the probe current signals.« less

  17. Influence of droplet charge on the chemical stability of citral in oil-in-water emulsions.

    PubMed

    Choi, Seung Jun; Decker, Eric Andrew; Henson, Lulu; Popplewell, L Michael; McClements, David Julian

    2010-08-01

    The chemical stability of citral, a flavor component widely used in beverage, food, and fragrance products, in oil-in-water emulsions stabilized by surfactants with different charge characteristics was investigated. Emulsions were prepared using cationic (lauryl alginate, LAE), non-ionic (polyoxyethylene (23) lauryl ether, Brij 35), and anionic (sodium dodecyl sulfate, SDS) surfactants at pH 3.5. The citral concentration decreased over time in all the emulsions, but the rate of decrease depended on surfactant type. After 7 d storage, the citral concentrations remaining in the emulsions were around 60% for LAE- or Brij 35-stabilized emulsions and 10% for SDS-stabilized emulsions. An increase in the local proton (H(+)) concentration around negatively charged droplet surfaces may account for the more rapid citral degradation observed in SDS-stabilized emulsions. A strong metal ion chelator (EDTA), which has previously been shown to be effective at increasing the oxidative stability of labile components, had no effect on citral stability in LAE- or Brij 35-stabilized emulsions, but it slightly decreased the initial rate of citral degradation in SDS-stabilized emulsions. These results suggest the surfactant type used to prepare emulsions should be controlled to improve the chemical stability of citral in emulsion systems.

  18. A simple method for the fast calculation of charge redistribution of solutes in an implicit solvent model

    NASA Astrophysics Data System (ADS)

    Dias, L. G.; Shimizu, K.; Farah, J. P. S.; Chaimovich, H.

    2002-09-01

    We propose and demonstrate the usefulness of a method, defined as generalized Born electronegativity equalization method (GBEEM) to estimate solvent-induced charge redistribution. The charges obtained by GBEEM, in a representative series of small organic molecules, were compared to PM3-CM1 charges in vacuum and in water. Linear regressions with appropriate correlation coefficients and standard deviations between GBEEM and PM3-CM1 methods were obtained ( R=0.94,SD=0.15, Ftest=234, N=32, in vacuum; R=0.94,SD=0.16, Ftest=218, N=29, in water). In order to test the GBEEM response when intermolecular interactions are involved we calculated a water dimer in dielectric water using both GBEEM and PM3-CM1 and the results were similar. Hence, the method developed here is comparable to established calculation methods.

  19. Spectroscopic studies of multiple charge transfer complexes of p-toluidine with π-acceptor picric acid in different polar solvents

    NASA Astrophysics Data System (ADS)

    Singh, Neeti; Ahmad, Afaq

    2010-04-01

    The charge transfer complexes of the donor p-toluidine with π-acceptor picric acid have been studied spectrophotometrically in various solvents such as acetone, ethanol, and methanol at room temperature using absorption spectrophotometer. The results indicate that formation of CTC in less polar solvent is high. The stoichiometry of the complex was found to be 1: 1 ratio by straight line method between donor and acceptor with maximum absorption bands. The data are discussed in terms of formation constant ( K CT), molar extinction coefficient (ɛCT), standard free energy (Δ G°), oscillator strength ( f), transition dipole moment (μEN), resonance energy ( R N) and ionization potential ( I D). The results indicate that the formation constant ( K CT) for the complex were shown to be dependent upon the nature of electron acceptor, donor and polarity of solvents which were used.

  20. Direct current dielectrophoretic manipulation of the ionic liquid droplets in water.

    PubMed

    Zhao, Kai; Li, Dongqing

    2018-07-13

    The ionic liquids (ILs) as the environmentally benign solvents show great potentials in microemulsion carrier systems and have been widely used in the biochemical and pharmaceutical fields. In the work, the ionic liquid-in-water microemulsions were fabricated by using two kinds of hydrophobic ionic liquid, 1-Butyl-3-methylimidazolium hexafluorophosphate [Bmim][PF 6 ] and 1-Hexyl-3-methylimidazolium hexafluorophosphate [Hmim][PF 6 ] with Tween 20. The ionic liquid droplets in water experience the dielectrophoretic (DEP) forces induced by applying electrical field via a nano-orifice and a micron orifice on the opposite channel walls of a microchannel. The dielectrophoretic behaviors of the ionic liquid-in-water emulsion droplets were investigated under direct current (DC) electric field. The positive and negative DEP behaviors of the ionic liquid-in-water droplets varying with the electrical conductivity of the suspending medium were investigated and two kinds of the ionic liquid droplets of similar sizes were separated by their different DEP behaviors. In addition, the separation of the ionic liquid-in-water droplets by size was conducted. This paper, for the first time to our knowledge, presents the DC-DEP manipulation of the ionic liquid-in-water emulsion droplets by size and by type. This method provides a platform to manipulate the ionic liquid droplets individually. Copyright © 2018 Elsevier B.V. All rights reserved.

  1. Solute-Solvent Charge-Transfer Excitations and Optical Absorption of Hydrated Hydroxide from Time-Dependent Density-Functional Theory.

    PubMed

    Opalka, Daniel; Sprik, Michiel

    2014-06-10

    The electronic structure of simple hydrated ions represents one of the most challenging problems in electronic-structure theory. Spectroscopic experiments identified the lowest excited state of the solvated hydroxide as a charge-transfer-to-solvent (CTTS) state. In the present work we report computations of the absorption spectrum of the solvated hydroxide ion, treating both solvent and solute strictly at the same level of theory. The average absorption spectrum up to 25 eV has been computed for samples taken from periodic ab initio molecular dynamics simulations. The experimentally observed CTTS state near the onset of the absorption threshold has been analyzed at the generalized-gradient approximation (GGA) and with a hybrid density-functional. Based on results for the lowest excitation energies computed with the HSE hybrid functional and a Davidson diagonalization scheme, the CTTS transition has been found 0.6 eV below the first absorption band of liquid water. The transfer of an electron to the solvent can be assigned to an excitation from the solute 2pπ orbitals, which are subject to a small energetic splitting due to the asymmetric solvent environment, to the significantly delocalized lowest unoccupied orbital of the solvent. The distribution of the centers of the excited state shows that CTTS along the OH(-) axis of the hydroxide ion is avoided. Furthermore, our simulations indicate that the systematic error arising in the calculated spectrum at the GGA originates from a poor description of the valence band energies in the solution.

  2. Performance of droplet generator and droplet collector in liquid droplet radiator under microgravity

    NASA Astrophysics Data System (ADS)

    Totani, T.; Itami, M.; Nagata, H.; Kudo, I.; Iwasaki, A.; Hosokawa, S.

    2002-06-01

    The Liquid Droplet Radiator (LDR) has an advantage over comparable conventional radiators in terms of the rejected heat power-weight ratio. Therefore, the LDR has attracted attention as an advanced radiator for high-power space systems that will be prerequisite for large space structures. The performance of the LDR under microgravity condition has been studied from the viewpoint of operational space use of the LDR in the future. In this study, the performances of a droplet generator and a droplet collector in the LDR are investigated using drop shafts in Japan: MGLAB and JAMIC. As a result, it is considered that (1) the droplet generator can produce uniform droplet streams in the droplet diameter range from 200 to 280 [µm] and the spacing range from 400 to 950 [µm] under microgravity condition, (2) the droplet collector with the incidence angle of 35 degrees can prevent a uniform droplet stream, in which droplet diameter is 250 [µm] and the velocity is 16 [m/s], from splashing under microgravity condition, whereas splashes may occur at the surface of the droplet collector in the event that a nonuniform droplet stream collides against it.

  3. Thermophoresis in nanoliter droplets to quantify aptamer binding.

    PubMed

    Seidel, Susanne A I; Markwardt, Niklas A; Lanzmich, Simon A; Braun, Dieter

    2014-07-21

    Biomolecule interactions are central to pharmacology and diagnostics. These interactions can be quantified by thermophoresis, the directed molecule movement along a temperature gradient. It is sensitive to binding induced changes in size, charge, or conformation. Established capillary measurements require at least 0.5 μL per sample. We cut down sample consumption by a factor of 50, using 10 nL droplets produced with acoustic droplet robotics (Labcyte). Droplets were stabilized in an oil-surfactant mix and locally heated with an IR laser. Temperature increase, Marangoni flow, and concentration distribution were analyzed by fluorescence microscopy and numerical simulation. In 10 nL droplets, we quantified AMP-aptamer affinity, cooperativity, and buffer dependence. Miniaturization and the 1536-well plate format make the method high-throughput and automation friendly. This promotes innovative applications for diagnostic assays in human serum or label-free drug discovery screening. © 2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.

  4. Molecular Dynamics Simulation Study of Solvent and State of Charge Effects on Solid-Phase Structure and Counterion Binding in a Nitroxide Radical Containing Polymer Energy Storage Material

    DOE PAGES

    Kemper, Travis W.; Gennett, Thomas; Larsen, Ross E.

    2016-10-19

    Here we performed molecular dynamics simulations to understand the effects of solvent swelling and state of charge (SOC) on the redox active, organic radical cathode material poly(2,2,6,6-tetramethylpiperidinyloxy methacrylate) (PTMA). We show that the polar solvent acetonitrile primarily solvates the nitroxide radical without disrupting the packing of the (2,2,6,6-tetramethylpiperidin-1-yl)oxyl (TEMPO) pendant groups of PTMA. We also simulated bulk PTMA in different SOC, 25%, 50%, 75%, and 100%, by converting the appropriate number of TEMPO groups to the cation charge state and adding BF 4 - counterions to the simulation. At each SOC the packing of PTMA, the solvent, and the counterionsmore » were examined. The binding of the anion to the nitroxide cation site was examined using the potential of mean force and found to be on the order of tens of meV, with a binding energy that decreased with increasing SOC. Additionally, we found that the cation state is stabilized by the presence of a nearby anion by more than 1 eV, and the implications of this stabilization on charge transport are discussed. Finally, we describe the implications of our results for how the SOC of an organic electrode affects electron and anion charge transport during the charging and discharging processes.« less

  5. Effect of viscosity on droplet-droplet collisional interaction

    NASA Astrophysics Data System (ADS)

    Finotello, Giulia; Padding, Johan T.; Deen, Niels G.; Jongsma, Alfred; Innings, Fredrik; Kuipers, J. A. M.

    2017-06-01

    A complete knowledge of the effect of droplet viscosity on droplet-droplet collision outcomes is essential for industrial processes such as spray drying. When droplets with dispersed solids are dried, the apparent viscosity of the dispersed phase increases by many orders of magnitude, which drastically changes the outcome of a droplet-droplet collision. However, the effect of viscosity on the droplet collision regime boundaries demarcating coalescence and reflexive and stretching separation is still not entirely understood and a general model for collision outcome boundaries is not available. In this work, the effect of viscosity on the droplet-droplet collision outcome is studied using direct numerical simulations employing the volume of fluid method. The role of viscous energy dissipation is analysed in collisions of droplets with different sizes and different physical properties. From the simulations results, a general phenomenological model depending on the capillary number (Ca, accounting for viscosity), the impact parameter (B), the Weber number (We), and the size ratio (Δ) is proposed.

  6. Understanding Solvent Manipulation of Morphology in Bulk-Heterojunction Organic Solar Cells.

    PubMed

    Chen, Yuxia; Zhan, Chuanlang; Yao, Jiannian

    2016-10-06

    Film morphology greatly influences the performance of bulk-heterojunction (BHJ)-structure-based solar cells. It is known that an interpenetrating bicontinuous network with nanoscale-separated donor and acceptor phases for charge transfer, an ordered molecular packing for exciton diffusion and charge transport, and a vertical compositionally graded structure for charge collection are prerequisites for achieving highly efficient BHJ organic solar cells (OSCs). Therefore, control of the morphology to obtain an ideal structure is a key problem. For this solution-processing BHJ system, the solvent participates fully in film processing. Its involvement is critical in modifying the nanostructure of BHJ films. In this review, we discuss the effects of solvent-related methods on the morphology of BHJ films, including selection of the casting solvent, solvent mixture, solvent vapor annealing, and solvent soaking. On the basis of a discussion on interaction strength and time between solvent and active materials, we believe that the solvent-morphology-performance relationship will be clearer and that solvent selection as a means to manipulate the morphology of BHJ films will be more rational. © 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.

  7. Electrostatic charge characteristics of jet nebulized aerosols.

    PubMed

    Kwok, Philip Chi Lip; Trietsch, Sebastiaan J; Kumon, Michiko; Chan, Hak-Kim

    2010-06-01

    Liquid droplets can be spontaneously charged in the absence of applied electric fields by spraying. It has been shown by computational simulation that charges may influence particle deposition in the airways. The electrostatic properties of jet nebulized aerosols and their potential effects on lung deposition have hardly been studied. A modified electrical low pressure impactor (ELPI) was employed to characterize the aerosol charges generated from jet nebulized commercial products. The charge and size measurements were conducted at 50% RH and 22 degrees C with a modified ELPI. Ventolin, Bricanyl, and Atrovent were nebulized using PARI LC Plus jet nebulizers coupled to a DeVilbiss Pulmo-Aide compressor. The aerosols were sampled in 30-sec durations. The drug deposits on the impactor stages were assayed chemically using high-performance liquid chromatography (HPLC). The charges of nebulized deionized water, isotonic saline, and the three commercial products diluted with saline were also measured to analyze the contributions of the major nebule ingredients on charging. No mass assays were performed on these runs. All three commercial nebules generated net negative charges. The magnitude of the charges reduced over the period of nebulization. Ventolin and Bricanyl yielded similar charge profiles. Highly variable charges were produced from deionized water. On the other hand, nebulized saline reproducibly generated net positive charges. Diluted commercial nebules showed charge polarity inversion. The charge profiles of diluted salbutamol and terbutaline solutions resembled those of saline, while the charges from diluted ipratropium solutions fluctuated near neutrality. The charge profiles were shown to be influenced by the concentration and physicochemical properties of the drugs, as well as the history of nebulization. The drugs may have unique isoelectric concentrations in saline at which the nebulized droplets would carry near-zero charges. According to results from

  8. Evaluating stratiform cloud base charge remotely

    NASA Astrophysics Data System (ADS)

    Harrison, R. Giles; Nicoll, Keri A.; Aplin, Karen L.

    2017-06-01

    Stratiform clouds acquire charge at their upper and lower horizontal boundaries due to vertical current flow in the global electric circuit. Cloud charge is expected to influence microphysical processes, but understanding is restricted by the infrequent in situ measurements available. For stratiform cloud bases below 1 km in altitude, the cloud base charge modifies the surface electric field beneath, allowing a new method of remote determination. Combining continuous cloud height data during 2015-2016 from a laser ceilometer with electric field mill data, cloud base charge is derived using a horizontal charged disk model. The median daily cloud base charge density found was -0.86 nC m-2 from 43 days' data. This is consistent with a uniformly charged region 40 m thick at the cloud base, now confirming that negative cloud base charge is a common feature of terrestrial layer clouds. This technique can also be applied to planetary atmospheres and volcanic plumes.Plain Language SummaryThe idea that clouds in the atmosphere can <span class="hlt">charge</span> electrically has been appreciated since the time of Benjamin Franklin, but it is less widely recognized that it is not just thunderclouds which contain electric <span class="hlt">charge</span>. For example, water <span class="hlt">droplets</span> in simple layer clouds, that are abundant and often responsible for an overcast day, carry electric <span class="hlt">charges</span>. The <span class="hlt">droplet</span> <span class="hlt">charging</span> arises at the upper and lower edges of the layer cloud. This occurs because the small <span class="hlt">droplets</span> at the edges draw <span class="hlt">charge</span> from the air outside the cloud. Understanding how strongly layer clouds <span class="hlt">charge</span> is important in evaluating electrical effects on the development of such clouds, for example, how thick the cloud becomes and whether it generates rain. Previously, cloud <span class="hlt">charge</span> measurement has required direct measurements within the cloud using weather balloons or aircraft. This work has monitored the lower cloud <span class="hlt">charge</span> continuously using instruments placed at the surface beneath</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1089501','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1089501"><span>Mass Transfer And Hydraulic Testing Of The V-05 And V-10 Contactors With The Next Generation <span class="hlt">Solvent</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>Herman, D. T.; Duignan, M. R.; Williams, M. R.</p> <p></p> <p>The Modular Caustic Side <span class="hlt">Solvent</span> Extraction (CSSX) Unit (MCU) facility is actively pursuing the transition from the current BOBCalixC6 based <span class="hlt">solvent</span> to the Next Generation <span class="hlt">Solvent</span> (NGS)-MCU <span class="hlt">solvent</span>. To support this integration of NGS into the MCU facilities, Savannah River Remediation (SRR) requested that Savannah River National Laboratory (SRNL) perform testing of a blend of the NGS (MaxCalix based <span class="hlt">solvent</span>) with the current <span class="hlt">solvent</span> (BOBCalixC6 based <span class="hlt">solvent</span>) for the removal of cesium (Cs) from the liquid salt waste stream. This testing differs from prior testing by utilizing a blend of BOBCalixC6 based <span class="hlt">solvent</span> and the NGS with the full (0.05more » M) concentration of the MaxCalix as well as a new suppressor, tris(3,7dimethyloctyl) guanidine. Single stage tests were conducted using the full size V-05 and V-10 centrifugal contactors installed at SRNL. These tests were designed to determine the mass transfer and hydraulic characteristics with the NGS <span class="hlt">solvent</span> blended with the projected heel of the BOBCalixC6 based <span class="hlt">solvent</span> that will exist in MCU at time of transition. The test program evaluated the amount of organic carryover and the <span class="hlt">droplet</span> size of the organic carryover phases using several analytical methods. Stage efficiency and mass distribution ratios were determined by measuring Cs concentration in the aqueous and organic phases during single contactor testing. The nominal cesium distribution ratio, D(Cs) measured for extraction ranged from 37-60. The data showed greater than 96% stage efficiency for extraction. No significant differences were noted for operations at 4, 8 or 12 gpm aqueous salt simulant feed flow rates. The first scrub test (contact with weak caustic solution) yielded average scrub D(Cs) values of 3.3 to 5.2 and the second scrub test produced an average value of 1.8 to 2.3. For stripping behavior, the “first stage” D Cs) values ranged from 0.04 to 0.08. The efficiency of the low flow (0.27 gpm aqueous) was calculated to be 82.7%. The</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1998JChPh.109.3077B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1998JChPh.109.3077B"><span>Cooperative effects in the structuring of fluoride water clusters: Ab initio hybrid quantum mechanical/molecular mechanical model incorporating polarizable fluctuating <span class="hlt">charge</span> <span class="hlt">solvent</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bryce, Richard A.; Vincent, Mark A.; Malcolm, Nathaniel O. J.; Hillier, Ian H.; Burton, Neil A.</p> <p>1998-08-01</p> <p>A new hybrid quantum mechanical/molecular mechanical model of solvation is developed and used to describe the structure and dynamics of small fluoride/water clusters, using an ab initio wave function to model the ion and a fluctuating <span class="hlt">charge</span> potential to model the waters. Appropriate parameters for the water-water and fluoride-water interactions are derived, with the fluoride anion being described by density functional theory and a large Gaussian basis. The role of <span class="hlt">solvent</span> polarization in determining the structure and energetics of F(H2O)4- clusters is investigated, predicting a slightly greater stability of the interior compared to the surface structure, in agreement with ab initio studies. An extended Lagrangian treatment of the polarizable water, in which the water atomic <span class="hlt">charges</span> fluctuate dynamically, is used to study the dynamics of F(H2O)4- cluster. A simulation using a fixed <span class="hlt">solvent</span> <span class="hlt">charge</span> distribution indicates principally interior, solvated states for the cluster. However, a preponderance of trisolvated configurations is observed using the polarizable model at 300 K, which involves only three direct fluoride-water hydrogen bonds. Ab initio calculations confirm this trisolvated species as a thermally accessible state at room temperature, in addition to the tetrasolvated interior and surface structures. Extension of this polarizable water model to fluoride clusters with five and six waters gave less satisfactory agreement with experimental energies and with ab initio geometries. However, our results do suggest that a quantitative model of <span class="hlt">solvent</span> polarization is fundamental for an accurate understanding of the properties of anionic water clusters.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16649769','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16649769"><span>Porous structures of polymer films prepared by spin coating with mixed <span class="hlt">solvents</span> under humid condition.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Park, Min Soo; Joo, Wonchul; Kim, Jin Kon</p> <p>2006-05-09</p> <p>We investigate the effects of interfacial energy between water and <span class="hlt">solvent</span> as well as polymer concentration on the formation of porous structures of polymer films prepared by spin coating of cellulose acetate butyrate (CAB) in mixed <span class="hlt">solvent</span> of tetrahydrofuran (THF) and chloroform under humid condition. The interfacial energy between water and the <span class="hlt">solvent</span> was gradually changed by the addition of chloroform to the <span class="hlt">solvent</span>. At a high polymer concentration (0.15 g/cm3 in THF), porous structures were limited only at the top surfaces of CAB films, regardless of interfacial energies, due to the high viscosity of the solution. At a medium concentration (approximately 0.08 g/cm3 in THF), CAB film had relatively uniform pores at the top surface and very small pores inside the film because of the mixing of the water <span class="hlt">droplets</span> with THF solution. When chloroform was added to THF, pores at the inner CAB film had a comparable size with those at the top surface because of the reduced degree of the mixing between the water <span class="hlt">droplets</span> and the mixed <span class="hlt">solvent</span>. A further decrease in polymer concentration (0.05 g/cm3 in THF) caused the final films to have a two-layer porous structure, and the size of pores at each layer was almost the same.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20110000778','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20110000778"><span><span class="hlt">Droplet</span> microactuator system</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Pamula, Vamsee K. (Inventor); Pollack, Michael G. (Inventor); Eckhardt, Allen E. (Inventor); Paik, Philip Y. (Inventor); Srinivasan, Vijay (Inventor)</p> <p>2010-01-01</p> <p>The present invention relates to a <span class="hlt">droplet</span> microactuator system. According to one embodiment, the <span class="hlt">droplet</span> microactuator system includes: (a) a <span class="hlt">droplet</span> microactuator configured to conduct <span class="hlt">droplet</span> operations; (b) a magnetic field source arranged to immobilize magnetically responsive beads in a <span class="hlt">droplet</span> during <span class="hlt">droplet</span> operations; (c) a sensor configured in a sensing relationship with the <span class="hlt">droplet</span> microactuator, such that the sensor is capable of sensing a signal from and/or a property of one or more <span class="hlt">droplets</span> on the <span class="hlt">droplet</span> microactuator; and (d) one or more processors electronically coupled to the <span class="hlt">droplet</span> microactuator and programmed to control electrowetting-mediated <span class="hlt">droplet</span> operations on the <span class="hlt">droplet</span> actuator and process electronic signals from the sensor.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29673767','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29673767"><span>Ferrofluid of magnetic clay and menthol based deep eutectic <span class="hlt">solvent</span>: Application in directly suspended <span class="hlt">droplet</span> microextraction for enrichment of some emerging contaminant explosives in water and soil samples.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zarei, Ali Reza; Nedaei, Maryam; Ghorbanian, Sohrab Ali</p> <p>2018-06-08</p> <p>In this work, for the first time, ferrofluid of magnetic montmorillonite nanoclay and deep eutectic <span class="hlt">solvent</span> was prepared and coupled with directly suspended <span class="hlt">droplet</span> microextraction. Incorporation of ferrofluid in a miniaturized sample preparation technique resulted in achieving high extraction efficiency while developing a green analytical method. The prepared ferrofluid has strong sorbing properties and hydrophobic characteristics. In this method, a micro-<span class="hlt">droplet</span> of ferrofluid was suspended into the vortex of a stirring aqueous solution and after completing the extraction process, was easily separated from the solution by a magnetic rod without any operational problems. The predominant experimental variables affecting the extraction efficiency of explosives were evaluated. Under optimal conditions, the limits of detection were in the range 0.22-0.91 μg L -1 . The enrichment factors were between 23 and 93 and the relative standard deviations were <10%. The relative recoveries were ranged from 88 to 104%. This method was successfully applied for the extraction and preconcentration of explosives in water and soil samples, followed their determination by high performance liquid chromatography with ultraviolet detection (HPLC-UV). Copyright © 2018 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21749137','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21749137"><span>Bubbles in <span class="hlt">solvent</span> microextraction: the influence of intentionally introduced bubbles on extraction efficiency.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Williams, D Bradley G; George, Mosotho J; Meyer, Riaan; Marjanovic, Ljiljana</p> <p>2011-09-01</p> <p>Significant improvements to microdrop extractions of triazine pesticides are realized by the intentional incorporation of an air bubble into the <span class="hlt">solvent</span> microdroplet used in this microextraction technique. The increase is attributed partly to greater <span class="hlt">droplet</span> surface area resulting from the air bubble being incorporated into the <span class="hlt">solvent</span> <span class="hlt">droplet</span> as opposed to it sitting thereon and partly to thin film phenomena. The method is useful at nanogram/liter levels (LOD 0.002-0.012 μg/L, LOQ 0.007-0.039 μg/L), is precise (7-12% at 10 μg/L concentration level), and is validated against certified reference materials containing 0.5 and 5.0 μg/L analyte. It tolerates water and fruit juice as matrixes without serious matrix effects. This new development brings a simple, inexpensive, and efficient preconcentration technique to bear which rivals solid phase microextraction methods.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..DFDE11005K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..DFDE11005K"><span>Deposition dynamics of multi-<span class="hlt">solvent</span> bioinks</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kaneelil, Paul; Pack, Min; Cui, Chunxiao; Han, Li-Hsin; Sun, Ying</p> <p>2017-11-01</p> <p>Inkjet printing cellular scaffolds using bioinks is gaining popularity due to the advancement of printing technology as well as the growing demands of regenerative medicine. Numerous studies have been conducted on printing scaffolds of biomimetic structures that support the cell production of human tissues. However, the underlying physics of the deposition dynamics of bioinks remains elusive. Of particular interest is the unclear deposition dynamics of multi-<span class="hlt">solvent</span> bioinks, which is often used to tune the micro-architecture formation. Here we systematically studied the effects of jetting frequency, <span class="hlt">solvent</span> properties, substrate wettability, and temperature on the three-dimensional deposition patterns of bioinks made of Methacrylated Gelatin and Carboxylated Gelatin. The microflows inside the inkjet-printed picolitre drops were visualized using fluorescence tracer particles to decipher the complex processes of multi-<span class="hlt">solvent</span> evaporation and solute self-assembly. The evolution of <span class="hlt">droplet</span> shape was observed using interferometry. With the integrated techniques, the interplay of <span class="hlt">solvent</span> evaporation, biopolymer deposition, and multi-drop interactions were directly observed for various ink and substrate properties, and printing conditions. Such knowledge enables the design and fabrication of a variety of tissue engineering scaffolds for potential use in regenerative medicine.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26921608','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26921608"><span>Negatively-<span class="hlt">charged</span> residues in the polar carboxy-terminal region in FSP27 are indispensable for expanding lipid <span class="hlt">droplets</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Tamori, Yoshikazu; Tateya, Sanshiro; Ijuin, Takeshi; Nishimoto, Yuki; Nakajima, Shinsuke; Ogawa, Wataru</p> <p>2016-03-01</p> <p>FSP27 has an important role in large lipid <span class="hlt">droplet</span> (LD) formation because it exchanges lipids at the contact site between LDs. In the present study, we clarify that the amino-terminal domain of FSP27 (amino acids 1-130) is dispensable for LD enlargement, although it accelerates LD growth. LD expansion depends on the carboxy-terminal domain of FSP27 (amino acids 131-239). Especially, the negative <span class="hlt">charge</span> of the acidic residues (D215, E218, E219 and E220) in the polar carboxy-terminal region (amino acids 202-239) is essential for the enlargement of LD. We propose that the carboxy-terminal domain of FSP27 has a crucial role in LD expansion, whereas the amino-terminal domain only has a supportive role. © 2016 Federation of European Biochemical Societies.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4914070','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4914070"><span>Transformation of eutectic emulsion to nanosuspension fabricating with <span class="hlt">solvent</span> evaporation and ultrasonication technique</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Phaechamud, Thawatchai; Tuntarawongsa, Sarun</p> <p>2016-01-01</p> <p>Eutectic <span class="hlt">solvent</span> can solubilize high amount of some therapeutic compounds. Volatile eutectic <span class="hlt">solvent</span> is interesting to be used as <span class="hlt">solvent</span> in the preparation of nanosuspension with emulsion <span class="hlt">solvent</span> evaporation technique. The mechanism of transformation from the eutectic emulsion to nanosuspension was investigated in this study. The 30% w/w ibuprofen eutectic solution was used as the internal phase, and the external phase is composed of Tween 80 as emulsifier. Ibuprofen nanosuspension was prepared by eutectic emulsion <span class="hlt">solvent</span> evaporating method followed with ultrasonication. During evaporation process, the ibuprofen concentration in emulsion <span class="hlt">droplets</span> was increased leading to a drug supersaturation but did not immediately recrystallize because of low glass transition temperature (Tg) of ibuprofen. The contact angle of the internal phase on ibuprofen was apparently lower than that of the external phase at all times of evaporation, indicating that the ibuprofen crystals were preferentially wetted by the internal phase than the external phase. From calculated dewetting value ibuprofen crystallization occurred in the <span class="hlt">droplet</span>. Crystallization of the drug was initiated with external mechanical force, and the particle size of the drug was larger due to Ostwald ripening. Cavitation force from ultrasonication minimized the ibuprofen crystals to the nanoscale. Particle size and zeta potential of formulated ibuprofen nanosuspension were 330.87±51.49 nm and −31.1±1.6 mV, respectively, and exhibited a fast dissolution. Therefore, the combination of eutectic emulsion <span class="hlt">solvent</span> evaporation method with ultrasonication was favorable for fabricating an ibuprofen nanosuspension, and the transformation mechanism was attained successfully. PMID:27366064</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27366064','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27366064"><span>Transformation of eutectic emulsion to nanosuspension fabricating with <span class="hlt">solvent</span> evaporation and ultrasonication technique.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Phaechamud, Thawatchai; Tuntarawongsa, Sarun</p> <p>2016-01-01</p> <p>Eutectic <span class="hlt">solvent</span> can solubilize high amount of some therapeutic compounds. Volatile eutectic <span class="hlt">solvent</span> is interesting to be used as <span class="hlt">solvent</span> in the preparation of nanosuspension with emulsion <span class="hlt">solvent</span> evaporation technique. The mechanism of transformation from the eutectic emulsion to nanosuspension was investigated in this study. The 30% w/w ibuprofen eutectic solution was used as the internal phase, and the external phase is composed of Tween 80 as emulsifier. Ibuprofen nanosuspension was prepared by eutectic emulsion <span class="hlt">solvent</span> evaporating method followed with ultrasonication. During evaporation process, the ibuprofen concentration in emulsion <span class="hlt">droplets</span> was increased leading to a drug supersaturation but did not immediately recrystallize because of low glass transition temperature (T g) of ibuprofen. The contact angle of the internal phase on ibuprofen was apparently lower than that of the external phase at all times of evaporation, indicating that the ibuprofen crystals were preferentially wetted by the internal phase than the external phase. From calculated dewetting value ibuprofen crystallization occurred in the <span class="hlt">droplet</span>. Crystallization of the drug was initiated with external mechanical force, and the particle size of the drug was larger due to Ostwald ripening. Cavitation force from ultrasonication minimized the ibuprofen crystals to the nanoscale. Particle size and zeta potential of formulated ibuprofen nanosuspension were 330.87±51.49 nm and -31.1±1.6 mV, respectively, and exhibited a fast dissolution. Therefore, the combination of eutectic emulsion <span class="hlt">solvent</span> evaporation method with ultrasonication was favorable for fabricating an ibuprofen nanosuspension, and the transformation mechanism was attained successfully.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28918933','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28918933"><span>Structural proteomics: Topology and relative accessibility of plant lipid <span class="hlt">droplet</span> associated proteins.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jolivet, Pascale; Aymé, Laure; Giuliani, Alexandre; Wien, Frank; Chardot, Thierry; Gohon, Yann</p> <p>2017-10-03</p> <p>Lipid <span class="hlt">droplets</span> are the major stock of lipids in oleaginous plant seeds. Despite their economic importance for oil production and biotechnological issues (biofuels, lubricants and plasticizers), numerous questions about their formation, structure and regulation are still unresolved. To determine water accessible domains of protein coating at lipid <span class="hlt">droplets</span> surface, a structural proteomic approach has been performed. This technique relies on the millisecond timescale production of hydroxyl radicals by the radiolysis of water using Synchrotron X-ray white beam. Thanks to the evolution of mass spectrometry analysis techniques this approach allows the creation of a map of the <span class="hlt">solvent</span> accessibility for proteins difficult to study by other means. Using these results, a S3 oleosin water accessibility map is proposed. This is the first time that such a map on an oleosin co-purified with plant lipid <span class="hlt">droplets</span> and other associated protein is presented. Lipid <span class="hlt">droplet</span> associated proteins function is linked to stability, structure and probably formation and lipid mobilization of <span class="hlt">droplets</span>. Structure of these proteins in their native environment, at the interface between bulk water and the lipidic core of these organelles is only based on hydrophobicity plot. Using hydroxyl radical footprinting and proteomics approaches we studied water accessibility of one major protein in these <span class="hlt">droplets</span>: S3 oleosin of Arabidopsis thaliana seeds. Copyright © 2017 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27118921','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27118921"><span>On well-posedness of variational models of <span class="hlt">charged</span> drops.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Muratov, Cyrill B; Novaga, Matteo</p> <p>2016-03-01</p> <p>Electrified liquids are well known to be prone to a variety of interfacial instabilities that result in the onset of apparent interfacial singularities and liquid fragmentation. In the case of electrically conducting liquids, one of the basic models describing the equilibrium interfacial configurations and the onset of instability assumes the liquid to be equipotential and interprets those configurations as local minimizers of the energy consisting of the sum of the surface energy and the electrostatic energy. Here we show that, surprisingly, this classical geometric variational model is mathematically ill-posed irrespective of the degree to which the liquid is electrified. Specifically, we demonstrate that an isolated spherical <span class="hlt">droplet</span> is never a local minimizer, no matter how small is the total <span class="hlt">charge</span> on the <span class="hlt">droplet</span>, as the energy can always be lowered by a smooth, arbitrarily small distortion of the <span class="hlt">droplet</span>'s surface. This is in sharp contrast to the experimental observations that a critical amount of <span class="hlt">charge</span> is needed in order to destabilize a spherical <span class="hlt">droplet</span>. We discuss several possible regularization mechanisms for the considered free boundary problem and argue that well-posedness can be restored by the inclusion of the entropic effects resulting in finite screening of free <span class="hlt">charges</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_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://pubs.er.usgs.gov/publication/70118563','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70118563"><span>Control of <span class="hlt">droplet</span> morphology for inkjet-printed TIPS-pentacene transistors</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Lee, Myung Won; Ryu, Gi Seong; Lee, Young Uk; Pearson, Christopher; Petty, Michael C.; Song, Chung Kun</p> <p>2012-01-01</p> <p>We report on methods to control the morphology of <span class="hlt">droplets</span> of 6,13-bis(triisopropyl-silylethynyl) pentacene (TIPS-PEN), which are then used in the fabrication of organic thin film transistors (OTFTs). The grain size and distribution of the TIPS-PEN were found to depend on the temperature of the <span class="hlt">droplets</span> during drying. The performance of the OTFTs could be improved by heating the substrate and also by changing the relative positions of the inkjet-printed <span class="hlt">droplets</span>. In our experiments, the optimum substrate temperature was 46 °C in air. Transistors with the TIPS-PEN grain boundaries parallel to the current flow between the source and drain electrodes exhibited <span class="hlt">charge</span> carrier mobilities of 0.44 ± 0.08 cm2/V s.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19603834','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19603834"><span>Imaging and estimating the surface heterogeneity on a <span class="hlt">droplet</span> containing cosolvents.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Fang, Xiaohua; Li, Bingquan; Wu, Jun; Maldarelli, Charles; Sokolov, Jonathan C; Rafailovich, Miriam H; Somasundaran, Ponisseril</p> <p>2009-07-23</p> <p>Cosolvents have numerous applications in many industries as well as scientific research. The shortage in the knowledge of the structures in a cosolvent system is significant. In this work, we display the spatial as well as the kinetic distribution of the cosolvents using <span class="hlt">droplets</span> as paradigms. When an alcohol/water-containing sessile <span class="hlt">droplet</span> evaporates on a substrate, it phase segregates into a water-enriched core and a thin alcohol prevailing shell. This is considered to be due to the different escaping rate of <span class="hlt">solvents</span> out of the liquid-vapor (l-v) interfaces. In between the core and shell phases, there exists a rough and solid-like liquid-liquid (l-l) wall interface as marked by the fluorescent polystyrene spheres and imaged by a confocal microscope. Holes and patches of beads are observed to form on this phase boundary. The water-dispersed beads prefer to partition within the core. The shell prevails in the <span class="hlt">droplet</span> during most of the drying and shrinks with the l-v boundary. By monitoring the morphological progression of the <span class="hlt">droplet</span>, the composition of the cosolvent at the liquid-vapor interface is obtained.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013PhDT.......151G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013PhDT.......151G"><span>Lysozyme pattern formation in evaporating <span class="hlt">droplets</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gorr, Heather Meloy</p> <p></p> <p>Liquid <span class="hlt">droplets</span> containing suspended particles deposited on a solid, flat surface generally form ring-like structures due to the redistribution of solute during evaporation (the "coffee ring effect"). The forms of the deposited patterns depend on complex interactions between solute(s), <span class="hlt">solvent</span>, and substrate in a rapidly changing, far from equilibrium system. Solute self-organization during evaporation of colloidal sessile <span class="hlt">droplets</span> has attracted the attention of researchers over the past few decades due to a variety of technological applications. Recently, pattern formation during evaporation of various biofluids has been studied due to potential applications in medical screening and diagnosis. Due to the complexity of 'real' biological fluids and other multicomponent systems, a comprehensive understanding of pattern formation during <span class="hlt">droplet</span> evaporation of these fluids is lacking. In this PhD dissertation, the morphology of the patterns remaining after evaporation of <span class="hlt">droplets</span> of a simplified model biological fluid (aqueous lysozyme solutions + NaCl) are examined by atomic force microscopy (AFM) and optical microscopy. Lysozyme is a globular protein found in high concentration, for example, in human tears and saliva. The drop diameters, D, studied range from the micro- to the macro- scale (1 microm -- 2 mm). In this work, the effect of evaporation conditions, solution chemistry, and heat transfer within the <span class="hlt">droplet</span> on pattern formation is examined. In micro-scale deposits of aqueous lysozyme solutions (1 microm < D < 50 microm), the protein motion and the resulting dried residue morphology are highly influenced by the decreased evaporation time of the drop. The effect of electrolytes on pattern formation is also investigated by adding varying concentrations NaCl to the lysozyme solutions. Finally, a novel pattern recognition program is described and implemented which classifies deposit images by their solution chemistries. The results presented in this Ph</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22978785','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22978785"><span>Accounting for changes in particle <span class="hlt">charge</span>, dry mass and composition occurring during studies of single levitated particles.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Haddrell, Allen E; Davies, James F; Yabushita, Akihiro; Reid, Jonathan P</p> <p>2012-10-11</p> <p>The most used instrument in single particle hygroscopic analysis over the past thirty years has been the electrodynamic balance (EDB). Two general assumptions are made in hygroscopic studies involving the EDB. First, it is assumed that the net <span class="hlt">charge</span> on the <span class="hlt">droplet</span> is invariant over the time scale required to record a hygroscopic growth cycle. Second, it is assumed that the composition of the <span class="hlt">droplet</span> is constant (aside from the addition and removal of water). In this study, we demonstrate that these assumptions cannot always be made and may indeed prove incorrect. The presence of net <span class="hlt">charge</span> in the humidified vapor phase reduces the total net <span class="hlt">charge</span> retained by the <span class="hlt">droplet</span> over prolonged levitation periods. The gradual reduction in <span class="hlt">charge</span> limits the reproducibility of hygroscopicity measurements made on repeated RH cycles with a single particle, or prolonged experiments in which the particle is held at a high relative humidity. Further, two contrasting examples of the influence of changes in chemical composition changes are reported. In the first, simple acid-base chemistry in the <span class="hlt">droplet</span> leads to the irreversible removal of gaseous ammonia from a <span class="hlt">droplet</span> containing an ammonium salt on a time scale that is shorter than the hygroscopicity measurement. In the second example, the net <span class="hlt">charge</span> on the <span class="hlt">droplet</span> (<100 fC) is high enough to drive redox chemistry within the <span class="hlt">droplet</span>. This is demonstrated by the reduction of iodic acid in a <span class="hlt">droplet</span> made solely of iodic acid and water to form iodine and an iodate salt.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25937106','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25937106"><span>Air-assisted liquid-liquid microextraction by solidifying the floating organic <span class="hlt">droplets</span> for the rapid determination of seven fungicide residues in juice samples.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>You, Xiangwei; Xing, Zhuokan; Liu, Fengmao; Zhang, Xu</p> <p>2015-05-22</p> <p>A novel air assisted liquid-liquid microextraction using the solidification of a floating organic <span class="hlt">droplet</span> method (AALLME-SFO) was developed for the rapid and simple determination of seven fungicide residues in juice samples, using the gas chromatography with electron capture detector (GC-ECD). This method combines the advantages of AALLME and dispersive liquid-liquid microextraction based on the solidification of floating organic <span class="hlt">droplets</span> (DLLME-SFO) for the first time. In this method, a low-density <span class="hlt">solvent</span> with a melting point near room temperature was used as the extraction <span class="hlt">solvent</span>, and the emulsion was rapidly formed by pulling in and pushing out the mixture of aqueous sample solution and extraction <span class="hlt">solvent</span> for ten times repeatedly using a 10-mL glass syringe. After centrifugation, the extractant <span class="hlt">droplet</span> could be easily collected from the top of the aqueous samples by solidifying it at a temperature lower than the melting point. Under the optimized conditions, good linearities with the correlation coefficients (γ) higher than 0.9959 were obtained and the limits of detection (LOD) varied between 0.02 and 0.25 μgL(-1). The proposed method was applied to determine the target fungicides in juice samples and acceptable recoveries ranged from 72.6% to 114.0% with the relative standard deviations (RSDs) of 2.3-13.0% were achieved. Compared with the conventional DLLME method, the newly proposed method will neither require a highly toxic chlorinated <span class="hlt">solvent</span> for extraction nor an organic dispersive <span class="hlt">solvent</span> in the application process; hence, it is more environmentally friendly. Copyright © 2015 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017NatMa..16..722L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017NatMa..16..722L"><span><span class="hlt">Charge</span>-transfer dynamics and nonlocal dielectric permittivity tuned with metamaterial structures as <span class="hlt">solvent</span> analogues</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lee, Kwang Jin; Xiao, Yiming; Woo, Jae Heun; Kim, Eunsun; Kreher, David; Attias, André-Jean; Mathevet, Fabrice; Ribierre, Jean-Charles; Wu, Jeong Weon; André, Pascal</p> <p>2017-07-01</p> <p><span class="hlt">Charge</span> transfer (CT) is a fundamental and ubiquitous mechanism in biology, physics and chemistry. Here, we evidence that CT dynamics can be altered by multi-layered hyperbolic metamaterial (HMM) substrates. Taking triphenylene:perylene diimide dyad supramolecular self-assemblies as a model system, we reveal longer-lived CT states in the presence of HMM structures, with both <span class="hlt">charge</span> separation and recombination characteristic times increased by factors of 2.4 and 1.7--that is, relative variations of 140 and 73%, respectively. To rationalize these experimental results in terms of driving force, we successfully introduce image dipole interactions in Marcus theory. The non-local effect herein demonstrated is directly linked to the number of metal-dielectric pairs, can be formalized in the dielectric permittivity, and is presented as a solid analogue to local <span class="hlt">solvent</span> polarity effects. This model and extra PH3T:PC60BM results show the generality of this non-local phenomenon and that a wide range of kinetic tailoring opportunities can arise from substrate engineering. This work paves the way toward the design of artificial substrates to control CT dynamics of interest for applications in optoelectronics and chemistry.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..DFD.Q5010D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..DFD.Q5010D"><span>Spray Formation from a <span class="hlt">Charged</span> Liquid Jet of a Dielectric Fluid</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Doak, William; de Bellis, Victor; Chiarot, Paul; Microfluidics; Multiphase Flow Laboratory Team</p> <p>2017-11-01</p> <p>Atomization of a dielectric micro-jet is achieved via an electrohydrodynamic <span class="hlt">charge</span> injection process. The atomizer is comprised of a grounded nozzle housing (ground electrode) and an internal probe (high voltage electrode) that is concentric with the emitting orifice. The internal probe is held at electric potentials ranging from 1-10 kV. A pressurized reservoir drives a dielectric fluid at a desired flow rate through the 100-micrometer diameter orifice. The fluid fills the cavity between the electrodes as it passes through the atomizer, impeding the transport of electrons. This process injects <span class="hlt">charge</span> into the flowing fluid. Upon exiting the orifice, the emitted jet is highly <span class="hlt">charged</span> and it deforms via a bending instability that is qualitatively similar to the behavior observed in the electrospinning of fibers. We observed bulging regions, or nodes, of highly <span class="hlt">charged</span> fluid forming along the bent, rotating jet. These nodes separate into highly <span class="hlt">charged</span> <span class="hlt">droplets</span> that emit satellite <span class="hlt">droplets</span>. The remaining ligaments break up due to capillarity in a process that produces additional satellites. All of the <span class="hlt">droplets</span> possess a normal (inertial) and radial (electrically-driven) momentum component. The radial component is responsible for the formation of a conical spray envelope. Our research focuses on the jet, its break up, and the <span class="hlt">droplet</span> dynamics of this system. This research supported by the American Chemical Society.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016APS..DFDG16004M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016APS..DFDG16004M"><span>A multiphase ion-transport analysis of the electrostatic disjoining pressure: implications for binary <span class="hlt">droplet</span> coalescence</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mason, Lachlan; Gebauer, Felix; Bart, Hans-Jörg; Stevens, Geoffrey; Harvie, Dalton</p> <p>2016-11-01</p> <p>Understanding the physics of emulsion coalescence is critical for the robust simulation of industrial <span class="hlt">solvent</span> extraction processes, in which loaded organic and raffinate phases are separated via the coalescence of dispersed <span class="hlt">droplets</span>. At the <span class="hlt">droplet</span> scale, predictive collision-outcome models require an accurate description of the repulsive surface forces arising from electrical-double-layer interactions. The conventional disjoining-pressure treatment of double-layer forces, however, relies on assumptions which do not hold generally for deformable <span class="hlt">droplet</span> collisions: namely, low interfacial curvature and negligible advection of ion species. This study investigates the validity bounds of the disjoining pressure approximation for low-inertia <span class="hlt">droplet</span> interactions. A multiphase ion-transport model, based on a coupling of <span class="hlt">droplet</span>-scale Nernst-Planck and Navier-Stokes equations, predicts ion-concentration fields that are consistent with the equilibrium Boltzmann distribution; indicating that the disjoining-pressure approach is valid for both static and dynamic interactions in low-Reynolds-number settings. The present findings support the development of coalescence kernels for application in macro-scale population balance modelling.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013APS..MARG34014M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013APS..MARG34014M"><span>Arrested of coalescence of emulsion <span class="hlt">droplets</span> of arbitrary size</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mbanga, Badel L.; Burke, Christopher; Blair, Donald W.; Atherton, Timothy J.</p> <p>2013-03-01</p> <p>With applications ranging from food products to cosmetics via targeted drug delivery systems, structured anisotropic colloids provide an efficient way to control the structure, properties and functions of emulsions. When two fluid emulsion <span class="hlt">droplets</span> are brought in contact, a reduction of the interfacial tension drives their coalescence into a larger <span class="hlt">droplet</span> of the same total volume and reduced exposed area. This coalescence can be partially or totally hindered by the presence of nano or micron-size particles that coat the interface as in Pickering emulsions. We investigate numerically the dependance of the mechanical stability of these arrested shapes on the particles size, their shape anisotropy, their polydispersity, their interaction with the <span class="hlt">solvent</span>, and the particle-particle interactions. We discuss structural shape changes that can be induced by tuning the particles interactions after arrest occurs, and provide design parameters for the relevant experiments.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014APS..GECFT1005M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014APS..GECFT1005M"><span>Non-thermal equilibrium plasma-liquid interactions with femtolitre <span class="hlt">droplets</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Maguire, Paul; Mahony, Charles; Bingham, Andrew; Patel, Jenish; Rutherford, David; McDowell, David; Mariotti, Davide; Bennet, Euan; Potts, Hugh; Diver, Declan</p> <p>2014-10-01</p> <p>Plasma-induced non-equilibrium liquid chemistry is little understood. It depends on a complex interplay of interface and near surface processes, many involving energy-dependent electron-induced reactions and the transport of transient species such as hydrated electrons. Femtolitre liquid <span class="hlt">droplets</span>, with an ultra-high ratio of surface area to volume, were transported through a low-temperature atmospheric pressure RF microplasma with transit times of 1--10 ms. Under a range of plasma operating conditions, we observe a number of non-equilibrium chemical processes that are dominated by energetic electron bombardment. Gas temperature and plasma parameters (ne ~ 1013 cm-3, Te < 4 eV) were determined while size and <span class="hlt">droplet</span> velocity profiles were obtained using a microscope coupled to a fast ICCD camera under low light conditions. Laminar mixed-phase <span class="hlt">droplet</span> flow is achieved and the plasma is seen to significantly deplete only the slower, smaller <span class="hlt">droplet</span> component due possibly to the interplay between evaporation, Rayleigh instabilities and <span class="hlt">charge</span> emission. Funding from EPSRC acknowledged (Grants EP/K006088/1 and EP/K006142/1).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ExFl...59...17F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ExFl...59...17F"><span>The dynamics of milk <span class="hlt">droplet-droplet</span> collisions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Finotello, Giulia; Kooiman, Roeland F.; Padding, Johan T.; Buist, Kay A.; Jongsma, Alfred; Innings, Fredrik; Kuipers, J. A. M.</p> <p>2018-01-01</p> <p>Spray drying is an important industrial process to produce powdered milk, in which concentrated milk is atomized into small <span class="hlt">droplets</span> and dried with hot gas. The characteristics of the produced milk powder are largely affected by agglomeration, combination of dry and partially dry particles, which in turn depends on the outcome of a collision between <span class="hlt">droplets</span>. The high total solids (TS) content and the presence of milk proteins cause a relatively high viscosity of the fed milk concentrates, which is expected to largely influence the collision outcomes of drops inside the spray. It is therefore of paramount importance to predict and control the outcomes of binary <span class="hlt">droplet</span> collisions. Only a few studies report on <span class="hlt">droplet</span> collisions of high viscous liquids and no work is available on <span class="hlt">droplet</span> collisions of milk concentrates. The current study therefore aims to obtain insight into the effect of viscosity on the outcome of binary collisions between <span class="hlt">droplets</span> of milk concentrates. To cover a wide range of viscosity values, three milk concentrates (20, 30 and 46% TS content) are investigated. An experimental set-up is used to generate two colliding <span class="hlt">droplet</span> streams with consistent <span class="hlt">droplet</span> size and spacing. A high-speed camera is used to record the trajectories of the <span class="hlt">droplets</span>. The recordings are processed by <span class="hlt">Droplet</span> Image Analysis in MATLAB to determine the relative velocities and the impact geometries for each individual collision. The collision outcomes are presented in a regime map dependent on the dimensionless impact parameter and Weber ( We) number. The Ohnesorge ( Oh) number is introduced to describe the effect of viscosity from one liquid to another and is maintained constant for each regime map by using a constant <span class="hlt">droplet</span> diameter ( d ˜ 700 μ m). In this work, a phenomenological model is proposed to describe the boundaries demarcating the coalescence-separation regimes. The collision dynamics and outcome of milk concentrates are compared with aqueous glycerol</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25166683','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25166683"><span><span class="hlt">Charge</span> fluctuations in nanoscale capacitors.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Limmer, David T; Merlet, Céline; Salanne, Mathieu; Chandler, David; Madden, Paul A; van Roij, René; Rotenberg, Benjamin</p> <p>2013-09-06</p> <p>The fluctuations of the <span class="hlt">charge</span> on an electrode contain information on the microscopic correlations within the adjacent fluid and their effect on the electronic properties of the interface. We investigate these fluctuations using molecular dynamics simulations in a constant-potential ensemble with histogram reweighting techniques. This approach offers, in particular, an efficient, accurate, and physically insightful route to the differential capacitance that is broadly applicable. We demonstrate these methods with three different capacitors: pure water between platinum electrodes and a pure as well as a <span class="hlt">solvent</span>-based organic electrolyte each between graphite electrodes. The total <span class="hlt">charge</span> distributions with the pure <span class="hlt">solvent</span> and <span class="hlt">solvent</span>-based electrolytes are remarkably Gaussian, while in the pure ionic liquid the total <span class="hlt">charge</span> distribution displays distinct non-Gaussian features, suggesting significant potential-driven changes in the organization of the interfacial fluid.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013PhRvL.111j6102L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013PhRvL.111j6102L"><span><span class="hlt">Charge</span> Fluctuations in Nanoscale Capacitors</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Limmer, David T.; Merlet, Céline; Salanne, Mathieu; Chandler, David; Madden, Paul A.; van Roij, René; Rotenberg, Benjamin</p> <p>2013-09-01</p> <p>The fluctuations of the <span class="hlt">charge</span> on an electrode contain information on the microscopic correlations within the adjacent fluid and their effect on the electronic properties of the interface. We investigate these fluctuations using molecular dynamics simulations in a constant-potential ensemble with histogram reweighting techniques. This approach offers, in particular, an efficient, accurate, and physically insightful route to the differential capacitance that is broadly applicable. We demonstrate these methods with three different capacitors: pure water between platinum electrodes and a pure as well as a <span class="hlt">solvent</span>-based organic electrolyte each between graphite electrodes. The total <span class="hlt">charge</span> distributions with the pure <span class="hlt">solvent</span> and <span class="hlt">solvent</span>-based electrolytes are remarkably Gaussian, while in the pure ionic liquid the total <span class="hlt">charge</span> distribution displays distinct non-Gaussian features, suggesting significant potential-driven changes in the organization of the interfacial fluid.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4841488','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4841488"><span>On well-posedness of variational models of <span class="hlt">charged</span> drops</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Novaga, Matteo</p> <p>2016-01-01</p> <p>Electrified liquids are well known to be prone to a variety of interfacial instabilities that result in the onset of apparent interfacial singularities and liquid fragmentation. In the case of electrically conducting liquids, one of the basic models describing the equilibrium interfacial configurations and the onset of instability assumes the liquid to be equipotential and interprets those configurations as local minimizers of the energy consisting of the sum of the surface energy and the electrostatic energy. Here we show that, surprisingly, this classical geometric variational model is mathematically ill-posed irrespective of the degree to which the liquid is electrified. Specifically, we demonstrate that an isolated spherical <span class="hlt">droplet</span> is never a local minimizer, no matter how small is the total <span class="hlt">charge</span> on the <span class="hlt">droplet</span>, as the energy can always be lowered by a smooth, arbitrarily small distortion of the <span class="hlt">droplet</span>'s surface. This is in sharp contrast to the experimental observations that a critical amount of <span class="hlt">charge</span> is needed in order to destabilize a spherical <span class="hlt">droplet</span>. We discuss several possible regularization mechanisms for the considered free boundary problem and argue that well-posedness can be restored by the inclusion of the entropic effects resulting in finite screening of free <span class="hlt">charges</span>. PMID:27118921</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29170768','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29170768"><span>Counterion effects on the ultrafast dynamics of <span class="hlt">charge-transfer-to-solvent</span> electrons.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Rivas, N; Moriena, G; Domenianni, L; Hodak, J H; Marceca, E</p> <p>2017-12-06</p> <p>We performed femtosecond transient absorption (TA) experiments to monitor the solvation dynamics of <span class="hlt">charge-transfer-to-solvent</span> (CTTS) electrons originating from UV photoexcitation of ammoniated iodide in close proximity to the counterions. Solutions of KI were prepared in liquid ammonia and TA experiments were carried out at different temperatures and densities, along the liquid-gas coexistence curve of the fluid. The results complement previous femtosecond TA work by P. Vöhringer's group in neat ammonia via multiphoton ionization. The dynamics of CTTS-detached electrons in ammonia was found to be strongly affected by ion pairing. Geminate recombination time constants as well as escape probabilities were determined from the measured temporal profiles and analysed as a function of the medium density. A fast unresolved (τ < 250 fs) increase of absorption related to the creation/thermalization of solvated electron species was followed by two decay components: one with a characteristic time around 10 ps, and a slower one that remains active for hundreds of picoseconds. While the first process is attributed to an early recombination of (I, e - ) pairs, the second decay and its asymptote reflects the effect of the K + counterion on the geminate recombination dynamics, rate and yield. The cation basically acts as an electron anchor that restricts the ejection distance, leading to <span class="hlt">solvent</span>-separated counterion-electron species. The formation of (K + , NH 3 , e - ) pairs close to the parent iodine atom brings the electron escape probability to very low values. Transient spectra of the electron species have also been estimated as a function of time by probing the temporal profiles at different wavelengths.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JChPh.146i4702S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JChPh.146i4702S"><span>Density functional description of size-dependent effects at nucleation on neutral and <span class="hlt">charged</span> nanoparticles</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shchekin, Alexander K.; Lebedeva, Tatiana S.</p> <p>2017-03-01</p> <p>A numerical study of size-dependent effects in the thermodynamics of a small <span class="hlt">droplet</span> formed around a solid nanoparticle has been performed within the square-gradient density functional theory. The Lennard-Jones fluid with the Carnahan-Starling model for the hard-sphere contribution to intermolecular interaction in liquid and vapor phases and interfaces has been used for description of the condensate. The intermolecular forces between the solid core and condensate molecules have been taken into account with the help of the Lennard-Jones part of the total molecular potential of the core. The influence of the electric <span class="hlt">charge</span> of the particle has been considered under assumption of the central Coulomb potential in the medium with dielectric permittivity depending on local condensate density. The condensate density profiles and equimolecular radii for equilibrium <span class="hlt">droplets</span> at different values of the condensate chemical potential have been computed in the cases of an uncharged solid core with the molecular potential, a <span class="hlt">charged</span> core without molecular potential, and a core with joint action of the Coulomb and molecular potentials. The appearance of stable equilibrium <span class="hlt">droplets</span> even in the absence of the electric <span class="hlt">charge</span> has been commented. As a next step, the capillary, disjoining pressure, and electrostatic contributions to the condensate chemical potential have been considered and compared with the predictions of classical thermodynamics in a wide range of values of the <span class="hlt">droplet</span> and the particle equimolecular radii. With the help of the found dependence of the condensate chemical potential in <span class="hlt">droplet</span> on the <span class="hlt">droplet</span> size, the activation barrier for nucleation on uncharged and <span class="hlt">charged</span> particles has been computed as a function of the vapor supersaturation. Finally, the work of <span class="hlt">droplet</span> formation and the work of wetting the particle have been found as functions of the <span class="hlt">droplet</span> size.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27438227','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27438227"><span>Development of a Tandem Electrodynamic Trap Apparatus for Merging <span class="hlt">Charged</span> <span class="hlt">Droplets</span> and Spectroscopic Characterization of Resultant Dried Particles.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kohno, Jun-Ya; Higashiura, Tetsu; Eguchi, Takaaki; Miura, Shumpei; Ogawa, Masato</p> <p>2016-08-11</p> <p>Materials work in multicomponent forms. A wide range of compositions must be tested to obtain the optimum composition for a specific application. We propose optimization using a series of small levitated single particles. We describe a tandem-trap apparatus for merging liquid <span class="hlt">droplets</span> and analyzing the merged <span class="hlt">droplets</span> and/or dried particles that are produced from the merged <span class="hlt">droplets</span> under levitation conditions. <span class="hlt">Droplet</span> merging was confirmed by Raman spectroscopic studies of the levitated particles. The tandem-trap apparatus enables the synthesis of a particle and spectroscopic investigation of its properties. This provides a basis for future investigation of the properties of levitated single particles.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4163133','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4163133"><span>What Protein <span class="hlt">Charging</span> (and Supercharging) Reveal about the Mechanism of Electrospray Ionization</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Loo, Rachel R. Ogorzalek; Lakshmanan, Rajeswari</p> <p>2014-01-01</p> <p>Understanding the <span class="hlt">charging</span> mechanism of electrospray ionization is central to overcoming shortcomings such as ion suppression or limited dynamic range and explaining phenomena such as supercharging. Towards that end, we explore what accumulated observations reveal about the mechanism of electrospray. We introduce the idea of an intermediate region for electrospray ionization (and other ionization methods) to account for the facts that solution <span class="hlt">charge</span> state distributions (CSDs) do not correlate to those observed by ESI– MS (the latter bear more <span class="hlt">charge</span>) and that gas phase reactions can reduce, but not increase the extent of <span class="hlt">charging</span>. This region incorporates properties, e.g., basicities, intermediate between solution and gas phase. Assuming that <span class="hlt">droplet</span> species polarize within the high electric field leads to equations describing ion emission resembling those from the equilibrium partitioning model. The equations predict many trends successfully, including CSD shifts to higher m/z for concentrated analytes and shifts to lower m/z for sprays employing smaller emitter opening diameters. From this view, a single mechanism can be formulated to explain how reagents that promote analyte <span class="hlt">charging</span> (“supercharging”) such as m–NBA, sulfolane, and 3–nitrobenzonitrile increase analyte <span class="hlt">charge</span> from “denaturing” and “native” <span class="hlt">solvent</span> systems. It is suggested that additives’ Brønsted basicities are inversely correlated to their ability to shift CSDs to lower m/z in positive ESI, as are Brønsted acidities for negative ESI. Because supercharging agents reduce an analyte's solution ionization, excess spray <span class="hlt">charge</span> is bestowed on evaporating ions carryingfewer opposing <span class="hlt">charges</span>. Brønsted basicity (or acidity) determines how much ESI <span class="hlt">charge</span> is lost to the agent (unavailable to evaporating analyte). PMID:25135609</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19905021','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19905021"><span>Effect of headgroup size, <span class="hlt">charge</span>, and <span class="hlt">solvent</span> structure on polymer-micelle interactions, studied by molecular dynamics simulations.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Shang, Barry Z; Wang, Zuowei; Larson, Ronald G</p> <p>2009-11-19</p> <p>We performed atomistic molecular dynamics simulations of anionic and cationic micelles in the presence of poly(ethylene oxide) (PEO) to understand why nonionic water-soluble polymers such as PEO interact strongly with anionic micelles but only weakly with cationic micelles. Our micelles include sodium n-dodecyl sulfate (SDS), n-dodecyl trimethylammonium chloride (DTAC), n-dodecyl ammonium chloride (DAC), and micelles in which we artificially reverse the sign of partial <span class="hlt">charges</span> in SDS and DTAC. We observe that the polymer interacts hydrophobically with anionic SDS but only weakly with cationic DTAC and DAC, in agreement with experiment. However, the polymer also interacts with the artificial anionic DTAC but fails to interact hydrophobically with the artificial cationic SDS, illustrating that large headgroup size does not explain the weak polymer interaction with cationic micelles. In addition, we observe through simulation that this preference for interaction with anionic micelles still exists in a dipolar "dumbbell" <span class="hlt">solvent</span>, indicating that water structure and hydrogen bonding alone cannot explain this preferential interaction. Our simulations suggest that direct electrostatic interactions between the micelle and polymer explain the preference for interaction with anionic micelles, even though the polymer overall carries no net <span class="hlt">charge</span>. This is possible given the asymmetric distribution of negative <span class="hlt">charges</span> on smaller atoms and positive <span class="hlt">charges</span> on larger units in the polymer chain.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016APS..DFD.M1003B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016APS..DFD.M1003B"><span>Acoustic <span class="hlt">droplet</span> vaporization of vascular <span class="hlt">droplets</span> in gas embolotherapy</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>2016-11-01</p> <p>This work is primarily motivated by a developmental gas embolotherapy technique for cancer treatment. In this methodology, infarction of tumors is induced by selectively formed vascular gas bubbles that arise from the acoustic vaporization of vascular <span class="hlt">droplets</span>. Additionally, micro- or nano-<span class="hlt">droplets</span> may be used as vehicles for localized drug delivery, with or without flow occlusion. In this talk, we examine the dynamics of acoustic <span class="hlt">droplet</span> vaporization through experiments and theoretical/computational fluid mechanics models, and investigate the bioeffects of acoustic <span class="hlt">droplet</span> vaporization on endothelial cells and in vivo. Functionalized <span class="hlt">droplets</span> that are targeted to tumor vasculature are examined. The influence of fluid mechanical and acoustic parameters, as well as <span class="hlt">droplet</span> functionalization, is explored. This work was supported by NIH Grant R01EB006476.</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('http://hdl.handle.net/2060/20110020390','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20110020390"><span>Modular <span class="hlt">droplet</span> actuator drive</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Pollack, Michael G. (Inventor); Paik, Philip (Inventor)</p> <p>2011-01-01</p> <p>A <span class="hlt">droplet</span> actuator drive including a detection apparatus for sensing a property of a <span class="hlt">droplet</span> on a <span class="hlt">droplet</span> actuator; circuitry for controlling the detection apparatus electronically coupled to the detection apparatus; a <span class="hlt">droplet</span> actuator cartridge connector arranged so that when a <span class="hlt">droplet</span> actuator cartridge electronically is coupled thereto: the <span class="hlt">droplet</span> actuator cartridge is aligned with the detection apparatus; and the detection apparatus can sense the property of the <span class="hlt">droplet</span> on a <span class="hlt">droplet</span> actuator; circuitry for controlling a <span class="hlt">droplet</span> actuator coupled to the <span class="hlt">droplet</span> actuator connector; and the <span class="hlt">droplet</span> actuator circuitry may be coupled to a processor.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1992CP....160..265M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1992CP....160..265M"><span>Intramolecular <span class="hlt">charge</span> transfer and trans-cis isomerization of the DCM styrene dye in polar <span class="hlt">solvents</span>. A CS INDO MRCI study</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Marguet, S.; Mialocq, J. C.; Millie, P.; Berthier, G.; Momicchioli, F.</p> <p>1992-03-01</p> <p>The <span class="hlt">solvent</span>-induced changes of trans-cis isomerization efficiency and electronic structure of the excited state of the DCM dye have been considered by means of CS INDO MRCI calculations. The potential energy curves, dipole moments and atomic <span class="hlt">charge</span> densities as a function of two internal coordinates, namely the rotation angle about the central "double" bond and the twisting of the dimethylamino group, have been obtained for the ground state and the lowest excited states. The structural requirements for the existence of ICT (intramolecular <span class="hlt">charge</span> transfer) excited states have been investigated by considering internal rotations about three single bonds. The reliability of the potential surfaces and of the solvation models has been discussed with reference to test-calculations on the DMABN molecule. In the first excited singlet state of DCM, the low-energy barrier for the trans-cis isomerization has been found unaffected by the <span class="hlt">solvent</span> polarity. The only singlet excited state presenting a large ICT character has been found to be the S 2 state for a perpendicularly twisted conformation of the dimethylamino group (TICT state). The assumption of a deactivation of the trans-isomer in the locally excited state through the TICT funnel has been largely discussed with reference to the simplifications of the present theoretical approach.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26616926','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26616926"><span>Bactericidal action mechanism of negatively <span class="hlt">charged</span> food grade clove oil nanoemulsions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Majeed, Hamid; Liu, Fei; Hategekimana, Joseph; Sharif, Hafiz Rizwan; Qi, Jing; Ali, Barkat; Bian, Yuan-Yuan; Ma, Jianguo; Yokoyama, Wallace; Zhong, Fang</p> <p>2016-04-15</p> <p>Clove oil (CO) anionic nanoemulsions were prepared with varying ratios of CO to canola oil (CA), emulsified and stabilized with purity gum ultra (PGU), a newly developed succinylated waxy maize starch. Interfacial tension measurements showed that CO acted as a co-surfactant and there was a gradual decrease in interfacial tension which favored the formation of small <span class="hlt">droplet</span> sizes on homogenization until a critical limit (5:5% v/v CO:CA) was reached. Antimicrobial activity of the negatively <span class="hlt">charged</span> CO nanoemulsion was determined against Gram positive GPB (Listeria monocytogenes and Staphylococcus aureus) and Gram negative GNB (Escherichia coli) bacterial strains using minimum inhibitory concentration (MIC) and a time kill dynamic method. Negatively <span class="hlt">charged</span> PGU emulsified CO nanoemulsion showed prolonged antibacterial activities against Gram positive bacterial strains. We concluded that negatively <span class="hlt">charged</span> CO nanoemulsion <span class="hlt">droplets</span> self-assemble with GPB cell membrane, and facilitated interaction with cellular components of bacteria. Moreover, no electrostatic interaction existed between negatively <span class="hlt">charged</span> <span class="hlt">droplets</span> and the GPB membrane. Copyright © 2015 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25409489','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25409489"><span>Secondary organic aerosol formation during evaporation of <span class="hlt">droplets</span> containing atmospheric aldehydes, amines, and ammonium sulfate.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Galloway, Melissa M; Powelson, Michelle H; Sedehi, Nahzaneen; Wood, Stephanie E; Millage, Katherine D; Kononenko, Julia A; Rynaski, Alec D; De Haan, David O</p> <p>2014-12-16</p> <p>Reactions of carbonyl compounds in cloudwater produce organic aerosol mass through in-cloud oxidation and during postcloud evaporation. In this work, postcloud evaporation was simulated in laboratory experiments on evaporating <span class="hlt">droplets</span> that contain mixtures of common atmospheric aldehydes with ammonium sulfate (AS), methylamine, or glycine. Aerosol diameters were measured during monodisperse <span class="hlt">droplet</span> drying experiments and during polydisperse <span class="hlt">droplet</span> equilibration experiments at 75% relative humidity, and condensed-phase mass was measured in bulk thermogravimetric experiments. The evaporation of water from a <span class="hlt">droplet</span> was found to trigger aldehyde reactions that increased residual particle volumes by a similar extent in room-temperature experiments, regardless of whether AS, methylamine, or glycine was present. The production of organic aerosol volume was highest from <span class="hlt">droplets</span> containing glyoxal, followed by similar production from methylglyoxal or hydroxyacetone. Significant organic aerosol production was observed for glycolaldehyde, acetaldehyde, and formaldehyde only at elevated temperatures in thermogravimetric experiments. In many experiments, the amount of aerosol produced was greater than the sum of all solutes plus nonvolatile <span class="hlt">solvent</span> impurities, indicating the additional presence of trapped water, likely caused by increasing aerosol-phase viscosity due to oligomer formation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26994584','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26994584"><span>A QSPR study on the <span class="hlt">solvent</span>-induced frequency shifts of acetone and dimethyl sulfoxide in organic <span class="hlt">solvents</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ou, Yu Heng; Chang, Chia Ming; Chen, Ying Shao</p> <p>2016-06-05</p> <p>In this study, <span class="hlt">solvent</span>-induced frequency shifts (SIFS) in the infrared spectrum of acetone and dimethyl sulfoxide in organic <span class="hlt">solvents</span> were investigated by using four types of quantum-chemical reactivity descriptors. The results showed that the SIFS of acetone is mainly affected by the electron-acceptance chemical potential and the maximum nucleophilic condensed local softness of organic <span class="hlt">solvents</span>, which represent the electron flow and the polarization between acetone and <span class="hlt">solvent</span> molecules. On the other hand, the SIFS of dimethyl sulfoxide changes with the maximum positive <span class="hlt">charge</span> of hydrogen atom and the inverse of apolar surface area of <span class="hlt">solvent</span> molecules, showing that the electrostatic and hydrophilic interactions are main mechanisms between dimethyl sulfoxide and <span class="hlt">solvent</span> molecules. The introduction of the four-element theory model-based quantitative structure-property relationship approach improved the assessing quality and provided a basis for interpreting the solute-<span class="hlt">solvent</span> interactions. Copyright © 2016 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16732657','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16732657"><span><span class="hlt">Droplet</span> size effects on film drainage between <span class="hlt">droplet</span> and substrate.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Steinhaus, Benjamin; Spicer, Patrick T; Shen, Amy Q</p> <p>2006-06-06</p> <p>When a <span class="hlt">droplet</span> approaches a solid surface, the thin liquid film between the <span class="hlt">droplet</span> and the surface drains until an instability forms and then ruptures. In this study, we utilize microfluidics to investigate the effects of film thickness on the time to film rupture for water <span class="hlt">droplets</span> in a flowing continuous phase of silicone oil deposited on solid poly(dimethylsiloxane) (PDMS) surfaces. The water <span class="hlt">droplets</span> ranged in size from millimeters to micrometers, resulting in estimated values of the film thickness at rupture ranging from 600 nm down to 6 nm. The Stefan-Reynolds equation is used to model film drainage beneath both millimeter- and micrometer-scale <span class="hlt">droplets</span>. For millimeter-scale <span class="hlt">droplets</span>, the experimental and analytical film rupture times agree well, whereas large differences are observed for micrometer-scale <span class="hlt">droplets</span>. We speculate that the differences in the micrometer-scale data result from the increases in the local thin film viscosity due to confinement-induced molecular structure changes in the silicone oil. A modified Stefan-Reynolds equation is used to account for the increased thin film viscosity of the micrometer-scale <span class="hlt">droplet</span> drainage case.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27535608','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27535608"><span><span class="hlt">Droplet</span> size influences division of mammalian cell factories in <span class="hlt">droplet</span> microfluidic cultivation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Periyannan Rajeswari, Prem Kumar; Joensson, Haakan N; Andersson-Svahn, Helene</p> <p>2017-01-01</p> <p>The potential of using <span class="hlt">droplet</span> microfluidics for screening mammalian cell factories has been limited by the difficulty in achieving continuous cell division during cultivation in <span class="hlt">droplets</span>. Here, we report the influence of <span class="hlt">droplet</span> size on mammalian cell division and viability during cultivation in <span class="hlt">droplets</span>. Chinese Hamster Ovary (CHO) cells, the most widely used mammalian host cells for biopharmaceuticals production were encapsulated and cultivated in 33, 180 and 320 pL <span class="hlt">droplets</span> for 3 days. Periodic monitoring of the <span class="hlt">droplets</span> during incubation showed that the cell divisions in 33 pL <span class="hlt">droplets</span> stopped after 24 h, whereas continuous cell division was observed in 180 and 320 pL <span class="hlt">droplets</span> for 72 h. The viability of the cells cultivated in the 33 pL <span class="hlt">droplets</span> also dropped to about 50% in 72 h. In contrast, the viability of the cells in the larger <span class="hlt">droplets</span> was above 90% even after 72 h of cultivation, making them a more suitable <span class="hlt">droplet</span> size for 72-h cultivation. This study shows a direct correlation of microfluidic <span class="hlt">droplet</span> size to the division and viability of mammalian cells. This highlights the importance of selecting suitable <span class="hlt">droplet</span> size for mammalian cell factory screening assays. © 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28460920','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28460920"><span>Heteroaggregation of lipid <span class="hlt">droplets</span> coated with sodium caseinate and lactoferrin.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>de Figueiredo Furtado, Guilherme; Michelon, Mariano; de Oliveira, Davi Rocha Bernardes; da Cunha, Rosiane Lopes</p> <p>2016-11-01</p> <p>Formation and characterization of <span class="hlt">droplet</span> heteroaggregates were investigated by mixing two emulsions previously stabilized by proteins oppositely <span class="hlt">charged</span>. Emulsions were composed of 5vol.% of sunflower oil and 95vol.% of sodium caseinate or lactoferrin aqueous dispersions. They were produced using ultrasound with fixed power (300W) and sonication time (6min). Different volume ratios (0-100%) of sodium caseinate-stabilized emulsion (<span class="hlt">droplet</span> diameter around 1.75μm) to lactoferrin-stabilized emulsion (<span class="hlt">droplet</span> diameter around 1.55μm) were mixed under conditions that both proteins showed opposite <span class="hlt">charges</span> (pH7). Influence of ionic strength (0-400mM NaCl) on the heteroaggregates stability was also evaluated. Creaming stability, zeta potential, microstructure, mean particle diameter and rheological properties of the heteroaggregates were measured. These properties depended on the volume ratio (0-100%) of sodium caseinate to lactoferrin-stabilized emulsion (C:L) and the ionic strength. In the absence of salt, different zeta potential values were obtained, rheological properties (viscosity and elastic moduli) were improved and the largest heteroaggregates were formed at higher content of lactoferrin-stabilized emulsion (60-80%). The system containing 40 and 60vol.% of sodium caseinate and lactoferrin stabilized emulsion, respectively, presented good stability against phase separation besides showing enhanced rheological and size properties due to extensive <span class="hlt">droplets</span> aggregation. Phase separation was observed only in the absence of sodium caseinate, demonstrating the higher susceptibility of lactoferrin to NaCl. The heteroaggregates produced may be useful functional agents for texture modification and controlled release since different rheological properties and sizes can be achieved depending on protein concentrations. Copyright © 2016 Elsevier Ltd. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1096257','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1096257"><span>An interface tracking model for <span class="hlt">droplet</span> electrocoalescence.</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>Erickson, Lindsay Crowl</p> <p></p> <p>This report describes an Early Career Laboratory Directed Research and Development (LDRD) project to develop an interface tracking model for <span class="hlt">droplet</span> electrocoalescence. Many fluid-based technologies rely on electrical fields to control the motion of <span class="hlt">droplets</span>, e.g. microfluidic devices for high-speed <span class="hlt">droplet</span> sorting, solution separation for chemical detectors, and purification of biodiesel fuel. Precise control over <span class="hlt">droplets</span> is crucial to these applications. However, electric fields can induce complex and unpredictable fluid dynamics. Recent experiments (Ristenpart et al. 2009) have demonstrated that oppositely <span class="hlt">charged</span> <span class="hlt">droplets</span> bounce rather than coalesce in the presence of strong electric fields. A transient aqueous bridge forms betweenmore » approaching drops prior to pinch-off. This observation applies to many types of fluids, but neither theory nor experiments have been able to offer a satisfactory explanation. Analytic hydrodynamic approximations for interfaces become invalid near coalescence, and therefore detailed numerical simulations are necessary. This is a computationally challenging problem that involves tracking a moving interface and solving complex multi-physics and multi-scale dynamics, which are beyond the capabilities of most state-of-the-art simulations. An interface-tracking model for electro-coalescence can provide a new perspective to a variety of applications in which interfacial physics are coupled with electrodynamics, including electro-osmosis, fabrication of microelectronics, fuel atomization, oil dehydration, nuclear waste reprocessing and solution separation for chemical detectors. We present a conformal decomposition finite element (CDFEM) interface-tracking method for the electrohydrodynamics of two-phase flow to demonstrate electro-coalescence. CDFEM is a sharp interface method that decomposes elements along fluid-fluid boundaries and uses a level set function to represent the interface.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018PhFl...30c2002L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018PhFl...30c2002L"><span>Internal flow inside <span class="hlt">droplets</span> within a concentrated emulsion during <span class="hlt">droplet</span> rearrangement</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Leong, Chia Min; Gai, Ya; Tang, Sindy K. Y.</p> <p>2018-03-01</p> <p><span class="hlt">Droplet</span> microfluidics, in which each <span class="hlt">droplet</span> serves as a micro-reactor, has found widespread use in high-throughput biochemical screening applications. These <span class="hlt">droplets</span> are often concentrated at various steps to form a concentrated emulsion. As part of a serial interrogation and sorting process, such concentrated emulsions are typically injected into a tapered channel leading to a constriction that fits one drop at a time for the probing of <span class="hlt">droplet</span> content in a serial manner. The flow physics inside the <span class="hlt">droplets</span> under these flow conditions are not well understood but are critical for predicting and controlling the mixing of reagents inside the <span class="hlt">droplets</span> as reactors. Here we investigate the flow field inside <span class="hlt">droplets</span> of a concentrated emulsion flowing through a tapered microchannel using micro-particle image velocimetry. The confining geometry of the channel forces the number of rows of drops to reduce by one at specific and uniformly spaced streamwise locations, which are referred to as <span class="hlt">droplet</span> rearrangement zones. Within each rearrangement zone, the phase-averaged velocity results show that the motion of the <span class="hlt">droplets</span> involved in the rearrangement process, also known as a T1 event, creates vortical structures inside themselves and their adjacent <span class="hlt">droplets</span>. These flow structures increase the circulation inside <span class="hlt">droplets</span> up to 2.5 times the circulation in <span class="hlt">droplets</span> at the constriction. The structures weaken outside of the rearrangement zones suggesting that the flow patterns created by the T1 process are transient. The time scale of circulation is approximately the same as the time scale of a T1 event. Outside of the rearrangement zones, flow patterns in the <span class="hlt">droplets</span> are determined by the relative velocity between the continuous and disperse phases.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AcSpA.188...72F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AcSpA.188...72F"><span>Air-assisted liquid-liquid microextraction using floating organic <span class="hlt">droplet</span> solidification for simultaneous extraction and spectrophotometric determination of some drugs in biological samples through chemometrics methods</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Farahmand, Farnaz; Ghasemzadeh, Bahar; Naseri, Abdolhossein</p> <p>2018-01-01</p> <p>An air assisted liquid-liquid microextraction by applying the solidification of a floating organic <span class="hlt">droplet</span> method (AALLME-SFOD) coupled with a multivariate calibration method, namely partial least squares (PLS), was introduced for the fast and easy determination of Atenolol (ATE), Propanolol (PRO) and Carvedilol (CAR) in biological samples via a spectrophotometric approach. The analytes would be extracted from neutral aqueous solution into 1-dodecanol as an organic <span class="hlt">solvent</span>, using AALLME. In this approach a low-density <span class="hlt">solvent</span> with a melting point close to room temperature was applied as the extraction <span class="hlt">solvent</span>. The emulsion was immediately formed by repeatedly pulling in and pushing out the aqueous sample solution and extraction <span class="hlt">solvent</span> mixture via a 10-mL glass syringe for ten times. After centrifugation, the extractant <span class="hlt">droplet</span> could be simply collected from the aqueous samples by solidifying the emulsion at a lower than the melting point temperature. In the next step, analytes were back extracted simultaneously into the acidic aqueous solution. Derringer and Suich multi-response optimization were utilized for simultaneous optimizing the parameters of three analytes. This method incorporates the benefits of AALLME and dispersive liquid-liquid microextraction considering the solidification of floating organic <span class="hlt">droplets</span> (DLLME-SFOD). Calibration graphs under optimized conditions were linear in the range of 0.30-6.00, 0.32-2.00 and 0.30-1.40 μg mL- 1 for ATE, CAR and PRO, respectively. Other analytical parameters were obtained as follows: enrichment factors (EFs) were found to be 11.24, 16.55 and 14.90, and limits of detection (LODs) were determined to be 0.09, 0.10 and 0.08 μg mL- 1 for ATE, CAR and PRO, respectively. The proposed method will require neither a highly toxic chlorinated <span class="hlt">solvent</span> for extraction nor an organic dispersive <span class="hlt">solvent</span> in the application process; hence, it is more environmentally friendly.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3526397','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3526397"><span><span class="hlt">Droplet</span> centrifugation, <span class="hlt">droplet</span> DNA extraction, and rapid <span class="hlt">droplet</span> thermocycling for simpler and faster PCR assay using wire-guided manipulations</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>2012-01-01</p> <p>A computer numerical control (CNC) apparatus was used to perform <span class="hlt">droplet</span> centrifugation, <span class="hlt">droplet</span> DNA extraction, and rapid <span class="hlt">droplet</span> thermocycling on a single superhydrophobic surface and a multi-chambered PCB heater. <span class="hlt">Droplets</span> were manipulated using “wire-guided” method (a pipette tip was used in this study). This methodology can be easily adapted to existing commercial robotic pipetting system, while demonstrated added capabilities such as vibrational mixing, high-speed centrifuging of <span class="hlt">droplets</span>, simple DNA extraction utilizing the hydrophobicity difference between the tip and the superhydrophobic surface, and rapid thermocycling with a moving <span class="hlt">droplet</span>, all with wire-guided <span class="hlt">droplet</span> manipulations on a superhydrophobic surface and a multi-chambered PCB heater (i.e., not on a 96-well plate). Serial dilutions were demonstrated for diluting sample matrix. Centrifuging was demonstrated by rotating a 10 μL <span class="hlt">droplet</span> at 2300 round per minute, concentrating E. coli by more than 3-fold within 3 min. DNA extraction was demonstrated from E. coli sample utilizing the disposable pipette tip to cleverly attract the extracted DNA from the <span class="hlt">droplet</span> residing on a superhydrophobic surface, which took less than 10 min. Following extraction, the 1500 bp sequence of Peptidase D from E. coli was amplified using rapid <span class="hlt">droplet</span> thermocycling, which took 10 min for 30 cycles. The total assay time was 23 min, including <span class="hlt">droplet</span> centrifugation, <span class="hlt">droplet</span> DNA extraction and rapid <span class="hlt">droplet</span> thermocycling. Evaporation from of 10 μL <span class="hlt">droplets</span> was not significant during these procedures, since the longest time exposure to air and the vibrations was less than 5 min (during DNA extraction). The results of these sequentially executed processes were analyzed using gel electrophoresis. Thus, this work demonstrates the adaptability of the system to replace many common laboratory tasks on a single platform (through re-programmability), in rapid succession (using <span class="hlt">droplets</span>), and with a high level of</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007PhDT.......102C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007PhDT.......102C"><span>Thermodynamic and kinetic control of <span class="hlt">charged</span>, amphiphilic triblock copolymer assembly via interaction with organic counterions in <span class="hlt">solvent</span> mixtures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cui, Honggang</p> <p>2007-12-01</p> <p>Amphiphilic block copolymers, consisting of at least two types of monomers with different affinity to the dissolving <span class="hlt">solvent(s</span>), have been recognized as a molecular building unit for their chemical tunability and design flexibility. Amphiphilic block copolymers with a chargeable block have structural features of polyelectrolytes, block copolymers and surfactants. The combination of these different features offers great flexibility for developing novel assembled morphologies at the nanoscale and outstanding ability to control and manipulate those morphologies. The nanostructures, formed from the spontaneous association of amphiphilic block copolymer in selective <span class="hlt">solvents</span>, show promise for applications in nanotechnology and pharmaceuticals, including drug delivery, tissue engineering and bio-imaging. A basic knowledge of their modes of self-assembly and their correspondence to application-related properties is just now being developed and poses a considerable scientific challenge. The goal of this dissertation is to investigate the associative behavior of <span class="hlt">charged</span>, amphiphilic block copolymers in <span class="hlt">solvent</span> mixtures while in the presence of organic counterions. Self-assembly of poly (acrylic acid)- block-poly (methyl acrylate)-block-polystyrene (PAA- b-PMA-b-PS) triblock copolymers produces nanodomains in THF/water solution specifically through the interaction with organic counterions (polyamines). These assembled structures can include classic micelles (spheres, cylinders and vesicles), but, more importantly, include non-classic micelles (disks, toroids, branched micelles and segmented micelles). Each micelle structure is stable and reproducible at different assembly conditions. The assembled micellar structures depend on not only solution components (thermodynamics) but also mixing procedure and consequent self-assembly pathway (kinetics). The key factors that determine the thermodynamic interactions that partially define the assembled structures and the kinetic</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28239060','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28239060"><span>Polydimethylsiloxane <span class="hlt">Droplets</span> Exhibit Extraordinarily High Antioxidative Effects in Deep-Frying.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Totani, Nagao; Yazaki, Naoko; Yawata, Miho</p> <p>2017-04-03</p> <p>The addition of more than about 1 ppm polydimethylsiloxane (PDMS) into oil results in PDMS forming both a layer at the oil-air interface and <span class="hlt">droplets</span> suspended in the oil. It is widely accepted that the extraordinarily strong and stable antioxidative effects of PDMS are due to the PDMS layer. However, the PDMS layer showed no antioxidative effects when canola oil did not contain <span class="hlt">droplets</span> but rather was covered with a layer of PDMS, then subjected to heating under high agitation to mimic deep-frying. Furthermore, no antioxidative effect was exhibited by oil-soluble methylphenylsiloxane (PMPS) in canola oil or by PDMS in PDMS-soluble canola oil fatty acid ester during heating, suggesting that PDMS must be insoluble and <span class="hlt">droplets</span> in oil in order for PDMS to exhibit an antioxidative effect during deep-frying. The zeta potential of PDMS <span class="hlt">droplets</span> suspended in canola oil was very high and thus the negatively <span class="hlt">charged</span> PDMS <span class="hlt">droplets</span> should attract nearby low molecular weight compounds. It was suggested that this attraction disturbed the motion of oxygen molecules and prevented their attack against unsaturated fatty acid moiety. This would be the reason in the deep-frying why PDMS suppressed the oxidation reaction of oil. PDMS <span class="hlt">droplets</span> also attracted volatile compounds (molecular weight below 125 Da) generated by heating canola oil. Thus, adding PDMS to oil after heating the oil resulted in the heated oil smelling less than heated oil without PDMS.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20120015397','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20120015397"><span><span class="hlt">Droplet</span> Deformation Prediction With the <span class="hlt">Droplet</span> Deformation and Breakup Model (DDB)</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Vargas, Mario</p> <p>2012-01-01</p> <p>The <span class="hlt">Droplet</span> Deformation and Breakup Model was used to predict deformation of <span class="hlt">droplets</span> approaching the leading edge stagnation line of an airfoil. The quasi-steady model was solved for each position along the <span class="hlt">droplet</span> path. A program was developed to solve the non-linear, second order, ordinary differential equation that governs the model. A fourth order Runge-Kutta method was used to solve the equation. Experimental slip velocities from <span class="hlt">droplet</span> breakup studies were used as input to the model which required slip velocity along the particle path. The center of mass displacement predictions were compared to the experimental measurements from the <span class="hlt">droplet</span> breakup studies for <span class="hlt">droplets</span> with radii in the range of 200 to 700 mm approaching the airfoil at 50 and 90 m/sec. The model predictions were good for the displacement of the center of mass for small and medium sized <span class="hlt">droplets</span>. For larger <span class="hlt">droplets</span> the model predictions did not agree with the experimental results.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3311332','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3311332"><span><span class="hlt">Droplet</span> formation and scaling in dense suspensions</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Miskin, Marc Z.; Jaeger, Heinrich M.</p> <p>2012-01-01</p> <p>When a dense suspension is squeezed from a nozzle, <span class="hlt">droplet</span> detachment can occur similar to that of pure liquids. While in pure liquids the process of <span class="hlt">droplet</span> detachment is well characterized through self-similar profiles and known scaling laws, we show here the simple presence of particles causes suspensions to break up in a new fashion. Using high-speed imaging, we find that detachment of a suspension drop is described by a power law; specifically we find the neck minimum radius, rm, scales like near breakup at time τ = 0. We demonstrate data collapse in a variety of particle/liquid combinations, packing fractions, <span class="hlt">solvent</span> viscosities, and initial conditions. We argue that this scaling is a consequence of particles deforming the neck surface, thereby creating a pressure that is balanced by inertia, and show how it emerges from topological constraints that relate particle configurations with macroscopic Gaussian curvature. This new type of scaling, uniquely enforced by geometry and regulated by the particles, displays memory of its initial conditions, fails to be self-similar, and has implications for the pressure given at generic suspension interfaces. PMID:22392979</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3838431','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3838431"><span>Calculating the binding free energies of <span class="hlt">charged</span> species based on explicit-<span class="hlt">solvent</span> simulations employing lattice-sum methods: An accurate correction scheme for electrostatic finite-size effects</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Rocklin, Gabriel J.; Mobley, David L.; Dill, Ken A.; Hünenberger, Philippe H.</p> <p>2013-01-01</p> <p> calculations for a given system, its dependence on the box size being analytical. The latter scheme also provides insight into the physical origin of the finite-size effects. These two schemes also encompass a correction for discrete <span class="hlt">solvent</span> effects that persists even in the limit of infinite box sizes. Application of either scheme essentially eliminates the size dependence of the corrected <span class="hlt">charging</span> free energies (maximal deviation of 1.5 kJ mol−1). Because it is simple to apply, the analytical correction scheme offers a general solution to the problem of finite-size effects in free-energy calculations involving <span class="hlt">charged</span> solutes, as encountered in calculations concerning, e.g., protein-ligand binding, biomolecular association, residue mutation, pKa and redox potential estimation, substrate transformation, solvation, and <span class="hlt">solvent-solvent</span> partitioning. PMID:24320250</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24320250','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24320250"><span>Calculating the binding free energies of <span class="hlt">charged</span> species based on explicit-<span class="hlt">solvent</span> simulations employing lattice-sum methods: an accurate correction scheme for electrostatic finite-size effects.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Rocklin, Gabriel J; Mobley, David L; Dill, Ken A; Hünenberger, Philippe H</p> <p>2013-11-14</p> <p> calculations for a given system, its dependence on the box size being analytical. The latter scheme also provides insight into the physical origin of the finite-size effects. These two schemes also encompass a correction for discrete <span class="hlt">solvent</span> effects that persists even in the limit of infinite box sizes. Application of either scheme essentially eliminates the size dependence of the corrected <span class="hlt">charging</span> free energies (maximal deviation of 1.5 kJ mol(-1)). Because it is simple to apply, the analytical correction scheme offers a general solution to the problem of finite-size effects in free-energy calculations involving <span class="hlt">charged</span> solutes, as encountered in calculations concerning, e.g., protein-ligand binding, biomolecular association, residue mutation, pKa and redox potential estimation, substrate transformation, solvation, and <span class="hlt">solvent-solvent</span> partitioning.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013JChPh.139r4103R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013JChPh.139r4103R"><span>Calculating the binding free energies of <span class="hlt">charged</span> species based on explicit-<span class="hlt">solvent</span> simulations employing lattice-sum methods: An accurate correction scheme for electrostatic finite-size effects</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rocklin, Gabriel J.; Mobley, David L.; Dill, Ken A.; Hünenberger, Philippe H.</p> <p>2013-11-01</p> <p> calculations for a given system, its dependence on the box size being analytical. The latter scheme also provides insight into the physical origin of the finite-size effects. These two schemes also encompass a correction for discrete <span class="hlt">solvent</span> effects that persists even in the limit of infinite box sizes. Application of either scheme essentially eliminates the size dependence of the corrected <span class="hlt">charging</span> free energies (maximal deviation of 1.5 kJ mol-1). Because it is simple to apply, the analytical correction scheme offers a general solution to the problem of finite-size effects in free-energy calculations involving <span class="hlt">charged</span> solutes, as encountered in calculations concerning, e.g., protein-ligand binding, biomolecular association, residue mutation, pKa and redox potential estimation, substrate transformation, solvation, and <span class="hlt">solvent-solvent</span> partitioning.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22945858','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22945858"><span>Rebounding <span class="hlt">droplet-droplet</span> collisions on superhydrophobic surfaces: from the phenomenon to <span class="hlt">droplet</span> logic.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Mertaniemi, Henrikki; Forchheimer, Robert; Ikkala, Olli; Ras, Robin H A</p> <p>2012-11-08</p> <p>When water <span class="hlt">droplets</span> impact each other while traveling on a superhydrophobic surface, we demonstrate that they are able to rebound like billiard balls. We present elementary Boolean logic operations and a flip-flop memory based on these rebounding water <span class="hlt">droplet</span> collisions. Furthermore, bouncing or coalescence can be easily controlled by process parameters. Thus by the controlled coalescence of reactive <span class="hlt">droplets</span>, here using the quenching of fluorescent metal nanoclusters as a model reaction, we also demonstrate an elementary operation for programmable chemistry. Copyright © 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.</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/2014APS..DFD.F1028A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014APS..DFD.F1028A"><span>Dynamics of skirting <span class="hlt">droplets</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Akers, Caleb; Hale, Jacob</p> <p>2014-11-01</p> <p>It has been observed that non-coalescence between a <span class="hlt">droplet</span> and pool of like fluid can be prolonged or inhibited by sustained relative motion between the two fluids. In this study, we quantitatively describe the motion of freely moving <span class="hlt">droplets</span> that skirt across the surface of a still pool of like fluid. <span class="hlt">Droplets</span> of different sizes and small Weber number were directed horizontally onto the pool surface. After stabilization of the <span class="hlt">droplet</span> shape after impact, the <span class="hlt">droplets</span> smoothly moved across the surface, slowing until coalescence. Using high-speed imaging, we recorded the <span class="hlt">droplet</span>'s trajectory from a top-down view as well as side views both slightly above and below the fluid surface. The <span class="hlt">droplets</span>' speed is observed to decrease exponentially, with the smaller <span class="hlt">droplets</span> slowing down at a greater rate. <span class="hlt">Droplets</span> infused with neutral density micro beads showed that the <span class="hlt">droplet</span> rolls along the surface of the pool. A qualitative model of this motion is presented.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1986PhDT........78M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1986PhDT........78M"><span><span class="hlt">Droplet</span> Growth</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Marder, Michael Paolo</p> <p></p> <p>When a mixture of two materials, such as aluminum and tin, or alcohol and water, is cooled below a certain temperature, the two components begin to separate. If one component is dilute in the other, it may separate out in the form of small spheres, and these will begin to enlarge, depleting the supersaturated material around them. If the dynamics is sufficiently slow, thermodynamics gives one considerable information about how the <span class="hlt">droplets</span> grow. Two types of experiment have explored this behavior and given puzzling results. Nucleation experiments measure the rate at which <span class="hlt">droplets</span> initially appear from a seemingly homogeneous mixture. Near the critical point in binary liquids, experiments conducted in the 1960's and early 1970's showed that nucleation was vastly slower than theory seemed to predict. The resolution of this problem arises by considering in detail the dynamics of growing <span class="hlt">droplets</span> and comparing it with what experiments actually measure. Here will be presented a more detailed comparison of theory and experiment than has before been completed, obtaining satisfactory agreement with no free parameters needed. A second type of experiment measures <span class="hlt">droplet</span> size distributions after long times. In the late stage, <span class="hlt">droplets</span> compete with each other for material, a few growing at the expense of others. A theory first proposed by Lifshitz and Slyozov claims that this distribution, properly scaled, should be universal, and independent of properties of materials. Yet experimental measurements consistently find distributions that are more broad and squat than the theory would predict. Satisfactory agreement with experiment can be achieved by considering two points. First, one must study the complete time development of <span class="hlt">droplet</span> size distributions, to understand when the asymptotic regime obtains. Second, <span class="hlt">droplet</span> size distributions are spread by correlations between <span class="hlt">droplets</span>. If one finds a small <span class="hlt">droplet</span>, it is small because large <span class="hlt">droplets</span> nearby are competing with it</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29806941','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29806941"><span>Magnetically Responsive Superhydrophobic Surface: In Situ Reversible Switching of Water <span class="hlt">Droplet</span> Wettability and Adhesion for <span class="hlt">Droplet</span> Manipulation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yang, Chao; Wu, Lei; Li, Gang</p> <p>2018-06-13</p> <p>A smart, magnetically responsive superhydrophobic surface was facilely prepared by combining spray coating and magnetic-field-directed self-assembly. The surface comprised a dense array of magnetorheological elastomer micropillars (MREMPs). Benefitting from the magnetic field-stiffening effect of the MREMPs, the surface exhibited reversible switching of the wettability and adhesion that was responsive to an on/off magnetic field. The wettability and adhesion properties of the surfaces with MREMPs were investigated under different magnetic fields. The results revealed that the adhesion force and sliding behaviors of these surfaces were strongly dependent on the intensity of the applied magnetic field and the mixing ratio of poly(dimethylsiloxane) (PDMS), iron particles, and <span class="hlt">solvent</span> (in solution) used for preparation of the magnetically responsive superhydrophobic surfaces. The adhesion transition was attributed to the tunable mechanical properties of the MREMPs, which was easily controlled by an external magnetic field. It was also demonstrated that the magnetically responsive superhydrophobic surface can be used as a "mechanical hand" for no-loss liquid <span class="hlt">droplet</span> transportation. This magnetically responsive superhydrophobic surface not only provides a novel interface for microfluidic control and <span class="hlt">droplet</span> transportation, but also opens up new avenues for achieving smart liquid-repellent skin, programmable fluid collection and transport, and smart microfluidic devices.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010AcSpA..75.1347S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010AcSpA..75.1347S"><span>Spectrophotometric and spectroscopic studies of <span class="hlt">charge</span> transfer complexes of p-toluidine as an electron donor with picric acid as an electron acceptor in different <span class="hlt">solvents</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Singh, Neeti; Khan, Ishaat M.; Ahmad, Afaq</p> <p>2010-04-01</p> <p>The <span class="hlt">charge</span> transfer complexes of the donor p-toluidine with π-acceptor picric acid have been studied spectrophotometrically in various <span class="hlt">solvents</span> such as carbon tetrachloride, chloroform, dichloromethane acetone, ethanol, and methanol at room temperature using absorption spectrophotometer. The results indicate that formation of CTC in non-polar <span class="hlt">solvent</span> is high. The stoichiometry of the complex was found to be 1:1 ratio by straight-line method between donor and acceptor with maximum absorption bands. The data are discussed in terms of formation constant ( KCT), molar extinction coefficient ( ɛCT), standard free energy (Δ Go), oscillator strength ( f), transition dipole moment ( μEN), resonance energy ( RN) and ionization potential ( ID). The results indicate that the formation constant ( KCT) for the complex was shown to be dependent upon the nature of electron acceptor, donor and polarity of <span class="hlt">solvents</span> that were used.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19950002783','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19950002783"><span>The effects of turbulence on <span class="hlt">droplet</span> drag and secondary <span class="hlt">droplet</span> breakup</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Song, Y.-H.; Coy, E.; Greenfield, S.; Ondas, M.; Prevish, T.; Spegar, T.; Santavicca, D.</p> <p>1994-01-01</p> <p>The objective of this research is to obtain an improved understanding of the behavior of <span class="hlt">droplets</span> in vaporizing sprays, particularly under conditions typical of those in high pressure rocket sprays. Experiments are conducted in a variety of high pressure, high temperature, optically-accessible flow systems, including one which is capable of operation at pressures up to 70 atm, temperatures up to 600 K, gas velocities up to 30 m/sec and turbulence intensities up to 40 percent. Single <span class="hlt">droplets</span>, 50 to 500 micron in diameter, are produced by an aerodynamic <span class="hlt">droplet</span> generator and transversely injected into the flow. Measurements are made of the <span class="hlt">droplet</span> position, size, velocity and temperature and of the <span class="hlt">droplet</span>'s vapor wake from which <span class="hlt">droplet</span> drag, dispersion, heating, vaporization and breakup are characterized.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24375950','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24375950"><span>Altering the self-organization of dyes on titania with dyeing <span class="hlt">solvents</span> to tune the <span class="hlt">charge</span>-transfer dynamics of sensitized solar cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Yinglin; Yang, Lin; Zhang, Jing; Li, Renzhi; Zhang, Min; Wang, Peng</p> <p>2014-04-14</p> <p>Herein we selected the model organic donor-acceptor dye C218 and modulated the self-organization of dye molecules on the surface of titania by changing the dyeing <span class="hlt">solvent</span> from chlorobenzene to a mixture of acetonitrile and tert-butanol. We further unveiled the relationship between the microstructure of a dye layer and the multichannel <span class="hlt">charge</span>-transfer dynamics that underlie the photovoltaic performance of dye-sensitized solar cells. © 2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22026433','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22026433"><span>Silica-coated quantum dots for optical evaluation of perfluorocarbon <span class="hlt">droplet</span> interactions with cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gorelikov, Ivan; Martin, Amanda L; Seo, Minseok; Matsuura, Naomi</p> <p>2011-12-20</p> <p>There has been recent interest in developing new, targeted, perfluorocarbon (PFC) <span class="hlt">droplet</span>-based contrast agents for medical imaging (e.g., magnetic resonance imaging, X-ray/computed tomography, and ultrasound imaging). However, due to the large number of potential PFCs and <span class="hlt">droplet</span> stabilization strategies available, it is challenging to determine in advance the PFC <span class="hlt">droplet</span> formulation that will result in the optimal in vivo behavior and imaging performance required for clinical success. We propose that the integration of fluorescent quantum dots (QDs) into new PFC <span class="hlt">droplet</span> agents can help to rapidly screen new PFC-based candidate agents for biological compatibility early in their development. QD labels can allow the interaction of PFC <span class="hlt">droplets</span> with single cells to be assessed at high sensitivity and resolution using optical methods in vitro, complementing the deeper depth penetration but lower resolution provided by PFC <span class="hlt">droplet</span> imaging using in vivo medical imaging systems. In this work, we introduce a simple and robust method to miscibilize silica-coated nanoparticles into hydrophobic and lipophobic PFCs through fluorination of the silica surface via a hydrolysis-condensation reaction with 1H,1H,2H,2H-perfluorodecyltriethoxysilane. Using CdSe/ZnS core/shell QDs, we show that nanoscale, QD-labeled PFC <span class="hlt">droplets</span> can be easily formed, with similar sizes and surface <span class="hlt">charges</span> as unlabeled PFC <span class="hlt">droplets</span>. The QD label can be used to determine the PFC <span class="hlt">droplet</span> uptake into cells in vitro by fluorescence microscopy and flow cytometry, and can be used to validate the fate of PFC <span class="hlt">droplets</span> in vivo in small animals via fluorescence microscopy of histological tissue sections. This is demonstrated in macrophage and cancer cells, and in rabbits, respectively. This work reveals the potential of using QD labels for rapid, preclinical, optical assessment of different PFC <span class="hlt">droplet</span> formulations for their future use in patients. © 2011 American Chemical Society</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26522982','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26522982"><span>Double emulsion <span class="hlt">solvent</span> evaporation techniques used for drug encapsulation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Iqbal, Muhammad; Zafar, Nadiah; Fessi, Hatem; Elaissari, Abdelhamid</p> <p>2015-12-30</p> <p>Double emulsions are complex systems, also called "emulsions of emulsions", in which the <span class="hlt">droplets</span> of the dispersed phase contain one or more types of smaller dispersed <span class="hlt">droplets</span> themselves. Double emulsions have the potential for encapsulation of both hydrophobic as well as hydrophilic drugs, cosmetics, foods and other high value products. Techniques based on double emulsions are commonly used for the encapsulation of hydrophilic molecules, which suffer from low encapsulation efficiency because of rapid drug partitioning into the external aqueous phase when using single emulsions. The main issue when using double emulsions is their production in a well-controlled manner, with homogeneous <span class="hlt">droplet</span> size by optimizing different process variables. In this review special attention has been paid to the application of double emulsion techniques for the encapsulation of various hydrophilic and hydrophobic anticancer drugs, anti-inflammatory drugs, antibiotic drugs, proteins and amino acids and their applications in theranostics. Moreover, the optimized ratio of the different phases and other process parameters of double emulsions are discussed. Finally, the results published regarding various types of <span class="hlt">solvents</span>, stabilizers and polymers used for the encapsulation of several active substances via double emulsion processes are reported. Copyright © 2015 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18680362','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18680362"><span>Water <span class="hlt">droplets</span> as template for next-generation self-assembled poly-(etheretherketone) with cardo membranes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gugliuzza, Annarosa; Aceto, Marianna Carmela; Macedonio, Francesca; Drioli, Enrico</p> <p>2008-08-28</p> <p>Next generation PEEK-WC membranes have been fabricated by using an innovative self-assembly technique. Patterned architectures have been achieved via a <span class="hlt">solvent</span>-reduced and water-assisted process, resulting in honeycomb packed geometry. The membranes exhibit monodisperse pores with size and shape comparable to those left by templating water <span class="hlt">droplets</span>. Influencing factors for the formation of self-assembled poly-(etheretherketone) with Cardo [PEEK-WC] membranes have been evaluated, identifying the critical parameters for nucleation, growth, and propagation of the <span class="hlt">droplet</span>-mobile arrays through the overall films. Structure-transport relationships have been discussed according to the results achieved from the implementation of membrane distillation processes, yielding indication about the suitability of self-assembled PEEK-WC films to work as interfaces in contactor operations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018SMaS...27g4007A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018SMaS...27g4007A"><span>Frequency domain analysis of <span class="hlt">droplet</span>-based electrostatic transducers</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Allegretto, Graham; Dobashi, Yuta; Dixon, Katelyn; Wyss, Justin; Yao, Dickson; Madden, John D. W.</p> <p>2018-07-01</p> <p>Squeezing a water <span class="hlt">droplet</span> between two electrodes can generate a potential difference by converting some of the mechanical energy in vibrations into electrical energy. By utilizing the high capacitance inherent to electric double layers, and the surface <span class="hlt">charging</span> at a polymer/water interface, we demonstrate a sensor that generates up to 892 mV peak-to-peak between 1 and 100 Hz, in response to a 250 μm deformation. This frequency response is described and explained using a linearized model in which the interfacial <span class="hlt">charge</span> acts as the priming voltage, removing the need for external <span class="hlt">charging</span> normally required in capacitive generators. The model suggests how to design the cell for maximum power output and provides an intuitive understanding of the high pass nature of the sensor. It successfully predicts the point of maximum power transfer.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015JASMS..26.1538B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015JASMS..26.1538B"><span>Mass Spectrometric Imaging Using Laser Ablation and <span class="hlt">Solvent</span> Capture by Aspiration (LASCA)</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Brauer, Jonathan I.; Beech, Iwona B.; Sunner, Jan</p> <p>2015-09-01</p> <p>A novel interface for ambient, laser ablation-based mass spectrometric imaging (MSI) referred to as laser ablation and <span class="hlt">solvent</span> capture by aspiration (LASCA) is presented and its performance demonstrated using selected, unaltered biological materials. LASCA employs a pulsed 2.94 μm laser beam for specimen ablation. Ablated materials in the laser plumes are collected on a hanging <span class="hlt">solvent</span> <span class="hlt">droplet</span> with electric field-enhanced trapping, followed by aspiration of <span class="hlt">droplets</span> and remaining plume material in the form of a coarse aerosol into a collection capillary. The gas and liquid phases are subsequently separated in a 10 μL-volume separatory funnel, and the solution is analyzed with electrospray ionization in a high mass resolution Q-ToF mass spectrometer. The LASCA system separates the sampling and ionization steps in MSI and combines high efficiencies of laser plume sampling and of electrospray ionization (ESI) with high mass resolution MS. Up to 2000 different compounds are detected from a single ablation spot (pixel). Using the LASCA platform, rapid (6 s per pixel), high sensitivity, high mass-resolution ambient imaging of "as-received" biological material is achieved routinely and reproducibly.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JMoSt1131..114G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JMoSt1131..114G"><span>UV-Vis spectroscopy and density functional study of <span class="hlt">solvent</span> effect on the <span class="hlt">charge</span> transfer band of the n → σ* complexes of 2-Methylpyridine and 2-Chloropyridine with molecular iodine</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gogoi, Pallavi; Mohan, Uttam; Borpuzari, Manash Protim; Boruah, Abhijit; Baruah, Surjya Kumar</p> <p>2017-03-01</p> <p>UV-Vis spectroscopy has established that Pyridine substitutes form n→σ* <span class="hlt">charge</span> transfer (CT) complexes with molecular Iodine. This study is a combined approach of purely experimental UV-Vis spectroscopy, Multiple linear regression theory and Computational chemistry to analyze the effect of <span class="hlt">solvent</span> upon the <span class="hlt">charge</span> transfer band of 2-Methylpyridine-I2 and 2-Chloropyridine-I2 complexes. Regression analysis verifies the dependence of the CT band upon different <span class="hlt">solvent</span> parameters. Dielectric constant and refractive index are considered among the bulk <span class="hlt">solvent</span> parameters and Hansen, Kamlet and Catalan parameters are taken into consideration at the molecular level. Density Functional Theory results explain well the blue shift of the CT bands in polar medium as an outcome of stronger donor acceptor interaction. A logarithmic relation between the bond length of the bridging atoms of the donor and the acceptor with the dielectric constant of the medium is established. Tauc plot and TDDFT study indicates a non-vertical electronic transition in the complexes. Buckingham and Lippert Mataga equations are applied to check the Polarizability effect on the CT band.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19820015570','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19820015570"><span>A new <span class="hlt">droplet</span> generator</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Slack, W. E.</p> <p>1982-01-01</p> <p>A new <span class="hlt">droplet</span> generator is described. A loud speaker driven extractor needle was immersed in a pendant drop. Pulsing the speaker extracted the needle forming a fluid ligament which will decay into a <span class="hlt">droplet</span>. The <span class="hlt">droplets</span> were sized by stroboscopic photographs. The <span class="hlt">droplet</span>'s size was changed by varying the amplitude of the speaker pulses and the extractor needle diameter. The mechanism of <span class="hlt">droplet</span> formation is discussed and photographs of ligament decay are presented. The <span class="hlt">droplet</span> generator worked well on both oil and water based pesticide formulations. Current applications and results are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5993475','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5993475"><span>Electrostatically driven fog collection using space <span class="hlt">charge</span> injection</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Damak, Maher; Varanasi, Kripa K.</p> <p>2018-01-01</p> <p>Fog collection can be a sustainable solution to water scarcity in many regions around the world. Most proposed collectors are meshes that rely on inertial collision for <span class="hlt">droplet</span> capture and are inherently limited by aerodynamics. We propose a new approach in which we introduce electrical forces that can overcome aerodynamic drag forces. Using an ion emitter, we introduce a space <span class="hlt">charge</span> into the fog to impart a net <span class="hlt">charge</span> to the incoming fog <span class="hlt">droplets</span> and direct them toward a collector using an imposed electric field. We experimentally measure the collection efficiency on single wires, two-wire systems, and meshes and propose a physical model to quantify it. We identify the regimes of optimal collection and provide insights into designing effective fog harvesting systems. PMID:29888324</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1032987','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/1032987"><span>Lossless <span class="hlt">droplet</span> transfer of <span class="hlt">droplet</span>-based microfluidic analysis</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Kelly, Ryan T [West Richland, WA; Tang, Keqi [Richland, WA; Page, Jason S [Kennewick, WA; Smith, Richard D [Richland, WA</p> <p>2011-11-22</p> <p>A transfer structure for <span class="hlt">droplet</span>-based microfluidic analysis is characterized by a first conduit containing a first stream having at least one immiscible <span class="hlt">droplet</span> of aqueous material and a second conduit containing a second stream comprising an aqueous fluid. The interface between the first conduit and the second conduit can define a plurality of apertures, wherein the apertures are sized to prevent exchange of the first and second streams between conduits while allowing lossless transfer of <span class="hlt">droplets</span> from the first conduit to the second conduit through contact between the first and second streams.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19990075873','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19990075873"><span>Electrostatic Hazard Considerations for ODC <span class="hlt">Solvent</span> Replacement Selection Testing</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Fairbourn, Brad</p> <p>1999-01-01</p> <p>ODC <span class="hlt">solvents</span> are used to clean many critical substrates during solid rocket motor production operations. Electrostatic <span class="hlt">charge</span> generation incidental to these cleaning operations can pose a major safety issue. Therefore, while determining the acceptability of various ODC replacement cleaners, one aspect of the selection criteria included determining the extent of electric <span class="hlt">charge</span> generation during a typical <span class="hlt">solvent</span> cleaning operation. A total of six candidate replacement cleaners, sixteen critical substrates, and two types of cleaning swatch materials were studied in simulated cleaning operations. <span class="hlt">Charge</span> generation and accumulation effects were investigated by measuring the peak voltage and brush discharging effects associated with each cleaning process combination. In some cases, <span class="hlt">charge</span> generation was found to be very severe. Using the conductivity information for each cleaner, the peak voltage data could in some cases, be qualitatively predicted. Test results indicated that severe <span class="hlt">charging</span> effects could result in brush discharges that could potentially result in flash fire hazards when occurring in close proximity to flammable vapor/air mixtures. Process controls to effectively mitigate these hazards are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017JASMS..28..332S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017JASMS..28..332S"><span><span class="hlt">Charging</span> of Proteins in Native Mass Spectrometry</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Susa, Anna C.; Xia, Zijie; Tang, Henry Y. H.; Tainer, John A.; Williams, Evan R.</p> <p>2017-02-01</p> <p>Factors that influence the <span class="hlt">charging</span> of protein ions formed by electrospray ionization from aqueous solutions in which proteins have native structures and function were investigated. Protein ions ranging in molecular weight from 12.3 to 79.7 kDa and pI values from 5.4 to 9.6 were formed from different solutions and reacted with volatile bases of gas-phase basicities higher than that of ammonia in the cell of a Fourier-transform ion cyclotron resonance mass spectrometer. The <span class="hlt">charge</span>-state distribution of cytochrome c ions formed from aqueous ammonium or potassium acetate is the same. Moreover, ions formed from these two solutions do not undergo proton transfer to 2-fluoropyridine, which is 8 kcal/mol more basic than ammonia. These results provide compelling evidence that proton transfer between ammonia and protein ions does not limit protein ion <span class="hlt">charge</span> in native electrospray ionization. Both circular dichroism and ion mobility measurements indicate that there are differences in conformations of proteins in pure water and aqueous ammonium acetate, and these differences can account for the difference in the extent of <span class="hlt">charging</span> and proton-transfer reactivities of protein ions formed from these solutions. The extent of proton transfer of the protein ions with higher gas-phase basicity bases trends with how closely the protein ions are <span class="hlt">charged</span> to the value predicted by the Rayleigh limit for spherical water <span class="hlt">droplets</span> approximately the same size as the proteins. These results indicate that <span class="hlt">droplet</span> <span class="hlt">charge</span> limits protein ion <span class="hlt">charge</span> in native mass spectrometry and are consistent with these ions being formed by the <span class="hlt">charged</span> residue mechanism.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28692870','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28692870"><span>Air-assisted liquid-liquid microextraction using floating organic <span class="hlt">droplet</span> solidification for simultaneous extraction and spectrophotometric determination of some drugs in biological samples through chemometrics methods.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Farahmand, Farnaz; Ghasemzadeh, Bahar; Naseri, Abdolhossein</p> <p>2018-01-05</p> <p>An air assisted liquid-liquid microextraction by applying the solidification of a floating organic <span class="hlt">droplet</span> method (AALLME-SFOD) coupled with a multivariate calibration method, namely partial least squares (PLS), was introduced for the fast and easy determination of Atenolol (ATE), Propanolol (PRO) and Carvedilol (CAR) in biological samples via a spectrophotometric approach. The analytes would be extracted from neutral aqueous solution into 1-dodecanol as an organic <span class="hlt">solvent</span>, using AALLME. In this approach a low-density <span class="hlt">solvent</span> with a melting point close to room temperature was applied as the extraction <span class="hlt">solvent</span>. The emulsion was immediately formed by repeatedly pulling in and pushing out the aqueous sample solution and extraction <span class="hlt">solvent</span> mixture via a 10-mL glass syringe for ten times. After centrifugation, the extractant <span class="hlt">droplet</span> could be simply collected from the aqueous samples by solidifying the emulsion at a lower than the melting point temperature. In the next step, analytes were back extracted simultaneously into the acidic aqueous solution. Derringer and Suich multi-response optimization were utilized for simultaneous optimizing the parameters of three analytes. This method incorporates the benefits of AALLME and dispersive liquid-liquid microextraction considering the solidification of floating organic <span class="hlt">droplets</span> (DLLME-SFOD). Calibration graphs under optimized conditions were linear in the range of 0.30-6.00, 0.32-2.00 and 0.30-1.40μg mL -1 for ATE, CAR and PRO, respectively. Other analytical parameters were obtained as follows: enrichment factors (EFs) were found to be 11.24, 16.55 and 14.90, and limits of detection (LODs) were determined to be 0.09, 0.10 and 0.08μg mL -1 for ATE, CAR and PRO, respectively. The proposed method will require neither a highly toxic chlorinated <span class="hlt">solvent</span> for extraction nor an organic dispersive <span class="hlt">solvent</span> in the application process; hence, it is more environmentally friendly. Copyright © 2017 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..MARL12011W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..MARL12011W"><span><span class="hlt">Droplets</span> on bent fibers</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Weyer, Floriane; Pan, Zhao; Pitt, William; Truscott, Tadd; Vandewalle, Nicolas</p> <p></p> <p><span class="hlt">Droplets</span> on fibers are part of our everyday lives. Many phenomena involve drops and fibers such as the formation of dew <span class="hlt">droplets</span> on a spiderweb, the trapping of water <span class="hlt">droplets</span> on cactus spines or the motion of <span class="hlt">droplets</span> on wetted moss hairs. These topics have been widely studied. In particular, Lorenceau et al. determined the critical volume of a water <span class="hlt">droplet</span> hanging on a horizontal fiber. Here, we address a similar question : we try to find out the maximum <span class="hlt">droplet</span> size on bent fibers, which are able to hold significantly more water than horizontal fibers. Indeed, we noticed that, in nature, some specific plants can hold large rain <span class="hlt">droplets</span> thanks to their Y-shaped leaves. We try to mimic these structures with nylon fibers, of different diameters, bent with various angles. For each set-up, the critical water volume is determined. Finally, we propose models of the physics involved in determining <span class="hlt">droplet</span> size that could be implemented in future fiber-based microfluidic devices.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28760972','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28760972"><span>Printed <span class="hlt">droplet</span> microfluidics for on demand dispensing of picoliter <span class="hlt">droplets</span> and cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cole, Russell H; Tang, Shi-Yang; Siltanen, Christian A; Shahi, Payam; Zhang, Jesse Q; Poust, Sean; Gartner, Zev J; Abate, Adam R</p> <p>2017-08-15</p> <p>Although the elementary unit of biology is the cell, high-throughput methods for the microscale manipulation of cells and reagents are limited. The existing options either are slow, lack single-cell specificity, or use fluid volumes out of scale with those of cells. Here we present printed <span class="hlt">droplet</span> microfluidics, a technology to dispense picoliter <span class="hlt">droplets</span> and cells with deterministic control. The core technology is a fluorescence-activated <span class="hlt">droplet</span> sorter coupled to a specialized substrate that together act as a picoliter <span class="hlt">droplet</span> and single-cell printer, enabling high-throughput generation of intricate arrays of <span class="hlt">droplets</span>, cells, and microparticles. Printed <span class="hlt">droplet</span> microfluidics provides a programmable and robust technology to construct arrays of defined cell and reagent combinations and to integrate multiple measurement modalities together in a single assay.</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/2017PNAS..114.8728C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PNAS..114.8728C"><span>Printed <span class="hlt">droplet</span> microfluidics for on demand dispensing of picoliter <span class="hlt">droplets</span> and cells</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cole, Russell H.; Tang, Shi-Yang; Siltanen, Christian A.; Shahi, Payam; Zhang, Jesse Q.; Poust, Sean; Gartner, Zev J.; Abate, Adam R.</p> <p>2017-08-01</p> <p>Although the elementary unit of biology is the cell, high-throughput methods for the microscale manipulation of cells and reagents are limited. The existing options either are slow, lack single-cell specificity, or use fluid volumes out of scale with those of cells. Here we present printed <span class="hlt">droplet</span> microfluidics, a technology to dispense picoliter <span class="hlt">droplets</span> and cells with deterministic control. The core technology is a fluorescence-activated <span class="hlt">droplet</span> sorter coupled to a specialized substrate that together act as a picoliter <span class="hlt">droplet</span> and single-cell printer, enabling high-throughput generation of intricate arrays of <span class="hlt">droplets</span>, cells, and microparticles. Printed <span class="hlt">droplet</span> microfluidics provides a programmable and robust technology to construct arrays of defined cell and reagent combinations and to integrate multiple measurement modalities together in a single assay.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009CP....358..111Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009CP....358..111Z"><span>Photo-dynamics of roseoflavin and riboflavin in aqueous and organic <span class="hlt">solvents</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zirak, P.; Penzkofer, A.; Mathes, T.; Hegemann, P.</p> <p>2009-03-01</p> <p>Roseoflavin (8-dimethylamino-8-demethyl- D-riboflavin) and riboflavin in aqueous and organic <span class="hlt">solvents</span> are studied by optical absorption spectroscopy, fluorescence spectroscopy, and fluorescence decay kinetics. <span class="hlt">Solvent</span> polarity dependent absorption shifts are observed. The fluorescence quantum yields are <span class="hlt">solvent</span> dependent. For roseoflavin the fluorescence decay shows a bi-exponential dependence (ps to sub-ps time constant, and 100 ps to a few ns time constant). The roseoflavin photo-dynamics is explained in terms of fast intra-molecular <span class="hlt">charge</span> transfer (diabatic electron transfer) from the dimethylamino electron donor group to the pteridin carbonyl electron acceptor followed by intra-molecular <span class="hlt">charge</span> recombination. The fast fluorescence component is due to direct locally-excited-state emission, and the slow fluorescence component is due to delayed locally-excited-state emission and <span class="hlt">charge</span> transfer state emission. The fluorescence decay of riboflavin is mono-exponential. The S 1-state potential energy surface is determined by vibronic relaxation and solvation dynamics due to excited-state dipole moment changes (adiabatic optical electron transfer).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20030005566&hterms=splash&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dsplash','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20030005566&hterms=splash&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dsplash"><span>Splashing <span class="hlt">Droplets</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>VanderWal, Randall L.; Kizito, John Patrick; Berger, Gordon M.; Iwan, J.; Alexander, D.; Tryggvason, Gretar</p> <p>2002-01-01</p> <p>Current data on <span class="hlt">droplet</span> breakup is scarce for the sizes and velocities typical of practical applications such as in spray combustion processes and coating processes. While much more representative of practical applications, the small spatial scales and rapid time-scales prevent detailed measurement of the internal fluid dynamics and liquid property gradients produced by impinging upon surfaces. Realized through the extended spatial and temporal scales afforded by a microgravity environment, an improved understanding of drop breakup dynamics is sought to understand and ultimately control the impingement dynamics of <span class="hlt">droplets</span> upon surfaces in practical situations. The primary objective of this research will be to mark the onset of different 'splashing modes' and to determine their temperature, pressure and angle dependence for impinging <span class="hlt">droplets</span> representative of practical fluids. In addition, we are modeling the evolution of <span class="hlt">droplets</span> that do not initially splash but rather undergo a 'fingering' evolution observed on the spreading fluid front and the transformation of these fingers into splashed products. An example of our experimental data is presented below. These images are of Isopar V impacting a mirror-polished surface. They were acquired using a high-speed camera at 1000 frames per second. They show the spreading of a single <span class="hlt">droplet</span> after impact and ensuing finger instabilities. Normal gravity experimental data such as this will guide low gravity measurements in the 2.2 second drop tower and KC-135 aircraft as available. Presently we are in the process of comparing the experimental data of <span class="hlt">droplet</span> shape evolution to numerical models, which can also capture the internal fluid dynamics and liquid property gradients such as produced by impingement upon a heated surface. To-date isothermal numerical data has been modeled using direct numerical simulations of representative splashing <span class="hlt">droplets</span>. The data obtained so far indicates that the present model describes well</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4325842','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4325842"><span>Seipin performs dissectible functions in promoting lipid <span class="hlt">droplet</span> biogenesis and regulating <span class="hlt">droplet</span> morphology</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Cartwright, Bethany R.; Binns, Derk D.; Hilton, Christopher L.; Han, Sungwon; Gao, Qiang; Goodman, Joel M.</p> <p>2015-01-01</p> <p>Seipin is necessary for both adipogenesis and lipid <span class="hlt">droplet</span> (LD) organization in nonadipose tissues; however, its molecular function is incompletely understood. Phenotypes in the seipin-null mutant of Saccharomyces cerevisiae include aberrant <span class="hlt">droplet</span> morphology (endoplasmic reticulum–<span class="hlt">droplet</span> clusters and size heterogeneity) and sensitivity of <span class="hlt">droplet</span> size to changes in phospholipid synthesis. It has not been clear, however, whether seipin acts in initiation of <span class="hlt">droplet</span> synthesis or at a later step. Here we utilize a system of de novo <span class="hlt">droplet</span> formation to show that the absence of seipin results in a delay in <span class="hlt">droplet</span> appearance with concomitant accumulation of neutral lipid in membranes. We also demonstrate that seipin is required for vectorial budding of <span class="hlt">droplets</span> toward the cytoplasm. Furthermore, we find that the normal rate of <span class="hlt">droplet</span> initiation depends on 14 amino acids at the amino terminus of seipin, deletion of which results in fewer, larger <span class="hlt">droplets</span> that are consistent with a delay in initiation but are otherwise normal in morphology. Importantly, other functions of seipin, namely vectorial budding and resistance to inositol, are retained in this mutant. We conclude that seipin has dissectible roles in both promoting early LD initiation and in regulating LD morphology, supporting its importance in LD biogenesis. PMID:25540432</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25138541','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25138541"><span><span class="hlt">Solvent</span> effects on polymer sorting of carbon nanotubes with applications in printed electronics.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Huiliang; Hsieh, Bing; Jiménez-Osés, Gonzalo; Liu, Peng; Tassone, Christopher J; Diao, Ying; Lei, Ting; Houk, Kendall N; Bao, Zhenan</p> <p>2015-01-07</p> <p>Regioregular poly(3-alkylthiophene) (P3AT) polymers have been previously reported for the selective, high-yield dispersion of semiconducting single-walled carbon nanotubes (SWCNTs) in toluene. Here, five alternative <span class="hlt">solvents</span> are investigated, namely, tetrahydrofuran, decalin, tetralin, m-xylene, and o-xylene, for the dispersion of SWCNTs by poly(3-dodecylthiophene) P3DDT. The dispersion yield could be increased to over 40% using decalin or o-xylene as the <span class="hlt">solvents</span> while maintaining high selectivity towards semiconducting SWCNTs. Molecular dynamics (MD) simulations in explicit <span class="hlt">solvents</span> are used to explain the improved sorting yield. In addition, a general mechanism is proposed to explain the selective dispersion of semiconducting SWCNTs by conjugated polymers. The possibility to perform selective sorting of semiconducting SWCNTs using various <span class="hlt">solvents</span> provides a greater diversity of semiconducting SWCNT ink properties, such as boiling point, viscosity, and surface tension as well as toxicity. The efficacy of these new semiconducting SWCNT inks is demonstrated by using the high boiling point and high viscosity <span class="hlt">solvent</span> tetralin for inkjet-printed transistors, where <span class="hlt">solvent</span> properties are more compatible with the inkjet printing head and improved <span class="hlt">droplet</span> formation. © 2014 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27288575','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27288575"><span>Efficient demulsification of oil-in-water emulsions using a zeolitic imidazolate framework: Adsorptive removal of oil <span class="hlt">droplets</span> from water.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lin, Kun-Yi Andrew; Chen, Yu-Chien; Phattarapattamawong, Songkeart</p> <p>2016-09-15</p> <p>To demulsify oil-in-water (O/W) emulsions, a zinc-based zeolitic imidazolate framework (ZIF-8) was employed for the first time to remove oil <span class="hlt">droplets</span> from water. ZIF-8 exhibits a high surface area and positive surface <span class="hlt">charges</span>, making it a suitable adsorbent to adsorb negatively-<span class="hlt">charged</span> oil <span class="hlt">droplets</span>. Adsorption behaviors of oil <span class="hlt">droplets</span> to ZIF-8 were studied by analyzing the adsorption kinetics and isotherm with theoretical models. The activation energy of adsorption of oil <span class="hlt">droplets</span> to ZIF-8 was determined as 24.1kJmol(-1). The Langmuir-Freundlich (L-F) model was found to be most applicable to interpret the isotherm data and the predicated maximum adsorption capacity of ZIF-8 can reach 6633mgg(-1), revealing a promising capability of ZIF-8 for demulsification. Factors influencing the adsorption of oil <span class="hlt">droplets</span> to ZIF-8 were investigated including temperature, pH, salt and surfactants. The adsorption capacity of ZIF-8 for oil was improved at elevated temperatures, whereas alkaline condition was unfavorable for the adsorption of oil <span class="hlt">droplets</span> due to the electrostatic repulsion at high pH. The adsorption capacity of ZIF-8 remained similar in the presence of NaCl but it was reduced in the presence of surfactants. ZIF-8 was regenerated by a simple ethanol-washing method; the regenerated ZIF-8 exhibited more than 85% of regeneration efficiency over six cycles. Its crystalline structure also remained intact after the regeneration. These characteristics indicate that ZIF-8 can be a promising and effective adsorbent to remove oil <span class="hlt">droplets</span> for demulsification of O/W emulsions. Copyright © 2016 Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5565430','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5565430"><span>Printed <span class="hlt">droplet</span> microfluidics for on demand dispensing of picoliter <span class="hlt">droplets</span> and cells</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Cole, Russell H.; Tang, Shi-Yang; Siltanen, Christian A.; Shahi, Payam; Zhang, Jesse Q.; Poust, Sean; Gartner, Zev J.; Abate, Adam R.</p> <p>2017-01-01</p> <p>Although the elementary unit of biology is the cell, high-throughput methods for the microscale manipulation of cells and reagents are limited. The existing options either are slow, lack single-cell specificity, or use fluid volumes out of scale with those of cells. Here we present printed <span class="hlt">droplet</span> microfluidics, a technology to dispense picoliter <span class="hlt">droplets</span> and cells with deterministic control. The core technology is a fluorescence-activated <span class="hlt">droplet</span> sorter coupled to a specialized substrate that together act as a picoliter <span class="hlt">droplet</span> and single-cell printer, enabling high-throughput generation of intricate arrays of <span class="hlt">droplets</span>, cells, and microparticles. Printed <span class="hlt">droplet</span> microfluidics provides a programmable and robust technology to construct arrays of defined cell and reagent combinations and to integrate multiple measurement modalities together in a single assay. PMID:28760972</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22790308','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22790308"><span><span class="hlt">Droplet</span> based microfluidics.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Seemann, Ralf; Brinkmann, Martin; Pfohl, Thomas; Herminghaus, Stephan</p> <p>2012-01-01</p> <p><span class="hlt">Droplet</span> based microfluidics is a rapidly growing interdisciplinary field of research combining soft matter physics, biochemistry and microsystems engineering. Its applications range from fast analytical systems or the synthesis of advanced materials to protein crystallization and biological assays for living cells. Precise control of <span class="hlt">droplet</span> volumes and reliable manipulation of individual <span class="hlt">droplets</span> such as coalescence, mixing of their contents, and sorting in combination with fast analysis tools allow us to perform chemical reactions inside the <span class="hlt">droplets</span> under defined conditions. In this paper, we will review available drop generation and manipulation techniques. The main focus of this review is not to be comprehensive and explain all techniques in great detail but to identify and shed light on similarities and underlying physical principles. Since geometry and wetting properties of the microfluidic channels are crucial factors for <span class="hlt">droplet</span> generation, we also briefly describe typical device fabrication methods in <span class="hlt">droplet</span> based microfluidics. Examples of applications and reaction schemes which rely on the discussed manipulation techniques are also presented, such as the fabrication of special materials and biophysical experiments.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20138576','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20138576"><span>Spectrophotometric and spectroscopic studies of <span class="hlt">charge</span> transfer complexes of p-toluidine as an electron donor with picric acid as an electron acceptor in different <span class="hlt">solvents</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Singh, Neeti; Khan, Ishaat M; Ahmad, Afaq</p> <p>2010-04-01</p> <p>The <span class="hlt">charge</span> transfer complexes of the donor p-toluidine with pi-acceptor picric acid have been studied spectrophotometrically in various <span class="hlt">solvents</span> such as carbon tetrachloride, chloroform, dichloromethane acetone, ethanol, and methanol at room temperature using absorption spectrophotometer. The results indicate that formation of CTC in non-polar <span class="hlt">solvent</span> is high. The stoichiometry of the complex was found to be 1:1 ratio by straight-line method between donor and acceptor with maximum absorption bands. The data are discussed in terms of formation constant (K(CT)), molar extinction coefficient (epsilon(CT)), standard free energy (DeltaG(o)), oscillator strength (f), transition dipole moment (mu(EN)), resonance energy (R(N)) and ionization potential (I(D)). The results indicate that the formation constant (K(CT)) for the complex was shown to be dependent upon the nature of electron acceptor, donor and polarity of <span class="hlt">solvents</span> that were used. Copyright 2010 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20100015714','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20100015714"><span><span class="hlt">Droplet</span> transport system and methods</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Neitzel, G. Paul (Inventor)</p> <p>2010-01-01</p> <p>Embodiments of <span class="hlt">droplet</span> transport systems and methods are disclosed for levitating and transporting single or encapsulated <span class="hlt">droplets</span> using thermocapillary convection. One method embodiment, among others comprises providing a <span class="hlt">droplet</span> of a first liquid; and applying thermocapillary convection to the <span class="hlt">droplet</span> to levitate and move the <span class="hlt">droplet</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26110977','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26110977"><span>A Comprehensive Model of Electric-Field-Enhanced Jumping-<span class="hlt">Droplet</span> Condensation on Superhydrophobic Surfaces.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Birbarah, Patrick; Li, Zhaoer; Pauls, Alexander; Miljkovic, Nenad</p> <p>2015-07-21</p> <p>Superhydrophobic micro/nanostructured surfaces for dropwise condensation have recently received significant attention due to their potential to enhance heat transfer performance by shedding positively <span class="hlt">charged</span> water <span class="hlt">droplets</span> via coalescence-induced <span class="hlt">droplet</span> jumping at length scales below the capillary length and allowing the use of external electric fields to enhance <span class="hlt">droplet</span> removal and heat transfer, in what has been termed electric-field-enhanced (EFE) jumping-<span class="hlt">droplet</span> condensation. However, achieving optimal EFE conditions for enhanced heat transfer requires capturing the details of transport processes that is currently lacking. While a comprehensive model has been developed for condensation on micro/nanostructured surfaces, it cannot be applied for EFE condensation due to the dynamic <span class="hlt">droplet</span>-vapor-electric field interactions. In this work, we developed a comprehensive physical model for EFE condensation on superhydrophobic surfaces by incorporating individual <span class="hlt">droplet</span> motion, electrode geometry, jumping frequency, field strength, and condensate vapor-flow dynamics. As a first step toward our model, we simulated jumping <span class="hlt">droplet</span> motion with no external electric field and validated our theoretical <span class="hlt">droplet</span> trajectories to experimentally obtained trajectories, showing excellent temporal and spatial agreement. We then incorporated the external electric field into our model and considered the effects of jumping <span class="hlt">droplet</span> size, electrode size and geometry, condensation heat flux, and <span class="hlt">droplet</span> jumping direction. Our model suggests that smaller jumping <span class="hlt">droplet</span> sizes and condensation heat fluxes require less work input to be removed by the external fields. Furthermore, the results suggest that EFE electrodes can be optimized such that the work input is minimized depending on the condensation heat flux. To analyze overall efficiency, we defined an incremental coefficient of performance and showed that it is very high (∼10(6)) for EFE condensation. We finally proposed mechanisms</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010JChPh.132r5101V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010JChPh.132r5101V"><span>The electrostatics of <span class="hlt">solvent</span> and membrane interfaces and the role of electronic polarizability</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Vorobyov, Igor; Allen, Toby W.</p> <p>2010-05-01</p> <p>The electrostatics of <span class="hlt">solvent</span> and lipid bilayer interfaces are investigated with the aim of understanding the interaction of ions and <span class="hlt">charged</span> peptides with biological membranes. We overcome the lacking dielectric response of hydrocarbon by carrying out atomistic molecular dynamics simulations using a polarizable model. For air-<span class="hlt">solvent</span> or <span class="hlt">solvent-solvent</span> interfaces, the effect of polarizability itself is small, yet changes in the fixed atomic <span class="hlt">charge</span> distribution are responsible for substantial changes in the potential. However, when electrostatics is probed by finite solutes, a cancellation of dominant quadrupolar terms from the macroscopic and microscopic (solute-<span class="hlt">solvent</span>) interfaces eliminates this dependence and leads to small net contributions to partitioning thermodynamics. In contrast, the membrane dipole potential exhibits considerable dependence on lipid electronic polarizability, due to its dominant dipolar contribution. We report the dipole potential for a polarizable lipid hydrocarbon membrane model of 480-610 mV, in better accord with experimental measurements.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011SPIE.7899E..3HS','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011SPIE.7899E..3HS"><span>Optical <span class="hlt">droplet</span> vaporization of micron-sized perfluorocarbon <span class="hlt">droplets</span> and their photoacoustic detection</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Strohm, Eric; Rui, Min; Gorelikov, Ivan; Matsuura, Naomi; Kolios, Michael</p> <p>2011-03-01</p> <p>An acoustic and photoacoustic characterization of micron-sized perfluorocarbon (PFC) <span class="hlt">droplets</span> is presented. PFC <span class="hlt">droplets</span> are currently being investigated as acoustic and photoacoustic contrast agents and as cancer therapy agents. Pulse echo measurements at 375 MHz were used to determine the diameter, ranging from 3.2 to 6.5 μm, and the sound velocity, ranging from 311 to 406 m/s of nine <span class="hlt">droplets</span>. An average sound velocity of 379 +/- 18 m/s was calculated for <span class="hlt">droplets</span> larger than the ultrasound beam width of 4.0 μm. Optical <span class="hlt">droplet</span> vaporization, where vaporization of a single <span class="hlt">droplet</span> occurred upon laser irradiation of sufficient intensity, was verified using pulse echo acoustic methods. The ultrasonic backscatter amplitude, acoustic impedance and attenuation increased after vaporization, consistent with a phase change from a liquid to gas core. Photoacoustic measurements were used to compare the spectra of three <span class="hlt">droplets</span> ranging in diameter from 3.0 to 6.2 μm to a theoretical model. Good agreement in the spectral features was observed over the bandwidth of the 375 MHz transducer.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014EGUGA..16..483S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014EGUGA..16..483S"><span>Homogeneous Freezing of Water <span class="hlt">Droplets</span> and its Dependence on <span class="hlt">Droplet</span> Size</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schmitt, Thea; Möhler, Ottmar; Höhler, Kristina; Leisner, Thomas</p> <p>2014-05-01</p> <p>The formulation and parameterisation of microphysical processes in tropospheric clouds, such as phase transitions, is still a challenge for weather and climate models. This includes the homogeneous freezing of supercooled water <span class="hlt">droplets</span>, since this is an important process in deep convective systems, where almost pure water <span class="hlt">droplets</span> may stay liquid until homogeneous freezing occurs at temperatures around 238 K. Though the homogeneous ice nucleation in supercooled water is considered to be well understood, recent laboratory experiments with typical cloud <span class="hlt">droplet</span> sizes showed one to two orders of magnitude smaller nucleation rate coefficients than previous literature results, including earlier results from experiments with single levitated water <span class="hlt">droplets</span> and from cloud simulation experiments at the AIDA (Aerosol Interaction and Dynamics in the Atmosphere) facility. This motivated us to re-analyse homogeneous <span class="hlt">droplet</span> freezing experiments conducted during the previous years at the AIDA cloud chamber. This cloud chamber has a volume of 84m3 and operates under atmospherically relevant conditions within wide ranges of temperature, pressure and humidity, whereby investigations of both tropospheric mixed-phase clouds and cirrus clouds can be realised. By controlled adiabatic expansions, the ascent of an air parcel in the troposphere can be simulated. According to our new results and their comparison to the results from single levitated <span class="hlt">droplet</span> experiments, the homogeneous freezing of water <span class="hlt">droplets</span> seems to be a volume-dependent process, at least for <span class="hlt">droplets</span> as small as a few micrometers in diameter. A contribution of surface induced freezing can be ruled out, in agreement to previous conclusions from the single <span class="hlt">droplet</span> experiments. The obtained volume nucleation rate coefficients are in good agreement, within error bars, with some previous literature data, including our own results from earlier AIDA experiments, but they do not agree with recently published lower volume</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2008AGUFM.H51I0983H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2008AGUFM.H51I0983H"><span>Infiltration of Liquid <span class="hlt">Droplets</span> Into Porous Media: Effects of Dynamic Contact Angle and Contact Angle Hysteresis</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hilpert, M.</p> <p>2008-12-01</p> <p>Infiltration of liquid <span class="hlt">droplets</span> into dry porous media occurs when rain drops fall onto soil, when accidentally spilling organic liquid (e.g., gasoline and chlorinated <span class="hlt">solvents</span>) onto ground, or when aerosol pesticides are not intercepted by the vegetation and then released to soils. If harmful chemicals are released from the <span class="hlt">droplet</span> into the atmosphere through evaporation, it is important to know the time of infiltration. We developed a theory for infiltration, which accounts for a general model for the dynamic contact angle between the <span class="hlt">droplet</span> and the porous medium as well as contact angle hysteresis. Our theory assumes the <span class="hlt">droplet</span> to have the shape of a spherical cap and the pressure within the <span class="hlt">droplet</span> to be uniform. The theory shows that <span class="hlt">droplet</span> infiltration involves three phases due to contact angle hysteresis: (1) an increasing drawing area (IDA) phase during which the interface between the <span class="hlt">droplet</span> and the porous medium increases, (2) a constant drawing area (CDA) phase during which the contact line of the <span class="hlt">droplet</span> remains pinned, and (3) a decreasing drawing area (DDA) phase. We find that infiltration always consists of a cascade process formed by the IDA, CDA, and DDA phases, where the entire process may begin or end in any of the three phases. The entire process is formulated with four nondimensional parameters: three contact angles (initial, advancing, and receding) and a porous permeability parameter that depends on porous medium geometry. The total time of infiltration and the time dependence of drawing area are critically affected by the occurrence of the IDA, CDA, and DDA phases as well as by the permeability. In general, the IDA and DDA phases are described by integro-differential equations. With ordinary differential equations (ODEs), we are able to approximate the IDA phase and to describe exactly infiltration processes that starts out with the CDA or DDA phase.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JAP...123p3102W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JAP...123p3102W"><span>Electrohydrodynamic assisted <span class="hlt">droplet</span> alignment for lens fabrication by <span class="hlt">droplet</span> evaporation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, Guangxu; Deng, Jia; Guo, Xing</p> <p>2018-04-01</p> <p>Lens fabrication by <span class="hlt">droplet</span> evaporation has attracted a lot of attention since the fabrication approach is simple and moldless. <span class="hlt">Droplet</span> position accuracy is a critical parameter in this approach, and thus it is of great importance to use accurate methods to realize the <span class="hlt">droplet</span> position alignment. In this paper, we propose an electrohydrodynamic (EHD) assisted <span class="hlt">droplet</span> alignment method. An electrostatic force was induced at the interface between materials to overcome the surface tension and gravity. The deviation of <span class="hlt">droplet</span> position from the center region was eliminated and alignment was successfully realized. We demonstrated the capability of the proposed method theoretically and experimentally. First, we built a simulation model coupled with the three-phase flow formulations and the EHD equations to study the three-phase flowing process in an electric field. Results show that it is the uneven electric field distribution that leads to the relative movement of the <span class="hlt">droplet</span>. Then, we conducted experiments to verify the method. Experimental results are consistent with the numerical simulation results. Moreover, we successfully fabricated a crater lens after applying the proposed method. A light emitting diode module packaging with the fabricated crater lens shows a significant light intensity distribution adjustment compared with a spherical cap lens.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25923721','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25923721"><span>Surfactant-adsorption-induced initial depinning behavior in evaporating water and nanofluid sessile <span class="hlt">droplets</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhong, Xin; Duan, Fei</p> <p>2015-05-19</p> <p>A surfactant-induced autophobic effect has been observed to initiate an intense depinning behavior at the initial stage of evaporation in both pure water and nanofluid sessile <span class="hlt">droplets</span>. The cationic surfactant adsorbing to the negatively <span class="hlt">charged</span> silicon wafer makes the solid surface more hydrophobic. The autophobing-induced depinning behavior, leading to an enlarged contact angle and a shortened base diameter, takes place only when the surfactant concentration is below its critical micelle concentration (cmc). The initial spreading degree right before the <span class="hlt">droplet</span> retraction, the retracting velocity of the contact line, and the duration of the initial <span class="hlt">droplet</span> retraction are shown to depend negatively on the surfactant concentration below the cmc. An unexpected enhancement in the initial depinning has been found in the nanofluid <span class="hlt">droplets</span>, possibly resulting from the hydrophilic interplay between the graphite nanoparticle deposition and the surfactant molecules. Such promotion of the initial depinning due to the nanoparticle deposition makes the <span class="hlt">droplet</span> retract even at a surfactant concentration higher than the cmc (1.5 cmc). The resulting deposition formed in the presence of the depinning behavior has great enhancement for coffee-ring formation as compared to the one free of surfactant, implying that the formation of a coffee ring does not require the pinning of the contact line during the entire drying process.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23406138','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23406138"><span>Thermodynamic and kinetic theory of nucleation, deliquescence and efflorescence transitions in the ensemble of <span class="hlt">droplets</span> on soluble particles.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Shchekin, Alexander K; Shabaev, Ilya V; Hellmuth, Olaf</p> <p>2013-02-07</p> <p>Thermodynamic and kinetic peculiarities of nucleation, deliquescence and efflorescence transitions in the ensemble of <span class="hlt">droplets</span> formed on soluble condensation nuclei from a <span class="hlt">solvent</span> vapor have been considered. The interplay of the effects of solubility and the size of condensation nuclei has been analyzed. Activation barriers for the deliquescence and phase transitions and for the reverse efflorescence transition have been determined as functions of the relative humidity of the vapor-gas atmosphere, initial size, and solubility of condensation nuclei. It has been demonstrated that, upon variations in the relative humidity of the atmosphere, the crossover in thermodynamically stable and unstable variables of the <span class="hlt">droplet</span> state takes place. The physical meaning of stable and unstable variables has been clarified. The kinetic equations for establishing equilibrium and steady distributions of binary <span class="hlt">droplets</span> have been solved. The specific times for relaxation, deliquescence and efflorescence transitions have been calculated.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1421799-charging-proteins-native-mass-spectrometry','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1421799-charging-proteins-native-mass-spectrometry"><span><span class="hlt">Charging</span> of Proteins in Native Mass Spectrometry</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Susa, Anna C.; Xia, Zijie; Tang, Henry Y. H.; ...</p> <p>2016-10-12</p> <p>Factors that influence the <span class="hlt">charging</span> of protein ions formed by electrospray ionization from aqueous solutions in which proteins have native structures and function were investigated. Protein ions ranging in molecular weight from 12.3 to 79.7 kDa and pI values from 5.4 to 9.6 were formed from different solutions and reacted with volatile bases of gas-phase basicities higher than that of ammonia in the cell of a Fourier-transform ion cyclotron resonance mass spectrometer. The <span class="hlt">charge</span>-state distribution of cytochrome c ions formed from aqueous ammonium or potassium acetate is the same. Moreover, ions formed from these two solutions do not undergo protonmore » transfer to 2-fluoropyridine, which is 8 kcal/mol more basic than ammonia. These results provide compelling evidence that proton transfer between ammonia and protein ions does not limit protein ion <span class="hlt">charge</span> in native electrospray ionization. Both circular dichroism and ion mobility measurements indicate that there are differences in conformations of proteins in pure water and aqueous ammonium acetate, and these differences can account for the difference in the extent of <span class="hlt">charging</span> and proton-transfer reactivities of protein ions formed from these solutions. The extent of proton transfer of the protein ions with higher gas-phase basicity bases trends with how closely the protein ions are <span class="hlt">charged</span> to the value predicted by the Rayleigh limit for spherical water <span class="hlt">droplets</span> approximately the same size as the proteins. These results indicate that <span class="hlt">droplet</span> <span class="hlt">charge</span> limits protein ion <span class="hlt">charge</span> in native mass spectrometry and are consistent with these ions being formed by the <span class="hlt">charged</span> residue mechanism.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApPhL.112s4102S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApPhL.112s4102S"><span>Fast electric control of the <span class="hlt">droplet</span> size in a microfluidic T-junction <span class="hlt">droplet</span> generator</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shojaeian, Mostafa; Hardt, Steffen</p> <p>2018-05-01</p> <p>The effect of DC electric fields on the generation of <span class="hlt">droplets</span> of water and xanthan gum solutions in sunflower oil at a microfluidic T-junction is experimentally studied. The electric field leads to a significant reduction of the <span class="hlt">droplet</span> diameter, by about a factor of 2 in the case of water <span class="hlt">droplets</span>. The <span class="hlt">droplet</span> size can be tuned by varying the electric field strength, an effect that can be employed to produce a stream of <span class="hlt">droplets</span> with a tailor-made size sequence. Compared to the case of purely hydrodynamic <span class="hlt">droplet</span> production without electric fields, the electric control has about the same effect on the <span class="hlt">droplet</span> size if the electric stress at the liquid/liquid interface is the same as the hydrodynamic stress.</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('http://adsabs.harvard.edu/abs/2004PhRvE..69f1508C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004PhRvE..69f1508C"><span>Can a <span class="hlt">droplet</span> break up under flow without elongating? Fragmentation of smectic monodisperse <span class="hlt">droplets</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Courbin, L.; Engl, W.; Panizza, P.</p> <p>2004-06-01</p> <p>We study the fragmentation under shear flow of smectic monodisperse <span class="hlt">droplets</span> at high volume fraction. Using small angle light scattering and optical microscopy, we reveal the existence of a break-up mechanism for which the <span class="hlt">droplets</span> burst into daughter <span class="hlt">droplets</span> of the same size. Surprisingly, this fragmentation process, which is strain controlled and occurs homogeneously in the cell, does not require any transient elongation of the <span class="hlt">droplets</span>. Systematic experiments as a function of the initial <span class="hlt">droplet</span> size and the applied shear rate show that the rupture is triggered by an instability of the inner <span class="hlt">droplet</span> structure.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4849166','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4849166"><span>Floating <span class="hlt">Droplet</span> Array: An Ultrahigh-Throughput Device for <span class="hlt">Droplet</span> Trapping, Real-time Analysis and Recovery</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Labanieh, Louai; Nguyen, Thi N.; Zhao, Weian; Kang, Dong-Ku</p> <p>2016-01-01</p> <p>We describe the design, fabrication and use of a dual-layered microfluidic device for ultrahigh-throughput <span class="hlt">droplet</span> trapping, analysis, and recovery using <span class="hlt">droplet</span> buoyancy. To demonstrate the utility of this device for digital quantification of analytes, we quantify the number of <span class="hlt">droplets</span>, which contain a β-galactosidase-conjugated bead among more than 100,000 immobilized <span class="hlt">droplets</span>. In addition, we demonstrate that this device can be used for <span class="hlt">droplet</span> clustering and real-time analysis by clustering several <span class="hlt">droplets</span> together into microwells and monitoring diffusion of fluorescein, a product of the enzymatic reaction of β-galactosidase and its fluorogenic substrate FDG, between <span class="hlt">droplets</span>. PMID:27134760</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018NanoC...5...12C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018NanoC...5...12C"><span>Dual-nozzle microfluidic <span class="hlt">droplet</span> generator</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Choi, Ji Wook; Lee, Jong Min; Kim, Tae Hyun; Ha, Jang Ho; Ahrberg, Christian D.; Chung, Bong Geun</p> <p>2018-05-01</p> <p>The <span class="hlt">droplet</span>-generating microfluidics has become an important technique for a variety of applications ranging from single cell analysis to nanoparticle synthesis. Although there are a large number of methods for generating and experimenting with <span class="hlt">droplets</span> on microfluidic devices, the dispensing of <span class="hlt">droplets</span> from these microfluidic devices is a challenge due to aggregation and merging of <span class="hlt">droplets</span> at the interface of microfluidic devices. Here, we present a microfluidic dual-nozzle device for the generation and dispensing of uniform-sized <span class="hlt">droplets</span>. The first nozzle of the microfluidic device is used for the generation of the <span class="hlt">droplets</span>, while the second nozzle can accelerate the <span class="hlt">droplets</span> and increase the spacing between them, allowing for facile dispensing of <span class="hlt">droplets</span>. Computational fluid dynamic simulations were conducted to optimize the design parameters of the microfluidic device.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29745386','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29745386"><span>Optical calorimetry in microfluidic <span class="hlt">droplets</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Chamoun, Jacob; Pattekar, Ashish; Afshinmanesh, Farzaneh; Martini, Joerg; Recht, Michael I</p> <p>2018-05-29</p> <p>A novel microfluidic calorimeter that measures the enthalpy change of reactions occurring in 100 μm diameter aqueous <span class="hlt">droplets</span> in fluoropolymer oil has been developed. The aqueous reactants flow into a microfluidic <span class="hlt">droplet</span> generation chip in separate fluidic channels, limiting contact between the streams until immediately before they form the <span class="hlt">droplet</span>. The diffusion-driven mixing of reactants is predominantly restricted to within the <span class="hlt">droplet</span>. The temperature change in <span class="hlt">droplets</span> due to the heat of reaction is measured optically by recording the reflectance spectra of encapsulated thermochromic liquid crystals (TLC) that are added to one of the reactant streams. As the <span class="hlt">droplets</span> travel through the channel, the spectral characteristics of the TLC represent the internal temperature, allowing optical measurement with a precision of ≈6 mK. The microfluidic chip and all fluids are temperature controlled, and the reaction heat within <span class="hlt">droplets</span> raises their temperature until thermal diffusion dissipates the heat into the surrounding oil and chip walls. Position resolved optical temperature measurement of the <span class="hlt">droplets</span> allows calculation of the heat of reaction by analyzing the <span class="hlt">droplet</span> temperature profile over time. Channel dimensions, <span class="hlt">droplet</span> generation rate, <span class="hlt">droplet</span> size, reactant stream flows and oil flow rate are carefully balanced to provide rapid diffusional mixing of reactants compared to thermal diffusion, while avoiding thermal "quenching" due to contact between the <span class="hlt">droplets</span> and the chip walls. Compared to conventional microcalorimetry, which has been used in this work to provide reference measurements, this new continuous flow <span class="hlt">droplet</span> calorimeter has the potential to perform titrations ≈1000-fold faster while using ≈400-fold less reactants per titration.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28985113','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28985113"><span><span class="hlt">Droplets</span> As Liquid Robots.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Čejková, Jitka; Banno, Taisuke; Hanczyc, Martin M; Štěpánek, František</p> <p>2017-01-01</p> <p>Liquid <span class="hlt">droplets</span> are very simple objects present in our everyday life. They are extremely important for many natural phenomena as well as for a broad variety of industrial processes. The conventional research areas in which the <span class="hlt">droplets</span> are studied include physical chemistry, fluid mechanics, chemical engineering, materials science, and micro- and nanotechnology. Typical studies include phenomena such as condensation and <span class="hlt">droplet</span> formation, evaporation of <span class="hlt">droplets</span>, or wetting of surfaces. The present article reviews the recent literature that employs <span class="hlt">droplets</span> as animated soft matter. It is argued that <span class="hlt">droplets</span> can be considered as liquid robots possessing some characteristics of living systems, and such properties can be applied to unconventional computing through maze solving or operation in logic gates. In particular, the lifelike properties and behavior of liquid robots, namely (i) movement, (ii) self-division, and (iii) group dynamics, will be discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4452895','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4452895"><span>Expanding roles for lipid <span class="hlt">droplets</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>Welte, Michael A.</p> <p>2015-01-01</p> <p>Summary Lipid <span class="hlt">droplets</span> are the intracellular sites for neutral lipid storage. They are critical for lipid metabolism and energy homeostasis, and their dysfunction has been linked to many diseases. Accumulating evidence suggests that the roles lipid <span class="hlt">droplets</span> play in biology are significantly broader than previously anticipated. Lipid <span class="hlt">droplets</span> are the source of molecules important in the nucleus: they can sequester transcription factors and chromatin components and generate the lipid ligands for certain nuclear receptors. Lipid <span class="hlt">droplets</span> have also emerged as important nodes for fatty acid trafficking, both inside the cell and between cells. In immunity, new roles for <span class="hlt">droplets</span>, not directly linked to lipid metabolism, have been uncovered, as assembly platforms for specific viruses and as reservoirs for proteins that fight intracellular pathogens. Until recently, knowledge about <span class="hlt">droplets</span> in the nervous system has been minimal, but now there are multiple links between lipid <span class="hlt">droplets</span> and neurodegeneration: Many candidate genes for Hereditary Spastic Paraplegia also have central roles in lipid-<span class="hlt">droplet</span> formation and maintenance, and mitochondrial dysfunction in neurons can lead to transient accumulating of lipid <span class="hlt">droplets</span> in neighboring glial cells, an event that may, in turn, contribute to neuronal damage. As the cell biology and biochemistry of lipid <span class="hlt">droplets</span> are increasingly well understood, the next few years should yield many new mechanistic insights into these novel functions of lipid <span class="hlt">droplets</span>. PMID:26035793</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26289536','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26289536"><span>Sensitive determination of methadone in human serum and urine by dispersive liquid-liquid microextraction based on the solidification of a floating organic <span class="hlt">droplet</span> followed by HPLC-UV.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Taheri, Salman; Jalali, Fahimeh; Fattahi, Nazir; Jalili, Ronak; Bahrami, Gholamreza</p> <p>2015-10-01</p> <p>Dispersive liquid-liquid microextraction based on solidification of floating organic <span class="hlt">droplet</span> was developed for the extraction of methadone and determination by high-performance liquid chromatography with UV detection. In this method, no microsyringe or fiber is required to support the organic microdrop due to the usage of an organic <span class="hlt">solvent</span> with a low density and appropriate melting point. Furthermore, the extractant <span class="hlt">droplet</span> can be collected easily by solidifying it at low temperature. 1-Undecanol and methanol were chosen as extraction and disperser <span class="hlt">solvents</span>, respectively. Parameters that influence extraction efficiency, i.e. volumes of extracting and dispersing <span class="hlt">solvents</span>, pH, and salt effect, were optimized by using response surface methodology. Under optimal conditions, enrichment factor for methadone was 134 and 160 in serum and urine samples, respectively. The limit of detection was 3.34 ng/mmL in serum and 1.67 ng/mL in urine samples. Compared with the traditional dispersive liquid-liquid microextraction, the proposed method obtained lower limit of detection. Moreover, the solidification of floating organic <span class="hlt">solvent</span> facilitated the phase transfer. And most importantly, it avoided using high-density and toxic <span class="hlt">solvents</span> of traditional dispersive liquid-liquid microextraction method. The proposed method was successfully applied to the determination of methadone in serum and urine samples of an addicted individual under methadone therapy. © 2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..DFDKP1066Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..DFDKP1066Y"><span>A Molecular Dynamics Study on Selective Cation Depletion from an Ionic Liquid <span class="hlt">Droplet</span> under an Electric Field</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yang, Yudong; Ahn, Myungmo; Im, Dojin; Oh, Jungmin; Kang, Inseok</p> <p>2017-11-01</p> <p>General electrohydrodynamic behavior of ionic liquid <span class="hlt">droplets</span> under an electric field is investigated using MD simulations. Especially, a unique behavior of ion depletion of an ionic liquid <span class="hlt">droplet</span> under a uniform electric field is studied. Shape deformation due to electric stress and ion distributions inside the <span class="hlt">droplet</span> are calculated to understand the ionic motion of imidazolium-based ionic liquid <span class="hlt">droplets</span> with 200 ion pairs of 2 kinds of ionic liquids: EMIM-NTf2 and EMIM-ES. The intermolecular force between cations and anions can be significantly different due to the nature of the structure and <span class="hlt">charge</span> distribution of the ions. Together with an analytical interpretation of the conducting <span class="hlt">droplet</span> in an electric field, the MD simulation successfully explains the mechanism of selective ion depletion of an ionic liquid <span class="hlt">droplet</span> in an electric field. The selective ion depletion phenomenon has been adopted to explain the experimentally observed retreating motion of a <span class="hlt">droplet</span> in a uniform electric field. The effect of anions on the cation depletion phenomenon can be accounted for from a direct approach to the intermolecular interaction. This research was supproted by the National Research Foundation of Korea (NRF) Grant funded by the Korea government (MSIP) (No. 2017R1D1A1B05035211).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25171210','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25171210"><span>How coalescing <span class="hlt">droplets</span> jump.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Enright, Ryan; Miljkovic, Nenad; Sprittles, James; Nolan, Kevin; Mitchell, Robert; Wang, Evelyn N</p> <p>2014-10-28</p> <p>Surface engineering at the nanoscale is a rapidly developing field that promises to impact a range of applications including energy production, water desalination, self-cleaning and anti-icing surfaces, thermal management of electronics, microfluidic platforms, and environmental pollution control. As the area advances, more detailed insights of dynamic wetting interactions on these surfaces are needed. In particular, the coalescence of two or more <span class="hlt">droplets</span> on ultra-low adhesion surfaces leads to <span class="hlt">droplet</span> jumping. Here we show, through detailed measurements of jumping <span class="hlt">droplets</span> during water condensation coupled with numerical simulations of binary <span class="hlt">droplet</span> coalescence, that this process is fundamentally inefficient with only a small fraction of the available excess surface energy (≲ 6%) convertible into translational kinetic energy. These findings clarify the role of internal fluid dynamics during the jumping <span class="hlt">droplet</span> coalescence process and underpin the development of systems that can harness jumping <span class="hlt">droplets</span> for a wide range of applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26877555','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26877555"><span>STABILITY OF A CYLINDRICAL SOLUTE-<span class="hlt">SOLVENT</span> INTERFACE: EFFECT OF GEOMETRY, ELECTROSTATICS, AND HYDRODYNAMICS.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Li, B O; Sun, Hui; Zhou, Shenggao</p> <p></p> <p>The solute-<span class="hlt">solvent</span> interface that separates biological molecules from their surrounding aqueous <span class="hlt">solvent</span> characterizes the conformation and dynamics of such molecules. In this work, we construct a <span class="hlt">solvent</span> fluid dielectric boundary model for the solvation of <span class="hlt">charged</span> molecules and apply it to study the stability of a model cylindrical solute-<span class="hlt">solvent</span> interface. The motion of the solute-<span class="hlt">solvent</span> interface is defined to be the same as that of <span class="hlt">solvent</span> fluid at the interface. The <span class="hlt">solvent</span> fluid is assumed to be incompressible and is described by the Stokes equation. The solute is modeled simply by the ideal-gas law. All the viscous force, hydrostatic pressure, solute-<span class="hlt">solvent</span> van der Waals interaction, surface tension, and electrostatic force are balanced at the solute-<span class="hlt">solvent</span> interface. We model the electrostatics by Poisson's equation in which the solute-<span class="hlt">solvent</span> interface is treated as a dielectric boundary that separates the low-dielectric solute from the high-dielectric <span class="hlt">solvent</span>. For a cylindrical geometry, we find multiple cylindrically shaped equilibrium interfaces that describe polymodal (e.g., dry and wet) states of hydration of an underlying molecular system. These steady-state solutions exhibit bifurcation behavior with respect to the <span class="hlt">charge</span> density. For their linearized systems, we use the projection method to solve the fluid equation and find the dispersion relation. Our asymptotic analysis shows that, for large wavenumbers, the decay rate is proportional to wavenumber with the proportionality half of the ratio of surface tension to <span class="hlt">solvent</span> viscosity, indicating that the <span class="hlt">solvent</span> viscosity does affect the stability of a solute-<span class="hlt">solvent</span> interface. Consequences of our analysis in the context of biomolecular interactions are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27714352','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27714352"><span>Weaving colloidal webs around <span class="hlt">droplets</span>: spontaneous assembly of extended colloidal networks encasing microfluidic <span class="hlt">droplet</span> ensembles.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zheng, Lu; Ho, Leon Yoon; Khan, Saif A</p> <p>2016-10-26</p> <p>The ability to form transient, self-assembling solid networks that 'cocoon' emulsion <span class="hlt">droplets</span> on-demand allows new possibilities in the rapidly expanding area of microfluidic <span class="hlt">droplet</span>-based materials science. In this communication, we demonstrate the spontaneous formation of extended colloidal networks that encase large microfluidic <span class="hlt">droplet</span> ensembles, thus completely arresting <span class="hlt">droplet</span> motion and effectively isolating each <span class="hlt">droplet</span> from others in the ensemble. To do this, we employ molecular inclusion complexes of β-cyclodextrin, which spontaneously form and assemble into colloidal solids at the <span class="hlt">droplet</span> interface and beyond, via the outward diffusion of a guest molecule (dichloromethane) from the <span class="hlt">droplets</span>. We illustrate the advantage of such transient network-based <span class="hlt">droplet</span> stabilization in the area of pharmaceutical crystallization, where we are able to fabricate monodisperse spherical crystalline microgranules of 5-methyl-2-[(2-nitrophenyl)amino]-3-thiophenecarbonitrile (ROY), a model hydrophobic drug, with a dramatic enhancement of particle properties compared to conventional methods.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19900001074','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19900001074"><span>Theory of <span class="hlt">droplet</span>. Part 1: Renormalized laws of <span class="hlt">droplet</span> vaporization in non-dilute sprays</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Chiu, H. H.</p> <p>1989-01-01</p> <p>The vaporization of a <span class="hlt">droplet</span>, interacting with its neighbors in a non-dilute spray environment is examined as well as a vaporization scaling law established on the basis of a recently developed theory of renormalized <span class="hlt">droplet</span>. The interacting <span class="hlt">droplet</span> consists of a centrally located <span class="hlt">droplet</span> and its vapor bubble which is surrounded by a cloud of <span class="hlt">droplets</span>. The distribution of the <span class="hlt">droplets</span> and the size of the cloud are characterized by a pair-distribution function. The vaporization of a <span class="hlt">droplet</span> is retarded by the collective thermal quenching, the vapor concentration accumulated in the outer sphere, and by the limited percolative passages for mass, momentum and energy fluxes. The retardation is scaled by the local collective interaction parameters (group combustion number of renormalized <span class="hlt">droplet</span>, <span class="hlt">droplet</span> spacing, renormalization number and local ambient conditions). The numerical results of a selected case study reveal that the vaporization correction factor falls from unity monotonically as the group combustion number increases, and saturation is likely to occur when the group combustion number reaches 35 to 40 with interdroplet spacing of 7.5 diameters and an environment temperature of 500 K. The scaling law suggests that dense sprays can be classified into: (1) a diffusively dense cloud characterized by uniform thermal quenching in the cloud; (2) a stratified dense cloud characterized by a radial stratification in temperature by the differential thermal quenching of the cloud; or (3) a sharply dense cloud marked by fine structure in the quasi-<span class="hlt">droplet</span> cloud and the corresponding variation in the correction factor due to the variation in the topological structure of the cloud characterized by a pair-distribution function of quasi-<span class="hlt">droplets</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015APS..MARQ46002L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015APS..MARQ46002L"><span>Precise measurements of <span class="hlt">droplet-droplet</span> contact forces in quasi-2D emulsions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lowensohn, Janna; Orellana, Carlos; Weeks, Eric</p> <p>2015-03-01</p> <p>We use microscopy to visualize a quasi-2D oil-in-water emulsion confined between two parallel slides. We then use the <span class="hlt">droplet</span> shapes to infer the forces they exert on each other. To calibrate our force law, we set up an emulsion in a tilted sample chamber so that the <span class="hlt">droplets</span> feel a known buoyant force. By correlating radius of the <span class="hlt">droplet</span> and length of contacts with the buoyant forces, we validate our empirical force law. We improve upon prior work in our lab by using a high-resolution camera to image each <span class="hlt">droplet</span> multiple times, thus providing sub-pixel resolution and reducing the noise. Our new technique identifies contact forces with only a 1% uncertainty, five times better than prior work. We demonstrate the utility of our technique by examining the normal modes of the <span class="hlt">droplet</span> contact network in our samples.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19970021428','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19970021428"><span>Interaction and Aggregation of Colloidal Biological Particles and <span class="hlt">Droplets</span> in Electrically-Driven Flows</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Davis, Robert H.; Loewenberg, Michael</p> <p>1997-01-01</p> <p>The primary objective of this research was to develop a fundamental understanding of aggregation and coalescence processes during electrically-driven migration of cells, particles and <span class="hlt">droplets</span>. The process by which <span class="hlt">charged</span> cells, particles, molecules, or drops migrate in a weak electric field is known as electrophoresis. If the migrating species have different <span class="hlt">charges</span> or surface potentials, they will migrate at different speeds and thus may collide and aggregate or coalesce. Aggregation and coalescence are undesirable, if the goal is to separate the different species on the basis of their different electrophoretic mobilities.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..DFDA15003N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..DFDA15003N"><span>Magnetic water-in-water <span class="hlt">droplet</span> microfluidics</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Navi, Maryam; Abbasi, Niki; Tsai, Scott</p> <p>2017-11-01</p> <p>Aqueous two-phase systems (ATPS) have shown to be ideal candidates for replacing the conventional water-oil systems used in <span class="hlt">droplet</span> microfluidics. We use an ATPS of Polyethylene Glycol (PEG) and Dextran (DEX) for microfluidic generation of magnetic water-in-water <span class="hlt">droplets</span>. As ferrofluid partitions to DEX phase, there is no significant diffusion of ferrofluid at the interface of the <span class="hlt">droplets</span>, rendering generation of magnetic DEX <span class="hlt">droplets</span> in a non-magnetic continuous phase of PEG possible. In this system, both phases are water-based and highly biocompatible. We microfluidically generate magnetic DEX <span class="hlt">droplets</span> at a flow-focusing junction in a jetting regime. We sort the <span class="hlt">droplets</span> based on their size by placing a permanent magnet downstream of the <span class="hlt">droplet</span> generation region, and show that the deflection of <span class="hlt">droplets</span> is in good agreement with a mathematical model. We also show that the magnetic DEX <span class="hlt">droplets</span> can be stabilized by lysozyme and be used for separation of single cell containing water-in-water <span class="hlt">droplets</span>. This system of magnetic water-in-water <span class="hlt">droplet</span> manipulation may find biomedical applications such as single-cell studies and drug delivery.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004MeScT..15..509L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004MeScT..15..509L"><span>Investigation of spray characteristics from a low-pressure common rail injector for use in a homogeneous <span class="hlt">charge</span> compression ignition engine</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lee, Kihyung; Reitz, Rolf D.</p> <p>2004-03-01</p> <p>Homogeneous <span class="hlt">charge</span> compression ignition (HCCI) combustion provides extremely low levels of pollutant emissions, and thus is an attractive alternative for future IC engines. In order to achieve a uniform mixture distribution within the engine cylinder, the characteristics of the fuel spray play an important role in the HCCI engine concept. It is well known that high-pressure common rail injection systems, mainly used in diesel engines, achieve poor mixture formation because of the possibility of direct fuel impingement on the combustion chamber surfaces. This paper describes spray characteristics of a low-pressure common rail injector which is intended for use in an HCCI engine. Optical diagnostics including laser diffraction and phase Doppler methods, and high-speed camera photography, were applied to measure the spray drop diameter and to investigate the spray development process. The drop sizing results of the laser diffraction method were compared with those of a phase Doppler particle analyser (PDPA) to validate the accuracy of the experiments. In addition, the effect of fuel properties on the spray characteristics was investigated using n-heptane, Stoddard <span class="hlt">solvent</span> (gasoline surrogate) and diesel fuel because HCCI combustion is sensitive to the fuel composition. The results show that the injector forms a hollow-cone sheet spray rather than a liquid jet, and the atomization efficiency is high (small <span class="hlt">droplets</span> are produced). The <span class="hlt">droplet</span> SMD ranged from 15 to 30 µm. The spray break-up characteristics were found to depend on the fuel properties. The break-up time for n-heptane is shorter and the drop SMD is smaller than that of Stoddard <span class="hlt">solvent</span> and diesel fuel.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AcSpA.196...16B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AcSpA.196...16B"><span>Detailed <span class="hlt">solvent</span>, structural, quantum chemical study and antimicrobial activity of isatin Schiff base</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Brkić, Dominik R.; Božić, Aleksandra R.; Marinković, Aleksandar D.; Milčić, Miloš K.; Prlainović, Nevena Ž.; Assaleh, Fathi H.; Cvijetić, Ilija N.; Nikolić, Jasmina B.; Drmanić, Saša Ž.</p> <p>2018-05-01</p> <p>The ratios of E/Z isomers of sixteen synthesized 1,3-dihydro-3-(substituted phenylimino)-2H-indol-2-one were studied using experimental and theoretical methodology. Linear solvation energy relationships (LSER) rationalized <span class="hlt">solvent</span> influence of the <span class="hlt">solvent</span>-solute interactions on the UV-Vis absorption maxima shifts (νmax) of both geometrical isomers using the Kamlet-Taft equation. Linear free energy relationships (LFER) in the form of single substituent parameter equation (SSP) was used to analyze substituent effect on pKa, NMR chemical shifts and νmax values. Electron <span class="hlt">charge</span> density was obtained by the use of Quantum Theory of Atoms in Molecules, i.e. Bader's analysis. The substituent and <span class="hlt">solvent</span> effect on intramolecular <span class="hlt">charge</span> transfer (ICT) were interpreted with the aid of time-dependent density functional (TD-DFT) method. Additionally, the results of TD-DFT calculations quantified the efficiency of ICT from the calculated <span class="hlt">charge</span>-transfer distance (DCT) and amount of transferred <span class="hlt">charge</span> (QCT). The antimicrobial activity was evaluated using broth microdilution method. 3D QSAR modeling was used to demonstrate the influence of substituents effect as well as molecule geometry on antimicrobial activity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17536901','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17536901"><span>Transport characteristics of expiratory <span class="hlt">droplets</span> and <span class="hlt">droplet</span> nuclei in indoor environments with different ventilation airflow patterns.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wan, M P; Chao, C Y H</p> <p>2007-06-01</p> <p>Expiratory <span class="hlt">droplets</span> and <span class="hlt">droplet</span> nuclei can be pathogen carriers for airborne diseases. Their transport characteristics were studied in detail in two idealized floor-supply-type ventilation flow patterns: Unidirectional-upward and single-side-floor, using a multiphase numerical model. The model was validated by running interferometric Mie imaging experiments using test <span class="hlt">droplets</span> with nonvolatile content, which formed <span class="hlt">droplet</span> nuclei, ultimately, in a class-100 clean-room chamber. By comparing the <span class="hlt">droplet</span> dispersion and removal characteristics with data of two other ceiling-supply ventilation systems collected from a previous work, deviations from the perfectly mixed ventilation condition were found to exist in various cases to different extent. The unidirectional-upward system was found to be more efficient in removing the smallest <span class="hlt">droplet</span> nuclei (formed from 1.5 mum <span class="hlt">droplets</span>) by air extraction, but it became less effective for larger <span class="hlt">droplets</span> and <span class="hlt">droplet</span> nuclei. Instead, the single-side-floor system was shown to be more favorable in removing these large <span class="hlt">droplets</span> and <span class="hlt">droplet</span> nuclei. In the single-side-floor system, the lateral overall dispersion coefficients for the small <span class="hlt">droplets</span> and nuclei (initial size </=45 mum) were about an order of magnitude higher than those in the unidirectional-upward system. It indicated that bulk lateral airflow transport in the single-side-floor system was much stronger than the lateral dispersion mechanism induced mainly by air turbulence in the unidirectional-upward system. The time required for the <span class="hlt">droplets</span> and <span class="hlt">droplet</span> nuclei to be transported to the exhaust vent or deposition surfaces for removal varied with different ventilation flow patterns. Possible underestimation of exposure level existed if the perfectly mixed condition was assumed. For example, the weak lateral dispersion in the unidirectional ventilation systems made expiratory <span class="hlt">droplets</span> and <span class="hlt">droplet</span> nuclei stay at close distance to the source leading to highly nonuniform</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29529464','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29529464"><span>Extended hierarchical <span class="hlt">solvent</span> perturbations from curved surfaces of mesoporous silica particles in a deep eutectic <span class="hlt">solvent</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hammons, Joshua A; Zhang, Fan; Ilavsky, Jan</p> <p>2018-06-15</p> <p>Many applications of deep eutectic <span class="hlt">solvents</span> (DES) rely on exploitation of their unique yet complex liquid structures. Due to the ionic nature of the DES components, their diffuse structures are perturbed in the presence of a <span class="hlt">charged</span> surface. We hypothesize that it is possible to perturb the bulk DES structure far (>100 nm) from a curved, <span class="hlt">charged</span> surface with mesoscopic dimensions. We performed in situ, synchrotron-based ultra-small angle X-ray scattering (USAXS) experiments to study the <span class="hlt">solvent</span> distribution near the surface of <span class="hlt">charged</span> mesoporous silica particles (MPS) (≈0.5 µm in diameter) suspended in both water and a common type of DES (1:2 choline Cl-:ethylene glycol). A careful USAXS analysis reveals that the perturbation of electron density distribution within the DES extends ≈1 μm beyond the particle surface, and that this perturbation can be manipulated by the addition of salt ions (AgCl). The concentration of the pore-filling fluid is greatly reduced in the DES. Notably, we extracted the real-space structures of these fluctuations from the USAXS data using a simulated annealing approach that does not require a priori knowledge about the scattering form factor, and can be generalized to a wide range of complex small-angle scattering problems. Copyright © 2018 Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21468418','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21468418"><span>Effect of <span class="hlt">solvent</span> composition on oxide morphology during flame spray pyrolysis of metal nitrates.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Strobel, Reto; Pratsinis, Sotiris E</p> <p>2011-05-28</p> <p>The effect of <span class="hlt">solvent</span> composition on particle formation during flame spray pyrolysis of inexpensive metal-nitrates has been investigated for alumina, iron oxide, cobalt oxide, zinc oxide and magnesium oxide. The as-prepared materials were characterized by electron microscopy, nitrogen adsorption, X-ray diffraction (XRD) and disc centrifugation (XDC). The influence of <span class="hlt">solvent</span> parameters such as boiling point, combustion enthalpy and chemical reactivity on formation of either homogeneous nanoparticles by evaporation/nucleation/coagulation (gas-to-particle conversion) or large particles through precipitation and conversion within the sprayed <span class="hlt">droplets</span> (<span class="hlt">droplet</span>-to-particle conversion) is discussed. For Al(2)O(3), Fe(2)O(3), Co(3)O(4) and partly also MgO, the presence of a carboxylic acid in the FSP solution resulted in homogeneous nanoparticles. This is attributed to formation of volatile metal carboxylates in solution as evidenced by attenuated total reflectance spectroscopy (ATR). For ZnO and MgO rather homogeneous nanoparticles were formed regardless of <span class="hlt">solvent</span> composition. For ZnO this is attributed to its relatively low dissociation temperature compared to other oxides. While for MgO this is traced to the high decomposition temperature of Mg(NO(3))(2) together with Mg(OH)(2)↔MgO transformations. Cobalt oxide (Co(3)O(4)) nanoparticles made by FSP were not aggregated but rather loosely agglomerated as determined by the excellent agreement between XRD- and XDC-derived crystallite and particle sizes, respectively, pointing out the potential of FSP to make non-aggregated particles. This journal is © the Owner Societies 2011</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('https://www.ncbi.nlm.nih.gov/pubmed/27130116','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27130116"><span>Encapsulation of single cells into monodisperse <span class="hlt">droplets</span> by fluorescence-activated <span class="hlt">droplet</span> formation on a microfluidic chip.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hu, Rui; Liu, Pian; Chen, Pu; Wu, Liang; Wang, Yao; Feng, Xiaojun; Liu, Bi-Feng</p> <p>2016-06-01</p> <p>Random compartmentalization of cells by common <span class="hlt">droplet</span> formation methods, i.e., T-junction and flow-focusing, results in low occupancy of <span class="hlt">droplets</span> by single cells. To resolve this issue, a fluorescence-activated <span class="hlt">droplet</span> formation method was developed for the on-command generation of <span class="hlt">droplets</span> and encapsulation of single cells. In this method, <span class="hlt">droplets</span> containing one cell were generated by switching on/off a two-phase hydrodynamic gating valve upon optical detection of single cells. To evaluate the developed method, flow visualization experiments were conducted with fluorescein. Results indicated that picoliter <span class="hlt">droplets</span> of uniform sizes (RSD<4.9%) could be generated. Encapsulation of single fluorescent polystyrene beads demonstrated an average of 94.3% <span class="hlt">droplets</span> contained one bead. Further application of the developed methods to the compartmentalization of individual HeLa cells indicated 82.5% occupancy of <span class="hlt">droplets</span> by single cells, representing a 3 fold increase in comparison to random compartmentalization. Copyright © 2016 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25934429','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25934429"><span>Asymmetric lipid-polymer particles (LIPOMER) by modified nanoprecipitation: role of non-<span class="hlt">solvent</span> composition.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jindal, Anil B; Devarajan, Padma V</p> <p>2015-07-15</p> <p>Asymmetric lipid polymer nanostructures (LIPOMER) comprising glyceryl monostearate (GMS) as lipid and Gantrez AN 119 (Gantrez) as polymer, revealed enhanced splenic accumulation. In the present paper, we attempt to explain the formation of asymmetric GMS LIPOMER using real time imaging. Particles were prepared by precipitation under static conditions using different non-<span class="hlt">solvent</span> phase compositions. The process was video recorded and the videos converted to time elapsed images using the FFmpeg 0.10.2 software at 25 frames/sec. Non-<span class="hlt">solvent</span> compositions comprising >30% of IPA/Acetone revealed significant stranding of the <span class="hlt">solvent</span> phase and slower onset of precipitation(2-6s). At lower concentrations of IPA and acetone, and in non-<span class="hlt">solvent</span> compositions comprising ethanol/water the stranding phenomenon was not evident. Further, rapid precipitation(<1 s) was evident. Nanoprecipitation based on the Marangoni effect is a result of diffusion stranding, interfacial turbulence, and mass transfer of <span class="hlt">solvent</span> and non-<span class="hlt">solvent</span> resulting in solute precipitation. Enhanced diffusion stranding favored by high interaction of GMS and Gantrez(low ΔPol), and the low solubility parameter(Δδtotal) and high mixing enthalpy(ΔHM) of GMS in IPA resulted in <span class="hlt">droplets</span> with random shapes analogous to an amoeba with pseudopodia, which on precipitation formed asymmetric particles. Asymmetric particles could be readily designed through appropriate selection of solutes and non-<span class="hlt">solvent</span> phase by modified nanoprecipitation. Copyright © 2015 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22678963-charge-transfer-solvent-reactions-from-sup-water-methanol-ethanol-studied-time-resolved-photoelectron-spectroscopy-liquids','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22678963-charge-transfer-solvent-reactions-from-sup-water-methanol-ethanol-studied-time-resolved-photoelectron-spectroscopy-liquids"><span><span class="hlt">Charge-transfer-to-solvent</span> reactions from I{sup −} to water, methanol, and ethanol studied by time-resolved photoelectron spectroscopy of liquids</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>Okuyama, Haruki; Karashima, Shutaro; Suzuki, Toshinori, E-mail: suzuki@kuchem.kyoto-u.ac.jp</p> <p></p> <p>The <span class="hlt">charge-transfer-to-solvent</span> (CTTS) reactions from iodide (I{sup −}) to H{sub 2}O, D{sub 2}O, methanol, and ethanol were studied by time-resolved photoelectron spectroscopy of liquid microjets using a magnetic bottle time-of-flight spectrometer with variable pass energy. Photoexcited iodide dissociates into a weak complex (a contact pair) of a solvated electron and an iodine atom in similar reaction times, 0.3 ps in H{sub 2}O and D{sub 2}O and 0.5 ps in methanol and ethanol, which are much shorter than their dielectric relaxation times. The results indicate that solvated electrons are formed with minimal <span class="hlt">solvent</span> reorganization in the long-range <span class="hlt">solvent</span> polarization field createdmore » for I{sup −}. The photoelectron spectra for CTTS in H{sub 2}O and D{sub 2}O—measured with higher accuracy than in our previous study [Y. I. Suzuki et al., Chem. Sci. 2, 1094 (2011)]—indicate that internal conversion yields from the photoexcited I{sup −*} (CTTS) state are less than 10%, while alcohols provide 2–3 times greater yields of internal conversion from I{sup −*}. The overall geminate recombination yields are found to be in the order of H{sub 2}O > D{sub 2}O > methanol > ethanol, which is opposite to the order of the mutual diffusion rates of an iodine atom and a solvated electron. This result is consistent with the transition state theory for an adiabatic outer-sphere electron transfer process, which predicts that the recombination reaction rate has a pre-exponential factor inversely proportional to a longitudinal <span class="hlt">solvent</span> relaxation time.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015EPJD...69..280L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015EPJD...69..280L"><span>Association of amino acids embedded in helium <span class="hlt">droplets</span> detected by mass spectrometry</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lalanne, Matthieu R.; Achazi, Georg; Reichwald, Sebastian; Lindinger, Albrecht</p> <p>2015-12-01</p> <p>Amino acids were embedded in helium <span class="hlt">droplets</span>. The electron impact ionization allows for detecting positively <span class="hlt">charged</span> glycine, valine, histidine, tryptophan and their principal fragments. Monomers and polymers with up to four amino acids are reported. Heterodimers of tryptophan and valine or histidine are observed as well as heterodimers of included fragments. The ability of these associations of molecules to form complexes with water is examined.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20356198','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20356198"><span>Effect of <span class="hlt">solvent</span> and subsequent thermal annealing on the performance of phenylenevinylene copolymer: PCBM solar cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Sharma, G D; Suresh, P; Sharma, S S; Vijay, Y K; Mikroyannidis, John A</p> <p>2010-02-01</p> <p>The morphology of the photoactive layer used in the bulk heterojunction photovoltaic devices is crucial for efficient <span class="hlt">charge</span> generation and their collection at the electrodes. We investigated the <span class="hlt">solvent</span> vapor annealing and thermal annealing effect of an alternating phenylenevinylene copolymer P:PCBM blend on its morphology and optical properties. The UV-visible absorption spectroscopy shows that both <span class="hlt">solvent</span> and thermal annealing can result in self-assembling of copolymer P to form an ordered structure, leading to enhanced absorption in the red region and hole transport enhancement. By combining the <span class="hlt">solvent</span> and thermal annealing of the devices, the power conversion efficiency is improved. This feature was attributed to the fact that the PCBM molecules begin to diffuse into aggregates and together with the ordered copolymer P phase form bicontinuous pathways in the entire layer for efficient <span class="hlt">charge</span> separation and transport. Furthermore, the measured photocurrent also suggests that the space <span class="hlt">charges</span> no longer limit the values of the short circuit current (J(sc)) and fill factor (FF) for <span class="hlt">solvent</span>-treated and thermally annealed devices. These results indicate that the higher J(sc) and PCE for the <span class="hlt">solvent</span>-treated and thermally annealed devices can be attributed to the phase separation of active layers, which leads to a balanced carrier mobility. The overall PCE of the device based on the combination of <span class="hlt">solvent</span> annealing and thermal annealing is about 3.7 %.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19990009689&hterms=Quasi+experiment&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3DQuasi%2Bexperiment','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19990009689&hterms=Quasi+experiment&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3DQuasi%2Bexperiment"><span>Fiber-Supported <span class="hlt">Droplet</span> Combustion. Experiment 32</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Dietrich, Daniel L.; Haggard, John B., Jr.; Nayagam, Vedha; Dryer, Frederick L.; Williams, Forman A.; Shaw, Ben D.</p> <p>1998-01-01</p> <p>Individual <span class="hlt">droplets</span> with diameters ranging from about 2 mm to 5 mm were burned under microgravity conditions in air at 1 bar with an ambient temperature of 300 K. Each <span class="hlt">droplet</span> was tethered by a silicon carbide fiber of 80 mm or 150 mm diameter to keep it in view of video recording, and, in some tests, a forced air flow was applied in a direction parallel to the fiber axis. Methanol, two methanol-water mixtures, two methanol-dodecanol mixtures, and two heptane-hexadecane mixtures were the fuels. <span class="hlt">Droplet</span> diameters were measured as functions of time and compared with existing theoretical predictions. The prediction that methanol <span class="hlt">droplets</span> extinguish at diameters that increase with increasing initial <span class="hlt">droplet</span> diameter is verified by these experiments. In addition, the quasi-steady burning rate constant of the heptane-hexadecane mixtures appears to decrease with increasing <span class="hlt">droplet</span> diameter; obscuration consistent with very heavy sooting, but without the formation of soot shells, is observed for the largest of these <span class="hlt">droplets</span>. Forced convective flow around methanol <span class="hlt">droplets</span> was found to increase the burning rate and to produce a ratio of downstream-to-upstream flame radius that remained constant as the <span class="hlt">droplet</span> size decreased, a trend in agreement with earlier results obtained at higher convective velocities for smaller <span class="hlt">droplets</span> having larger flame standoff ratios. There are a number of implications of the experimental results regarding <span class="hlt">droplet</span>-combustion theory.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012PhDT........23Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012PhDT........23Y"><span>Geometries in Soft Matter From Geometric Frustration, Liquid <span class="hlt">Droplets</span> to Electrostatics in Solution</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yao, Zhenwei</p> <p></p> <p>This thesis explores geometric aspects of soft matter systems. The topics covered fall into three categories: (i) geometric frustrations, including the interplay of geometry and topological defects in two dimensional systems, and the frustration of a planar sheet attached to a curved surface; (ii) geometries of liquid <span class="hlt">droplets</span>, including the curvature driven instabilities of toroidal liquid <span class="hlt">droplets</span> and the self-propulsion of <span class="hlt">droplets</span> on a spatially varying surface topography; (iii) the study of the electric double layer structure around <span class="hlt">charged</span> spherical interfaces by a geometric method. In (i), we study the crystalline order on capillary bridges with varying Gaussian curvature. Energy requires the appearance of topological defects on the surface, which are natural spots for biological activity and chemical functionalization. We further study how liquid crystalline order deforms flexible structured vesicles. In particular we find faceted tetrahedral vesicle as the ground state, which may lead to the design of supra-molecular structures with tetrahedral symmetry and new classes of nano-carriers. Furthermore, by a simple paper model we explore the geometric frustration on a planar sheet when brought to a negative curvature surface in a designed elasto-capillary system. In (ii), motivated by the idea of realizing crystalline order on a stable toroidal <span class="hlt">droplet</span> and a beautiful experiment on toroidal <span class="hlt">droplets</span>, we study the Rayleigh instability and the shrinking instability of thin and fat toroidal <span class="hlt">droplets</span>, where the toroidal geometry plays an essential role. In (iii), by a geometric mapping we construct an approximate analytic spherical solution to the nonlinear Poisson-Boltzmann equation, and identify the applicability regime of the solution. The derived geometric solution enables further analytical study of spherical electrostatic systems such as colloidal suspensions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/6258274-does-harcus-hush-theory-really-work-solvent-dependence-intervalence-charge-transfer-energetics-nh-sub-sub-ru-sup-ii-bipyridine-ru-sup-iii-nh-sub-sub-sup-limit-infinite-dilution','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/6258274-does-harcus-hush-theory-really-work-solvent-dependence-intervalence-charge-transfer-energetics-nh-sub-sub-ru-sup-ii-bipyridine-ru-sup-iii-nh-sub-sub-sup-limit-infinite-dilution"><span>Does Harcus-Hush theory really work The <span class="hlt">solvent</span> dependence of intervalence <span class="hlt">charge</span>-transfer energetics in (NH[sub 3])[sub 5]Ru[sup II]-4,4'-bipyridine-Ru[sup III](NH[sub 3] )[sub 5][sup 5+] in the limit of infinite dilution</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>Hupp, J.T.; Dong, Y.; Blackbourn, R.L.</p> <p>1993-04-01</p> <p>Because of concern about ion-pairing artifacts, the <span class="hlt">solvent</span> dependence of the intervalence <span class="hlt">charge</span>-transfer absorption energy for a prototypical mixed-valence system, (NH[sub 3])[sub 5]Ru[sup III]-4,4'-bipyridine-Ru[sup II](NH[sub 3])[sub 5][sup 5+], has been reexamined in the limit of infinite dilution. New data are reported for 14 <span class="hlt">solvents</span>. While one of these (hexamethylphosphoramide) yields anomalous energetics, the absorption energies for the remaining 13 <span class="hlt">solvents</span> agree qualitatively with the predictions of the Marcus-Hush theory (i.e., two-sphere dielectric continuum theory). On a quantitative basis, however, there is substantial disagreement with theory, at least when the <span class="hlt">charge</span>-transfer distance is equated with the metal-to-metal separation distance (as conventionallymore » done). Replacement of this distance with a much shorter distance inferred from by electronic Stark-effect spectroscopy leads to a 3-fold decrease in the magnitude of calculated <span class="hlt">solvent</span> reorganizational contributions to the overall intervalence energy (and therefore, very good agreement with experiment). Unfortunately, the use of such a short <span class="hlt">charge</span>-transfer distance (d = 5.1 [+-] 0.7 A) also leads to a violation of one of the boundary conditions for use of the two-sphere model. Reformulation of the problem in terms of a generalized dipole-inversion, dielectric cavity problem, however, leads to nearly perfect agreement between theory and experiment. Additional analysis shows that experiment now also agrees reasonably well with theory regarding the magnitude of <span class="hlt">solvent</span>-independent energy contributions. Finally, it is noted that downward revision in the estimated <span class="hlt">charge</span>-transfer distance (from 11.3 to 5.1 A) leads to a substantial upward revision in the experimental (i.e., oscillator-strength based) estimate of the electronic coupling element, H[sub if], for intervalence transfer. 33 refs., 3 figs., 2 tabs.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009mss..confEFA02W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009mss..confEFA02W"><span>Intramolecular <span class="hlt">Charge</span> Transfer States in the Condensed Phase</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Williams, C. F.; Herbert, J. M.</p> <p>2009-06-01</p> <p>Time-Dependent Density Functional Theory (TDDFT) with long range corrected functionals can give accurate results for the energies of electronically excited states involving Intramolecular <span class="hlt">Charge</span> Transfer (ICT) in large molecules. If this is combined with a Molecular Mechanics (MM) representation of the surrounding <span class="hlt">solvent</span> this technique can be used to interpret the results of condensed phase UV-Vis Spectroscopy. Often the MM region is represented by a set of point <span class="hlt">charges</span>, however this means that the <span class="hlt">solvent</span> cannot repolarize to adapt to the new <span class="hlt">charge</span> distribution as a result of ICT and so the excitation energies to ICT states are overestimated. To solve this problem an algorithm that interfaces TDDFT with the polarizable force-field AMOEBA is presented; the effect of solvation on <span class="hlt">charge</span> transfer in species such as 4,4'dimethylaminobenzonitrile (DMABN) is discussed. M.A. Rohrdanz, K.M. Martins, and J.M. Herbert, J. Chem. Phys. 130 034107 (2008).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28205642','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28205642"><span>Evaporation of inclined water <span class="hlt">droplets</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Kim, Jin Young; Hwang, In Gyu; Weon, Byung Mook</p> <p>2017-02-16</p> <p>When a drop is placed on a flat substrate tilted at an inclined angle, it can be deformed by gravity and its initial contact angle divides into front and rear contact angles by inclination. Here we study on evaporation dynamics of a pure water <span class="hlt">droplet</span> on a flat solid substrate by controlling substrate inclination and measuring mass and volume changes of an evaporating <span class="hlt">droplet</span> with time. We find that complete evaporation time of an inclined <span class="hlt">droplet</span> becomes longer as gravitational influence by inclination becomes stronger. The gravity itself does not change the evaporation dynamics directly, whereas the gravity-induced <span class="hlt">droplet</span> deformation increases the difference between front and rear angles, which quickens the onset of depinning and consequently reduces the contact radius. This result makes the evaporation rate of an inclined <span class="hlt">droplet</span> to be slow. This finding would be important to improve understanding on evaporation dynamics of inclined <span class="hlt">droplets</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5311959','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5311959"><span>Evaporation of inclined water <span class="hlt">droplets</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>Kim, Jin Young; Hwang, In Gyu; Weon, Byung Mook</p> <p>2017-01-01</p> <p>When a drop is placed on a flat substrate tilted at an inclined angle, it can be deformed by gravity and its initial contact angle divides into front and rear contact angles by inclination. Here we study on evaporation dynamics of a pure water <span class="hlt">droplet</span> on a flat solid substrate by controlling substrate inclination and measuring mass and volume changes of an evaporating <span class="hlt">droplet</span> with time. We find that complete evaporation time of an inclined <span class="hlt">droplet</span> becomes longer as gravitational influence by inclination becomes stronger. The gravity itself does not change the evaporation dynamics directly, whereas the gravity-induced <span class="hlt">droplet</span> deformation increases the difference between front and rear angles, which quickens the onset of depinning and consequently reduces the contact radius. This result makes the evaporation rate of an inclined <span class="hlt">droplet</span> to be slow. This finding would be important to improve understanding on evaporation dynamics of inclined <span class="hlt">droplets</span>. PMID:28205642</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22493215-counterintuitive-electron-localisation-from-density-functional-theory-polarisable-solvent-models','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22493215-counterintuitive-electron-localisation-from-density-functional-theory-polarisable-solvent-models"><span>Counterintuitive electron localisation from density-functional theory with polarisable <span class="hlt">solvent</span> models</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>Dale, Stephen G., E-mail: sdale@ucmerced.edu; Johnson, Erin R., E-mail: erin.johnson@dal.ca</p> <p>2015-11-14</p> <p>Exploration of the solvated electron phenomena using density-functional theory (DFT) generally results in prediction of a localised electron within an induced <span class="hlt">solvent</span> cavity. However, it is well known that DFT favours highly delocalised <span class="hlt">charges</span>, rendering the localisation of a solvated electron unexpected. We explore the origins of this counterintuitive behaviour using a model Kevan-structure system. When a polarisable-continuum <span class="hlt">solvent</span> model is included, it forces electron localisation by introducing a strong energetic bias that favours integer <span class="hlt">charges</span>. This results in the formation of a large energetic barrier for <span class="hlt">charge</span>-hopping and can cause the self-consistent field to become trapped in local minimamore » thus converging to stable solutions that are higher in energy than the ground electronic state. Finally, since the bias towards integer <span class="hlt">charges</span> is caused by the polarisable continuum, these findings will also apply to other classical polarisation corrections, as in combined quantum mechanics and molecular mechanics (QM/MM) methods. The implications for systems beyond the solvated electron, including cationic DNA bases, are discussed.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20050177216','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20050177216"><span>Fiber-Supported <span class="hlt">Droplet</span> Combustion Experiment-2</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Colantonio, Renato O.</p> <p>1998-01-01</p> <p>A major portion of the energy produced in the world today comes from the burning of liquid hydrocarbon fuels in the form of <span class="hlt">droplets</span>. Understanding the fundamental physical processes involved in <span class="hlt">droplet</span> combustion is not only important in energy production but also in propulsion, in the mitigation of combustion-generated pollution, and in the control of the fire hazards associated with handling liquid combustibles. Microgravity makes spherically symmetric combustion possible, allowing investigators to easily validate their <span class="hlt">droplet</span> models without the complicating effects of gravity. The Fiber-Supported <span class="hlt">Droplet</span> Combustion (FSDC-2) investigation was conducted in the Microgravity Glovebox facility of the shuttles' Spacelab during the reflight of the Microgravity Science Laboratory (MSL- 1R) on STS-94 in July 1997. FSDC-2 studied fundamental phenomena related to liquid fuel <span class="hlt">droplet</span> combustion in air. Pure fuels and mixtures of fuels were burned as isolated single and duo <span class="hlt">droplets</span> with and without forced air convection. FSDC-2 is sponsored by the NASA Lewis Research Center, whose researchers are working in cooperation with several investigators from industry and academia. The rate at which a <span class="hlt">droplet</span> burns is important in many commercial applications. The classical theory of <span class="hlt">droplet</span> burning assumes that, for an isolated, spherically symmetric, single-fuel <span class="hlt">droplet</span>, the gas-phase combustion processes are much faster than the <span class="hlt">droplet</span> surface regression rate and that the liquid phase is at a uniform temperature equal to the boiling point. Recent, more advanced models predict that both the liquid and gas phases are unsteady during a substantial portion of the <span class="hlt">droplet</span>'s burning history, thus affecting the instantaneous and average burning rates, and that flame radiation is a dominant mechanism that can extinguish flames in a microgravity environment. FSDC-2 has provided well-defined, symmetric <span class="hlt">droplet</span> burning data including radiative emissions to validate these theoretical</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24175664','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24175664"><span>Development of novel zein-sodium caseinate nanoparticle (ZP)-stabilized emulsion films for improved water barrier properties via emulsion/<span class="hlt">solvent</span> evaporation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wang, Li-Juan; Yin, Ye-Chong; Yin, Shou-Wei; Yang, Xiao-Quan; Shi, Wei-Jian; Tang, Chuan-He; Wang, Jin-Mei</p> <p>2013-11-20</p> <p>This work attempted to develop novel high barrier zein/SC nanoparticle (ZP)-stabilized emulsion films through microfluidic emulsification (ZPE films) or in combination with <span class="hlt">solvent</span> (ethyl acetate) evaporation techniques (ZPE-EA films). Some physical properties, including tensile and optical properties, water vapor permeability (WVP), and surface hydrophobicity, as well as the microstructure of ZP-stabilized emulsion films were evaluated and compared with SC emulsion (SCE) films. The emulsion/<span class="hlt">solvent</span> evaporation approach reduced lipid <span class="hlt">droplets</span> of ZP-stabilized emulsions, and lipid <span class="hlt">droplets</span> of ZP-stabilized emulsions were similar to or slightly lower than that of SC emulsions. However, ZP- and SC-stabilized emulsion films exhibited a completely different microstructure, nanoscalar lipid <span class="hlt">droplets</span> were homogeneously distributed in the ZPE film matrix and interpenetrating protein-oil complex networks occurred within ZPE-EA films, whereas SCE films presented a heterogeneous microstructure. The different stabilization mechanisms against creaming or coalescence during film formation accounted for the preceding discrepancy of the microstructures between ZP-and SC-stabilized emulsion films. Interestingly, ZP-stabilized emulsion films exhibited a better water barrier efficiency, and the WVP values were only 40-50% of SCE films. A schematic representation for the formation of ZP-stabilized emulsion films was proposed to relate the physical performance of the films with their microstructure and to elucidate the possible forming mechanism of the films.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29747499','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29747499"><span>Interfaces <span class="hlt">Charged</span> by a Nonionic Surfactant.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lee, Joohyung; Zhou, Zhang-Lin; Behrens, Sven Holger</p> <p>2018-05-24</p> <p>Highly hydrophobic, water-insoluble nonionic surfactants are often considered irrelevant to the ionization of interfaces at which they adsorb, despite observations that suggest otherwise. In the present study, we provide unambiguous evidence for the participation of a water-insoluble surfactant in interfacial ionization by conducting electrophoresis experiments for surfactant-stabilized nonpolar oil <span class="hlt">droplets</span> in aqueous continuous phase. It was found that the surfactant with amine headgroup positively <span class="hlt">charged</span> the surface of oil suspended in aqueous continuous phase (oil/water interface), which is consistent with its basic nature. In nonpolar oil continuous phase, the same surfactant positively <span class="hlt">charged</span> the surface of solid silica (solid/oil interface) which is often considered acidic. The latter observation is exactly opposite to what the traditional acid-base mechanism of surface <span class="hlt">charging</span> would predict, most clearly suggesting the possibility for another <span class="hlt">charging</span> mechanism.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1557553','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1557553"><span>The Denaturation Transition of DNA in Mixed <span class="hlt">Solvents</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>Hammouda, Boualem; Worcester, David</p> <p>2006-01-01</p> <p>The helix-to-coil denaturation transition in DNA has been investigated in mixed <span class="hlt">solvents</span> at high concentration using ultraviolet light absorption spectroscopy and small-angle neutron scattering. Two <span class="hlt">solvents</span> have been used: water and ethylene glycol. The “melting” transition temperature was found to be 94°C for 4% mass fraction DNA/d-water and 38°C for 4% mass fraction DNA/d-ethylene glycol. The DNA melting transition temperature was found to vary linearly with the <span class="hlt">solvent</span> fraction in the mixed <span class="hlt">solvents</span> case. Deuterated <span class="hlt">solvents</span> (d-water and d-ethylene glycol) were used to enhance the small-angle neutron scattering signal and 0.1M NaCl (or 0.0058 g/g mass fraction) salt concentration was added to screen <span class="hlt">charge</span> interactions in all cases. DNA structural information was obtained by small-angle neutron scattering, including a correlation length characteristic of the inter-distance between the hydrogen-containing (desoxyribose sugar-amine base) groups. This correlation length was found to increase from 8.5 to 12.3 Å across the melting transition. Ethylene glycol and water mixed <span class="hlt">solvents</span> were found to mix randomly in the solvation region in the helix phase, but nonideal <span class="hlt">solvent</span> mixing was found in the melted coil phase. In the coil phase, <span class="hlt">solvent</span> mixtures are more effective solvating agents than either of the individual <span class="hlt">solvents</span>. Once melted, DNA coils behave like swollen water-soluble synthetic polymer chains. PMID:16815902</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://www.ars.usda.gov/research/publications/publication/?seqNo115=283699','TEKTRAN'); return false;" href="http://www.ars.usda.gov/research/publications/publication/?seqNo115=283699"><span>Airspeed and orifice size affect spray <span class="hlt">droplet</span> spectra from an aerial electrostatic nozzle for rotary-wing applications</span></a></p> <p><a target="_blank" href="https://www.ars.usda.gov/research/publications/find-a-publication/">USDA-ARS?s Scientific Manuscript database</a></p> <p></p> <p></p> <p>The aerial electrostatic spraying system patented by the USDA-ARS is a unique aerial application system which inductively <span class="hlt">charges</span> spray <span class="hlt">droplets</span> for the purpose of increasing deposition and efficacy. While this system has many potential benefits, no published data exits which describe how changes i...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/494112','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/494112"><span>Uniform-<span class="hlt">droplet</span> spray forming</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>Blue, C.A.; Sikka, V.K.; Chun, Jung-Hoon</p> <p>1997-04-01</p> <p>The uniform-<span class="hlt">droplet</span> process is a new method of liquid-metal atomization that results in single <span class="hlt">droplets</span> that can be used to produce mono-size powders or sprayed-on to substrates to produce near-net shapes with tailored microstructure. The mono-sized powder-production capability of the uniform-<span class="hlt">droplet</span> process also has the potential of permitting engineered powder blends to produce components of controlled porosity. Metal and alloy powders are commercially produced by at least three different methods: gas atomization, water atomization, and rotating disk. All three methods produce powders of a broad range in size with a very small yield of fine powders with single-sized <span class="hlt">droplets</span> thatmore » can be used to produce mono-size powders or sprayed-on substrates to produce near-net shapes with tailored microstructures. The economical analysis has shown the process to have the potential of reducing capital cost by 50% and operating cost by 37.5% when applied to powder making. For the spray-forming process, a 25% savings is expected in both the capital and operating costs. The project is jointly carried out at Massachusetts Institute of Technology (MIT), Tuffs University, and Oak Ridge National Laboratory (ORNL). Preliminary interactions with both finished parts and powder producers have shown a strong interest in the uniform-<span class="hlt">droplet</span> process. Systematic studies are being conducted to optimize the process parameters, understand the solidification of <span class="hlt">droplets</span> and spray deposits, and develop a uniform-<span class="hlt">droplet</span>-system (UDS) apparatus appropriate for processing engineering alloys.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28528432','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28528432"><span>Lipid <span class="hlt">Droplets</span>: Formation to Breakdown.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Meyers, Alex; Weiskittel, Taylor M; Dalhaimer, Paul</p> <p>2017-06-01</p> <p>One of the most exciting areas of cell biology during the last decade has been the study of lipid <span class="hlt">droplets</span>. Lipid <span class="hlt">droplets</span> allow cells to store non-polar molecules such as neutral lipids in specific compartments where they are sequestered from the aqueous environment of the cell yet can be accessed through regulated mechanisms. These structures are highly conserved, appearing in organisms throughout the phylogenetic tree. Until somewhat recently, lipid <span class="hlt">droplets</span> were widely regarded as inert, however progress in the field has continued to demonstrate their vast roles in a number of cellular processes in both mitotic and post-mitotic cells. No doubt the increase in the attention given to lipid <span class="hlt">droplet</span> research is due to their central role in current pressing human diseases such as obesity, type-2 diabetes, and atherosclerosis. This review provides a mechanistic timeline from neutral lipid synthesis through lipid <span class="hlt">droplet</span> formation and size augmentation to <span class="hlt">droplet</span> breakdown.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20010004354','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20010004354"><span>Vibration-Induced <span class="hlt">Droplet</span> Atomization</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Smith, M. K.; James, A.; Vukasinovic, B.; Glezer, A.</p> <p>1999-01-01</p> <p>Thermal management is critical to a number of technologies used in a microgravity environment and in Earth-based systems. Examples include electronic cooling, power generation systems, metal forming and extrusion, and HVAC (heating, venting, and air conditioning) systems. One technique that can deliver the large heat fluxes required for many of these technologies is two-phase heat transfer. This type of heat transfer is seen in the boiling or evaporation of a liquid and in the condensation of a vapor. Such processes provide very large heat fluxes with small temperature differences. Our research program is directed toward the development of a new, two-phase heat transfer cell for use in a microgravity environment. In this paper, we consider the main technology used in this cell, a novel technique for the atomization of a liquid called vibration-induced <span class="hlt">droplet</span> atomization. In this process, a small liquid <span class="hlt">droplet</span> is placed on a thin metal diaphragm that is made to vibrate by an attached piezoelectric transducer. The vibration induces capillary waves on the free surface of the <span class="hlt">droplet</span> that grow in amplitude and then begin to eject small secondary <span class="hlt">droplets</span> from the wave crests. In some situations, this ejection process develops so rapidly that the entire <span class="hlt">droplet</span> seems to burst into a small cloud of atomized <span class="hlt">droplets</span> that move away from the diaphragm at speeds of up to 50 cm/s. By incorporating this process into a heat transfer cell, the active atomization and transport of the small liquid <span class="hlt">droplets</span> could provide a large heat flux capability for the device. Experimental results are presented that document the behavior of the diaphragm and the <span class="hlt">droplet</span> during the course of a typical bursting event. In addition, a simple mathematical model is presented that qualitatively reproduces all of the essential features we have seen in a burst event. From these two investigations, we have shown that delayed <span class="hlt">droplet</span> bursting results when the system passes through a resonance</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/2017NatSR...740059B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017NatSR...740059B"><span>Stochastic kinetics reveal imperative role of anisotropic interfacial tension to determine morphology and evolution of nucleated <span class="hlt">droplets</span> in nematogenic films</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bhattacharjee, Amit Kumar</p> <p>2017-01-01</p> <p>For isotropic fluids, classical nucleation theory predicts the nucleation rate, barrier height and critical <span class="hlt">droplet</span> size by ac- counting for the competition between bulk energy and interfacial tension. The nucleation process in liquid crystals is less understood. We numerically investigate nucleation in monolayered nematogenic films using a mesoscopic framework, in par- ticular, we study the morphology and kinetic pathway in spontaneous formation and growth of <span class="hlt">droplets</span> of the stable phase in the metastable background. The parameter κ that quantifies the anisotropic elastic energy plays a central role in determining the geometric structure of the <span class="hlt">droplets</span>. Noncircular nematic <span class="hlt">droplets</span> with homogeneous director orientation are nucleated in a background of supercooled isotropic phase for small κ. For large κ, noncircular <span class="hlt">droplets</span> with integer topological <span class="hlt">charge</span>, accompanied by a biaxial ring at the outer surface, are nucleated. The isotropic <span class="hlt">droplet</span> shape in a superheated nematic background is found to depend on κ in a similar way. Identical growth laws are found in the two cases, although an unusual two-stage mechanism is observed in the nucleation of isotropic <span class="hlt">droplets</span>. Temporal distributions of successive events indi- cate the relevance of long-ranged elasticity-mediated interactions within the isotropic domains. Implications for a theoretical description of nucleation in anisotropic fluids are discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016RJPCA..90..130T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016RJPCA..90..130T"><span>The theoretical investigation of <span class="hlt">solvent</span> effects on the relative stability and 15N NMR shielding of antidepressant heterocyclic drug</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tahan, Arezoo; Khojandi, Mahya; Salari, Ali Akbar</p> <p>2016-01-01</p> <p>The density functional theory (DFT) and Tomasi's polarized continuum model (PCM) were used for the investigation of <span class="hlt">solvent</span> polarity and its dielectric constant effects on the relative stability and NMR shielding tensors of antidepressant mirtazapine (MIR). The obtained results indicated that the relative stability in the polar <span class="hlt">solvents</span> is higher than that in non-polar <span class="hlt">solvents</span> and the most stable structure was observed in the water at the B3LYP/6-311++G ( d, p) level of theory. Also, natural bond orbital (NBO) interpretation demonstrated that by increase of <span class="hlt">solvent</span> dielectric constant, negative <span class="hlt">charge</span> on nitrogen atoms of heterocycles and resonance energy for LP(N10) → σ* and π* delocalization of the structure's azepine ring increase and the highest values of them were observed in water. On the other hand, NMR calculations showed that with an increase in negative <span class="hlt">charge</span> of nitrogen atoms, isotropic chemical shielding (σiso) around them increase and nitrogen of piperazine ring (N19) has the highest values of negative <span class="hlt">charge</span> and σiso among nitrogen atoms. NMR calculations also represented that direct <span class="hlt">solvent</span> effect on nitrogen of pyridine ring (N15) is more than other nitrogens, while its effect on N19 is less than other ones. Based on NMR data and NBO interpretation, it can be deduced that with a decrease in the negative <span class="hlt">charge</span> on nitrogen atoms, the intramolecular effects on them decrease, while direct <span class="hlt">solvent</span> effect increases.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/19231352','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/19231352"><span>A novel dispersive liquid-liquid microextraction based on solidification of floating organic <span class="hlt">droplet</span> method for determination of polycyclic aromatic hydrocarbons in aqueous samples.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Xu, Hui; Ding, Zongqing; Lv, Lili; Song, Dandan; Feng, Yu-Qi</p> <p>2009-03-16</p> <p>A new dispersive liquid-liquid microextraction based on solidification of floating organic <span class="hlt">droplet</span> method (DLLME-SFO) was developed for the determination of five kinds of polycyclic aromatic hydrocarbons (PAHs) in environmental water samples. In this method, no specific holder, such as the needle tip of microsyringe and the hollow fiber, is required for supporting the organic microdrop due to the using of organic <span class="hlt">solvent</span> with low density and proper melting point. Furthermore, the extractant <span class="hlt">droplet</span> can be collected easily by solidifying it in the lower temperature. 1-Dodecanol was chosen as extraction <span class="hlt">solvent</span> in this work. A series of parameters that influence extraction were investigated systematically. Under optimal conditions, enrichment factors (EFs) for PAHs were in the range of 88-118. The limit of detections (LODs) for naphthalene, diphenyl, acenaphthene, anthracene and fluoranthene were 0.045, 0.86, 0.071, 1.1 and 0.66ngmL(-1), respectively. Good reproducibility and recovery of the method were also obtained. Compared with the traditional liquid-phase microextraction (LPME) and dispersive liquid-liquid microextraction (DLLME) methods, the proposed method obtained about 2 times higher enrichment factor than those in LPME. Moreover, the solidification of floating organic <span class="hlt">solvent</span> facilitated the phase transfer. And most importantly, it avoided using high-density and toxic <span class="hlt">solvent</span> in the traditional DLLME method. The proposed method was successfully applied to determinate PAHs in the environmental water samples. The simple and low-cost method provides an alternative method for the analysis of non-polar compounds in complex environmental water.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/18988203','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18988203"><span>Using <span class="hlt">solvent</span>-free sample preparation to promote protonation of poly(ethylene oxide)s with labile end-groups in matrix-assisted laser desorption/ionisation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Mazarin, Michael; Phan, Trang N T; Charles, Laurence</p> <p>2008-12-01</p> <p>Protonation is usually required to observe intact ions during matrix-assisted laser desorption/ionization (MALDI) of polymers containing fragile end-groups while cation adduction induces chain-end degradation. These polymers, generally obtained via living free radical polymerization techniques, are terminated with a functionality in which a bond is prone to homolytic cleavage, as required by the polymerization process. A <span class="hlt">solvent</span>-free sample preparation method was used here to avoid salt contaminant from the <span class="hlt">solvent</span> traditionally used in the dried-<span class="hlt">droplet</span> MALDI procedure. <span class="hlt">Solvent</span>-based and <span class="hlt">solvent</span>-free sample preparations were compared for a series of three poly(ethylene oxide) polymers functionalized with a labile end-group in a nitroxide-mediated polymerization reaction, using 2,4,6-trihydroxyacetophenone (THAP) as the matrix without any added salt. Intact oligomer ions could only be produced as protonated molecules in <span class="hlt">solvent</span>-free MALDI while sodium adducts of degraded polymers were formed from the dried-<span class="hlt">droplet</span> samples. Although MALDI analysis was performed at the laser threshold, fragmentation of protonated macromolecules was still observed to occur. However, in contrast to sodiated molecules, dissociation of protonated oligomers does not involve the labile C--ON bond of the end-group. As the macromolecule size increased, protonation appeared to be less efficient and sodium adduction became the dominant ionization process, although no sodium salt was added in the preparation. Formation of sodiated degraded macromolecules would be dictated by increasing cation affinity as the size of the oligomers increases and would reveal the presence of salts at trace levels in the MALDI samples.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19970003810','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19970003810"><span><span class="hlt">Droplet</span> Combustion and Soot Formation in Microgravity</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Avedisian, C. Thomas</p> <p>1994-01-01</p> <p>One of the most complex processes involved in the combustion ot liquid fuels is the formation of soot. A well characterized flow field and simplified flame structure can improve considerably the understanding of soot formation processes. The simplest flame shape to analyze for a <span class="hlt">droplet</span> is spherical with its associated one-dimensional flow field. It is a fundamental limit and the oldest and most often analyzed configuration of <span class="hlt">droplet</span> combustion. Spherical symmetry in the <span class="hlt">droplet</span> burning process will arise when there is no relative motion between the <span class="hlt">droplet</span> and ambience or uneven heating around the <span class="hlt">droplet</span> periphery, and buoyancy effects are negligible. The flame and <span class="hlt">droplet</span> are then concentric with each other and there is no liquid circulation within the <span class="hlt">droplet</span>. An understanding of the effect of soot on <span class="hlt">droplet</span> combustion should therefore benefit from this simplified configuration. Soot formed during spherically symmetric <span class="hlt">droplet</span> combustion, however, has only recently drawn attention and it appears to be one of the few aspects associated with <span class="hlt">droplet</span> combustion which have not yet been thoroughly investigated. For this review, the broad subject of <span class="hlt">droplet</span> combustion is narrowed considerably by restricting attention specifically to soot combined with spherically symmetric <span class="hlt">droplet</span> burning processes that are promoted.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=MSFC-0300169&hterms=learn+better+video&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dlearn%2Bbetter%2Bvideo','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=MSFC-0300169&hterms=learn+better+video&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dlearn%2Bbetter%2Bvideo"><span>Fuel <span class="hlt">Droplet</span> Burning During <span class="hlt">Droplet</span> Combustion 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>2003-01-01</p> <p>Fuel ignites and burns in the <span class="hlt">Droplet</span> Combustion Experiment (DCE) on STS-94 on July 4 1997, MET:2/05:40 (approximate). The DCE was designed to investigate the fundamental combustion aspects of single, isolated <span class="hlt">droplets</span> under different pressures and ambient oxygen concentrations for a range of <span class="hlt">droplet</span> sizes varying between 2 and 5 mm. DCE used various fuels -- in drops ranging from 1 mm (0.04 inches) to 5 mm (0.2 inches) -- and mixtures of oxidizers and inert gases to learn more about the physics of combustion in the simplest burning configuration, a sphere. The experiment elapsed time is shown at the bottom of the composite image. The DCE principal investigator was Forman Williams, University of California, San Diego. The experiment was part of the space research investigations conducted during the Microgravity Science Laboratory-1R mission (STS-94, July 1-17 1997). Advanced combustion experiments will be a part of investigations plarned for the International Space Station. (1.4MB, 13-second MPEG, screen 320 x 240 pixels; downlinked video, higher quality not available)A still JPG composite of this movie is available at http://mix.msfc.nasa.gov/ABSTRACTS/MSFC-0300168.html.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20129107','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20129107"><span><span class="hlt">Droplet</span>-based gene expression analysis using a device with magnetic force-based-<span class="hlt">droplet</span>-handling system.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Okochi, Mina; Tsuchiya, Hiroyoshi; Kumazawa, Fumitaka; Shikida, Mitsuhiro; Honda, Hiroyuki</p> <p>2010-02-01</p> <p>A <span class="hlt">droplet</span>-based cell lysis and reverse transcription-polymerase chain reaction (PCR) were performed on-chip employing magnetic force-based-<span class="hlt">droplet</span>-handling system. The actuation with a magnet offers a simple system for <span class="hlt">droplet</span> manipulation; it does not need mechanical fluidic systems such as pumps and valves for handling solutions. It can be used as a powerful tool for various biochemical applications by moving and coalescing sample <span class="hlt">droplets</span> using magnetic beads immersed in mineral oil. The <span class="hlt">droplet</span> containing magnetic beads and the cells were manipulated with the magnet located underneath the channel, and coalesced with a <span class="hlt">droplet</span> of lysis buffer. Using K562 cells as the leukemia model, the cell lysis, cDNA synthesis, and amplification of WT1 gene that is known as the prognostic factor for acute leukemia were successfully performed from a single cell. Copyright (c) 2009 The Society for Biotechnology, Japan. Published by Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1342588-morphology-evolution-high-performance-polymer-solar-cells-processed-from-nonhalogenated-solvent','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1342588-morphology-evolution-high-performance-polymer-solar-cells-processed-from-nonhalogenated-solvent"><span>Morphology evolution in high-performance polymer solar cells processed from nonhalogenated <span class="hlt">solvent</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Cai, Wanzhu; Liu, Peng; Jin, Yaocheng; ...</p> <p>2015-05-26</p> <p>A new processing protocol based on non-halogenated <span class="hlt">solvent</span> and additive is developed to produce polymer solar cells with power conversion efficiencies better than those processed from commonly used halogenated <span class="hlt">solvent</span>-additive pair. Morphology studies show that good performance correlates with a finely distributed nanomorphology with a well-defined polymer fibril network structure, which leads to balanced <span class="hlt">charge</span> transport in device operation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/20723270-reorganization-energy-electron-transfer-nonpolar-solvents-molecular-level-treatment-solvent','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/20723270-reorganization-energy-electron-transfer-nonpolar-solvents-molecular-level-treatment-solvent"><span>The reorganization energy of electron transfer in nonpolar <span class="hlt">solvents</span>: Molecular level treatment of the <span class="hlt">solvent</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>Leontyev, I.V.; Tachiya, M.</p> <p></p> <p>The intermolecular electron transfer in a solute pair consisting of pyrene and dimethylaniline is investigated in a nonpolar <span class="hlt">solvent</span>, n-hexane. The earlier elaborated approach [M. Tachiya, J. Phys Chem. 97, 5911 (1993)] is used; this method provides a physically relevant background for separating inertial and inertialess polarization responses for both nonpolarizable and polarizable molecular level simulations. The molecular-dynamics technique was implemented for obtaining the equilibrium ensemble of <span class="hlt">solvent</span> configurations. The nonpolar <span class="hlt">solvent</span>, n-hexane, was treated in terms of OPLS-AA parametrization. Solute Lennard-Jones parameters were taken from the same parametrization. Solute <span class="hlt">charge</span> distributions of the initial and final states were determinedmore » using ab initio level [HF/6-31G(d,p)] quantum-chemical calculations. Configuration analysis was performed explicitly taking into account the anisotropic polarizability of n-hexane. It is shown that the Gaussian law well describes calculated distribution functions of the <span class="hlt">solvent</span> coordinate, therefore, the rate constant of the ET reaction can be characterized by the reorganization energy. Evaluated values of the reorganization energies are in a range of 0.03-0.11 eV and significant contribution (more then 40% of magnitude) comes from anisotropic polarizability. Investigation of the reorganization energy {lambda} dependence on the solute pair separation distance d revealed unexpected behavior. The dependence has a very sharp peak at the distance d=7 A where <span class="hlt">solvent</span> molecules are able to penetrate into the intermediate space between the solute pair. The reason for such behavior is clarified. This new effect has a purely molecular origin and cannot be described within conventional continuum <span class="hlt">solvent</span> models.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1236613-solvent-effects-time-dependent-self-consistent-field-methods-optical-response-calculations','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1236613-solvent-effects-time-dependent-self-consistent-field-methods-optical-response-calculations"><span><span class="hlt">Solvent</span> effects in time-dependent self-consistent field methods. I. Optical response calculations</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Bjorgaard, J. A.; Kuzmenko, V.; Velizhanin, K. A.; ...</p> <p>2015-01-22</p> <p>In this study, we implement and examine three excited state <span class="hlt">solvent</span> models in time-dependent self-consistent field methods using a consistent formalism which unambiguously shows their relationship. These are the linear response, state specific, and vertical excitation <span class="hlt">solvent</span> models. Their effects on energies calculated with the equivalent of COSMO/CIS/AM1 are given for a set of test molecules with varying excited state <span class="hlt">charge</span> transfer character. The resulting <span class="hlt">solvent</span> effects are explained qualitatively using a dipole approximation. It is shown that the fundamental differences between these <span class="hlt">solvent</span> models are reflected by the character of the calculated excitations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..DMP.K1063L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..DMP.K1063L"><span>Selfbound quantum <span class="hlt">droplets</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Langen, Tim; Wenzel, Matthias; Schmitt, Matthias; Boettcher, Fabian; Buehner, Carl; Ferrier-Barbut, Igor; Pfau, Tilman</p> <p>2017-04-01</p> <p>Self-bound many-body systems are formed through a balance of attractive and repulsive forces and occur in many physical scenarios. Liquid <span class="hlt">droplets</span> are an example of a self-bound system, formed by a balance of the mutual attractive and repulsive forces that derive from different components of the inter-particle potential. On the basis of the recent finding that an unstable bosonic dipolar gas can be stabilized by a repulsive many-body term, it was predicted that three-dimensional self-bound quantum <span class="hlt">droplets</span> of magnetic atoms should exist. Here we report on the observation of such <span class="hlt">droplets</span> using dysprosium atoms, with densities 108 times lower than a helium <span class="hlt">droplet</span>, in a trap-free levitation field. We find that this dilute magnetic quantum liquid requires a minimum, critical number of atoms, below which the liquid evaporates into an expanding gas as a result of the quantum pressure of the individual constituents. Consequently, around this critical atom number we observe an interaction-driven phase transition between a gas and a self-bound liquid in the quantum degenerate regime with ultracold atoms.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27309943','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27309943"><span>Effects of copolymer composition, film thickness, and <span class="hlt">solvent</span> vapor annealing time on dewetting of ultrathin block copolymer films.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Huang, Changchun; Wen, Gangyao; Li, Jingdan; Wu, Tao; Wang, Lina; Xue, Feifei; Li, Hongfei; Shi, Tongfei</p> <p>2016-09-15</p> <p>Effects of copolymer composition, film thickness, and <span class="hlt">solvent</span> vapor annealing time on dewetting of spin-coated polystyrene-block-poly(methyl methacrylate) (PS-b-PMMA) films (<20nm thick) were mainly investigated by atomic force microscopy. Surface chemical analysis of the ultrathin films annealed for different times were performed using X-ray photoelectron spectroscopy and contact angle measurement. With the annealing of acetone vapor, dewetting of the films with different thicknesses occur via the spinodal dewetting and the nucleation and growth mechanisms, respectively. The PS-b-PMMA films rupture into <span class="hlt">droplets</span> which first coalesce into large ones to reduce the surface free energy. Then the large <span class="hlt">droplets</span> rupture into small ones to increase the contact area between PMMA blocks and acetone molecules resulting from ultimate migration of PMMA blocks to <span class="hlt">droplet</span> surface, which is a novel dewetting process observed in spin-coated films for the first time. Copyright © 2016 Elsevier Inc. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27126222','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27126222"><span>Integrated <span class="hlt">Droplet</span>-Based Microextraction with ESI-MS for Removal of Matrix Interference in Single-Cell Analysis.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Xiao-Chao; Wei, Zhen-Wei; Gong, Xiao-Yun; Si, Xing-Yu; Zhao, Yao-Yao; Yang, Cheng-Dui; Zhang, Si-Chun; Zhang, Xin-Rong</p> <p>2016-04-29</p> <p>Integrating <span class="hlt">droplet</span>-based microfluidics with mass spectrometry is essential to high-throughput and multiple analysis of single cells. Nevertheless, matrix effects such as the interference of culture medium and intracellular components influence the sensitivity and the accuracy of results in single-cell analysis. To resolve this problem, we developed a method that integrated <span class="hlt">droplet</span>-based microextraction with single-cell mass spectrometry. Specific extraction <span class="hlt">solvent</span> was used to selectively obtain intracellular components of interest and remove interference of other components. Using this method, UDP-Glc-NAc, GSH, GSSG, AMP, ADP and ATP were successfully detected in single MCF-7 cells. We also applied the method to study the change of unicellular metabolites in the biological process of dysfunctional oxidative phosphorylation. The method could not only realize matrix-free, selective and sensitive detection of metabolites in single cells, but also have the capability for reliable and high-throughput single-cell analysis.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010PhDT.........4G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010PhDT.........4G"><span>Turbulent dispersion of slightly buoyant oil <span class="hlt">droplets</span> and turbulent breakup of crude oil <span class="hlt">droplets</span> mixed with dispersants</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Gopalan, Balaji</p> <p></p> <p>In part I, high speed in-line digital holographic cinematography is used for studying turbulent diffusion of slightly buoyant 0.5-1.2 mm diameter diesel <span class="hlt">droplets</span> (specific gravity of 0.85) and 50 mum diameter neutral density particles. Experiments are performed in a 50x50x70 mm3 sample volume in a controlled, nearly isotropic turbulence facility, which is characterized by 2-D PIV. An automated tracking program has been used for measuring velocity time history of more than 17000 <span class="hlt">droplets</span> and 15000 particles. The PDF's of <span class="hlt">droplet</span> velocity fluctuations are close to Gaussian for all turbulent intensities ( u'i ). The mean rise velocity of <span class="hlt">droplets</span> is enhanced or suppressed, compared to quiescent rise velocity (Uq), depending on Stokes number at lower turbulence levels, but becomes unconditionally enhanced at higher turbulence levels. The horizontal <span class="hlt">droplet</span> velocity rms exceeds the fluid velocity rms for most of the data, while the vertical ones are higher than the fluid only at the highest turbulence level. The scaled <span class="hlt">droplet</span> horizontal diffusion coefficient is higher than the vertical one, for 1 < u'i /Uq < 5, consistent with trends of the <span class="hlt">droplet</span> velocity fluctuations. Conversely, the scaled <span class="hlt">droplet</span> horizontal diffusion timescale is smaller than the vertical one due to crossing trajectories effect. The <span class="hlt">droplet</span> diffusion coefficients scaled by the product of turbulence intensity and an integral length scale is a monotonically increasing function of u'i /Uq. Part II of this work explains the formation of micron sized <span class="hlt">droplets</span> in turbulent flows from crude oil <span class="hlt">droplets</span> pre-mixed with dispersants. Experimental visualization shows that this breakup starts with the formation of very long and quite stable, single or multiple micro threads that trail behind millimeter sized <span class="hlt">droplets</span>. These threads form in regions with localized increase in concentration of surfactant, which in turn depends on the flow around the <span class="hlt">droplet</span>. The resulting reduction of local surface tension</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27268751','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27268751"><span>Intramolecular, Exciplex-Mediated, Proton-Coupled, <span class="hlt">Charge</span>-Transfer Processes in N,N-Dimethyl-3-(1-pyrenyl)propan-1-ammonium Cations: Influence of Anion, <span class="hlt">Solvent</span> Polarity, and Temperature.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Safko, Trevor M; Faleiros, Marcelo M; Atvars, Teresa D Z; Weiss, Richard G</p> <p>2016-06-16</p> <p>An intramolecular exciplex-mediated, proton-coupled, <span class="hlt">charge</span>-transfer (PCCT) process has been investigated for a series of N,N-dimethyl-3-(1-pyrenyl)propan-1-ammonium cations with different anions (PyS) in <span class="hlt">solvents</span> of low to intermediate polarity over a wide temperature range. <span class="hlt">Solvent</span> mediates both the equilibrium between conformations of the cation that place the pyrenyl and ammonium groups in proximity (conformation C) or far from each other (conformation O) and the ability of the ammonium group to transfer a proton adiabatically in the PyS excited singlet state. Thus, exciplex emission, concurrent with the PCCT process, was observed only in hydrogen-bond accepting <span class="hlt">solvents</span> of relatively low polarity (tetrahydrofuran, ethyl acetate, and 1,4-dioxane) and not in dichloromethane. From the exciplex emission and other spectroscopic and thermodynamic data, the acidity of the ammonium group in conformation C of the excited singlet state of PyS (pKa*) has been estimated to be ca. -3.4 in tetrahydrofuran. The ratios between the intensities of emission from the exciplex and the locally excited state (IEx/ILE) appear to be much more dependent on the nature of the anion than are the rates of exciplex formation and decay, although the excited state data do not provide a quantitative measure of the anion effect on the C-O equilibrium. The activation energies associated with exciplex formation in THF are calculated to be 0.08 to 0.15 eV lower than for the neutral amine, N,N-dimethyl-3-(1-pyrenyl)propan-1-amine. Decay of the exciplexes formed from the deprotonation of PyS is hypothesized to occur through <span class="hlt">charge</span>-recombination processes. To our knowledge, this is the first example in which photoacidity and intramolecular exciplex formation (i.e., a PCCT reaction) are coupled.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1136634','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/1136634"><span>Chip-based <span class="hlt">droplet</span> sorting</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>Beer, Neil Reginald; Lee, Abraham; Hatch, Andrew</p> <p>2014-07-01</p> <p>A non-contact system for sorting monodisperse water-in-oil emulsion <span class="hlt">droplets</span> in a microfluidic device based on the <span class="hlt">droplet</span>'s contents and their interaction with an applied electromagnetic field or by identification and sorting.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4927099','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4927099"><span>Contribution of <span class="hlt">Charged</span> Groups to the Enthalpic Stabilization of the Folded States of Globular 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>Dadarlat, Voichita M.; Post, Carol Beth</p> <p>2016-01-01</p> <p>In this paper we use the results from all atom MD simulations of proteins and peptides to assess individual contribution of <span class="hlt">charged</span> atomic groups to the enthalpic stability of the native state of globular proteins and investigate how the distribution of <span class="hlt">charged</span> atomic groups in terms of <span class="hlt">solvent</span> accessibility relates to protein enthalpic stability. The contributions of <span class="hlt">charged</span> groups is calculated using a comparison of nonbonded interaction energy terms from equilibrium simulations of <span class="hlt">charged</span> amino acid dipeptides in water (the “unfolded state”) and <span class="hlt">charged</span> amino acids in globular proteins (the “folded state”). Contrary to expectation, the analysis shows that many buried, <span class="hlt">charged</span> atomic groups contribute favorably to protein enthalpic stability. The strongest enthalpic contributions favoring the folded state come from the carboxylate (COO−) groups of either Glu or Asp. The contributions from Arg guanidinium groups are generally somewhat stabilizing, while NH3+ groups from Lys contribute little toward stabilizing the folded state. The average enthalpic gain due to the transfer of a methyl group in an apolar amino acid from solution to the protein interior is described for comparison. Notably, <span class="hlt">charged</span> groups that are less exposed to <span class="hlt">solvent</span> contribute more favorably to protein native-state enthalpic stability than <span class="hlt">charged</span> groups that are <span class="hlt">solvent</span> exposed. While <span class="hlt">solvent</span> reorganization/release has favorable contributions to folding for all <span class="hlt">charged</span> atomic groups, the variation in folded state stability among proteins comes mainly from the change in the nonbonded interaction energy of <span class="hlt">charged</span> groups between the unfolded and folded states. A key outcome is that the calculated enthalpic stabilization is found to be inversely proportional to the excess <span class="hlt">charge</span> density on the surface, in support of an hypothesis proposed previously. PMID:18303881</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5003119','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5003119"><span>Electropermanent magnet actuation for <span class="hlt">droplet</span> ferromicrofluidics</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Padovani, José I.; Jeffrey, Stefanie S.; Howe, Roger T.</p> <p>2016-01-01</p> <p><span class="hlt">Droplet</span> actuation is an essential mechanism for <span class="hlt">droplet</span>-based microfluidic systems. On-demand electromagnetic actuation is used in a ferrofluid-based microfluidic system for water <span class="hlt">droplet</span> displacement. Electropermanent magnets (EPMs) are used to induce 50 mT magnetic fields in a ferrofluid filled microchannel with gradients up to 6.4 × 104 kA/m2. Short 50 µs current pulses activate the electropermanent magnets and generate negative magnetophoretic forces that range from 10 to 70 nN on 40 to 80 µm water-in-ferrofluid <span class="hlt">droplets</span>. Maximum <span class="hlt">droplet</span> displacement velocities of up to 300 µm/s are obtained under flow and no-flow conditions. Electropermanent magnet-activated <span class="hlt">droplet</span> sorting under continuous flow is demonstrated using a split-junction microfluidic design. PMID:27583301</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhRvE..95c3114W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhRvE..95c3114W"><span>Sintering of polydisperse viscous <span class="hlt">droplets</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wadsworth, Fabian B.; Vasseur, Jérémie; Llewellin, Edward W.; Dingwell, Donald B.</p> <p>2017-03-01</p> <p>Sintering—or coalescence—of compacts of viscous <span class="hlt">droplets</span> is driven by the interfacial tension between the <span class="hlt">droplets</span> and the interstitial gas phase. The process, which occurs in a range of industrial and natural settings, such as the manufacture of ceramics and the welding of volcanic ash, causes the compact to densify, to become stronger, and to become less permeable. We investigate the role of <span class="hlt">droplet</span> polydispersivity in sintering dynamics by conducting experiments in which populations of glass spheres with different size distributions are heated to temperatures above the glass transition interval. We quantify the progress of sintering by tracking changes in porosity with time. The sintering dynamics is modeled by treating the system as a random distribution of interstitial gas bubbles shrinking under the action of interfacial tension only. We identify the scaling between the polydispersivity of the initial <span class="hlt">droplets</span> and the dynamics of bulk densification. The framework that we develop allows the sintering dynamics of arbitrary polydisperse populations of <span class="hlt">droplets</span> to be predicted if the initial <span class="hlt">droplet</span> (or particle) size distribution is known.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2002APS..DFD.FC009N','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2002APS..DFD.FC009N"><span>Yeast <span class="hlt">Droplets</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Nguyen, Baochi; Upadhyaya, Arpita; van Oudenaarden, Alexander; Brenner, Michael</p> <p>2002-11-01</p> <p>It is well known that the Young's law and surface tension govern the shape of liquid <span class="hlt">droplets</span> on solid surfaces. Here we address through experiments and theory the shape of growing aggregates of yeast on agar substrates, and assess whether these ideas still hold. Experiments are carried out on Baker's yeast, with different levels of expressions of an adhesive protein governing cell-cell and cell-substrate adhesion. Changing either the agar concentration or the expression of this protein modifies the local contact angle of a yeast <span class="hlt">droplet</span>. When the colony is small, the shape is a spherical cap with the contact angle obeying Young's law. However, above a critical volume this structure is unstable, and the <span class="hlt">droplet</span> becomes nonspherical. We present a theoretical model where this instability is caused by bulk elastic effects. The model predicts that the transition depends on both volume and contact angle, in a manner quantitatively consistent with our experiments.</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/2017Cryo...88...78K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017Cryo...88...78K"><span>Pulsed beam of extremely large helium <span class="hlt">droplets</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kuma, Susumu; Azuma, Toshiyuki</p> <p>2017-12-01</p> <p>We generated a pulsed helium <span class="hlt">droplet</span> beam with average <span class="hlt">droplet</span> diameters of up to 2 μ m using a solenoid pulsed valve operated at temperatures as low as 7 K. The <span class="hlt">droplet</span> diameter was controllable over two orders of magnitude, or six orders of the number of atoms per <span class="hlt">droplet</span>, by lowering the valve temperature from 21 to 7 K. A sudden <span class="hlt">droplet</span> size change attributed to the so-called ;supercritical expansion; was firstly observed in pulsed mode, which is necessary to obtain the micrometer-scale <span class="hlt">droplets</span>. This beam source is beneficial for experiments that require extremely large helium <span class="hlt">droplets</span> in intense, pulsed form.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22323144','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22323144"><span><span class="hlt">Droplet</span> microfluidics for single-cell analysis.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Brouzes, Eric</p> <p>2012-01-01</p> <p>This book chapter aims at providing an overview of all the aspects and procedures needed to develop a <span class="hlt">droplet</span>-based workflow for single-cell analysis (see Fig. 10.1). The surfactant system used to stabilize <span class="hlt">droplets</span> is a critical component of <span class="hlt">droplet</span> microfluidics; its properties define the type of <span class="hlt">droplet</span>-based assays and workflows that can be developed. The scope of this book chapter is limited to fluorinated surfactant systems that have proved to generate extremely stable <span class="hlt">droplets</span> and allow to easily retrieve the encapsulated material. The formulation section discusses how the experimental parameters influence the choice of the surfactant system to use. The circuit design section presents recipes to design and integrate different <span class="hlt">droplet</span> modules into a whole assay. The fabrication section describes the manufacturing of microfluidic chip including the surface treatment which is pivotal in <span class="hlt">droplet</span> microfluidics. Finally, the last section reviews the experimental setup for fluorescence detection with an emphasis on cell injection and incubation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3467560','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3467560"><span>Enhanced <span class="hlt">Droplet</span> Control by Transition Boiling</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Grounds, Alex; Still, Richard; Takashina, Kei</p> <p>2012-01-01</p> <p>A <span class="hlt">droplet</span> of water on a heated surface can levitate over a film of gas produced by its own evaporation in the Leidenfrost effect. When the surface is prepared with ratchet-like saw-teeth topography, these <span class="hlt">droplets</span> can self-propel and can even climb uphill. However, the extent to which the <span class="hlt">droplets</span> can be controlled is limited by the physics of the Leidenfrost effect. Here, we show that transition boiling can be induced even at very high surface temperatures and provide additional control over the <span class="hlt">droplets</span>. Ratchets with acute protrusions enable <span class="hlt">droplets</span> to climb steeper inclines while ratchets with sub-structures enable their direction of motion to be controlled by varying the temperature of the surface. The <span class="hlt">droplets</span>' departure from the Leidenfrost regime is assessed by analysing the sound produced by their boiling. We anticipate these techniques will enable the development of more sophisticated methods for controlling small <span class="hlt">droplets</span> and heat transfer. PMID:23056912</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012NatSR...2E.720G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012NatSR...2E.720G"><span>Enhanced <span class="hlt">Droplet</span> Control by Transition Boiling</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Grounds, Alex; Still, Richard; Takashina, Kei</p> <p>2012-10-01</p> <p>A <span class="hlt">droplet</span> of water on a heated surface can levitate over a film of gas produced by its own evaporation in the Leidenfrost effect. When the surface is prepared with ratchet-like saw-teeth topography, these <span class="hlt">droplets</span> can self-propel and can even climb uphill. However, the extent to which the <span class="hlt">droplets</span> can be controlled is limited by the physics of the Leidenfrost effect. Here, we show that transition boiling can be induced even at very high surface temperatures and provide additional control over the <span class="hlt">droplets</span>. Ratchets with acute protrusions enable <span class="hlt">droplets</span> to climb steeper inclines while ratchets with sub-structures enable their direction of motion to be controlled by varying the temperature of the surface. The <span class="hlt">droplets</span>' departure from the Leidenfrost regime is assessed by analysing the sound produced by their boiling. We anticipate these techniques will enable the development of more sophisticated methods for controlling small <span class="hlt">droplets</span> and heat transfer.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ShWav..28..599B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ShWav..28..599B"><span>Experimental study of detonation of large-scale powder-<span class="hlt">droplet</span>-vapor mixtures</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Bai, C.-H.; Wang, Y.; Xue, K.; Wang, L.-F.</p> <p>2018-05-01</p> <p>Large-scale experiments were carried out to investigate the detonation performance of a 1600-m3 ternary cloud consisting of aluminum powder, fuel <span class="hlt">droplets</span>, and vapor, which were dispersed by a central explosive in a cylindrically stratified configuration. High-frame-rate video cameras and pressure gauges were used to analyze the large-scale explosive dispersal of the mixture and the ensuing blast wave generated by the detonation of the cloud. Special attention was focused on the effect of the descending motion of the <span class="hlt">charge</span> on the detonation performance of the dispersed ternary cloud. The <span class="hlt">charge</span> was parachuted by an ensemble of apparatus from the designated height in order to achieve the required terminal velocity when the central explosive was detonated. A descending <span class="hlt">charge</span> with a terminal velocity of 32 m/s produced a cloud with discernably increased concentration compared with that dispersed from a stationary <span class="hlt">charge</span>, the detonation of which hence generates a significantly enhanced blast wave beyond the scaled distance of 6 m/kg^{1/3}. The results also show the influence of the descending motion of the <span class="hlt">charge</span> on the jetting phenomenon and the distorted shock front.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MicST..30..143C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MicST..30..143C"><span>Spreading of Annular <span class="hlt">Droplets</span> on a Horizontal Fiber</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chen, Xue; Ding, Zijing; Liu, Rong</p> <p>2018-05-01</p> <p>This paper investigates an annular <span class="hlt">droplet</span> on a horizontal fiber. The static state and the dynamic spreading process of the <span class="hlt">droplet</span> is analyzed. A full model describing the profile of a static <span class="hlt">droplet</span> is derived from the energy variation principle. To study the dynamical spreading of the <span class="hlt">droplet</span>, we derive a lubrication model which is verified by the full model. It indicates that the lubrication model is valid for a thin <span class="hlt">droplet</span>. Results of the static <span class="hlt">droplet</span> reveal that, when the fiber radius is very small, the <span class="hlt">droplet</span> tends to have a spherical shape; if the fiber radius is very large, the <span class="hlt">droplet</span> approaches to a parabolic profile. Furthermore, the time-evolution study is carried out to investigate the dynamical spreading of the <span class="hlt">droplet</span>. It is highlighted that when the fiber radius is small, the <span class="hlt">droplet</span> can breakup into small <span class="hlt">droplets</span> or contract into a sharp shape. For a large fiber radius, the <span class="hlt">droplet</span> spreads to a steady profile. In addition, the liquid viscosity is found to retard the deformation of the <span class="hlt">droplet</span> and the motion of the contact lines.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017MicST.tmp...62C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017MicST.tmp...62C"><span>Spreading of Annular <span class="hlt">Droplets</span> on a Horizontal Fiber</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chen, Xue; Ding, Zijing; Liu, Rong</p> <p>2017-12-01</p> <p>This paper investigates an annular <span class="hlt">droplet</span> on a horizontal fiber. The static state and the dynamic spreading process of the <span class="hlt">droplet</span> is analyzed. A full model describing the profile of a static <span class="hlt">droplet</span> is derived from the energy variation principle. To study the dynamical spreading of the <span class="hlt">droplet</span>, we derive a lubrication model which is verified by the full model. It indicates that the lubrication model is valid for a thin <span class="hlt">droplet</span>. Results of the static <span class="hlt">droplet</span> reveal that, when the fiber radius is very small, the <span class="hlt">droplet</span> tends to have a spherical shape; if the fiber radius is very large, the <span class="hlt">droplet</span> approaches to a parabolic profile. Furthermore, the time-evolution study is carried out to investigate the dynamical spreading of the <span class="hlt">droplet</span>. It is highlighted that when the fiber radius is small, the <span class="hlt">droplet</span> can breakup into small <span class="hlt">droplets</span> or contract into a sharp shape. For a large fiber radius, the <span class="hlt">droplet</span> spreads to a steady profile. In addition, the liquid viscosity is found to retard the deformation of the <span class="hlt">droplet</span> and the motion of the contact lines.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhFl...29l2105H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhFl...29l2105H"><span>Effect of surface roughness on <span class="hlt">droplet</span> splashing</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hao, Jiguang</p> <p>2017-12-01</p> <p>It is well known that rough surfaces trigger prompt splashing and suppress corona splashing on <span class="hlt">droplet</span> impact. Upon water <span class="hlt">droplet</span> impact, we experimentally found that a slightly rough substrate triggers corona splashing which is suppressed to prompt splashing by both further increase and further decrease of surface roughness. The nonmonotonic effect of surface roughness on corona splashing weakens with decreasing <span class="hlt">droplet</span> surface tension. The threshold velocities for prompt splashing and corona splashing are quantified under different conditions including surface roughness, <span class="hlt">droplet</span> diameter, and <span class="hlt">droplet</span> surface tension. It is determined that slight roughness significantly enhances both prompt splashing and corona splashing of a water <span class="hlt">droplet</span>, whereas it weakly affects low-surface-tension <span class="hlt">droplet</span> splashing. Consistent with previous studies, high roughness triggers prompt splashing and suppresses corona splashing. Further experiments on <span class="hlt">droplet</span> spreading propose that the mechanism of slight roughness enhancing water <span class="hlt">droplet</span> splashing is due to the decrease of the wetted area with increasing surface roughness.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19990054031&hterms=Acoustic+levitation&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3DAcoustic%2Blevitation','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19990054031&hterms=Acoustic+levitation&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3DAcoustic%2Blevitation"><span>Formation and Levitation of Unconfined <span class="hlt">Droplet</span> Clusters</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Liu, S.; Ruff, G. A.</p> <p>1999-01-01</p> <p>Combustion experiments using arrays of <span class="hlt">droplets</span> seek to provide a link between single <span class="hlt">droplet</span> combustion phenomena and the behavior of complex spray combustion systems. Both single <span class="hlt">droplet</span> and <span class="hlt">droplet</span> array studies have been conducted in microgravity to better isolate the <span class="hlt">droplet</span> interaction phenomena and eliminate or reduce the confounding effects of buoyancy-induced convection. In most experiments involving <span class="hlt">droplet</span> arrays, the <span class="hlt">droplets</span> are supported on fibers to keep them stationary and close together before the combustion event. The presence of the fiber, however, disturbs the combustion process by introducing a source of heat transfer and asymmetry into the configuration. As the number of drops in a <span class="hlt">droplet</span> array increases, supporting the drops on fibers becomes less practical because of the cumulative effect of the fibers on the combustion process. The overall objective of this research is to study the combustion of well-characterized drop clusters in a microgravity environment. Direct experimental observations and measurements of the combustion of <span class="hlt">droplet</span> clusters would fill a large gap in our current understanding of <span class="hlt">droplet</span> and spray combustion and provide unique experimental data for the verification and improvement of spray combustion models. This paper describes current work on the design and performance of an apparatus to generate and stabilize <span class="hlt">droplet</span> clusters using acoustic and electrostatic forces.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018MolPh.116.1003C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018MolPh.116.1003C"><span>3DRISM-HI-D2MSA: an improved analytic theory to compute <span class="hlt">solvent</span> structure around hydrophobic solutes with proper treatment of solute–<span class="hlt">solvent</span> electrostatic interactions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cao, Siqin; Zhu, Lizhe; Huang, Xuhui</p> <p>2018-04-01</p> <p>The 3D reference interaction site model (3DRISM) is a powerful tool to study the thermodynamic and structural properties of liquids. However, for hydrophobic solutes, the inhomogeneity of the <span class="hlt">solvent</span> density around them poses a great challenge to the 3DRISM theory. To address this issue, we have previously introduced the hydrophobic-induced density inhomogeneity theory (HI) for purely hydrophobic solutes. To further consider the complex hydrophobic solutes containing partial <span class="hlt">charges</span>, here we propose the D2MSA closure to incorporate the short-range and long-range interactions with the D2 closure and the mean spherical approximation, respectively. We demonstrate that our new theory can compute the <span class="hlt">solvent</span> distributions around real hydrophobic solutes in water and complex organic <span class="hlt">solvents</span> that agree well with the explicit <span class="hlt">solvent</span> molecular dynamics simulations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AcSpA.188..252G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AcSpA.188..252G"><span>Photophysics of a coumarin based Schiff base in <span class="hlt">solvents</span> of varying polarities</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ghosh, Saptarshi; Roy, Nayan; Singh, T. Sanjoy; Chattopadhyay, Nitin</p> <p>2018-01-01</p> <p>The present work reports detailed photophysics of a coumarin based Schiff base, namely, (E)-7-(((8-hydroxyquinolin-2-yl)methylene)amino)-4-methyl-2H-chromen-2-one (HMC) in different <span class="hlt">solvents</span> of varying polarity exploiting steady state absorption, fluorescence and time resolved fluorescence spectroscopy. The dominant photophysical features of HMC are discussed in terms of emission from an intramolecular <span class="hlt">charge</span> transfer (ICT) excited state. Molecular orbital (MO) diagrams as obtained from DFT based computational analysis confirms the occurrence of <span class="hlt">charge</span> transfer from 8‧-hydroxy quinoline moiety of the molecule to the coumarin part. The notable difference in the photophysical response of HMC from its analogous coumarin (C480) lies in a lower magnitude of fluorescence quantum yield of the former, particularly in the <span class="hlt">solvents</span> of low polarity, which is rationalized by considering the higher rate of non-radiative decay of HMC in apolar <span class="hlt">solvents</span>. Phosphorescence emission as well as phosphorescence lifetime of HMC has also been reported in 77 K frozen matrix.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20070035073','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20070035073"><span>Pyrolysis of Large Black Liquor <span class="hlt">Droplets</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Bartkus, Tadas P.; Dietrich, Daniel L.; T'ien, James S.; Wessel, Richard A.</p> <p>2007-01-01</p> <p>This paper presents the results of experiments involving the pyrolysis of large black liquor <span class="hlt">droplets</span> in the NASA KC-135 reduced gravity aircraft. The reduced gravity environment facilitated the study of <span class="hlt">droplets</span> up to 9 mm in diameter extending the results of previous studies to <span class="hlt">droplet</span> sizes that are similar to those encountered in recovery boilers. Single black liquor <span class="hlt">droplets</span> were rapidly inserted into a 923 K oven. The primary independent variables were the initial <span class="hlt">droplet</span> diameter (0.5 mm to 9 mm), the black liquor solids content (66.12% - 72.9% by mass), and the ambient oxygen mole fraction (0.0 - 0.21). Video records of the experiments provided size and shape of the <span class="hlt">droplets</span> as a function of time. The results show that the particle diameter at the end of the drying stage (D(sub DRY)) increases linearly with the initial particle diameter (D(sub O)). The results further show that the ratio of the maximum swollen diameter (D(sub MAX)) to D(sub O) decreases with increasing D(sub O) for <span class="hlt">droplets</span> with D(sub O) less than 4 mm. This ratio was independent of D(sub O) for <span class="hlt">droplets</span> with D(sub O) greater than 4 mm. The particle is most spherical at the end of drying, and least spherical at maximum swollen size, regardless of initial sphericity and <span class="hlt">droplet</span> size.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20070026246','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20070026246"><span>Pyrolysis of Large Black Liquor <span class="hlt">Droplets</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Bartkus, Tadas P.; T'ien, James S.; Dietrich, Daniel L.; Wessel, Richard A.</p> <p>2007-01-01</p> <p>This paper presents the results of experiments involving the pyrolysis of large black liquor <span class="hlt">droplets</span> in the NASA KC-135 reduced gravity aircraft. The reduced gravity environment facilitated the study of <span class="hlt">droplets</span> up to 9 mm in diameter extending the results of previous studies to <span class="hlt">droplet</span> sizes that are similar to those encountered in recovery boilers. Single black liquor <span class="hlt">droplets</span> were rapidly inserted into a 923 K oven. The primary independent variables were the initial <span class="hlt">droplet</span> diameter (0.5 mm to 9 mm), the black liquor solids content (66.12% - 72.9% by mass), and the ambient oxygen mole fraction (0.0 - 0.21). Video records of the experiments provided size and shape of the <span class="hlt">droplets</span> as a function of time. The results show that the particle diameter at the end of the drying stage (D(sub DRY) ) increases linearly with the initial particle diameter (D(sub O)). The results further show that the ratio of the maximum swollen diameter (D(sub MAX)) to D(sub O) decreases with increasing D(sub O) for <span class="hlt">droplets</span> with D(sub O) less than 4 mm. This ratio was independent of D(sub O) for <span class="hlt">droplets</span> with D(sub O) greater than 4 mm. The particle is most spherical at the end of drying, and least spherical at maximum swollen size, regardless of initial sphericity and <span class="hlt">droplet</span> size.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25641162','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25641162"><span>Molecular dynamics investigation of the ionic liquid/enzyme interface: application to engineering enzyme surface <span class="hlt">charge</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Burney, Patrick R; Nordwald, Erik M; Hickman, Katie; Kaar, Joel L; Pfaendtner, Jim</p> <p>2015-04-01</p> <p>Molecular simulations of the enzymes Candida rugosa lipase and Bos taurus α-chymotrypsin in aqueous ionic liquids 1-butyl-3-methylimidazolium chloride and 1-ethyl-3-methylimidazolium ethyl sulfate were used to study the change in enzyme-<span class="hlt">solvent</span> interactions induced by modification of the enzyme surface <span class="hlt">charge</span>. The enzymes were altered by randomly mutating lysine surface residues to glutamate, effectively decreasing the net surface <span class="hlt">charge</span> by two for each mutation. These mutations resemble succinylation of the enzyme by chemical modification, which has been shown to enhance the stability of both enzymes in ILs. After establishing that the enzymes were stable on the simulated time scales, we focused the analysis on the organization of the ionic liquid substituents about the enzyme surface. Calculated <span class="hlt">solvent</span> <span class="hlt">charge</span> densities show that for both enzymes and in both <span class="hlt">solvents</span> that changing positively <span class="hlt">charged</span> residues to negative <span class="hlt">charge</span> does indeed increase the <span class="hlt">charge</span> density of the <span class="hlt">solvent</span> near the enzyme surface. The radial distribution of IL constituents with respect to the enzyme reveals decreased interactions with the anion are prevalent in the modified systems when compared to the wild type, which is largely accompanied by an increase in cation contact. Additionally, the radial dependence of the <span class="hlt">charge</span> density and ion distribution indicates that the effect of altering enzyme <span class="hlt">charge</span> is confined to short range (≤1 nm) ordering of the IL. Ultimately, these results, which are consistent with that from prior experiments, provide molecular insight into the effect of enzyme surface <span class="hlt">charge</span> on enzyme stability in ILs. © 2015 Wiley Periodicals, Inc.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26196035','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26196035"><span>On-chip dilution in nanoliter <span class="hlt">droplets</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Thakur, Raviraj; Amin, Ahmed M; Wereley, Steve</p> <p>2015-09-07</p> <p><span class="hlt">Droplet</span> microfluidics is enabling reactions at nano- and picoliter scale, resulting in faster and cheaper biological and chemical analyses. However, varying concentrations of samples on a drop-to-drop basis is still a challenging task in <span class="hlt">droplet</span> microfluidics, primarily limited due to lack of control over individual <span class="hlt">droplets</span>. In this paper, we report an on-chip microfluidic <span class="hlt">droplet</span> dilution strategy using three-valve peristaltic pumps.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1996JChPh.10511335V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1996JChPh.10511335V"><span>Stretch-collapse transition of polyelectrolyte brushes in a poor <span class="hlt">solvent</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 Goeler, F.; Muthukumar, M.</p> <p>1996-12-01</p> <p>This paper describes the behavior of <span class="hlt">charged</span>, polymer brushes in electrolyte solutions of varying <span class="hlt">solvent</span> quality. The brush height, d, dependence on the chain length, L (=Nl, where l is the Kuhn length), the grafting density σ, and <span class="hlt">solvent</span> conditions is determined. We consider a monomer-monomer potential consisting of three components: (1) a long-ranged, screened Coulombic component of strength v¯/l (l is the Kuhn length) and range κ-1; (2) a short-ranged, two-body component of strength w¯l; and (3) a short-ranged, three-body component of strength ūl3. In particular, we examine the transition from a stretched state to a collapsed state in a poor <span class="hlt">solvent</span> (w¯<0) as the <span class="hlt">solvent</span> quality is decreased. Using dimensional analysis, Monte Carlo methods, and a variational technique, a first order transition is observed as predicted by the scaling arguments of Ross et al. and Borisov et al. for high <span class="hlt">charge</span>/grafting densities. Using a variational procedure, we derive an analytical expression for the brush size and determine, quantitatively, the critical conditions for a first order transition in terms of key dimensionless variables, vN5/2, κlN1/2, wN3/2, and uN2 (where v=2πσl2v¯, w=σl2w¯, and u=σ2l4ū).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011JCAMD..25..477S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011JCAMD..25..477S"><span>Docking glycosaminoglycans to proteins: analysis of <span class="hlt">solvent</span> inclusion</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Samsonov, Sergey A.; Teyra, Joan; Pisabarro, M. Teresa</p> <p>2011-05-01</p> <p>Glycosaminoglycans (GAGs) are anionic polysaccharides, which participate in key processes in the extracellular matrix by interactions with protein targets. Due to their <span class="hlt">charged</span> nature, accurate consideration of electrostatic and water-mediated interactions is indispensable for understanding GAGs binding properties. However, <span class="hlt">solvent</span> is often overlooked in molecular recognition studies. Here we analyze the abundance of <span class="hlt">solvent</span> in GAG-protein interfaces and investigate the challenges of adding explicit <span class="hlt">solvent</span> in GAG-protein docking experiments. We observe PDB GAG-protein interfaces being significantly more hydrated than protein-protein interfaces. Furthermore, by applying molecular dynamics approaches we estimate that about half of GAG-protein interactions are water-mediated. With a dataset of eleven GAG-protein complexes we analyze how <span class="hlt">solvent</span> inclusion affects Autodock 3, eHiTs, MOE and FlexX docking. We develop an approach to de novo place explicit <span class="hlt">solvent</span> into the binding site prior to docking, which uses the GRID program to predict positions of waters and to locate possible areas of <span class="hlt">solvent</span> displacement upon ligand binding. To investigate how <span class="hlt">solvent</span> placement affects docking performance, we compare these results with those obtained by taking into account information about the <span class="hlt">solvent</span> position in the crystal structure. In general, we observe that inclusion of <span class="hlt">solvent</span> improves the results obtained with these methods. Our data show that Autodock 3 performs best, though it experiences difficulties to quantitatively reproduce experimental data on specificity of heparin/heparan sulfate disaccharides binding to IL-8. Our work highlights the current challenges of introducing <span class="hlt">solvent</span> in protein-GAGs recognition studies, which is crucial for exploiting the full potential of these molecules for rational engineering.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3651735','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3651735"><span>Influence of <span class="hlt">droplet</span> size, pH and ionic strength on endotoxin-triggered ordering transitions in liquid crystalline <span class="hlt">droplets</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>Miller, Daniel S.; Abbott, Nicholas L.</p> <p>2012-01-01</p> <p>We report an investigation of ordering transitions that are induced in water-dispersed, micrometer-sized <span class="hlt">droplets</span> of a thermotropic liquid crystal (LC) by the bacterial lipopolysaccharide endotoxin. We reveal that the ordering transitions induced by endotoxin – from a bipolar state of the <span class="hlt">droplets</span> to a radial state – are strongly dependent on the size of the LC <span class="hlt">droplets</span>. Specifically, as the diameters of the LC <span class="hlt">droplets</span> increase from 2 μm to above 10 μm (in phosphate buffered saline with an ionic strength of 90 mM and a pH of 7.2), we measured the percentage of <span class="hlt">droplets</span> exhibiting a radial configuration in the presence of 100 pg/mL endotoxin to decrease from 98 ± 1 % to 3 ± 2 %. In addition, we measured a decrease in either the ionic strength or pH of the aqueous phase to reduce the percentage of <span class="hlt">droplets</span> exhibiting a radial configuration in the presence of endotoxin. These results, when interpreted within the context of a simple thermodynamic model that incorporates the contributions of elasticity and surface anchoring to the free energies of the LC <span class="hlt">droplets</span>, lead us to conclude that (i) the elastic constant K24 plays a central role in determining the size-dependent response of the LC <span class="hlt">droplets</span> to endotoxin, and (ii) endotoxin-triggered ordering transitions occur only under solution conditions (pH, ionic strength) where the combined contributions of elasticity and surface anchoring to the free energies of the bipolar and radial configurations of the LC <span class="hlt">droplets</span> are similar in magnitude. Our analysis also suggests that the presence of endotoxin perturbs the free energies of the LC <span class="hlt">droplets</span> by ~10−17 J/<span class="hlt">droplet</span>, which is comparable to the standard free energy of self-association of ~103 endotoxin molecules. These results, when combined with prior reports of localization of endotoxin at the center of LC <span class="hlt">droplets</span>, are consistent with the hypothesis that self-assembly of endotoxin within micrometer-sized LC <span class="hlt">droplets</span> provides the driving force for the ordering</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26247820','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26247820"><span>A programmable microfluidic static <span class="hlt">droplet</span> array for <span class="hlt">droplet</span> generation, transportation, fusion, storage, and retrieval.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jin, Si Hyung; Jeong, Heon-Ho; Lee, Byungjin; Lee, Sung Sik; Lee, Chang-Soo</p> <p>2015-01-01</p> <p>We present a programmable microfluidic static <span class="hlt">droplet</span> array (SDA) device that can perform user-defined multistep combinatorial protocols. It combines the passive storage of aqueous <span class="hlt">droplets</span> without any external control with integrated microvalves for discrete sample dispensing and dispersion-free unit operation. The addressable picoliter-volume reaction is systematically achieved by consecutively merging programmable sequences of reagent <span class="hlt">droplets</span>. The SDA device is remarkably reusable and able to perform identical enzyme kinetic experiments at least 30 times via automated cross-contamination-free removal of <span class="hlt">droplets</span> from individual hydrodynamic traps. Taking all these features together, this programmable and reusable universal SDA device will be a general microfluidic platform that can be reprogrammed for multiple applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21627144','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21627144"><span>Dynamics of <span class="hlt">droplet</span> motion under electrowetting actuation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Annapragada, S Ravi; Dash, Susmita; Garimella, Suresh V; Murthy, Jayathi Y</p> <p>2011-07-05</p> <p>The static shape of <span class="hlt">droplets</span> under electrowetting actuation is well understood. The steady-state shape of the <span class="hlt">droplet</span> is obtained on the basis of the balance of surface tension and electrowetting forces, and the change in the apparent contact angle is well characterized by the Young-Lippmann equation. However, the transient <span class="hlt">droplet</span> shape behavior when a voltage is suddenly applied across a <span class="hlt">droplet</span> has received less attention. Additional dynamic frictional forces are at play during this transient process. We present a model to predict this transient behavior of the <span class="hlt">droplet</span> shape under electrowetting actuation. The <span class="hlt">droplet</span> shape is modeled using the volume of fluid method. The electrowetting and dynamic frictional forces are included as an effective dynamic contact angle through a force balance at the contact line. The model is used to predict the transient behavior of water <span class="hlt">droplets</span> on smooth hydrophobic surfaces under electrowetting actuation. The predictions of the transient behavior of <span class="hlt">droplet</span> shape and contact radius are in excellent agreement with our experimental measurements. The internal fluid motion is explained, and the <span class="hlt">droplet</span> motion is shown to initiate from the contact line. An approximate mathematical model is also developed to understand the physics of the <span class="hlt">droplet</span> motion and to describe the overall <span class="hlt">droplet</span> motion and the contact line velocities. © 2011 American Chemical Society</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.ncbi.nlm.nih.gov/pubmed/17064095','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17064095"><span><span class="hlt">Solvent</span> effect on redox properties of hexanethiolate monolayer-protected gold nanoclusters.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Su, Bin; Zhang, Meiqin; Shao, Yuanhua; Girault, Hubert H</p> <p>2006-11-02</p> <p>The capacitance of monolayer-protected gold nanoclusters (MPCs), C(MPC), in solution has been theoretically reconsidered from an electrostatic viewpoint, in which an MPC is considered as an isolated <span class="hlt">charged</span> sphere within two dielectric layers, the intrinsic coating monolayer, and the bulk <span class="hlt">solvent</span>. The model predicts that the bulk <span class="hlt">solvent</span> provides an important contribution to C(MPC) and influences the redox properties of MPCs. This theoretical prediction is then examined experimentally by comparing the redox properties of MPCs in four organic <span class="hlt">solvents</span>: 1,2-dichloroethane (DCE), dichloromethane (DCM), chlorobenzene (CB), and toluene (TOL), in all of which MPCs have excellent solubility. Furthermore, this set of organic <span class="hlt">solvents</span> features a dielectric constant in a range from 10.37 (DCE) to 2.38 (TOL), which is wide enough to probe the <span class="hlt">solvent</span> effect. In these organic <span class="hlt">solvents</span>, tetrahexylammonium bis(trifluoromethylsulfonyl)imide (THATf2N) is used as the supporting electrolyte. Cyclic and differential pulse voltammetric results provide concrete evidence that, despite the monolayer protection, the <span class="hlt">solvent</span> plays a significant effect on the properties of MPCs in solution.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018NRL....13..108X','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018NRL....13..108X"><span>Oil Contact Angles in a Water-Decane-Silicon Dioxide System: Effects of Surface <span class="hlt">Charge</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Xu, Shijing; Wang, Jingyao; Wu, Jiazhong; Liu, Qingjie; Sun, Chengzhen; Bai, Bofeng</p> <p>2018-04-01</p> <p>Oil wettability in the water-oil-rock systems is very sensitive to the evolution of surface <span class="hlt">charges</span> on the rock surfaces induced by the adsorption of ions and other chemical agents in water flooding. Through a set of large-scale molecular dynamics simulations, we reveal the effects of surface <span class="hlt">charge</span> on the oil contact angles in an ideal water-decane-silicon dioxide system. The results show that the contact angles of oil nano-<span class="hlt">droplets</span> have a great dependence on the surface <span class="hlt">charges</span>. As the surface <span class="hlt">charge</span> density exceeds a critical value of 0.992 e/nm2, the contact angle reaches up to 78.8° and the water-wet state is very apparent. The variation of contact angles can be confirmed from the number density distributions of oil molecules. With increasing the surface <span class="hlt">charge</span> density, the adsorption of oil molecules weakens and the contact areas between nano-<span class="hlt">droplets</span> and silicon dioxide surface are reduced. In addition, the number density distributions, RDF distributions, and molecular orientations indicate that the oil molecules are adsorbed on the silicon dioxide surface layer-by-layer with an orientation parallel to the surface. However, the layered structure of oil molecules near the silicon dioxide surface becomes more and more obscure at higher surface <span class="hlt">charge</span> densities.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29675565','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29675565"><span>Oil Contact Angles in a Water-Decane-Silicon Dioxide System: Effects of Surface <span class="hlt">Charge</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Xu, Shijing; Wang, Jingyao; Wu, Jiazhong; Liu, Qingjie; Sun, Chengzhen; Bai, Bofeng</p> <p>2018-04-19</p> <p>Oil wettability in the water-oil-rock systems is very sensitive to the evolution of surface <span class="hlt">charges</span> on the rock surfaces induced by the adsorption of ions and other chemical agents in water flooding. Through a set of large-scale molecular dynamics simulations, we reveal the effects of surface <span class="hlt">charge</span> on the oil contact angles in an ideal water-decane-silicon dioxide system. The results show that the contact angles of oil nano-<span class="hlt">droplets</span> have a great dependence on the surface <span class="hlt">charges</span>. As the surface <span class="hlt">charge</span> density exceeds a critical value of 0.992 e/nm 2 , the contact angle reaches up to 78.8° and the water-wet state is very apparent. The variation of contact angles can be confirmed from the number density distributions of oil molecules. With increasing the surface <span class="hlt">charge</span> density, the adsorption of oil molecules weakens and the contact areas between nano-<span class="hlt">droplets</span> and silicon dioxide surface are reduced. In addition, the number density distributions, RDF distributions, and molecular orientations indicate that the oil molecules are adsorbed on the silicon dioxide surface layer-by-layer with an orientation parallel to the surface. However, the layered structure of oil molecules near the silicon dioxide surface becomes more and more obscure at higher surface <span class="hlt">charge</span> densities.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=MSFC-0300179&hterms=Shrink&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3DShrink','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=MSFC-0300179&hterms=Shrink&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3DShrink"><span>Two <span class="hlt">Droplets</span> on Wire Approaching Ignition</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>The Fiber-Supported <span class="hlt">Droplet</span> Combustion (FSDC) uses two <span class="hlt">droplets</span> positioned on the fiber wire, instead of the usual one. Two <span class="hlt">droplets</span> more closely simulates the environment in engines, which ignite many fuel <span class="hlt">droplets</span> at once. The behavior of the burning was also unexpected -- the <span class="hlt">droplets</span> moved together after ignition, generating quite a bit of data for understanding the interaction of fuel <span class="hlt">droplets</span> while they burn. This MPEG movie (1.3 MB) shows a time-lapse of this burn (3x speed). Because FSDC is backlit (the bright glow behind the drops), you carnot see the glow of the <span class="hlt">droplets</span> while they burn -- instead, you see them shrink! The small blobs left on the wire after the burn are the beads used to center the fuel <span class="hlt">droplet</span> on the wire. This image was taken on STS-94, July 12, 1997, MET:10/19:13 (approximate). FSDC-2 studied fundamental phenomena related to liquid fuel <span class="hlt">droplet</span> combustion in air. Pure fuels and mixtures of fuels were burned as isolated single and dual <span class="hlt">droplets</span> with and without forced air convection. The FSDC guest investigator was Forman Williams, University of California, San Diego. The experiment was part of the space research investigations conducted during the Microgravity Science Laboratory-1R mission (STS-94, July 1-17 1997). Advanced combustion experiments will be a part of investigations planned for the International Space Station. (1.3MB, 12-second MPEG, screen 320 x 240 pixels; downlinked video, higher quality not available) A still JPG composite of this movie is available at http://mix.msfc.nasa.gov/ABSTRACTS/MSFC-0300178.html.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29776036','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29776036"><span>Entropic lattice Boltzmann model for <span class="hlt">charged</span> leaky dielectric multiphase fluids in electrified jets.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Lauricella, Marco; Melchionna, Simone; Montessori, Andrea; Pisignano, Dario; Pontrelli, Giuseppe; Succi, Sauro</p> <p>2018-03-01</p> <p>We present a lattice Boltzmann model for <span class="hlt">charged</span> leaky dielectric multiphase fluids in the context of electrified jet simulations, which are of interest for a number of production technologies including electrospinning. The role of nonlinear rheology on the dynamics of electrified jets is considered by exploiting the Carreau model for pseudoplastic fluids. We report exploratory simulations of <span class="hlt">charged</span> <span class="hlt">droplets</span> at rest and under a constant electric field, and we provide results for <span class="hlt">charged</span> jet formation under electrospinning conditions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015APS..DFDA32009B','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015APS..DFDA32009B"><span>Dancing <span class="hlt">droplets</span>: Contact angle, drag, and confinement</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Benusiglio, Adrien; Cira, Nate; Prakash, Manu</p> <p>2015-11-01</p> <p>When deposited on a clean glass slide, a mixture of water and propylene glycol forms a <span class="hlt">droplet</span> of given contact angle, when both pure liquids spread. (Cira, Benusiglio, Prakash: Nature, 2015). The <span class="hlt">droplet</span> is stabilized by a gradient of surface tension due to evaporation that induces a Marangoni flow from the border to the apex of the <span class="hlt">droplets</span>. The apparent contact angle of the <span class="hlt">droplets</span> depends on both their composition and the external humidity as captured by simple models. These <span class="hlt">droplets</span> present remarkable properties such as lack of a large pinning force. We discuss the drag on these <span class="hlt">droplets</span> as a function of various parameters. We show theoretical and experimental results of how various confinement geometries change the vapor gradient and the dynamics of <span class="hlt">droplet</span> attraction.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1766322','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=1766322"><span>Reactions in <span class="hlt">Droplets</span> in Microfluidic Channels</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Song, Helen; Chen, Delai L.; Ismagilov, Rustem F.</p> <p>2006-01-01</p> <p>Fundamental and applied research in chemistry and biology benefits from opportunities provided by <span class="hlt">droplet</span>-based microfluidic systems. These systems enable the miniaturization of reactions by compartmentalizing reactions in <span class="hlt">droplets</span> of femoliter to microliter volumes. Compartmentalization in <span class="hlt">droplets</span> provides rapid mixing of reagents, control of the timing of reactions on timescales from milliseconds to months, control of interfacial properties, and the ability to synthesize and transport solid reagents and products. <span class="hlt">Droplet</span>-based microfluidics can help to enhance and accelerate chemical and biochemical screening, protein crystallization, enzymatic kinetics, and assays. Moreover, the control provided by <span class="hlt">droplets</span> in microfluidic devices can lead to new scientific methods and insights. PMID:17086584</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016APS..DFDD36010V','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016APS..DFDD36010V"><span><span class="hlt">Droplets</span> on porous hydrophobic surfaces perfused with gas: An air-table for <span class="hlt">droplets</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Vourdas, Nikolaos; Stathopoulos, Vassilis; Laboratory of Chemistry; Materials Technology Team</p> <p>2016-11-01</p> <p>Wetting phenomena on porous hydrophobic surfaces are strongly related to the volume and the pressure of gas pockets resided at the solid-liquid interface. When the porous medium is perfused with gas by means of backpressure an inherently sessile pinned <span class="hlt">droplet</span> undergoes various changes in its shape, contact angles and mobility. This provides an alternative method for active and controlled <span class="hlt">droplet</span> actuation, without use of electricity, magnetism, foreign particles etc. Superhydrophobicity is not a prerequisite, electrode fabrication is not needed, the liquid is not affected thermally or chemically etc. In this work we explore this method, study the pertinent underlying mechanisms, and propose some applications. The adequate backpressure for <span class="hlt">droplet</span> actuation has been measured for various hydrophobic porous surfaces. Backpressure for actuation may be as low as some tens of mbar for some cases, thus providing a rather low-energy demanding alternative. The <span class="hlt">droplet</span> actuation mechanism has been followed numerically; it entails depinning of the receding contact line and movement, by means of a forward wave propagation reaching on the front of the <span class="hlt">droplet</span>. Applications in valving water plugs inside open- or closed- channel fluidics will be provided.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010PhDT........66T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010PhDT........66T"><span>Development of <span class="hlt">droplet</span> microfluidic platforms for the synthesis of monodisperse lipid vesicles and polymer particles</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Teh, Shia-Yen</p> <p></p> <p>This body of work presents my approaches to the design and development of microfluidic platforms for synthesizing monodisperse polymer particles and phospholipid vesicles. There is interest in both of these particles for use in a variety of biomedical applications. Poly(D,L-lactide-co-glycolic acid) (PLGA) particles in particular have been sought out as vehicles for drug delivery due to their biocompatibility and because the rate of degradation -- hence cargo release - can be controlled. On the other hand, liposomes possess membrane structures resembling that of cells, an ability to adopt both hydrophilic and hydrophobic molecules, and are easily functionalized, which make lipid vesicles the ideal candidate for applications ranging from targeted therapeutic delivery to formation of artificial cells. However, current methods of production for both of these particles result in a wide range of sizes and poor cargo uptake efficiency. We address these challenges by utilizing a flow focusing <span class="hlt">droplet</span> generation design, which allows for fine control over <span class="hlt">droplet</span> size and improves encapsulation efficiencies. The size of these <span class="hlt">droplets</span> can be determined by channel geometry and the ratio of fluid flow rates. I will discuss the work I have done to improve upon current technologies to form nano- to micrometer sized PLGA particles and cell-sized lipid vesicles. <span class="hlt">Solvent</span> evaporation and <span class="hlt">solvent</span> extraction methods were implemented and tested in several device designs to optimize the formation process. The particles produced were characterized for their stability, size variation, and ability to encapsulate a model drug. The release profiles of PLGA particles were also measured to determine the length of delivery. In addition, I worked on the generation of monodisperse lipid vesicles to investigate the application of liposomes as an artificial cell. As a proof of principle, expression of green fluorescent protein (GFP) was successfully carried out in the lipid vesicles. This</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29188994','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29188994"><span>Particle Manipulation Methods in <span class="hlt">Droplet</span> Microfluidics.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Tenje, Maria; Fornell, Anna; Ohlin, Mathias; Nilsson, Johan</p> <p>2018-02-06</p> <p>This Feature describes the different particle manipulation techniques available in the <span class="hlt">droplet</span> microfluidics toolbox to handle particles encapsulated inside <span class="hlt">droplets</span> and to manipulate whole <span class="hlt">droplets</span>. We address the advantages and disadvantages of the different techniques to guide new users.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1981drbu.coll...19H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1981drbu.coll...19H"><span>Colliding <span class="hlt">droplets</span>: A short film presentation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hendricks, C. D.</p> <p>1981-12-01</p> <p>A series of experiments were performed in which liquid <span class="hlt">droplets</span> were caused to collide. Impact velocities to several meters per second and <span class="hlt">droplet</span> diameters up to 600 micrometers were used. The impact parameters in the collisions vary from zero to greater than the sum of the <span class="hlt">droplet</span> radii. Photographs of the collisions were taken with a high speed framing camera in order to study the impacts and subsequent behavior of the <span class="hlt">droplets</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013APS..DFDG36008J','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013APS..DFDG36008J"><span>Optofluidic <span class="hlt">droplet</span> coalescence on a microfluidic chip</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Jung, Jin Ho; Lee, Kyung Heon; Lee, Kang Soo; Cho, Hyunjun; Ha, Byung Hang; Destgeer, Ghulam; Sung, Hyung Jin</p> <p>2013-11-01</p> <p>Coalescence is the procedure that two or more <span class="hlt">droplets</span> fuse during contact to form a larger <span class="hlt">droplet</span>. Optofluidic <span class="hlt">droplet</span> coalescence on a microfluidic chip was demonstrated with theoretical and experimental approaches. <span class="hlt">Droplets</span> were produced in a T-junction geometry and their velocities and sizes were adjusted by flow rate. In order to bring them in a direct contact of coalescence, optical gradient force was used to trap the <span class="hlt">droplets</span>. A theoretical modeling of the coalescence was derived by combining the optical force and drag force on the <span class="hlt">droplet</span>. The analytical expression of the optical force on a sphere <span class="hlt">droplet</span> was employed to estimate the trapping efficiency in the ray optics regime. The drag force acting on the <span class="hlt">droplet</span> was calculated in terms of the fluid velocity, viscosity and the geometrical parameters of a microfluidic channel. The <span class="hlt">droplet</span> coalescence was conducted in a microfluidic setup equipped with a 1064 CW laser, focusing optics, a syringe pump, a custom-made stage and a sCMOS camera. The <span class="hlt">droplets</span> were successfully coalesced using the optical gradient force. The experimental data of coalescence were in good agreement with the prediction. This work was supported by the Creative Research Initiatives program (No.2013-003364) of the National Research Foundation of Korea (MSIP).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25059128','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25059128"><span>Detection of heavy-metal ions using liquid crystal <span class="hlt">droplet</span> patterns modulated by interaction between negatively <span class="hlt">charged</span> carboxylate and heavy-metal cations.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Han, Gyeo-Re; Jang, Chang-Hyun</p> <p>2014-10-01</p> <p>Herein, we demonstrated a simple, sensitive, and rapid label-free detection method for heavy-metal (HM) ions using liquid crystal (LC) <span class="hlt">droplet</span> patterns on a solid surface. Stearic-acid-doped LC <span class="hlt">droplet</span> patterns were spontaneously generated on an n-octyltrichlorosilane (OTS)-treated glass substrate by evaporating a solution of the nematic LC, 4-cyano-4'-pentylbiphenyl (5CB), dissolved in heptane. The optical appearance of the <span class="hlt">droplet</span> patterns was a dark crossed texture when in contact with air, which represents the homeotropic orientation of the LC. This was caused by the steric interaction between the LC molecules and the alkyl chains of the OTS-treated surface. The dark crossed appearance of the acid-doped LC patterns was maintained after the addition of phosphate buffered saline (PBS) solution (pH 8.1 at 25°C). The deprotonated stearic-acid molecules self-assembled through the LC/aqueous interface, thereby supporting the homeotropic anchoring of 5CB. However, the optical image of the acid-doped LC <span class="hlt">droplet</span> patterns incubated with PBS containing HM ions appeared bright, indicating a planar orientation of 5CB at the aqueous/LC <span class="hlt">droplet</span> interface. This dark to bright transition of the LC patterns was caused by HM ions attached to the deprotonated carboxylate moiety, followed by the sequential interruption of the self-assembly of the stearic acid at the LC/aqueous interface. The results showed that the acid-doped LC pattern system not only enabled the highly sensitive detection of HM ions at a sub-nanomolar concentration but it also facilitated rapid detection (<10 min) with simple procedures. Copyright © 2014 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012APS..MARP41010S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012APS..MARP41010S"><span>Nonlinear electrohydrodynamics of a viscous <span class="hlt">droplet</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Salipante, Paul; Vlahovska, Petia</p> <p>2012-02-01</p> <p>A classic result due to G.I.Taylor is that a drop placed in a uniform electric field adopts a prolate or oblate spheroidal shape, the flow and shape being axisymmetrically aligned with the applied field. We report an instability and transition to a nonaxisymmetric rotational flow in strong fields, similar to the rotation of solid dielectric spheres observed by Quincke in the 19th century. Our experiments reveal novel <span class="hlt">droplet</span> behaviors such as tumbling, oscillations and chaotic dynamics even under creeping flow conditions. A phase diagram demonstrates the dependence of these behaviors on drop size, viscosity ratio and electric field strength. The theoretical model, which includes anisotropy in the polarization relaxation, elucidates the interplay of interface deformation and <span class="hlt">charging</span> as the source of the rich nonlinear dynamics.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://pubs.er.usgs.gov/publication/70020138','USGSPUBS'); return false;" href="https://pubs.er.usgs.gov/publication/70020138"><span>Investigation of aggregation in <span class="hlt">solvent</span> extraction of lanthanides by acidic extractants (organophosphorus and naphthenic acid)</span></a></p> <p><a target="_blank" href="http://pubs.er.usgs.gov/pubs/index.jsp?view=adv">USGS Publications Warehouse</a></p> <p>Zhou, N.; Wu, J.; Yu, Z.; Neuman, R.D.; Wang, D.; Xu, G.</p> <p>1997-01-01</p> <p>Three acidic extractants (I) di(2-ethylhexyl) phosphoric acid (HDEHP), (II) 2-ethylhexyl phosphonic acid mono-2-ethylhexyl ester (HEHPEHE) and (III) naphthenic acid were employed in preparing the samples for the characterization of the coordination structure of lanthanide-extractant complexes and the physicochemical nature of aggregates formed in the organic diluent of the <span class="hlt">solvent</span> extraction systems. Photo correlation spectroscopy (PCS) results on the aggregates formed by the partially saponified HDEHP in n-heptane showed that the hydrodynamic radius of the aggregates was comparable to the molecular dimensions of HDEHP. The addition of 2-octanol into the diluent, by which the mixed <span class="hlt">solvent</span> was formed, increased the dimensions of the corresponding aggregates. Aggregates formed from the lanthanide ions and HDEHP in the organic phase of the extraction systems were found very unstable. In the case of naphthenic acid, PCS data showed the formation of w/o microemulsion from the saponified naphthenic acid in the mixed <span class="hlt">solvent</span>. The extraction of lanthanides by the saponified naphthenic acid in the mixed <span class="hlt">solvent</span> under the given experimental conditions was a process of destruction of the w/o microemulsion. A possible mechanism of the breakdown of the w/o microemulsion <span class="hlt">droplets</span> is discussed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23834593','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23834593"><span>Bütschli dynamic <span class="hlt">droplet</span> system.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Armstrong, Rachel; Hanczyc, Martin</p> <p>2013-01-01</p> <p>Dynamical oil-water systems such as <span class="hlt">droplets</span> display lifelike properties and may lend themselves to chemical programming to perform useful work, specifically with respect to the built environment. We present Bütschli water-in-oil <span class="hlt">droplets</span> as a model for further investigation into the development of a technology with living properties. Otto Bütschli first described the system in 1898, when he used alkaline water <span class="hlt">droplets</span> in olive oil to initiate a saponification reaction. This simple recipe produced structures that moved and exhibited characteristics that resembled, at least superficially, the amoeba. We reconstructed the Bütschli system and observed its life span under a light microscope, observing chemical patterns and <span class="hlt">droplet</span> behaviors in nearly three hundred replicate experiments. Self-organizing patterns were observed, and during this dynamic, embodied phase the <span class="hlt">droplets</span> provided a means of introducing temporal and spatial order in the system with the potential for chemical programmability. The authors propose that the discrete formation of dynamic <span class="hlt">droplets</span>, characterized by their lifelike behavior patterns, during a variable window of time (from 30 s to 30 min after the addition of alkaline water to the oil phase), qualify this system as an example of living technology. The analysis of the Bütschli <span class="hlt">droplets</span> suggests that a set of conditions may precede the emergence of lifelike characteristics and exemplifies the richness of this rudimentary chemical system, not only for artificial life investigations but also for possible real-world applications in architectural practice.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26486337','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26486337"><span><span class="hlt">Droplets</span>, Bubbles and Ultrasound Interactions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Shpak, Oleksandr; Verweij, Martin; de Jong, Nico; Versluis, Michel</p> <p>2016-01-01</p> <p>The interaction of <span class="hlt">droplets</span> and bubbles with ultrasound has been studied extensively in the last 25 years. Microbubbles are broadly used in diagnostic and therapeutic medical applications, for instance, as ultrasound contrast agents. They have a similar size as red blood cells, and thus are able to circulate within blood vessels. Perfluorocarbon liquid <span class="hlt">droplets</span> can be a potential new generation of microbubble agents as ultrasound can trigger their conversion into gas bubbles. Prior to activation, they are at least five times smaller in diameter than the resulting bubbles. Together with the violent nature of the phase-transition, the <span class="hlt">droplets</span> can be used for local drug delivery, embolotherapy, HIFU enhancement and tumor imaging. Here we explain the basics of bubble dynamics, described by the Rayleigh-Plesset equation, bubble resonance frequency, damping and quality factor. We show the elegant calculation of the above characteristics for the case of small amplitude oscillations by linearizing the equations. The effect and importance of a bubble coating and effective surface tension are also discussed. We give the main characteristics of the power spectrum of bubble oscillations. Preceding bubble dynamics, ultrasound propagation is introduced. We explain the speed of sound, nonlinearity and attenuation terms. We examine bubble ultrasound scattering and how it depends on the wave-shape of the incident wave. Finally, we introduce <span class="hlt">droplet</span> interaction with ultrasound. We elucidate the ultrasound-focusing concept within a <span class="hlt">droplets</span> sphere, <span class="hlt">droplet</span> shaking due to media compressibility and <span class="hlt">droplet</span> phase-conversion dynamics.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20040053530','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20040053530"><span><span class="hlt">Droplet</span> Vaporization In A Levitating Acoustic Field</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Ruff, G. A.; Liu, S.; Ciobanescu, I.</p> <p>2003-01-01</p> <p>Combustion experiments using arrays of <span class="hlt">droplets</span> seek to provide a link between single <span class="hlt">droplet</span> combustion phenomena and the behavior of complex spray combustion systems. Both single <span class="hlt">droplet</span> and <span class="hlt">droplet</span> array studies have been conducted in microgravity to better isolate the <span class="hlt">droplet</span> interaction phenomena and eliminate or reduce the effects of buoyancy-induced convection. In most experiments involving <span class="hlt">droplet</span> arrays, the <span class="hlt">droplets</span> are supported on fibers to keep them stationary and close together before the combustion event. The presence of the fiber, however, disturbs the combustion process by introducing a source of heat transfer and asymmetry into the configuration. As the number of drops in a <span class="hlt">droplet</span> array increases, supporting the drops on fibers becomes less practical because of the cumulative effect of the fibers on the combustion process. To eliminate the effect of the fiber, several researchers have conducted microgravity experiments using unsupported <span class="hlt">droplets</span>. Jackson and Avedisian investigated single, unsupported drops while Nomura et al. studied <span class="hlt">droplet</span> clouds formed by a condensation technique. The overall objective of this research is to extend the study of unsupported drops by investigating the combustion of well-characterized drop clusters in a microgravity environment. Direct experimental observations and measurements of the combustion of <span class="hlt">droplet</span> clusters would provide unique experimental data for the verification and improvement of spray combustion models. In this work, the formation of drop clusters is precisely controlled using an acoustic levitation system so that dilute, as well as dense clusters can be created and stabilized before combustion in microgravity is begun. While the low-gravity test facility is being completed, tests have been conducted in 1-g to characterize the effect of the acoustic field on the vaporization of single and multiple <span class="hlt">droplets</span>. This is important because in the combustion experiment, the <span class="hlt">droplets</span> will be formed and</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29655348','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29655348"><span>Mesoscopic electrohydrodynamic simulations of binary colloidal suspensions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Rivas, Nicolas; Frijters, Stefan; Pagonabarraga, Ignacio; Harting, Jens</p> <p>2018-04-14</p> <p>A model is presented for the solution of electrokinetic phenomena of colloidal suspensions in fluid mixtures. We solve the discrete Boltzmann equation with a Bhatnagar-Gross-Krook collision operator using the lattice Boltzmann method to simulate binary fluid flows. <span class="hlt">Solvent-solvent</span> and <span class="hlt">solvent</span>-solute interactions are implemented using a pseudopotential model. The Nernst-Planck equation, describing the kinetics of dissolved ion species, is solved using a finite difference discretization based on the link-flux method. The colloids are resolved on the lattice and coupled to the hydrodynamics and electrokinetics through appropriate boundary conditions. We present the first full integration of these three elements. The model is validated by comparing with known analytic solutions of ionic distributions at fluid interfaces, dielectric <span class="hlt">droplet</span> deformations, and the electrophoretic mobility of colloidal suspensions. Its possibilities are explored by considering various physical systems, such as breakup of <span class="hlt">charged</span> and neutral <span class="hlt">droplets</span> and colloidal dynamics at either planar or spherical fluid interfaces.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018JChPh.148n4101R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018JChPh.148n4101R"><span>Mesoscopic electrohydrodynamic simulations of binary colloidal suspensions</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rivas, Nicolas; Frijters, Stefan; Pagonabarraga, Ignacio; Harting, Jens</p> <p>2018-04-01</p> <p>A model is presented for the solution of electrokinetic phenomena of colloidal suspensions in fluid mixtures. We solve the discrete Boltzmann equation with a Bhatnagar-Gross-Krook collision operator using the lattice Boltzmann method to simulate binary fluid flows. <span class="hlt">Solvent-solvent</span> and <span class="hlt">solvent</span>-solute interactions are implemented using a pseudopotential model. The Nernst-Planck equation, describing the kinetics of dissolved ion species, is solved using a finite difference discretization based on the link-flux method. The colloids are resolved on the lattice and coupled to the hydrodynamics and electrokinetics through appropriate boundary conditions. We present the first full integration of these three elements. The model is validated by comparing with known analytic solutions of ionic distributions at fluid interfaces, dielectric <span class="hlt">droplet</span> deformations, and the electrophoretic mobility of colloidal suspensions. Its possibilities are explored by considering various physical systems, such as breakup of <span class="hlt">charged</span> and neutral <span class="hlt">droplets</span> and colloidal dynamics at either planar or spherical fluid interfaces.</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/14754227','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/14754227"><span>Theory of polyelectrolytes in <span class="hlt">solvents</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Chitanvis, Shirish M</p> <p>2003-12-01</p> <p>Using a continuum description, we account for fluctuations in the ionic <span class="hlt">solvent</span> surrounding a Gaussian, <span class="hlt">charged</span> chain and derive an effective short-ranged potential between the <span class="hlt">charges</span> on the chain. This potential is repulsive at short separations and attractive at longer distances. The chemical potential can be derived from this potential. When the chemical potential is positive, it leads to a meltlike state. For a vanishingly low concentration of segments, this state exhibits scaling behavior for long chains. The Flory exponent characterizing the radius of gyration for long chains is calculated to be approximately 0.63, close to the classical value obtained for second order phase transitions. For short chains, the radius of gyration varies linearly with N, the chain length, and is sensitive to the parameters in the interaction potential. The linear dependence on the chain length N indicates a stiff behavior. The chemical potential associated with this interaction changes sign, when the screening length in the ionic <span class="hlt">solvent</span> exceeds a critical value. This leads to condensation when the chemical potential is negative. In this state, it is shown using the mean-field approximation that spherical and toroidal condensed shapes can be obtained. The thickness of the toroidal polyelectrolyte is studied as a function of the parameters of the model, such as the ionic screening length. The predictions of this theory should be amenable to experimental verification.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26120062','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26120062"><span>Capillary <span class="hlt">droplet</span> propulsion on a fibre.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Haefner, Sabrina; Bäumchen, Oliver; Jacobs, Karin</p> <p>2015-09-21</p> <p>A viscous liquid film coating a fibre becomes unstable and decays into <span class="hlt">droplets</span> due to the Rayleigh-Plateau instability (RPI). Here, we report on the generation of uniform <span class="hlt">droplets</span> on a hydrophobized fibre by taking advantage of this effect. In the late stages of liquid column breakup, a three-phase contact line can be formed at one side of the <span class="hlt">droplet</span> by spontaneous rupture of the thinning film. The resulting capillary imbalance leads to <span class="hlt">droplet</span> propulsion along the fibre. We study the dynamics and the dewetting speed of the <span class="hlt">droplet</span> as a function of molecular weight as well as temperature and compare to a force balance model based on purely viscous dissipation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19990019814','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19990019814"><span>Fiber Supported <span class="hlt">Droplet</span> Combustion-2 (FSDC-2)</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Colantonio, Renato; Dietrich, Daniel; Haggard, John B., Jr.; Nayagan, Vedha; Dryer, Frederick L.; Shaw, Benjamin D.; Williams, Forman A.</p> <p>1998-01-01</p> <p>Experimental results for the burning characteristics of fiber supported, liquid <span class="hlt">droplets</span> in ambient Shuttle cabin air (21% oxygen, 1 bar pressure) were obtained from the Glove Box Facility aboard the STS-94/MSL-1 mission using the Fiber Supported <span class="hlt">Droplet</span> Combustion - 2 (FSDC-2) apparatus. The combustion of individual <span class="hlt">droplets</span> of methanol/water mixtures, ethanol, ethanol/water azeotrope, n-heptane, n-decane, and n-heptane/n-hexadecane mixtures were studied in quiescent air. The effects of low velocity, laminar gas phase forced convection on the combustion of individual <span class="hlt">droplets</span> of n-heptane and n-decane were investigated and interactions of two <span class="hlt">droplet</span>-arrays of n-heptane and n-decane <span class="hlt">droplets</span> were also studied with and without gas phase convective flow. Initial diameters ranging from about 2mm to over 6mm were burned on 80-100 micron silicon fibers. In addition to phenomenological observations, quantitative data were obtained in the form of backlit images of the burning <span class="hlt">droplets</span>, overall flame images, and radiometric combustion emission measurements as a function of the burning time in each experiment. In all, 124 of the 129 attempted experiments (or about twice the number of experiments originally planned for the STS-94/MSL-1 mission) were conducted successfully. The experimental results contribute new observations on the combustion properties of pure alkanes, binary alkane mixtures, and simple alcohols for <span class="hlt">droplet</span> sizes not studied previously, including measurements on individual <span class="hlt">droplets</span> and two-<span class="hlt">droplet</span> arrays, inclusive of the effects of forced gas phase convection. New phenomena characterized experimentally for the first time include radiative extinction of <span class="hlt">droplet</span> burning for alkanes and the "twin effect" which occurs as a result of interactions during the combustion of two-<span class="hlt">droplet</span> arrays. Numerical modeling of isolated <span class="hlt">droplet</span> combustion phenomenon has been conducted for methanol/water mixtures, n-heptane, and n-heptane/n-hexadecane mixtures, and results</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018HMT...tmp..152C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018HMT...tmp..152C"><span>Experimental study on the <span class="hlt">droplet</span> formation around pins of different geometry for the design of a compact falling-<span class="hlt">droplet</span> absorber</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cola, Fabrizio; Romagnoli, Alessandro; Hey, Jonathan</p> <p>2018-05-01</p> <p>Absorber downsizing for the development of compact absorption chillers is a known challenge of this type of refrigerator. Past studies have revealed how a <span class="hlt">droplet</span> flow regime can increase the interface area and enhance absorption rates, especially during the <span class="hlt">droplet</span> formation. This study proposes a space-efficient design for an adiabatic absorber based on a bank of solid pins coupled with a <span class="hlt">droplet</span> flow regime. Manufacturing through 3D printing technique is used to study the effect of different fin shapes during <span class="hlt">droplet</span> formation. <span class="hlt">Droplet</span> behavior is firstly studied analytically through a variational approach. Experiments on pure water are then carried out to validate the model and produce design guidelines for a H2O-LiBr absorber. Results show that the analytical model is more accurate in the regions close to the <span class="hlt">droplet</span> bottom. The rhomboidal geometry with 120° returned the smallest <span class="hlt">droplet</span> volume without allowing coalescence of more <span class="hlt">droplets</span>, ensuring the maintenance of <span class="hlt">droplet</span> flow and a high surface area for mass transfer. Disturbances in the <span class="hlt">droplet</span> profiles were observed, caused by the pin-<span class="hlt">droplet</span> interaction. A map has been then created to allow a quick sizing of the absorber and find its main geometrical and operational features.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19910067786&hterms=uranium&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Duranium','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19910067786&hterms=uranium&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3Duranium"><span>Uranium <span class="hlt">droplet</span> core nuclear rocket</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Anghaie, Samim</p> <p>1991-01-01</p> <p>Uranium <span class="hlt">droplet</span> nuclear rocket is conceptually designed to utilize the broad temperature range ofthe liquid phase of metallic uranium in <span class="hlt">droplet</span> configuration which maximizes the energy transfer area per unit fuel volume. In a baseline system dissociated hydrogen at 100 bar is heated to 6000 K, providing 2000 second of Isp. Fission fragments and intense radian field enhance the dissociation of molecular hydrogen beyond the equilibrium thermodynamic level. Uranium <span class="hlt">droplets</span> in the core are confined and separated by an axisymmetric vortex flow generated by high velocity tangential injection of hydrogen in the mid-core regions. <span class="hlt">Droplet</span> uranium flow to the core is controlled and adjusted by a twin flow nozzle injection system.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014APS..MARD22002S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014APS..MARD22002S"><span>Mechanical vibration of viscoelastic liquid <span class="hlt">droplets</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sharp, James; Harrold, Victoria</p> <p>2014-03-01</p> <p>The resonant vibrations of viscoelastic sessile <span class="hlt">droplets</span> supported on different substrates were monitored using a simple laser light scattering technique. In these experiments, laser light was reflected from the surfaces of <span class="hlt">droplets</span> of high Mw poly acrylamide-co-acrylic acid (PAA) dissolved in water. The scattered light was allowed to fall on the surface of a photodiode detector and a mechanical impulse was applied to the drops using a vibration motor mounted beneath the substrates. The mechanical impulse caused the <span class="hlt">droplets</span> to vibrate and the scattered light moved across the surface of the photodiode. The resulting time dependent photodiode signal was then Fourier transformed to obtain the mechanical vibrational spectra of the <span class="hlt">droplets</span>. The frequencies and widths of the resonant peaks were extracted for <span class="hlt">droplets</span> containing different concentrations of PAA and with a range of sizes. This was repeated for PAA loaded water drops on surfaces which displayed different values of the three phase contact angle. The results were compared to a simple model of <span class="hlt">droplet</span> vibration which considers the formation of standing wave states on the surface of a viscoelastic <span class="hlt">droplet</span>. We gratefully acknowledge the support of the Leverhulme trust under grant number RPG-2012-702.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22136217','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22136217"><span>Influence of film dimensions on film <span class="hlt">droplet</span> formation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Holmgren, Helene; Ljungström, Evert</p> <p>2012-02-01</p> <p>Aerosol particles may be generated from rupturing liquid films through a <span class="hlt">droplet</span> formation mechanism. The present work was undertaken with the aim to throw some light on the influence of film dimensions on <span class="hlt">droplet</span> formation with possible consequences for exhaled breath aerosol formation. The film <span class="hlt">droplet</span> formation process was mimicked by using a purpose-built device, where fluid films were spanned across holes of known diameters. As the films burst, <span class="hlt">droplets</span> were formed and the number and size distributions of the resulting <span class="hlt">droplets</span> were determined. No general relation could be found between hole diameter and the number of <span class="hlt">droplets</span> generated per unit surface area of fluid film. Averaged over all film sizes, a higher surface tension yielded higher concentrations of <span class="hlt">droplets</span>. Surface tension did not influence the resulting <span class="hlt">droplet</span> diameter, but it was found that smaller films generated smaller <span class="hlt">droplets</span>. This study shows that small fluid films generate <span class="hlt">droplets</span> as efficiently as large films, and that <span class="hlt">droplets</span> may well be generated from films with diameters below 1 mm. This has implications for the formation of film <span class="hlt">droplets</span> from reopening of closed airways because human terminal bronchioles are of similar dimensions. Thus, the results provide support for the earlier proposed mechanism where reopening of closed airways is one origin of exhaled particles.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010SPIE.7743E..0DS','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010SPIE.7743E..0DS"><span>Influence of palmitoyl pentapeptide and Ceramide III B on the <span class="hlt">droplet</span> size of nanoemulsion</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Sondari, Dewi; Haryono, Agus; Harmami, Sri Budi; Randy, Ahmad</p> <p>2010-05-01</p> <p>The influence of the Palmitoyl Pentapeptide (PPp) and Ceramide IIIB (Cm III B) as active ingredients on the <span class="hlt">droplet</span> size of nano-emulsion was studied using different kinds of oil (avocado oil, sweet almond oil, jojoba oil, mineral oil and squalene). The formation of nano-emulsions were prepared in water mixed non ionic surfactant/oils system using the spontaneous emulsification mechanism. The aqueous solution, which consist of water and Tween® 20 as a hydrophilic surfactant was mixed homogenously. The organic solution, which consist of oil and Span® 80 as a lipophilic surfactant was mixed homogenously in ethanol. Ethanol was used as a water miscible <span class="hlt">solvent</span>, which can help the formation of nano-emulsion. The oil phase (containing the blend of surfactant Span® 80, ethanol, oil and active ingredient) and the aqueous phase (containing water and Tween® 20) were separately prepared at room temperatures. The oil phase was slowly added into aqueous phase under continuous mechanical agitation (18000 rpm). All samples were subsequently homogenized with Ultra-Turrax for 30 minutes. The characterizations of nano-emulsion were carried out using photo-microscope and particle size analyzer. Addition of active ingredients on the formation of nano-emulsion gave smallest <span class="hlt">droplet</span> size compared without active ingredients addition on the formation of nano-emulsion. Squalene oil with Palmitoyl Pentapeptide (PPm) and Ceramide IIIB (Cm IIIB) gave smallest <span class="hlt">droplet</span> size (184.0 nm) compared without Palmitoyl Pentapeptide and Ceramide IIIB (214.9 nm), however the <span class="hlt">droplets</span> size of the emulsion prepared by the other oils still in the range of nano-emulsion (below 500 nm). The stability of nano-emulsion was observed using two methods. In one method, the stability of nano-emulsion was observed for three months at temperature of 5°C and 50°C, while in the other method, the stability nano-emulsion was observed by centrifuged at 12000 rpm for 30 minutes. Nanoemulsion with active ingredient</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21076437','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21076437"><span>The epididymis, cytoplasmic <span class="hlt">droplets</span> and male fertility.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cooper, Trevor G</p> <p>2011-01-01</p> <p>The potential of spermatozoa to become motile during post-testicular maturation, and the relationship between the cytoplasmic <span class="hlt">droplet</span> and fertilizing capacity are reviewed. Post-testicular maturation of spermatozoa involves the autonomous induction of motility, which can occur in vivo in testes with occluded excurrent ducts and in vitro in testicular explants, and artefactual changes in morphology that appear to occur in the testis in vitro. Both modifications may reflect time-dependent oxidation of disulphide bonds of head and tail proteins. Regulatory volume decrease (RVD), which counters sperm swelling at ejaculation, is discussed in relation to loss of cytoplasmic <span class="hlt">droplets</span> and consequences for fertility. It is postulated that: (i) fertile males possess spermatozoa with sufficient osmolytes to drive RVD at ejaculation, permitting the <span class="hlt">droplet</span> to round up and pinch off without membrane rupture; and (ii) infertile males possess spermatozoa with insufficient osmolytes so that RVD is inadequate, the <span class="hlt">droplet</span> swells and the resulting flagellar angulation prevents <span class="hlt">droplet</span> loss. <span class="hlt">Droplet</span> retention at ejaculation is a harbinger of infertility caused by failure of the spermatozoon to negotiate the uterotubal junction or mucous and reach the egg. In this hypothesis, the epididymis regulates fertility indirectly by the extent of osmolyte provision to spermatozoa, which influences RVD and therefore <span class="hlt">droplet</span> loss. Man is an exception, because ejaculated human spermatozoa retain their <span class="hlt">droplets</span>. This may reflect their short midpiece, approximating head length, permitting a swollen <span class="hlt">droplet</span> to extend along the entire midpiece; this not only obviates <span class="hlt">droplet</span> migration and flagellar angulation but also hampers <span class="hlt">droplet</span> loss.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2009ApPhL..95o4104S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2009ApPhL..95o4104S"><span>Discrete microfluidics: Reorganizing <span class="hlt">droplet</span> arrays at a bend</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Surenjav, Enkhtuul; Herminghaus, Stephan; Priest, Craig; Seemann, Ralf</p> <p>2009-10-01</p> <p>Microfluidic manipulation of densely packed <span class="hlt">droplet</span> arrangements (i.e., gel emulsions) using sharp microchannel bends was studied as a function of bend angle, <span class="hlt">droplet</span> volume fraction, <span class="hlt">droplet</span> size, and flow velocity. Emulsion reorganization was found to be specifically dependent on the pathlength that the <span class="hlt">droplets</span> are forced to travel as they navigate the bend under spatial confinement. We describe how bend-induced <span class="hlt">droplet</span> displacements might be exploited in complex, <span class="hlt">droplet</span>-based microfluidics.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22188180','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22188180"><span>Nanoliter <span class="hlt">droplet</span> vitrification for oocyte cryopreservation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zhang, Xiaohui; Khimji, Imran; Shao, Lei; Safaee, Hooman; Desai, Khanjan; Keles, Hasan Onur; Gurkan, Umut Atakan; Kayaalp, Emre; Nureddin, Aida; Anchan, Raymond M; Maas, Richard L; Demirci, Utkan</p> <p>2012-04-01</p> <p>Oocyte cryopreservation remains largely experimental, with live birth rates of only 2-4% per thawed oocyte. In this study, we present a nanoliter <span class="hlt">droplet</span> technology for oocyte vitrification. An ejector-based <span class="hlt">droplet</span> vitrification system was designed to continuously cryopreserve oocytes in nanoliter <span class="hlt">droplets</span>. Oocyte survival rates, morphologies and parthenogenetic development after each vitrification step were assessed in comparison with fresh oocytes. Oocytes were retrieved after cryoprotectant agent loading/unloading, and nanoliter <span class="hlt">droplet</span> encapsulation showed comparable survival rates to fresh oocytes after 24 h in culture. Also, oocytes recovered after vitrification/thawing showed similar morphologies to those of fresh oocytes. Additionally, the rate of oocyte parthenogenetic activation after nanoliter <span class="hlt">droplet</span> encapsulation was comparable with that observed for fresh oocytes. This nanoliter <span class="hlt">droplet</span> technology enables the vitrification of oocytes at higher cooling and warming rates using lower cryoprotectant agent levels (i.e., 1.4 M ethylene glycol, 1.1 M dimethyl sulfoxide and 1 M sucrose), thus making it a potential technology to improve oocyte cryopreservation outcomes.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23556952','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23556952"><span>Bouncing <span class="hlt">droplets</span> on a billiard table.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Shirokoff, David</p> <p>2013-03-01</p> <p>In a set of experiments, Couder et al. demonstrate that an oscillating fluid bed may propagate a bouncing <span class="hlt">droplet</span> through the guidance of the surface waves. I present a dynamical systems model, in the form of an iterative map, for a <span class="hlt">droplet</span> on an oscillating bath. I examine the <span class="hlt">droplet</span> bifurcation from bouncing to walking, and prescribe general requirements for the surface wave to support stable walking states. I show that in addition to walking, there is a region of large forcing that may support the chaotic motion of the <span class="hlt">droplet</span>. Using the map, I then investigate the <span class="hlt">droplet</span> trajectories in a square (billiard ball) domain. I show that in large domains, the long time trajectories are either non-periodic dense curves or approach a quasiperiodic orbit. In contrast, in small domains, at low forcing, trajectories tend to approach an array of circular attracting sets. As the forcing increases, the attracting sets break down and the <span class="hlt">droplet</span> travels throughout space.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17304941','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17304941"><span>Estimation of aerosol <span class="hlt">droplet</span> sizes by using a modified DC-III portable <span class="hlt">droplet</span> measurement system under laboratory and field conditions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Dennett, James A; Stark, Pamela M; Vessey, Nathan Y; Parsons, Ray E; Bueno, Rudy</p> <p>2006-12-01</p> <p>Modification of the DC-III portable <span class="hlt">droplet</span> measurement system, permitting its use under field conditions, is described. Under laboratory conditions, the system effectively sampled water <span class="hlt">droplets</span> from aerosols produced by a dry ice/water generator and high-pressure syringe. Seven <span class="hlt">droplet</span> sizes, totaling 71,053 <span class="hlt">droplets</span> within 22 tests (dry ice method), consisted of 1-, 2-, 6-, 11-, 18-, 25-, and 34-microm <span class="hlt">droplets</span> with individual (rounded) percentages of 45.25, 37.22, 13.85, 3.17, 0.45, 0.02, and 0.005, respectively, for each size. Cumulatively, 1-microm <span class="hlt">droplets</span> accounted for ca. 45.25% of the <span class="hlt">droplets</span> sampled; combined with 2-microm (ca. 82.48% together), 6-microm (ca. 96.33% together), and 11-microm <span class="hlt">droplets</span>, yielded ca. 99.51% of the <span class="hlt">droplets</span> sampled. The syringe produced 12 <span class="hlt">droplet</span> sizes, with 4,121 <span class="hlt">droplets</span> sampled, consisting of 1, 2, 6, 11, 18, 25, 34, 45, 56, 69, 83, and 99 microm with individual percentages of 15.43, 21.91, 24.58, 17.30, 10.62, 4.65, 2.93, 1.33, 0.63, 0.33, 0.16, 0.07, respectively, for each size. The 6-microm <span class="hlt">droplets</span> contributed the highest individual percentage, and cumulatively, these <span class="hlt">droplets</span> combined with 1- and 2-microm <span class="hlt">droplets</span>, yielding 61.93%, whereas 11- to 45-microm <span class="hlt">droplets</span> contributed 36.83%, for a total of 98.76%. <span class="hlt">Droplets</span> measuring 56-99 microm accounted for ca. 1.24% of <span class="hlt">droplets</span> sampled. Hand-fogger oil aerosols produced 12 <span class="hlt">droplet</span> sizes (1-38 microm) at test distances of 7.6 and 15.2 m, with 1,979 and 268 <span class="hlt">droplets</span> sampled, respectively, during 10 tests at each distance. With analysis of variance of transformed individual percentages for each size at both distances, no significant differences were observed for 7.6 and 15.2 m. Cumulatively, 1-, 2-, 3-, and 5-microm <span class="hlt">droplets</span> contributed 82.87 and 80.97%, whereas 8-, 11-, 14-, and 18-microm <span class="hlt">droplets</span> added 14.55% to totals at both 7.6 and 15.2 m, respectively. <span class="hlt">Droplets</span> measuring 22, 27, 32, and 38 microm contributed 2.57% and 4.47% to samples obtained at 7.6 and 15.2 m. The</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20010074091','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20010074091"><span>Combustion of Unconfined <span class="hlt">Droplet</span> Clusters in Microgravity</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Ruff, G. A.; Liu, S.</p> <p>2001-01-01</p> <p>Combustion experiments using arrays of <span class="hlt">droplets</span> seek to provide a link between single <span class="hlt">droplet</span> combustion phenomena and the behavior of complex spray combustion systems. Both single <span class="hlt">droplet</span> and <span class="hlt">droplet</span> array studies have been conducted in microgravity to better isolate the <span class="hlt">droplet</span> interaction phenomena and eliminate or reduce the confounding effects of buoyancy-induced convection. In most experiments involving <span class="hlt">droplet</span> arrays, the <span class="hlt">droplets</span> are supported on fibers to keep them stationary and close together before the combustion event. The presence of the fiber, however, disturbs the combustion process by introducing a source of heat transfer and asymmetry into the configuration. As the number of drops in a <span class="hlt">droplet</span> array increases, supporting the drops on fibers becomes less practical because of the cumulative effect of the fibers on the combustion process. To eliminate the effect of the fiber, several researchers have conducted microgravity experiments using unsupported <span class="hlt">droplets</span>. Jackson and Avedisian investigated single, unsupported drops while Nomura et al. studied <span class="hlt">droplet</span> clouds formed by a condensation technique. The overall objective of this research is to extend the study of unsupported drops by investigating the combustion of well-characterized drop clusters in a microgravity environment. Direct experimental observations and measurements of the combustion of <span class="hlt">droplet</span> clusters would fill a large gap in our current understanding of <span class="hlt">droplet</span> and spray combustion and provide unique experimental data for the verification and improvement of spray combustion models. In this work, the formation of drop clusters is precisely controlled using an acoustic levitation system so that dilute, as well as dense clusters can be created and stabilized before combustion in microgravity is begun. This paper describes the design and performance of the 1-g experimental apparatus, some preliminary 1-g results, and plans for testing in microgravity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27075732','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27075732"><span>A highly addressable static <span class="hlt">droplet</span> array enabling digital control of a single <span class="hlt">droplet</span> at pico-volume resolution.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jeong, Heon-Ho; Lee, Byungjin; Jin, Si Hyung; Jeong, Seong-Geun; Lee, Chang-Soo</p> <p>2016-04-26</p> <p><span class="hlt">Droplet</span>-based microfluidics enabling exquisite liquid-handling has been developed for diagnosis, drug discovery and quantitative biology. Compartmentalization of samples into a large number of tiny <span class="hlt">droplets</span> is a great approach to perform multiplex assays and to improve reliability and accuracy using a limited volume of samples. Despite significant advances in microfluidic technology, individual <span class="hlt">droplet</span> handling in pico-volume resolution is still a challenge in obtaining more efficient and varying multiplex assays. We present a highly addressable static <span class="hlt">droplet</span> array (SDA) enabling individual digital manipulation of a single <span class="hlt">droplet</span> using a microvalve system. In a conventional single-layer microvalve system, the number of microvalves required is dictated by the number of operation objects; thus, individual trap-and-release on a large-scale 2D array format is highly challenging. By integrating double-layer microvalves, we achieve a "balloon" valve that preserves the pressure-on state under released pressure; this valve can allow the selective releasing and trapping of 7200 multiplexed pico-<span class="hlt">droplets</span> using only 1 μL of sample without volume loss. This selectivity and addressability completely arranged only single-cell encapsulated <span class="hlt">droplets</span> from a mixture of <span class="hlt">droplet</span> compositions via repetitive selective trapping and releasing. Thus, it will be useful for efficient handling of miniscule volumes of rare or clinical samples in multiplex or combinatory assays, and the selective collection of samples.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018ApPhL.112f3701F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018ApPhL.112f3701F"><span>Binary particle separation in <span class="hlt">droplet</span> microfluidics using acoustophoresis</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fornell, Anna; Cushing, Kevin; Nilsson, Johan; Tenje, Maria</p> <p>2018-02-01</p> <p>We show a method for separation of two particle species with different acoustic contrasts originally encapsulated in the same <span class="hlt">droplet</span> in a continuous two-phase system. This was realized by using bulk acoustic standing waves in a 380 μm wide silicon-glass microfluidic channel. Polystyrene particles (positive acoustic contrast particles) and in-house synthesized polydimethylsiloxane (PDMS) particles (negative acoustic contrast particles) were encapsulated inside water-in-oil <span class="hlt">droplets</span> either individually or in a mixture. At acoustic actuation of the system at the fundamental resonance frequency, the polystyrene particles were moved to the center of the <span class="hlt">droplet</span> (pressure node), while the PDMS particles were moved to the sides of the <span class="hlt">droplet</span> (pressure anti-nodes). The acoustic particle manipulation step was combined in series with a trifurcation <span class="hlt">droplet</span> splitter, and as the original <span class="hlt">droplet</span> passed through the splitter and was divided into three daughter <span class="hlt">droplets</span>, the polystyrene particles were directed into the center daughter <span class="hlt">droplet</span>, while the PDMS particles were directed into the two side daughter <span class="hlt">droplets</span>. The presented method expands the <span class="hlt">droplet</span> microfluidics tool-box and offers new possibilities to perform binary particle separation in <span class="hlt">droplet</span> microfluidic systems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1998PhyB..249..762W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1998PhyB..249..762W"><span>Cyclotron resonance of interacting quantum Hall <span class="hlt">droplets</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Widmann, M.; Merkt, U.; Cortés, M.; Häusler, W.; Eberl, K.</p> <p>1998-06-01</p> <p>The line shape and position of cyclotron resonance in gated GaAs/GaAlAs heterojunctions with δ-doped layers of negatively <span class="hlt">charged</span> beryllium acceptors, that provide strong potential fluctuations in the channels of the quasi-two-dimensional electron systems, are examined. Specifically, the magnetic quantum limit is considered when the electrons are localized in separate quantum Hall <span class="hlt">droplets</span> in the valleys of the disorder potential. A model treating disorder and electron-electron interaction on an equal footing accounts for all of the principal experimental findings: blue shifts from the unperturbed cyclotron frequency that decrease when the electron density is reduced, surprisingly narrow lines in the magnetic quantum limit, and asymmetric lines due to additional oscillator strength on their high-frequency sides.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015APS..MAR.P1357F','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015APS..MAR.P1357F"><span>Particle-Laden Leidenfrost <span class="hlt">Droplets</span>: Final-Stage Observations</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Fang, Zecong; Xu, Jie</p> <p>2015-03-01</p> <p>Little interest has been paid to the final stage of a Leidenfrost <span class="hlt">droplet</span> until a recent study by Celestini et al [Phys. Rev. Lett. 109, 034501 (2012)] reporting an unexpected take-off phenomenon of micrometer sized pure liquid <span class="hlt">droplets</span> (Rl < R <Ri , where Rl is the take-off radius, and Ri is the critical radius above which <span class="hlt">droplets</span> start to lose sphericity). In our study, we first report an unexpected observation on millimeter sized water Leidenfrost <span class="hlt">droplets</span> (R >Ri), which behave quite differently from the previous study. While an originally micrometer sized Leidenfrost <span class="hlt">droplet</span> takes off due to breakdown of lubrication regime, and hovers above its vapor layer until disappearing in the final stage of evaporation, an originally millimetric Leidenfrost drop is observed to hover and oscillate, taking off and falling back consecutively. We further report another interesting observation on water <span class="hlt">droplets</span> containing micrometric glass beads. These <span class="hlt">droplets</span> spontaneously organize and buckle together during evaporation. In addition to oscillation just like pure <span class="hlt">droplets</span>, these particle-laden drops create an unexpected explosive shoot-up at the end of evaporation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016PhyA..443..486K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016PhyA..443..486K"><span>Quasistatic packings of <span class="hlt">droplets</span> in flat microfluidic channels</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kadivar, Erfan</p> <p>2016-02-01</p> <p>As observed in recent experiments, monodisperse <span class="hlt">droplets</span> self-assemble spontaneously in different ordered packings. In this work, we present a numerical study of the <span class="hlt">droplet</span> packings in the flat rectangular microfluidic channels. Employing the boundary element method, we numerically solve the Stokes equation in two-dimension and investigate the appearance of <span class="hlt">droplet</span> packing and transition between one and two-row packings of monodisperse emulsion <span class="hlt">droplets</span>. By calculating packing force applied on the <span class="hlt">droplet</span> interface, we investigate the effect of flow rate, <span class="hlt">droplet</span> size, and surface tension on the packing configurations of <span class="hlt">droplets</span> and transition between different topological packings.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1395759','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1395759"><span>Impinging Water <span class="hlt">Droplets</span> on Inclined Glass 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>Armijo, Kenneth Miguel; Lance, Blake; Ho, Clifford K.</p> <p></p> <p>Multiphase computational models and tests of falling water <span class="hlt">droplets</span> on inclined glass surfaces were developed to investigate the physics of impingement and potential of these <span class="hlt">droplets</span> to self-clean glass surfaces for photovoltaic modules and heliostats. A multiphase volume-of-fluid model was developed in ANSYS Fluent to simulate the impinging <span class="hlt">droplets</span>. The simulations considered different <span class="hlt">droplet</span> sizes (1 mm and 3 mm), tilt angles (0°, 10°, and 45°), <span class="hlt">droplet</span> velocities (1 m/s and 3 m/s), and wetting characteristics (wetting=47° contact angle and non-wetting = 93° contact angle). Results showed that the spread factor (maximum <span class="hlt">droplet</span> diameter during impact divided by the initialmore » <span class="hlt">droplet</span> diameter) decreased with increasing inclination angle due to the reduced normal force on the surface. The hydrophilic surface yielded greater spread factors than the hydrophobic surface in all cases. With regard to impact forces, the greater surface tilt angles yielded lower normal forces, but higher shear forces. Experiments showed that the experimentally observed spread factor (maximum <span class="hlt">droplet</span> diameter during impact divided by the initial <span class="hlt">droplet</span> diameter) was significantly larger than the simulated spread factor. Observed spread factors were on the order of 5 - 6 for <span class="hlt">droplet</span> velocities of ~3 m/s, whereas the simulated spread factors were on the order of 2. <span class="hlt">Droplets</span> were observed to be mobile following impact only for the cases with 45° tilt angle, which matched the simulations. An interesting phenomenon that was observed was that shortly after being released from the nozzle, the water <span class="hlt">droplet</span> oscillated (like a trampoline) due to the "snapback" caused by the surface tension of the water <span class="hlt">droplet</span> being released from the nozzle. This oscillation impacted the velocity immediately after the release. Future work should evaluate the impact of parameters such as tilt angle and surface wettability on the impact of particle/soiling uptake and removal to investigate ways that</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('https://www.ncbi.nlm.nih.gov/pubmed/18386855','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/18386855"><span>Ultrafast <span class="hlt">charge-transfer-to-solvent</span> dynamics of iodide in tetrahydrofuran. 2. Photoinduced electron transfer to counterions in solution.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bragg, Arthur E; Schwartz, Benjamin J</p> <p>2008-04-24</p> <p>The excited states of atomic anions in liquids are bound only by the polarization of the surrounding <span class="hlt">solvent</span>. Thus, the electron-detachment process following excitation to one of these <span class="hlt">solvent</span>-bound states, known as <span class="hlt">charge-transfer-to-solvent</span> (CTTS) states, provides a useful probe of <span class="hlt">solvent</span> structure and dynamics. These transitions and subsequent relaxation dynamics also are influenced by other factors that alter the solution environment local to the CTTS anion, including the presence of cosolutes, cosolvents, and other ions. In this paper, we examine the ultrafast CTTS dynamics of iodide in liquid tetrahydrofuran (THF) with a particular focus on how the <span class="hlt">solvent</span> dynamics and the CTTS electron-ejection process are altered in the presence of various counterions. In weakly polar <span class="hlt">solvents</span> such as THF, iodide salts can be strongly ion-paired in solution; the steady-state UV-visible absorption spectroscopy of various iodide salts in liquid THF indicates that the degree of ion-pairing changes from strong to weak to none as the counterion is switched from Na+ to tetrabutylammonium (t-BA+) to crown-ether-complexed Na+, respectively. In our ultrafast experiments, we have excited the I- CTTS transition of these various iodide salts at 263 nm and probed the dynamics of the CTTS-detached electrons throughout the visible and near-IR. In the previous paper of this series (Bragg, A. E.; Schwartz, B. J. J. Phys. Chem. B 2008, 112, 483-494), we found that for "counterion-free" I- (obtained by complexing Na+ with a crown ether) the CTTS electrons were ejected approximately 6 nm from their partner iodine atoms, the result of significant nonadiabatic coupling between the CTTS excited state and extended electronic states supported by the naturally existing <span class="hlt">solvent</span> cavities in liquid THF, which also serve as pre-existing electron traps. In contrast, for the highly ion-paired NaI/THF system, we find that approximately 90% of the CTTS electrons are "captured" by a nearby Na+ to form (Na</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012SPIE.8223E..2FS','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012SPIE.8223E..2FS"><span>Photoacoustic spectral characterization of perfluorocarbon <span class="hlt">droplets</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Strohm, Eric; Gorelikov, Ivan; Matsuura, Naomi; Kolios, Michael</p> <p>2012-02-01</p> <p>Perfluorocarbon <span class="hlt">droplets</span> containing optical absorbing nanoparticles have been developed for use as theranostic agents (for both imaging and therapy) and as dual-mode contrast agents. <span class="hlt">Droplets</span> can be used as photoacoustic contrast agents, vaporized via optical irradiation, then the resulting bubbles can be used as ultrasound imaging and therapeutic agents. The photoacoustic signals from micron-sized <span class="hlt">droplets</span> containing silica coated gold nanospheres were measured using ultra-high frequencies (100-1000 MHz). The spectra of <span class="hlt">droplets</span> embedded in a gelatin phantom were compared to a theoretical model which calculates the pressure wave from a spherical homogenous liquid undergoing thermoelastic expansion resulting from laser absorption. The location of the spectral features of the theoretical model and experimental spectra were in agreement after accounting for increases in the <span class="hlt">droplet</span> sound speed with frequency. The agreement between experiment and model indicate that <span class="hlt">droplets</span> (which have negligible optical absorption in the visible and infrared spectra by themselves) emitted pressure waves related to the <span class="hlt">droplet</span> composition and size, and was independent of the physical characteristics of the optical absorbing nanoparticles. The diameter of individual <span class="hlt">droplets</span> was calculated using three independent methods: the time domain photoacoustic signal, the time domain pulse echo ultrasound signal, and a fit to the photoacoustic model, then compared to the diameter as measured by optical microscopy. It was found the photoacoustic and ultrasound methods calculated diameters an average of 2.6% of each other, and 8.8% lower than that measured using optical microscopy. The discrepancy between the calculated diameters and the optical measurements may be due to the difficulty in resolving the <span class="hlt">droplet</span> edges after being embedded in the translucent gelatin medium.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19940018563','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19940018563"><span><span class="hlt">Droplet</span> turbulence interactions under subcritical and supercritical conditions</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Coy, E. B.; Greenfield, S. C.; Ondas, M. S.; Song, Y.-H.; Spegar, T. D.; Santavicca, D. A.</p> <p>1993-01-01</p> <p>The goal of this research is to experimentally characterize the behavior of <span class="hlt">droplets</span> in vaporizing liquid sprays under conditions typical of those encountered in high pressure combustion systems such as liquid fueled rocket engines. Of particular interest are measurements of <span class="hlt">droplet</span> drag, <span class="hlt">droplet</span> heating, <span class="hlt">droplet</span> vaporization, <span class="hlt">droplet</span> distortion, and secondary <span class="hlt">droplet</span> breakup, under both subcritical and supercritical conditions. The paper presents a brief description of the specific accomplishments which have been made over the past year.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25325619','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25325619"><span>A <span class="hlt">solvent</span> replenishment solution for managing evaporation of biochemical reactions in air-matrix digital microfluidics devices.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Jebrail, Mais J; Renzi, Ronald F; Sinha, Anupama; Van De Vreugde, Jim; Gondhalekar, Carmen; Ambriz, Cesar; Meagher, Robert J; Branda, Steven S</p> <p>2015-01-07</p> <p>Digital microfluidics (DMF) is a powerful technique for sample preparation and analysis for a broad range of biological and chemical applications. In many cases, it is desirable to carry out DMF on an open surface, such that the matrix surrounding the <span class="hlt">droplets</span> is ambient air. However, the utility of the air-matrix DMF format has been severely limited by problems with <span class="hlt">droplet</span> evaporation, especially when the <span class="hlt">droplet</span>-based biochemical reactions require high temperatures for long periods of time. We present a simple solution for managing evaporation in air-matrix DMF: just-in-time replenishment of the reaction volume using <span class="hlt">droplets</span> of <span class="hlt">solvent</span>. We demonstrate that this solution enables DMF-mediated execution of several different biochemical reactions (RNA fragmentation, first-strand cDNA synthesis, and PCR) over a range of temperatures (4-95 °C) and incubation times (up to 1 h or more) without use of oil, humidifying chambers, or off-chip heating modules. Reaction volumes and temperatures were maintained roughly constant over the course of each experiment, such that the reaction kinetics and products generated by the air-matrix DMF device were comparable to those of conventional benchscale reactions. This simple yet effective solution for evaporation management is an important advance in developing air-matrix DMF for a wide variety of new, high-impact applications, particularly in the biomedical sciences.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1225865','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1225865"><span>A <span class="hlt">solvent</span> replenishment solution for managing evaporation of biochemical reactions in air-matrix digital microfluidics devices</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>Jebrail, Mais J.; Renzi, Ronald F.; Sinha, Anupama</p> <p></p> <p>Digital microfluidics (DMF) is a powerful technique for sample preparation and analysis for a broad range of biological and chemical applications. In many cases, it is desirable to carry out DMF on an open surface, such that the matrix surrounding the <span class="hlt">droplets</span> is ambient air. However, the utility of the air-matrix DMF format has been severely limited by problems with <span class="hlt">droplet</span> evaporation, especially when the <span class="hlt">droplet</span>-based biochemical reactions require high temperatures for long periods of time. We present a simple solution for managing evaporation in air-matrix DMF: just-in-time replenishment of the reaction volume using <span class="hlt">droplets</span> of <span class="hlt">solvent</span>. We demonstrate thatmore » this solution enables DMF-mediated execution of several different biochemical reactions (RNA fragmentation, first-strand cDNA synthesis, and PCR) over a range of temperatures (4–95 °C) and incubation times (up to 1 h or more) without use of oil, humidifying chambers, or off-chip heating modules. Reaction volumes and temperatures were maintained roughly constant over the course of each experiment, such that the reaction kinetics and products generated by the air-matrix DMF device were comparable to those of conventional benchscale reactions. As a result, this simple yet effective solution for evaporation management is an important advance in developing air-matrix DMF for a wide variety of new, high-impact applications, particularly in the biomedical sciences.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1225865-solvent-replenishment-solution-managing-evaporation-biochemical-reactions-air-matrix-digital-microfluidics-devices','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1225865-solvent-replenishment-solution-managing-evaporation-biochemical-reactions-air-matrix-digital-microfluidics-devices"><span>A <span class="hlt">solvent</span> replenishment solution for managing evaporation of biochemical reactions in air-matrix digital microfluidics devices</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Jebrail, Mais J.; Renzi, Ronald F.; Sinha, Anupama; ...</p> <p>2014-10-01</p> <p>Digital microfluidics (DMF) is a powerful technique for sample preparation and analysis for a broad range of biological and chemical applications. In many cases, it is desirable to carry out DMF on an open surface, such that the matrix surrounding the <span class="hlt">droplets</span> is ambient air. However, the utility of the air-matrix DMF format has been severely limited by problems with <span class="hlt">droplet</span> evaporation, especially when the <span class="hlt">droplet</span>-based biochemical reactions require high temperatures for long periods of time. We present a simple solution for managing evaporation in air-matrix DMF: just-in-time replenishment of the reaction volume using <span class="hlt">droplets</span> of <span class="hlt">solvent</span>. We demonstrate thatmore » this solution enables DMF-mediated execution of several different biochemical reactions (RNA fragmentation, first-strand cDNA synthesis, and PCR) over a range of temperatures (4–95 °C) and incubation times (up to 1 h or more) without use of oil, humidifying chambers, or off-chip heating modules. Reaction volumes and temperatures were maintained roughly constant over the course of each experiment, such that the reaction kinetics and products generated by the air-matrix DMF device were comparable to those of conventional benchscale reactions. As a result, this simple yet effective solution for evaporation management is an important advance in developing air-matrix DMF for a wide variety of new, high-impact applications, particularly in the biomedical sciences.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5947303','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5947303"><span>Nucleophilic Substitution in Solution: Activation Strain Analysis of Weak and Strong <span class="hlt">Solvent</span> Effects</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Hamlin, Trevor A.; van Beek, Bas; Wolters, Lando P.</p> <p>2018-01-01</p> <p>Abstract We have quantum chemically studied the effect of various polar and apolar <span class="hlt">solvents</span> on the shape of the potential energy surface (PES) of a diverse collection of archetypal nucleophilic substitution reactions at carbon, silicon, phosphorus, and arsenic by using density functional theory at the OLYP/TZ2P level. In the gas phase, all our model SN2 reactions have single‐well PESs, except for the nucleophilic substitution reaction at carbon (SN2@C), which has a double‐well energy profile. The presence of the <span class="hlt">solvent</span> can have a significant effect on the shape of the PES and, thus, on the nature of the SN2 process. Solvation energies, <span class="hlt">charges</span> on the nucleophile or leaving group, and structural features are compared for the various SN2 reactions in a spectrum of <span class="hlt">solvents</span>. We demonstrate how solvation can change the shape of the PES, depending not only on the polarity of the <span class="hlt">solvent</span>, but also on how the <span class="hlt">charge</span> is distributed over the interacting molecular moieties during different stages of the reaction. In the case of a nucleophilic substitution at three‐coordinate phosphorus, the reaction can be made to proceed through a single‐well [no transition state (TS)], bimodal barrier (two TSs), and then through a unimodal transition state (one TS) simply by increasing the polarity of the <span class="hlt">solvent</span>. PMID:29457865</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1988JCrGr..90...39T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1988JCrGr..90...39T"><span>Mechanism of protein precipitation and stabilization by co-<span class="hlt">solvents</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Timasheff, Serge N.; Arakawa, Tsutomu</p> <p>1988-07-01</p> <p>The interactions between proteins and a number of substances which, when present at high concentration, stabilize or precipitate proteins, have been analyzed in terms of the preferential interactions of these co-<span class="hlt">solvents</span> with proteins. In all cases, stabilization or precipitation was accompanied by preferential exclusion of the co-<span class="hlt">solvent</span> from the immediate domain of the protein, i.e., preferential hydration of the protein. This means that addition of the co-<span class="hlt">solvent</span> to the aqueous protein solution increased the chemical potentials of both components. The thermodynamic interaction parameters derived from such data make it possible to calculate the salting out constant, Ks, as well as to construct a phase isotherm for any given <span class="hlt">solvent</span> mixture which indicates the limiting protein solubility. The salting-out effect can be decomposed into contributions from non-specific preferential exclusion and specific binding of the ligand to the protein, the balance leading to solubilization or precipitation. In reactions, such as denaturation, the effect of co-<span class="hlt">solvent</span> on the reaction depends on the difference in the preferential interactions of the two end states of the protein. Principal sources of preferential exclusion have been identified as steric exclusion, increase of the surface tension of water by the co-<span class="hlt">solvent</span>, repulsion by <span class="hlt">charged</span> loci on the protein and solvophobicity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1198576-lipidomic-proteomic-analysis-caenorhabditis-elegans-lipid-droplets-identification-acs-lipid-droplet-associated-protein','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1198576-lipidomic-proteomic-analysis-caenorhabditis-elegans-lipid-droplets-identification-acs-lipid-droplet-associated-protein"><span>Lipidomic and proteomic analysis of Caenorhabditis elegans lipid <span class="hlt">droplets</span> and identification of ACS-4 as a lipid <span class="hlt">droplet</span>-associated protein</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Vrablik, Tracy L.; Petyuk, Vladislav A.; Larson, Emily M.; ...</p> <p>2015-06-27</p> <p>Lipid <span class="hlt">droplets</span> are cytoplasmic organelles that store neutral lipids for membrane synthesis and energy reserves. In this study, we characterized the lipid and protein composition of purified Caenorhabditis elegans lipid <span class="hlt">droplets</span>. These lipid <span class="hlt">droplets</span> are composed mainly of triacylglycerols, surrounded by a phospholipid monolayer composed primarily of phosphatidylcholine and phosphatidylethanolamine. The fatty acid composition of the triacylglycerols is rich in fatty acid species obtained from the dietary Escherichia coli, including cyclopropane fatty acids and cis-vaccenic acid. Unlike other organisms, C. elegans lipid <span class="hlt">droplets</span> contain very little cholesterol or cholesterol esters. Comparison of the lipid <span class="hlt">droplet</span> proteomes of wild type andmore » high-fat daf-2 mutant strains shows a very similar proteome in both strains, except that the most abundant protein in the C. elegans lipid <span class="hlt">droplet</span> proteome, MDT-28, is relatively less abundant in lipid <span class="hlt">droplets</span> isolated from daf-2 mutants. Functional analysis of lipid <span class="hlt">droplet</span> proteins identified in our proteomic studies indicated an enrichment of proteins required for growth and fat homeostasis in C. elegans. Finally, we confirmed the localization of one of the newly identified lipid <span class="hlt">droplet</span> proteins, ACS-4. We found that ACS-4 localizes to the surface of lipid <span class="hlt">droplets</span> in the C. elegans intestine and skin. This study bolsters C. elegans as a model to study the dynamics and functions of lipid <span class="hlt">droplets</span> in a multicellular organism.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19890010942','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19890010942"><span><span class="hlt">Droplet</span> combustion at reduced gravity</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Dryer, F. L.; Williams, F. A.</p> <p>1988-01-01</p> <p>The current work involves theoretical analyses of the effects identified, experiments in the NASA Lewis drop towers performed in the middeck areas of the Space Shuttle. In addition, there is laboratory work associated with the design of the flight apparatus. Calculations have shown that some of the test-matrix data can be obtained in drop towers, and some are achievable only in the space experiments. The apparatus consists of a <span class="hlt">droplet</span> dispensing device (syringes), a <span class="hlt">droplet</span> positioning device (opposing, retractable, hollow needles), a <span class="hlt">droplet</span> ignition device (two matched pairs of retractable spark electrodes), gas and liquid handling systems, a data acquisition system (mainly giving motion-picture records of the combustion in two orthogonal views, one with backlighting for <span class="hlt">droplet</span> resolution), and associated electronics.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29457909','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29457909"><span><span class="hlt">Droplet</span> Translation Actuated by Photoelectrowetting.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Palma, Cesar; Deegan, Robert D</p> <p>2018-03-13</p> <p>In traditional electrowetting-on-dielectric (EWOD) devices, <span class="hlt">droplets</span> are moved about a substrate using electric fields produced by an array of discrete electrodes. In this study, we show that a drop can be driven across a substrate with a localized light beam by exploiting the photoelectrowetting (PEW) effect, a light-activated variant of EWOD. <span class="hlt">Droplet</span> transport actuated by PEW eliminates the need for electrode arrays and the complexities entailed in their fabrication and control, and offers a new approach for designing lab-on-a-chip applications. We report measurements of the maximum <span class="hlt">droplet</span> speed as a function of frequency and magnitude of the applied bias, intensity of illumination, volume of the <span class="hlt">droplet</span>, and viscosity and also introduce a model that reproduces these data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28222256','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28222256"><span>Dynamic Melting of Freezing <span class="hlt">Droplets</span> on Ultraslippery Superhydrophobic Surfaces.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Chu, Fuqiang; Wu, Xiaomin; Wang, Lingli</p> <p>2017-03-08</p> <p>Condensed <span class="hlt">droplet</span> freezing and freezing <span class="hlt">droplet</span> melting phenomena on the prepared ultraslippery superhydrophobic surface were observed and discussed in this study. Although the freezing delay performance of the surface is common, the melting of the freezing <span class="hlt">droplets</span> on the surface is quite interesting. Three self-propelled movements of the melting <span class="hlt">droplets</span> (ice- water mixture) were found including the <span class="hlt">droplet</span> rotating, the <span class="hlt">droplet</span> jumping, and the <span class="hlt">droplet</span> sliding. The melting <span class="hlt">droplet</span> rotating, which means that the melting <span class="hlt">droplet</span> rotates spontaneously on the superhydrophobic surface like a spinning top, is first reported in this study and may have some potential applications in various engineering fields. The melting <span class="hlt">droplet</span> jumping and sliding are similar to those occurring during condensation but have larger size scale and motion scale, as the melting <span class="hlt">droplets</span> have extra-large specific surface area with much more surface energy available. These self-propelled movements make all the melting <span class="hlt">droplets</span> on the superhydrophobic surface dynamic, easily removed, which may be promising for the anti-icing/frosting applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19870034539&hterms=silicone+sheet&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dsilicone%2Bsheet','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19870034539&hterms=silicone+sheet&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dsilicone%2Bsheet"><span>Heat transfer studies on the liquid <span class="hlt">droplet</span> radiator</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Mattick, A. T.; Nelson, M.</p> <p>1987-01-01</p> <p>This paper examines radiation transfer in the <span class="hlt">droplet</span> sheet of a liquid <span class="hlt">droplet</span> radiator including non-isotropic scattering by the <span class="hlt">droplets</span>. Non-isotropic scattering becomes significant for small <span class="hlt">droplets</span> (diameter less than 0.1 mm) and for low emissivity liquids. For <span class="hlt">droplets</span> with an emittance of 0.1 and for a <span class="hlt">droplet</span> sheet optical depth or 5, the radiated power varies by about 12 percent, depending on whether scattering is predominantly forward or backward. An experimental measurement of the power emitted by a cylindrical cloud of heated <span class="hlt">droplets</span> of silicone fluid is also reported. The measured cloud emissivity correlates, within experimental error, with the analytical model.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018EML....14...79L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018EML....14...79L"><span>Electrophoretic kinetics of concentrated TiO2 nanoparticle suspensions in aprotic <span class="hlt">solvent</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, So-Yeon; Yim, Jung-Ryoul; Lee, Se-Hee; Choi, In-Suk; Nam, Ki Tae; Joo, Young-Chang</p> <p>2018-01-01</p> <p>We studied the dependences of the concentration of additive and particle size on the electrophoretic mobility of TiO2 nanoparticles. A high concentration of TiO2 nanoparticles was dispersed in aprotic <span class="hlt">solvent</span>, which is similar to the operating conditions of electrophoretic applications. Because spectroscopy has limits to measuring the electrophoretic mobility of concentrated suspensions in aprotic <span class="hlt">solvents</span>, we developed a new measurement to determine the electrophoretic mobility of particles using the reflectance change according to the motion of the particles. TiO2 nanoparticles with sizes of 31 nm to 164 nm were synthesized by hydrolysis and were dispersed in cyclohexanone with a dye (Sudan Black B) for use in the new measurement method. In a concentrated suspension in aprotic <span class="hlt">solvent</span>, the mobility of the particles was proportional to the dye concentration and was inversely proportional to the size of the particles. This infers that the particle size influences the drag force rather than the surface <span class="hlt">charge</span>, and therefore, to increase the mobility by changing the surface <span class="hlt">charge</span>, an additive is effective. [Figure not available: see fulltext.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19830012219','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19830012219"><span>Test results of modified electrical <span class="hlt">charged</span> particle generator for application to fog dispersal</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Frost, W.; Huang, K. H.</p> <p>1983-01-01</p> <p>Modifications to a <span class="hlt">charged</span> particle generator for use in fog dispersal applications were made and additional testing carried out. The modified nozzle, however, did not work as planned, and reported results are the unmodified nozzle. The addition of a positive displacement pump to supply the liquid water was highly successful. Measurements of the generator output current were made with a cylindrical collector system as well as with the needle probe used in previous studies. Measurements with the cylindrical collector and the needle probe showed identical agreement within the variability of the experiment. A high-voltage prove was purchased, and measurements of the corona voltage as well as the voltage variation in the <span class="hlt">charged</span> particle jet were made. Electric fields in the vertical direction on the order of 1,000,000 v/m were measured. The voltage distribution along the centerline of the jet was compared with the numerical solutions of the Poisson equation and showed very good agreement. Velocity measurements using a pitot tube were made. The resulting measurements were compared with theoretical and other reported experimental results. The measured data showed the appropriate trends and agreed well with reported results. Based on the measured current-to-mass ratio from the <span class="hlt">charged</span> particle generator, a calculation of the average <span class="hlt">droplet</span> size was made. <span class="hlt">Droplet</span> sizes were estimated to range between 0.8 and 0.4 microns. Using measured data, an analysis of the height to which the <span class="hlt">droplet</span> can be dispersed by the <span class="hlt">charged</span> particle generator was made. Although the mathematical model is highly simplified, the results indicated that particles would achieve heights on the order of 80 m.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhFl...29j3305C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhFl...29j3305C"><span>Collisions of <span class="hlt">droplets</span> on spherical particles</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Charalampous, Georgios; Hardalupas, Yannis</p> <p>2017-10-01</p> <p>Head-on collisions between <span class="hlt">droplets</span> and spherical particles are examined for water <span class="hlt">droplets</span> in the diameter range between 170 μm and 280 μm and spherical particles in the diameter range between 500 μm and 2000 μm. The <span class="hlt">droplet</span> velocities range between 6 m/s and 11 m/s, while the spherical particles are fixed in space. The Weber and Ohnesorge numbers and ratio of <span class="hlt">droplet</span> to particle diameter were between 92 < We < 1015, 0.0070 < Oh < 0.0089, and 0.09 < Ω < 0.55, respectively. The <span class="hlt">droplet</span>-particle collisions are first quantified in terms of the outcome. In addition to the conventional deposition and splashing regimes, a regime is observed in the intermediate region, where the <span class="hlt">droplet</span> forms a stable crown, which does not breakup but propagates along the particle surface and passes around the particle. This regime is prevalent when the <span class="hlt">droplets</span> collide on small particles. The characteristics of the collision at the onset of rim instability are also described in terms of the location of the film on the particle surface and the orientation and length of the ejected crown. Proper orthogonal decomposition identified that the first 2 modes are enough to capture the overall morphology of the crown at the splashing threshold.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016APS..DFDM36008K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016APS..DFDM36008K"><span>Three dimensional force balance of asymmetric <span class="hlt">droplets</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kim, Yeseul; Lim, Su Jin; Cho, Kun; Weon, Byung Mook</p> <p>2016-11-01</p> <p>An equilibrium contact angle of a <span class="hlt">droplet</span> is determined by a horizontal force balance among vapor, liquid, and solid, which is known as Young's law. Conventional wetting law is valid only for axis-symmetric <span class="hlt">droplets</span>, whereas real <span class="hlt">droplets</span> are often asymmetric. Here we show that three-dimensional geometry must be considered for a force balance for asymmetric <span class="hlt">droplets</span>. By visualizing asymmetric <span class="hlt">droplets</span> placed on a free-standing membrane in air with X-ray microscopy, we are able to identify that force balances in one side and in other side control pinning behaviors during evaporation of <span class="hlt">droplets</span>. We find that X-ray microscopy is powerful for realizing the three-dimensional force balance, which would be essential in interpretation and manipulation of wetting, spreading, and drying dynamics for asymmetric <span class="hlt">droplets</span>. This research was supported by Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (NRF-2016R1D1A1B01007133).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015APS..DFDE35001G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015APS..DFDE35001G"><span>Gel-like double-emulsion <span class="hlt">droplets</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Guzowski, Jan; Korczyk, Piotr; Garstecki, Piotr; Stone, Howard</p> <p>2015-11-01</p> <p>We experimentally study the problem of packing of micro-<span class="hlt">droplets</span> inside a <span class="hlt">droplet</span> of another immiscible liquid phase. We use microfluidics to encapsulate multiple monodisperse aqueous segments inside a drop of oil. For small numbers N (N<10) of the aqueous <span class="hlt">droplets</span> and at their volume fraction in oil exceeding the close-packing threshold we observe multiple metastable structures with well-defined point-group symmetries. We attribute the observed metastability to the deformability of the <span class="hlt">droplets</span> which leads to effective many-body interactions and energy barriers for rearrangement. By changing the composition of the oil phase we find that when the surface tensions of the <span class="hlt">droplets</span> and of the encapsulating phase are comparable, the energy barriers are high enough to trap elongated structures or even linear chains, independently of N. However, when the surface tension of the encapsulating phase is much larger than that of the <span class="hlt">droplets</span>, non-spherical morphologies are stable only at sufficiently high N. In such a case multiple internal interfaces can hold stresses and prevent relaxation of the global deformations which leads to a plastic, gel-like behavior. Our findings can serve as guidelines for synthesis of functional particles as well as for designing biomimetic materials, e.g. for tissue engineering. J.G. acknowledges financial support from Polish Ministry of Science provided within the framework Mobility Plus.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1415438','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1415438"><span>Methods for producing thin film <span class="hlt">charge</span> selective transport layers</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>Hammond, Scott Ryan; Olson, Dana C.; van Hest, Marinus Franciscus Antonius Maria</p> <p></p> <p>Methods for producing thin film <span class="hlt">charge</span> selective transport layers are provided. In one embodiment, a method for forming a thin film <span class="hlt">charge</span> selective transport layer comprises: providing a precursor solution comprising a metal containing reactive precursor material dissolved into a complexing <span class="hlt">solvent</span>; depositing the precursor solution onto a surface of a substrate to form a film; and forming a <span class="hlt">charge</span> selective transport layer on the substrate by annealing the film.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27983858','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27983858"><span>Interaction of <span class="hlt">Charged</span> Patchy Protein Models with Like-<span class="hlt">Charged</span> Polyelectrolyte Brushes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yigit, Cemil; Kanduč, Matej; Ballauff, Matthias; Dzubiella, Joachim</p> <p>2017-01-10</p> <p>We study the adsorption of <span class="hlt">charged</span> patchy particle models (CPPMs) on a thin film of a like-<span class="hlt">charged</span> and dense polyelectrolyte (PE) brush (of 50 monomers per chain) by means of implicit-<span class="hlt">solvent</span>, explicit-salt Langevin dynamics computer simulations. Our previously introduced set of CPPMs embraces well-defined one- and two-patched spherical globules, each of the same net <span class="hlt">charge</span> and (nanometer) size, with mono- and multipole moments comparable to those of small globular proteins. We focus on electrostatic effects on the adsorption far away from the isoelectric point of typical proteins, i.e., where <span class="hlt">charge</span> regulation plays no role. Despite the same net <span class="hlt">charge</span> of the brush and globule, we observe large binding affinities up to tens of the thermal energy, k B T, which are enhanced by decreasing salt concentration and increasing <span class="hlt">charge</span> of the patch(es). Our analysis of the distance-resolved potentials of mean force together with a phenomenological description of all leading interaction contributions shows that the attraction is strongest at the brush surface, driven by multipolar, Born (self-energy), and counterion-release contributions, dominating locally over the monopolar and steric repulsions.</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('http://hdl.handle.net/2060/19960008394','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19960008394"><span>Combustion of interacting <span class="hlt">droplet</span> arrays in a microgravity environment</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Dietrich, Daniel L.</p> <p>1995-01-01</p> <p>This research program involves the study of one and two dimensional arrays of <span class="hlt">droplets</span> in a buoyant-free environment. The purpose of the work is to extend the database and theories that exist for single <span class="hlt">droplets</span> into the regime where <span class="hlt">droplet</span> interactions are important. The eventual goal being to use the results of this work as inputs to models on spray combustion where <span class="hlt">droplets</span> seldom burn individually; instead the combustion history of a <span class="hlt">droplet</span> is strongly influenced by the presence of the neighboring <span class="hlt">droplets</span>. Throughout the course of the work, a number of related aspects of isolated <span class="hlt">droplet</span> combustion have also been investigated. This paper will review our progress in microgravity <span class="hlt">droplet</span> array combustion, advanced diagnostics (specifically L2) applied to isolated <span class="hlt">droplet</span> combustion, and radiative extinction large <span class="hlt">droplet</span> flames. A small-scale <span class="hlt">droplet</span> combustion experiment being developed for the Space Shuttle will also be described.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23404263','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23404263"><span>Encapsulation of single cells on a microfluidic device integrating <span class="hlt">droplet</span> generation with fluorescence-activated <span class="hlt">droplet</span> sorting.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wu, Liang; Chen, Pu; Dong, Yingsong; Feng, Xiaojun; Liu, Bi-Feng</p> <p>2013-06-01</p> <p>Encapsulation of single cells is a challenging task in <span class="hlt">droplet</span> microfluidics due to the random compartmentalization of cells dictated by Poisson statistics. In this paper, a microfluidic device was developed to improve the single-cell encapsulation rate by integrating <span class="hlt">droplet</span> generation with fluorescence-activated <span class="hlt">droplet</span> sorting. After cells were loaded into aqueous <span class="hlt">droplets</span> by hydrodynamic focusing, an on-flight fluorescence-activated sorting process was conducted to isolate <span class="hlt">droplets</span> containing one cell. Encapsulation of fluorescent polystyrene beads was investigated to evaluate the developed method. A single-bead encapsulation rate of more than 98 % was achieved under the optimized conditions. Application to encapsulate single HeLa cells was further demonstrated with a single-cell encapsulation rate of 94.1 %, which is about 200 % higher than those obtained by random compartmentalization. We expect this new method to provide a useful platform for encapsulating single cells, facilitating the development of high-throughput cell-based assays.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27568642','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27568642"><span>A poly(dimethylsiloxane) microfluidic sheet reversibly adhered on a glass plate for creation of emulsion <span class="hlt">droplets</span> for <span class="hlt">droplet</span> digital PCR.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Nakashoji, Yuta; Tanaka, Hironari; Tsukagoshi, Kazuhiko; Hashimoto, Masahiko</p> <p>2017-01-01</p> <p>A PDMS microfluidic chip with T-junction channel geometry, two inlet reservoirs, and one outlet reservoir was reversibly adhered on a glass plate through the viscoelastic properties of PDMS. This formed a detachable microfluidic device for creation of water-in-oil emulsion <span class="hlt">droplets</span> that were used as discrete reaction compartments for the <span class="hlt">droplet</span> digital PCR. The PDMS/glass device could continuously produce monodisperse <span class="hlt">droplets</span> without leakage of fluids using a vacuum-driven autonomous micropumping method. This <span class="hlt">droplet</span> preparation technique only required evacuation of air dissolved in the PDMS before loading of oil and aqueous phases into separate inlet reservoirs. Degassing of the PDMS chip at approximately 300 Pa for 1.5 h in a vacuum desiccator gave 40 000 <span class="hlt">droplets</span> in 80 min, which corresponded to a generation frequency of up to nine <span class="hlt">droplets</span> per second. Over multiple runs the <span class="hlt">droplet</span> creation was very reproducible, and the size reproducibility of generated <span class="hlt">droplets</span> (polydispersity of up to 4.1%) was comparable to that acquired using other microfluidic <span class="hlt">droplet</span> preparation techniques. Because the PDMS chip can be peeled off the glass plate, blocked channels can easily be fixed when they arise, and this extends the lifetime of the chip. Single DNA molecules partitioned into the <span class="hlt">droplets</span> were successfully amplified by PCR. In addition, the <span class="hlt">droplet</span> digital PCR platform allowed absolute quantification of low copy numbers of target DNA, and was robust against instrumental variance. © 2016 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3319864','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3319864"><span>Nanoliter <span class="hlt">droplet</span> vitrification for oocyte cryopreservation</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, Xiaohui; Khimji, Imran; Shao, Lei; Safaee, Hooman; Desai, Khanjan; Keles, Hasan Onur; Gurkan, Umut Atakan; Kayaalp, Emre; Nureddin, Aida; Anchan, Raymond M; Maas, Richard L; Demirci, Utkan</p> <p>2011-01-01</p> <p>Aim Oocyte cryopreservation remains largely experimental, with live birth rates of only 2–4% per thawed oocyte. In this study, we present a nanoliter <span class="hlt">droplet</span> technology for oocyte vitrification. Materials & methods An ejector-based <span class="hlt">droplet</span> vitrification system was designed to continuously cryopreserve oocytes in nanoliter <span class="hlt">droplets</span>. Oocyte survival rates, morphologies and parthenogenetic development after each vitrification step were assessed in comparison with fresh oocytes. Results Oocytes were retrieved after cryoprotectant agent loading/unloading, and nanoliter <span class="hlt">droplet</span> encapsulation showed comparable survival rates to fresh oocytes after 24 h in culture. Also, oocytes recovered after vitrification/thawing showed similar morphologies to those of fresh oocytes. Additionally, the rate of oocyte parthenogenetic activation after nanoliter <span class="hlt">droplet</span> encapsulation was comparable with that observed for fresh oocytes. This nanoliter <span class="hlt">droplet</span> technology enables the vitrification of oocytes at higher cooling and warming rates using lower cryoprotectant agent levels (i.e., 1.4 M ethylene glycol, 1.1 M dimethyl sulfoxide and 1 M sucrose), thus making it a potential technology to improve oocyte cryopreservation outcomes. PMID:22188180</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23020318','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23020318"><span>Affine-response model of molecular solvation of ions: Accurate predictions of asymmetric <span class="hlt">charging</span> free energies.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bardhan, Jaydeep P; Jungwirth, Pavel; Makowski, Lee</p> <p>2012-09-28</p> <p>Two mechanisms have been proposed to drive asymmetric <span class="hlt">solvent</span> response to a solute <span class="hlt">charge</span>: a static potential contribution similar to the liquid-vapor potential, and a steric contribution associated with a water molecule's structure and <span class="hlt">charge</span> distribution. In this work, we use free-energy perturbation molecular-dynamics calculations in explicit water to show that these mechanisms act in complementary regimes; the large static potential (∼44 kJ/mol/e) dominates asymmetric response for deeply buried <span class="hlt">charges</span>, and the steric contribution dominates for <span class="hlt">charges</span> near the solute-<span class="hlt">solvent</span> interface. Therefore, both mechanisms must be included in order to fully account for asymmetric solvation in general. Our calculations suggest that the steric contribution leads to a remarkable deviation from the popular "linear response" model in which the reaction potential changes linearly as a function of <span class="hlt">charge</span>. In fact, the potential varies in a piecewise-linear fashion, i.e., with different proportionality constants depending on the sign of the <span class="hlt">charge</span>. This discrepancy is significant even when the <span class="hlt">charge</span> is completely buried, and holds for solutes larger than single atoms. Together, these mechanisms suggest that implicit-<span class="hlt">solvent</span> models can be improved using a combination of affine response (an offset due to the static potential) and piecewise-linear response (due to the steric contribution).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3470608','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3470608"><span>Affine-response model of molecular solvation of ions: Accurate predictions of asymmetric <span class="hlt">charging</span> free energies</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Bardhan, Jaydeep P.; Jungwirth, Pavel; Makowski, Lee</p> <p>2012-01-01</p> <p>Two mechanisms have been proposed to drive asymmetric <span class="hlt">solvent</span> response to a solute <span class="hlt">charge</span>: a static potential contribution similar to the liquid-vapor potential, and a steric contribution associated with a water molecule's structure and <span class="hlt">charge</span> distribution. In this work, we use free-energy perturbation molecular-dynamics calculations in explicit water to show that these mechanisms act in complementary regimes; the large static potential (∼44 kJ/mol/e) dominates asymmetric response for deeply buried <span class="hlt">charges</span>, and the steric contribution dominates for <span class="hlt">charges</span> near the solute-<span class="hlt">solvent</span> interface. Therefore, both mechanisms must be included in order to fully account for asymmetric solvation in general. Our calculations suggest that the steric contribution leads to a remarkable deviation from the popular “linear response” model in which the reaction potential changes linearly as a function of <span class="hlt">charge</span>. In fact, the potential varies in a piecewise-linear fashion, i.e., with different proportionality constants depending on the sign of the <span class="hlt">charge</span>. This discrepancy is significant even when the <span class="hlt">charge</span> is completely buried, and holds for solutes larger than single atoms. Together, these mechanisms suggest that implicit-<span class="hlt">solvent</span> models can be improved using a combination of affine response (an offset due to the static potential) and piecewise-linear response (due to the steric contribution). PMID:23020318</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20040161236','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20040161236"><span>Bi-Component <span class="hlt">Droplet</span> Combustion in Reduced Gravity</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Shaw, Benjamin D.</p> <p>2004-01-01</p> <p>This research deals with reduced-gravity combustion of bi-component <span class="hlt">droplets</span> initially in the mm size range or larger. The primary objectives of the research are to study the effects of <span class="hlt">droplet</span> internal flows, thermal and solutal Marangoni stresses, and species volatility differences on liquid species transport and overall combustion phenomena (e.g., gas-phase unsteadiness, burning rates, sooting, radiation, and extinction). The research program utilizes a reduced gravity environment so that buoyancy effects are rendered negligible. Use of large <span class="hlt">droplets</span> also facilitates visualization of <span class="hlt">droplet</span> internal flows, which is important for this research. In the experiments, <span class="hlt">droplets</span> composed of low- and high-volatility species are burned. The low-volatility components are initially present in small amounts. As combustion of a <span class="hlt">droplet</span> proceeds, the liquid surface mass fraction of the low-volatility component will increase with time, resulting in a sudden and temporary decrease in <span class="hlt">droplet</span> burning rates as the <span class="hlt">droplet</span> rapidly heats to temperatures close to the boiling point of the low-volatility component. This decrease in burning rates causes a sudden and temporary contraction of the flame. The decrease in burning rates and the flame contraction can be observed experimentally. Measurements of burning rates as well as the onset time for flame contraction allow effective liquid-phase species diffusivities to be calculated, e.g., using asymptotic theory. It is planned that <span class="hlt">droplet</span> internal flows will be visualized in flight and ground-based experiments. In this way, effective liquid species diffusivities can be related to <span class="hlt">droplet</span> internal flow characteristics. This program is a continuation of extensive ground-based experimental and theoretical research on bi-component <span class="hlt">droplet</span> combustion that has been ongoing for several years. The focal point of this program is a flight experiment (Bi-Component <span class="hlt">Droplet</span> Combustion Experiment, BCDCE). This flight experiment is under</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20010074057','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20010074057"><span>Bi-Component <span class="hlt">Droplet</span> Combustion in Reduced Gravity</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Shaw, B. D.</p> <p>2001-01-01</p> <p>This research deals with reduced-gravity combustion of bi-component <span class="hlt">droplets</span> initially in the mm size range or larger. The primary objectives of the research are to study the effects of <span class="hlt">droplet</span> internal flows, thermal and solutal Marangoni stresses, and species volatility differences on liquid species transport and overall combustion phenomena (e.g., gas-phase unsteadiness, burning rates, sooting, radiation, and extinction). The research program utilizes a reduced-gravity environment so that buoyancy effects are rendered negligible. Use of large <span class="hlt">droplets</span> also facilitates visualization of <span class="hlt">droplet</span> internal flows, which is important for this research. In the experiments, <span class="hlt">droplets</span> composed of low- and high-volatility species are burned. The low-volatility components are initially present in small amounts. As combustion of a <span class="hlt">droplet</span> proceeds, the liquid surface mass fraction of the low-volatility component will increase with time, resulting in a sudden and temporary decrease in <span class="hlt">droplet</span> burning rates as the <span class="hlt">droplet</span> rapidly heats to temperatures close to the boiling point of the low-volatility component. This decrease in burning rates causes a sudden and temporary contraction of the flame. The decrease in burning rates and the flame contraction can be observed experimentally. Measurements of burning rates as well as the onset time for flame contraction allow effective liquid-phase species diffusivities to be calculated, e.g., using asymptotic theory. It is planned that <span class="hlt">droplet</span> internal flows will be visualized in future flight and ground-based experiments. In this way, effective liquid species diffusivities can be related to <span class="hlt">droplet</span> internal flow characteristics. This program is a continuation of extensive ground based experimental and theoretical research on bi-component <span class="hlt">droplet</span> combustion that has been ongoing for several years. The focal point of this program is a flight experiment (Bi-Component <span class="hlt">Droplet</span> Combustion Experiment, BCDCE). This flight experiment is under</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17626298','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17626298"><span>Evaluation of the influence of the internal aqueous <span class="hlt">solvent</span> structure on electrostatic interactions at the protein-<span class="hlt">solvent</span> interface by nonlocal continuum electrostatic approach.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Rubinstein, Alexander; Sherman, Simon</p> <p></p> <p>The dielectric properties of the polar <span class="hlt">solvent</span> on the protein-<span class="hlt">solvent</span> interface at small intercharge distances are still poorly explored. To deconvolute this problem and to evaluate the pair-wise electrostatic interaction (PEI) energies of the point <span class="hlt">charges</span> located at the protein-<span class="hlt">solvent</span> interface we used a nonlocal (NL) electrostatic approach along with a static NL dielectric response function of water. The influence of the aqueous <span class="hlt">solvent</span> microstructure (determined by a strong nonelectrostatic correlation effect between water dipoles within the orientational Debye polarization mode) on electrostatic interactions at the interface was studied in our work. It was shown that the PEI energies can be significantly higher than the energies evaluated by the classical (local) consideration, treating water molecules as belonging to the bulk <span class="hlt">solvent</span> with a high dielectric constant. Our analysis points to the existence of a rather extended, effective low-dielectric interfacial water shell on the protein surface. The main dielectric properties of this shell (effective thickness together with distance- and orientation-dependent dielectric permittivity function) were evaluated. The dramatic role of this shell was demonstrated when estimating the protein association rate constants.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/24341760','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/24341760"><span>Construction and manipulation of functional three-dimensional <span class="hlt">droplet</span> networks.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wauer, Tobias; Gerlach, Holger; Mantri, Shiksha; Hill, Jamie; Bayley, Hagan; Sapra, K Tanuj</p> <p>2014-01-28</p> <p>Previously, we reported the manual assembly of lipid-coated aqueous <span class="hlt">droplets</span> in oil to form two-dimensional (2D) networks in which the <span class="hlt">droplets</span> are connected through single lipid bilayers. Here we assemble lipid-coated <span class="hlt">droplets</span> in robust, freestanding 3D geometries: for example, a 14-<span class="hlt">droplet</span> pyramidal assembly. The networks are designed, and each <span class="hlt">droplet</span> is placed in a designated position. When protein pores are inserted in the bilayers between specific constituent <span class="hlt">droplets</span>, electrical and chemical communication pathways are generated. We further describe an improved means to construct 3D <span class="hlt">droplet</span> networks with defined organizations by the manipulation of aqueous <span class="hlt">droplets</span> containing encapsulated magnetic beads. The <span class="hlt">droplets</span> are maneuvered in a magnetic field to form simple construction modules, which are then used to form larger 2D and 3D structures including a 10-<span class="hlt">droplet</span> pyramid. A methodology to construct freestanding, functional 3D <span class="hlt">droplet</span> networks is an important step toward the programmed and automated manufacture of synthetic minimal tissues.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016APS..DFDH16007H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016APS..DFDH16007H"><span>Shock wave-<span class="hlt">droplet</span> interaction</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Habibi Khoshmehr, Hamed; Krechetnikov, Rouslan</p> <p>2016-11-01</p> <p>Disintegration of a liquid <span class="hlt">droplet</span> under the action of a shock wave is experimentally investigated. The shock wave-pulse is electromagnetically generated by discharging a high voltage capacitor into a flat spiral coil, above which an isolated circular metal membrane is placed in a close proximity. The Lorentz force arising due to the eddy current induced in the membrane abruptly accelerates it away from the spiral coil thus generating a shock wave. The liquid <span class="hlt">droplet</span> placed at the center of the membrane, where the maximum deflection occurs, is disintegrated in the process of interaction with the shock wave. The effects of <span class="hlt">droplet</span> viscosity and surface tension on the <span class="hlt">droplet</span> destruction are studied with high-speed photography. Water-glycerol solution at different concentrations is used for investigating the effect of viscosity and various concentrations of water-sugar and water-ethanol solution are used for studying the effect of surface tension. Here we report on how the metamorphoses, which a liquid drop undergoes in the process of interaction with a shock wave, are affected by varied viscosity and surface tension.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23473514','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23473514"><span>Ultrasound-assisted surfactant-enhanced emulsification microextraction based on the solidification of a floating organic <span class="hlt">droplet</span> used for the simultaneous determination of six fungicide residues in juices and red wine.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>You, Xiangwei; Wang, Suli; Liu, Fengmao; Shi, Kaiwei</p> <p>2013-07-26</p> <p>A novel ultrasound-assisted surfactant-enhanced emulsification microextraction technique based on the solidification of a floating organic <span class="hlt">droplet</span> followed by high performance liquid chromatography with diode array detection was developed for simultaneous determination of six fungicide residues in juices and red wine samples. The low-toxicity <span class="hlt">solvent</span>, 1-dodecanol, was used as an extraction <span class="hlt">solvent</span>. For its low density and proper melting point near room temperature, the extractant <span class="hlt">droplet</span> was collected easily by solidifying it at a low temperature. The surfactant, Tween 80, was used as an emulsifier to enhance the dispersion of the water-immiscible extraction <span class="hlt">solvent</span> into an aqueous phase, which hastened the mass-transfer of the analytes. Organic dispersive <span class="hlt">solvent</span> typically required in common dispersive liquid-liquid microextraction methods was not used in the proposed method. Some parameters (e.g., the type and volume of extraction <span class="hlt">solvent</span>, the type and concentration of surfactant, ultrasound extraction time, salt addition, and volume of samples) that affect the extraction efficiency were optimized. The proposed method showed a good linearity within the range of 5μgL(-1)-1000μgL(-1), with the correlation coefficients (γ) higher than 0.9969. The limits of detection for the method ranged from 0.4μgL(-1) to 1.4μgL(-1). Further, this simple, practical, sensitive, and environmentally friendly method was successfully applied to determine the target fungicides in juice and red wine samples. The recoveries of the target fungicides in red wine and fruit juice samples were 79.5%-113.4%, with relative standard deviations that ranged from 0.4% to 12.3%. Copyright © 2013 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016SMaS...25d5007H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016SMaS...25d5007H"><span>Hydrophobic polymer covered by a grating electrode for converting the mechanical energy of water <span class="hlt">droplets</span> into electrical energy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Helseth, L. E.; Guo, X. D.</p> <p>2016-04-01</p> <p>Water contact electric harvesting has a great potential as a new energy technology for powering small-scale electronics, but a better understanding of the dynamics governing the conversion from mechanical to electrical energy on the polymer surfaces is needed. Important questions are how current correlates with <span class="hlt">droplet</span> kinetic energy and what happens to the <span class="hlt">charge</span> dynamics when a large number of <span class="hlt">droplets</span> are incident on the polymer simultaneously. Here we address these questions by studying the current that is generated in an external electrical circuit when water <span class="hlt">droplets</span> impinge on hydrophobic fluorinated ethylene propylene film containing a grating electrode on the back side. <span class="hlt">Droplets</span> moving down an inclined polymer plane exhibit a characteristic periodic current time trace, and it is found that the peak current scales with sine of the inclination angle. For single <span class="hlt">droplets</span> in free fall impinging onto the polymer, it is found that the initial peak current scales with the height of the free fall. The transition from individual <span class="hlt">droplets</span> to a nearly continuous stream was investigated using the spectral density of the current signal. In both regimes, the high frequency content of the spectral density scales as f -2. For low frequencies, the low frequency content at low volume rates was noisy but nearly constant, whereas for high volume rates an increase with frequency is observed. It is demonstrated that the output signal from the system exposed to water <span class="hlt">droplets</span> from a garden hose can be rectified and harvested by a 33 μF capacitor, where the stored energy increases at a rate of about 20 μJ in 100 s.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/22398998-supersonic-laser-induced-jetting-aluminum-micro-droplets','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/22398998-supersonic-laser-induced-jetting-aluminum-micro-droplets"><span>Supersonic laser-induced jetting of aluminum micro-<span class="hlt">droplets</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>Zenou, M.; Additive Manufacturing Lab, Orbotech Ltd., P.O. Box 215, 81101 Yavne; Sa'ar, A.</p> <p></p> <p>The <span class="hlt">droplet</span> velocity and the incubation time of pure aluminum micro-<span class="hlt">droplets</span>, printed using the method of sub-nanosecond laser induced forward transfer, have been measured indicating the formation of supersonic laser-induced jetting. The incubation time and the <span class="hlt">droplet</span> velocity were extracted by measuring a transient electrical signal associated with <span class="hlt">droplet</span> landing on the surface of the acceptor substrate. This technique has been exploited for studying small volume <span class="hlt">droplets</span>, in the range of 10–100 femto-litters for which supersonic velocities were measured. The results suggest elastic propagation of the <span class="hlt">droplets</span> across the donor-to-acceptor gap, a nonlinear deposition dynamics on the surface of themore » acceptor and overall efficient energy transfer from the laser beam to the <span class="hlt">droplets</span>.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23098336','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23098336"><span><span class="hlt">Solvent</span>-driven reductive activation of carbon dioxide by gold anions.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Knurr, Benjamin J; Weber, J Mathias</p> <p>2012-11-14</p> <p>Catalytic activation and electrochemical reduction of CO(2) for the formation of chemically usable feedstock and fuel are central goals for establishing a carbon neutral fuel cycle. The role of <span class="hlt">solvent</span> molecules in catalytic processes is little understood, although <span class="hlt">solvent</span>-solute interactions can strongly influence activated intermediate species. We use vibrational spectroscopy of mass-selected Au(CO(2))(n)(-) cluster ions to probe the solvation of AuCO(2)(-) as a model for a reactive intermediate in the reductive activation of a CO(2) ligand by a single-atom catalyst. For the first few <span class="hlt">solvent</span> molecules, solvation of the complex preferentially occurs at the CO(2) moiety, enhancing reductive activation through polarization of the excess <span class="hlt">charge</span> onto the partially reduced ligand. At higher levels of solvation, direct interaction of additional <span class="hlt">solvent</span> molecules with the Au atom diminishes reduction. The results show how the solvation environment can enhance or diminish the effects of a catalyst, offering design criteria for single-atom catalyst engineering.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012AcAau..75...78T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AcAau..75...78T"><span>Non-equilibrium diffusion combustion of a fuel <span class="hlt">droplet</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tyurenkova, Veronika V.</p> <p>2012-06-01</p> <p>A mathematical model for the non-equilibrium combustion of <span class="hlt">droplets</span> in rocket engines is developed. This model allows to determine the divergence of combustion rate for the equilibrium and non-equilibrium model. Criterion for <span class="hlt">droplet</span> combustion deviation from equilibrium is introduced. It grows decreasing <span class="hlt">droplet</span> radius, accommodation coefficient, temperature and decreases on decreasing diffusion coefficient. Also divergence from equilibrium increases on reduction of <span class="hlt">droplet</span> radius. <span class="hlt">Droplet</span> burning time essentially increases under non-equilibrium conditions. Comparison of theoretical and experimental data shows that to have adequate solution for small <span class="hlt">droplets</span> it is necessary to use the non-equilibrium model.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19930040888&hterms=burnout&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dburnout','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19930040888&hterms=burnout&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3Dburnout"><span>Supercritical <span class="hlt">droplet</span> combustion and related transport phenomena</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Yang, Vigor; Hsieh, K. C.; Shuen, J. S.</p> <p>1993-01-01</p> <p>An overview of recent advances in theoretical analyses of supercritical <span class="hlt">droplet</span> vaporization and combustion is conducted. Both hydrocarbon and cryogenic liquid <span class="hlt">droplets</span> over a wide range of thermodynamic states are considered. Various important high-pressure effects on <span class="hlt">droplet</span> behavior, such as thermodynamic non-ideality, transport anomaly, and property variation, are reviewed. Results indicate that the ambient gas pressure exerts significant control of <span class="hlt">droplet</span> gasification and burning processes through its influence on fluid transport, gas-liquid interfacial thermodynamics, and chemical reactions. The <span class="hlt">droplet</span> gasification rate increases progressively with pressure. However, the data for the overall burnout time exhibit a considerable change in the combustion mechanism at the criticl pressure, mainly as a result of reduced mass diffusivity and latent heat of vaporization with increased pressure. The influence of <span class="hlt">droplet</span> size on the burning characteristics is also noted.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5055387','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5055387"><span>Combinatorial microfluidic <span class="hlt">droplet</span> engineering for biomimetic material synthesis</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Bawazer, Lukmaan A.; McNally, Ciara S.; Empson, Christopher J.; Marchant, William J.; Comyn, Tim P.; Niu, Xize; Cho, Soongwon; McPherson, Michael J.; Binks, Bernard P.; deMello, Andrew; Meldrum, Fiona C.</p> <p>2016-01-01</p> <p>Although <span class="hlt">droplet</span>-based systems are used in a wide range of technologies, opportunities for systematically customizing their interface chemistries remain relatively unexplored. This article describes a new microfluidic strategy for rapidly tailoring emulsion <span class="hlt">droplet</span> compositions and properties. The approach uses a simple platform for screening arrays of <span class="hlt">droplet</span>-based microfluidic devices and couples this with combinatorial selection of the <span class="hlt">droplet</span> compositions. Through the application of genetic algorithms over multiple screening rounds, <span class="hlt">droplets</span> with target properties can be rapidly generated. The potential of this method is demonstrated by creating <span class="hlt">droplets</span> with enhanced stability, where this is achieved by selecting carrier fluid chemistries that promote titanium dioxide formation at the <span class="hlt">droplet</span> interfaces. The interface is a mixture of amorphous and crystalline phases, and the resulting composite <span class="hlt">droplets</span> are biocompatible, supporting in vitro protein expression in their interiors. This general strategy will find widespread application in advancing emulsion properties for use in chemistry, biology, materials, and medicine. PMID:27730209</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015JChPh.143f4905Y','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015JChPh.143f4905Y"><span>Like-<span class="hlt">charged</span> protein-polyelectrolyte complexation driven by <span class="hlt">charge</span> patches</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Yigit, Cemil; Heyda, Jan; Ballauff, Matthias; Dzubiella, Joachim</p> <p>2015-08-01</p> <p>We study the pair complexation of a single, highly <span class="hlt">charged</span> polyelectrolyte (PE) chain (of 25 or 50 monomers) with like-<span class="hlt">charged</span> patchy protein models (CPPMs) by means of implicit-<span class="hlt">solvent</span>, explicit-salt Langevin dynamics computer simulations. Our previously introduced set of CPPMs embraces well-defined zero-, one-, and two-patched spherical globules each of the same net <span class="hlt">charge</span> and (nanometer) size with mono- and multipole moments comparable to those of globular proteins with similar size. We observe large binding affinities between the CPPM and the like-<span class="hlt">charged</span> PE in the tens of the thermal energy, kBT, that are favored by decreasing salt concentration and increasing <span class="hlt">charge</span> of the patch(es). Our systematic analysis shows a clear correlation between the distance-resolved potentials of mean force, the number of ions released from the PE, and CPPM orientation effects. In particular, we find a novel two-site binding behavior for PEs in the case of two-patched CPPMs, where intermediate metastable complex structures are formed. In order to describe the salt-dependence of the binding affinity for mainly dipolar (one-patched) CPPMs, we introduce a combined counterion-release/Debye-Hückel model that quantitatively captures the essential physics of electrostatic complexation in our systems.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/19930011006','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/19930011006"><span>Laser diagnostics for microgravity <span class="hlt">droplet</span> studies</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Winter, Michael</p> <p>1993-01-01</p> <p>Rapid advances have recently been made in numerical simulation of <span class="hlt">droplet</span> combustion under microgravity conditions, while experimental capabilities remain relatively primitive. Calculations can now provide detailed information on mass and energy transport, complex gas-phase chemistry, multi-component molecular diffusion, surface evaporation and heterogeneous reaction, which provides a clearer picture of both quasi-steady as well as dynamic behavior of <span class="hlt">droplet</span> combustion. Experiments concerning these phenomena typically result in pictures of the burning <span class="hlt">droplets</span>, and the data therefrom describe <span class="hlt">droplet</span> surface regression along with flame and soot shell position. With much more precise, detailed, experimental diagnostics, significant gains could be made on the dynamics and flame structural changes which occur during <span class="hlt">droplet</span> combustion. Since microgravity experiments become increasingly more expensive as they progress from drop towers and flights to spaceborne experiments, there is a great need to maximize the information content from these experiments. Sophisticated measurements using laser diagnostics on individual <span class="hlt">droplets</span> and combustion phenomena are now possible. These include measuring flow patterns and temperature fields within <span class="hlt">droplets</span>, vaporization rates and vaporization enhancement, radical species profiling in flames and gas-phase flow-tagging velocimetry. Although these measurements are sophisticated, they have undergone maturation to the degree where with some development, they are applicable to studies of microgravity <span class="hlt">droplet</span> combustion. This program beginning in September of 1992, will include a series of measurements in the NASA Learjet, KC-135 and Drop Tower facilities for investigating the range of applicability of these diagnostics while generating and providing fundamental data to ongoing NASA research programs in this area. This program is being conducted in collaboration with other microgravity investigators and is aimed toward supplementing</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('http://adsabs.harvard.edu/abs/2015EL....11054002S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015EL....11054002S"><span>Janus <span class="hlt">droplet</span> as a catalytic micromotor</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shklyaev, Sergey</p> <p>2015-06-01</p> <p>Self-propulsion of a Janus <span class="hlt">droplet</span> in a solution of surfactant, which reacts on a half of a drop surface, is studied theoretically. The <span class="hlt">droplet</span> acts as a catalytic motor creating a concentration gradient, which generates its surface-tension-driven motion; the self-propulsion speed is rather high, 60 μ \\text{m/s} and more. This catalytic motor has several advantages over other micromotors: simple manufacturing, easily attained neutral buoyancy. In contrast to a single-fluid <span class="hlt">droplet</span>, which demonstrates a self-propulsion as a result of symmetry breaking instability, for the Janus one no stability threshold exists; hence, the <span class="hlt">droplet</span> radius can be scaled down to micrometers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/21149135-transient-droplet-behavior-droplet-breakup-during-bulk-confined-shear-flow-blends-one-viscoelastic-component-experiments-modelling-simulations','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/21149135-transient-droplet-behavior-droplet-breakup-during-bulk-confined-shear-flow-blends-one-viscoelastic-component-experiments-modelling-simulations"><span>Transient <span class="hlt">Droplet</span> Behavior and <span class="hlt">Droplet</span> Breakup during Bulk and Confined Shear Flow in Blends with One Viscoelastic Component: Experiments, Modelling and Simulations</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>Cardinaels, Ruth; Verhulst, Kristof; Moldenaers, Paula</p> <p>2008-07-07</p> <p>The transient <span class="hlt">droplet</span> deformation and <span class="hlt">droplet</span> orientation after inception of shear, the shape relaxation after cessation of shear and <span class="hlt">droplet</span> breakup during shear, are microscopically studied, both under bulk and confined conditions. The studied blends contain one viscoelastic Boger fluid phase. A counter rotating setup, based on a Paar Physica MCR300, is used for the <span class="hlt">droplet</span> visualisation. For bulk shear flow, it is shown that the <span class="hlt">droplet</span> deformation during startup of shear flow and the shape relaxation after cessation of shear flow are hardly influenced by <span class="hlt">droplet</span> viscoelasticity, even at moderate to high capillary and Deborah numbers. The effects ofmore » <span class="hlt">droplet</span> viscoelasticity only become visible close to the critical conditions and a novel break-up mechanism is observed. Matrix viscoelasticity has a more pronounced effect, causing overshoots in the deformation and significantly inhibiting relaxation. However, different applied capillary numbers prior to cessation of shear flow, with the Deborah number fixed, still result in a single master curve for shape retraction, as in fully Newtonian systems. The long tail in the <span class="hlt">droplet</span> relaxation can be qualitatively described with a phenomenological model for <span class="hlt">droplet</span> deformation, when using a 5-mode Giesekus model for the fluid rheology. It is found that the shear flow history significantly affects the <span class="hlt">droplet</span> shape evolution and the breakup process in blends with one viscoelastic component. Confining a <span class="hlt">droplet</span> between two plates accelerates the <span class="hlt">droplet</span> deformation kinetics, similar to fully Newtonian systems. However, the increased <span class="hlt">droplet</span> deformation, due to wall effects, causes the steady state to be reached at a later instant in time. <span class="hlt">Droplet</span> relaxation is less sensitive to confinement, leading to slower relaxation kinetics only for highly confined <span class="hlt">droplets</span>. For the blend with a viscoelastic <span class="hlt">droplet</span>, a non-monotonous trend is found for the critical capillary number as a function of the confinement ratio. Finally</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017APS..DFDL16010W','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..DFDL16010W"><span>Interface-Resolving Simulation of Collision Efficiency of Cloud <span class="hlt">Droplets</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Wang, Lian-Ping; Peng, Cheng; Rosa, Bodgan; Onishi, Ryo</p> <p>2017-11-01</p> <p>Small-scale air turbulence could enhance the geometric collision rate of cloud <span class="hlt">droplets</span> while large-scale air turbulence could augment the diffusional growth of cloud <span class="hlt">droplets</span>. Air turbulence could also enhance the collision efficiency of cloud <span class="hlt">droplets</span>. Accurate simulation of collision efficiency, however, requires capture of the multi-scale <span class="hlt">droplet</span>-turbulence and <span class="hlt">droplet-droplet</span> interactions, which has only been partially achieved in the recent past using the hybrid direct numerical simulation (HDNS) approach. % where Stokes disturbance flow is assumed. The HDNS approach has two major drawbacks: (1) the short-range <span class="hlt">droplet-droplet</span> interaction is not treated rigorously; (2) the finite-Reynolds number correction to the collision efficiency is not included. In this talk, using two independent numerical methods, we will develop an interface-resolved simulation approach in which the disturbance flows are directly resolved numerically, combined with a rigorous lubrication correction model for near-field <span class="hlt">droplet-droplet</span> interaction. This multi-scale approach is first used to study the effect of finite flow Reynolds numbers on the <span class="hlt">droplet</span> collision efficiency in still air. Our simulation results show a significant finite-Re effect on collision efficiency when the <span class="hlt">droplets</span> are of similar sizes. Preliminary results on integrating this approach in a turbulent flow laden with <span class="hlt">droplets</span> will also be presented. This work is partially supported by the National Science Foundation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014PhFl...26j3302S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014PhFl...26j3302S"><span>Physics of puffing and microexplosion of emulsion fuel <span class="hlt">droplets</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shinjo, J.; Xia, J.; Ganippa, L. C.; Megaritis, A.</p> <p>2014-10-01</p> <p>The physics of water-in-oil emulsion <span class="hlt">droplet</span> microexplosion/puffing has been investigated using high-fidelity interface-capturing simulation. Varying the dispersed-phase (water) sub-<span class="hlt">droplet</span> size/location and the initiation location of explosive boiling (bubble formation), the <span class="hlt">droplet</span> breakup processes have been well revealed. The bubble growth leads to local and partial breakup of the parent oil <span class="hlt">droplet</span>, i.e., puffing. The water sub-<span class="hlt">droplet</span> size and location determine the after-puffing dynamics. The boiling surface of the water sub-<span class="hlt">droplet</span> is unstable and evolves further. Finally, the sub-<span class="hlt">droplet</span> is wrapped by boiled water vapor and detaches itself from the parent oil <span class="hlt">droplet</span>. When the water sub-<span class="hlt">droplet</span> is small, the detachment is quick, and the oil <span class="hlt">droplet</span> breakup is limited. When it is large and initially located toward the parent <span class="hlt">droplet</span> center, the <span class="hlt">droplet</span> breakup is more extensive. For microexplosion triggered by the simultaneous growth of multiple separate bubbles, each explosion is local and independent initially, but their mutual interactions occur at a later stage. The degree of breakup can be larger due to interactions among multiple explosions. These findings suggest that controlling microexplosion/puffing is possible in a fuel spray, if the emulsion-fuel blend and the ambient flow conditions such as heating are properly designed. The current study also gives us an insight into modeling the puffing and microexplosion of emulsion <span class="hlt">droplets</span> and sprays.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19890000034&hterms=multiple+emulsions&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dmultiple%2Bemulsions','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19890000034&hterms=multiple+emulsions&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dmultiple%2Bemulsions"><span>Photopolymerization Of Levitated <span class="hlt">Droplets</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Rembaum, Alan; Rhim, Won-Kyu; Hyson, Michael T.; Chang, Manchium</p> <p>1989-01-01</p> <p>Experimental containerless process combines two established techniques to make variety of polymeric microspheres. In single step, electrostatically-levitated monomer <span class="hlt">droplets</span> polymerized by ultraviolet light. Faster than multiple-step emulsion polymerization process used to make microspheres. <span class="hlt">Droplets</span> suspended in cylindrical quadrupole electrostatic levitator. Alternating electrostatic field produces dynamic potential along axis. Process enables tailoring of microspheres for medical, scientific, and industrial applications.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28447798','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28447798"><span>Spontaneous <span class="hlt">Droplet</span> Motion on a Periodically Compliant Substrate.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Liu, Tianshu; Nadermann, Nichole; He, Zhenping; Strogatz, Steven H; Hui, Chung-Yuen; Jagota, Anand</p> <p>2017-05-23</p> <p><span class="hlt">Droplet</span> motion arises in many natural phenomena, ranging from the familiar gravity-driven slip and arrest of raindrops on windows to the directed transport of <span class="hlt">droplets</span> for water harvesting by plants and animals under dry conditions. Deliberate transportation and manipulation of <span class="hlt">droplets</span> are also important in many technological applications, including <span class="hlt">droplet</span>-based microfluidic chemical reactors and for thermal management. <span class="hlt">Droplet</span> motion usually requires gradients of surface energy or temperature or external vibration to overcome contact angle hysteresis. Here, we report a new phenomenon in which a drying <span class="hlt">droplet</span> placed on a periodically compliant surface undergoes spontaneous, erratic motion in the absence of surface energy gradients and external stimuli such as vibration. By modeling the <span class="hlt">droplet</span> as a mass-spring system on a substrate with periodically varying compliance, we show that the stability of equilibrium depends on the size of the <span class="hlt">droplet</span>. Specifically, if the center of mass of the drop lies at a stable equilibrium point of the system, it will stay there until evaporation reduces its size and this fixed point becomes unstable; with any small perturbation, the <span class="hlt">droplet</span> then moves to one of its neighboring fixed points.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2013APS..DFD.D6005H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2013APS..DFD.D6005H"><span>Identification of viscous <span class="hlt">droplets</span>' physical properties that determine <span class="hlt">droplet</span> behaviors in inertial microfluidics</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hur, Soojung Claire</p> <p>2013-11-01</p> <p>Inertial effects in microfluidic systems have recently recognized as a robust and passive way of focusing and ordering microscale particles and cells continuously. Moreover, theoretical analysis has shown that there exists a force away from channel walls in Poiseuille flow that locates deformable particles closer to the channel center than rigid counterparts. Then, the particle deformability can be extrapolated from the positions of particles with known sizes in the channel. Here, behaviors of various viscous <span class="hlt">droplets</span> in inertial flow were investigated to identify critical properties determining their dynamic lateral position. Fluorinated oil solutions (μ = 1.7 mPas and 5 mPas) containing <span class="hlt">droplets</span> (1mPas< μ<1.3Pas) were injected into a microfluidic channel with a syringe pump (8 < Rc < 50). Interfacial tension between aqueous and oil phases were varied by adding controlled amount of a surfactant. The diameter, a, deformability, Def, and dynamic lateral position, Xeq, were determined using high-speed microscopy. Xeq, was found to correlate with the particle Capillary Number, CaP, regardless of <span class="hlt">droplet</span> viscosities when CaP <0.02 or CaP >0.2, suggesting that the viscous drag from the continuous phase and the interfacial tension were competing factors determining Xeq. Experimental results suggested that (i) interplay among <span class="hlt">droplet</span>'s viscosity, interfacial tension and inertia of carrier fluid determines dynamic lateral position of <span class="hlt">droplets</span> and (ii) the dominant property varies at a different regime.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23885414','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23885414"><span>Development of an imaging system for single <span class="hlt">droplet</span> characterization using a <span class="hlt">droplet</span> generator.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Minov, S Vulgarakis; Cointault, F; Vangeyte, J; Pieters, J G; Hijazi, B; Nuyttens, D</p> <p>2012-01-01</p> <p>The spray <span class="hlt">droplets</span> generated by agricultural nozzles play an important role in the application accuracy and efficiency of plant protection products. The limitations of the non-imaging techniques and the recent improvements in digital image acquisition and processing increased the interest in using high speed imaging techniques in pesticide spray characterisation. The goal of this study was to develop an imaging technique to evaluate the characteristics of a single spray <span class="hlt">droplet</span> using a piezoelectric single <span class="hlt">droplet</span> generator and a high speed imaging technique. Tests were done with different camera settings, lenses, diffusers and light sources. The experiments have shown the necessity for having a good image acquisition and processing system. Image analysis results contributed in selecting the optimal set-up for measuring <span class="hlt">droplet</span> size and velocity which consisted of a high speed camera with a 6 micros exposure time, a microscope lens at a working distance of 43 cm resulting in a field of view of 1.0 cm x 0.8 cm and a Xenon light source without diffuser used as a backlight. For measuring macro-spray characteristics as the <span class="hlt">droplet</span> trajectory, the spray angle and the spray shape, a Macro Video Zoom lens at a working distance of 14.3 cm with a bigger field of view of 7.5 cm x 9.5 cm in combination with a halogen spotlight with a diffuser and the high speed camera can be used.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010ffcd.confE.143K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010ffcd.confE.143K"><span>Fog, plant leaves and deposition of <span class="hlt">droplets</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Konrad, W.; Ebner, M.; Traiser, C.; Roth-Nebelsick, A.</p> <p>2010-07-01</p> <p>For various plants and animals, the accumulation of fog or dew <span class="hlt">droplets</span> constitutes an essential part of their water supply. Understanding how water <span class="hlt">droplets</span> deposited by fog or dew events interact with plant or animal surfaces is essential for gaining insight into the functionality of these surfaces. Besides being interesting within the realm of biology, this knowledge is indispensable for technical applications. Frequently, it is advantageous to know (i) the growth rate of a <span class="hlt">droplet</span> attached by surface tension to a surface which grows due to a given influx of fog particles, (ii) the maximum volume and (iii) the "lifespan" of a <span class="hlt">droplet</span> before it detaches from the surface or starts to slide down along the plant surface, driven by gravity. Starting from principles of physics, we calculate quantitative expressions addressing questions (i) to (iii) for <span class="hlt">droplets</span> which are attached to surfaces characterised by a high degree of symmetry, such as horizontally oriented or inclined planes, sections of spheres, cones and rotationally symmetric crevices. Furthermore, we treat the behaviour of <span class="hlt">droplets</span> attached to a surface of non-constant contact angle. Although real surfaces never meet their geometric idealisations, results based on these often represent suitable and useful approximations to reality. Finally, we apply our results to Stipagrostis sabulicola, a dune grass of the Namib desert which satisfies its water demand solely by capturing fog and dew <span class="hlt">droplets</span>. Pictures taken with a scanning electron microscope show that the stem of S. sabulicola is longitudinally built up by alternating elevated and countersunk strips. Filling gaps in the experimental observation with theoretical speculation, the following picture emerges: Assuming that the elevated strips exhibit a higher contact angle than the countersunk strips, water <span class="hlt">droplets</span> being deposited on the elevated strips are drawn towards the latter. The lower contact angle which prevails there increases the <span class="hlt">droplets</span></p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/20877790','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/20877790"><span>Uptake and withdrawal of <span class="hlt">droplets</span> from carbon nanotubes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Schebarchov, D; Hendy, S C</p> <p>2011-01-01</p> <p>We give an account of recent studies of <span class="hlt">droplet</span> uptake and withdrawal from carbon nanotubes using simple theoretical arguments and molecular dynamics simulations. Firstly, the thermodynamics of <span class="hlt">droplet</span> uptake and release is considered and tested via simulation. We show that the Laplace pressure acting on a <span class="hlt">droplet</span> assists capillary uptake, allowing sufficiently small non-wetting <span class="hlt">droplets</span> to be absorbed. We then demonstrate how the uptake and release of <span class="hlt">droplets</span> of non-wetting fluids can be exploited for the use of carbon nanotubes as nanopipettes. Finally, we extend the Lucas-Washburn model to deal with the dynamics of <span class="hlt">droplet</span> capillary uptake, and again test this by comparison with molecular dynamics simulations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011Nanos...3..134S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011Nanos...3..134S"><span>Uptake and withdrawal of <span class="hlt">droplets</span> from carbon nanotubes</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Schebarchov, D.; Hendy, S. C.</p> <p>2011-01-01</p> <p>We give an account of recent studies of <span class="hlt">droplet</span> uptake and withdrawal from carbon nanotubes using simple theoretical arguments and molecular dynamics simulations. Firstly, the thermodynamics of <span class="hlt">droplet</span> uptake and release is considered and tested via simulation. We show that the Laplace pressure acting on a <span class="hlt">droplet</span> assists capillary uptake, allowing sufficiently small non-wetting <span class="hlt">droplets</span> to be absorbed. We then demonstrate how the uptake and release of <span class="hlt">droplets</span> of non-wetting fluids can be exploited for the use of carbon nanotubes as nanopipettes. Finally, we extend the Lucas-Washburn model to deal with the dynamics of <span class="hlt">droplet</span> capillary uptake, and again test this by comparison with molecular dynamics simulations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28005864','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28005864"><span>Phase rainbow refractometry for accurate <span class="hlt">droplet</span> variation characterization.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Wu, Yingchun; Promvongsa, Jantarat; Saengkaew, Sawitree; Wu, Xuecheng; Chen, Jia; Gréhan, Gérard</p> <p>2016-10-15</p> <p>We developed a one-dimensional phase rainbow refractometer for the accurate trans-dimensional measurements of <span class="hlt">droplet</span> size on the micrometer scale as well as the tiny <span class="hlt">droplet</span> diameter variations at the nanoscale. The dependence of the phase shift of the rainbow ripple structures on the <span class="hlt">droplet</span> variations is revealed. The phase-shifting rainbow image is recorded by a telecentric one-dimensional rainbow imaging system. Experiments on the evaporating monodispersed <span class="hlt">droplet</span> stream show that the phase rainbow refractometer can measure the tiny <span class="hlt">droplet</span> diameter changes down to tens of nanometers. This one-dimensional phase rainbow refractometer is capable of measuring the <span class="hlt">droplet</span> refractive index and diameter, as well as variations.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/435754-modeling-metal-droplet-sprays-spray-forming','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/435754-modeling-metal-droplet-sprays-spray-forming"><span>Modeling metal <span class="hlt">droplet</span> sprays in spray forming</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>Muoio, N.G.; Crowe, C.T.; Fritsching, U.</p> <p>1995-12-31</p> <p>Spray casting is a process whereby a molten metal stream is atomized and deposited on a substrate. The rapid solidification of the metal <span class="hlt">droplets</span> gives rise to a fine grain structure and improved material properties. This paper presents a simulation for the fluid and thermal interaction of the fluid and <span class="hlt">droplets</span> in the spray and the effect on the <span class="hlt">droplet</span> spray pattern. Good agreement is obtained between the measured and predicted <span class="hlt">droplet</span> mass flux distribution in the spray.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016APS..DFDKP1030Z','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016APS..DFDKP1030Z"><span>Impact of Viscous <span class="hlt">Droplets</span> on Superamphiphobic Surfaces</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Zhao, Binyu; Chen, Longquan; Deng, Xu</p> <p>2016-11-01</p> <p>Superamphiphobic coating is promising for various applications in industry, e.g. self-cleaning windows, where the impingement of <span class="hlt">droplets</span> on surfaces is commonly encountered. In this work, we experimentally investigated the impact of <span class="hlt">droplets</span> with similar surface tension (63-72 mN/m) but much different viscosity (1-150 mPa s) on superamphiphobic surfaces. We found that <span class="hlt">droplets</span> can rebound from the superamphiphobic surfaces when the impact velocity is larger than a critical value, which linearly increases with the liquid viscosity. <span class="hlt">Droplet</span> with higher viscosity spreads, retracts slower, and eventually rebounds lower and fewer times than that of low viscous <span class="hlt">droplet</span>. These findings have important implications for surface engineers to use superamphiphobic coatings. Furthermore, we measured the maximum spreading factors for <span class="hlt">droplet</span> impact on superamphiphobic surfaces and proposed a simple model based on energy conversation to describe its relationship to the Weber number and Reynolds number.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21272076','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21272076"><span>Inhalation of expiratory <span class="hlt">droplets</span> in aircraft cabins.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Gupta, J K; Lin, C-H; Chen, Q</p> <p>2011-08-01</p> <p>Airliner cabins have high occupant density and long exposure time, so the risk of airborne infection transmission could be high if one or more passengers are infected with an airborne infectious disease. The <span class="hlt">droplets</span> exhaled by an infected passenger may contain infectious agents. This study developed a method to predict the amount of expiratory <span class="hlt">droplets</span> inhaled by the passengers in an airliner cabin for any flight duration. The spatial and temporal distribution of expiratory <span class="hlt">droplets</span> for the first 3 min after the exhalation from the index passenger was obtained using the computational fluid dynamics simulations. The perfectly mixed model was used for beyond 3 min after the exhalation. For multiple exhalations, the <span class="hlt">droplet</span> concentration in a zone can be obtained by adding the <span class="hlt">droplet</span> concentrations for all the exhalations until the current time with a time shift via the superposition method. These methods were used to determine the amount of <span class="hlt">droplets</span> inhaled by the susceptible passengers over a 4-h flight under three common scenarios. The method, if coupled with information on the viability and the amount of infectious agent in the <span class="hlt">droplet</span>, can aid in evaluating the infection risk. The distribution of the infectious agents contained in the expiratory <span class="hlt">droplets</span> of an infected occupant in an indoor environment is transient and non-uniform. The risk of infection can thus vary with time and space. The investigations developed methods to predict the spatial and temporal distribution of expiratory <span class="hlt">droplets</span>, and the inhalation of these <span class="hlt">droplets</span> in an aircraft cabin. The methods can be used in other indoor environments to assess the relative risk of infection in different zones, and suitable measures to control the spread of infection can be adopted. Appropriate treatment can be implemented for the zone identified as high-risk zones. © 2011 John Wiley & Sons A/S.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17640659','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17640659"><span>Microextraction in a tetrabutylammonium bromide/ammonium sulfate aqueous two-phase system and electrohydrodynamic generation of a micro-<span class="hlt">droplet</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Song, Young Soo; Choi, Young Hoon; Kim, Do Hyun</p> <p>2007-08-31</p> <p>Microextraction of methyl orange in the aqueous two-phase system (ATPS) formed by dissolving tetrabutylammonium bromide (TBAB) and ammonium sulfate (AS) is reported. Methyl orange was transported from the AS-rich phase to TBAB-rich phase across the interface of the two immiscible phases. The electrohydrodynamic effect on the shape of the interface of two immiscible flows was also observed by applying dc voltage at the T-junction of the microchannel and the generation of a <span class="hlt">droplet</span> of AS-rich phase was observed when the potential difference between positive and negative electrodes exceeds a threshold voltage. The minimum voltage necessary for the <span class="hlt">droplet</span> generation depends on pH due to the degree of dissociation and <span class="hlt">charge</span> accumulation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017AIPA....7c5014C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017AIPA....7c5014C"><span><span class="hlt">Droplet</span> ejection and sliding on a flapping film</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chen, Xi; Doughramaji, Nicole; Betz, Amy Rachel; Derby, Melanie M.</p> <p>2017-03-01</p> <p>Water recovery and subsequent reuse are required for human consumption as well as industrial, and agriculture applications. Moist air streams, such as cooling tower plumes and fog, represent opportunities for water harvesting. In this work, we investigate a flapping mechanism to increase <span class="hlt">droplet</span> shedding on thin, hydrophobic films for two vibrational cases (e.g., ± 9 mm and 11 Hz; ± 2 mm and 100 Hz). Two main mechanisms removed water <span class="hlt">droplets</span> from the flapping film: vibrational-induced coalescence/sliding and <span class="hlt">droplet</span> ejection from the surface. Vibrations mobilized <span class="hlt">droplets</span> on the flapping film, increasing the probability of coalescence with neighboring <span class="hlt">droplets</span> leading to faster <span class="hlt">droplet</span> growth. <span class="hlt">Droplet</span> departure sizes of 1-2 mm were observed for flapping films, compared to 3-4 mm on stationary films, which solely relied on gravity for <span class="hlt">droplet</span> removal. Additionally, flapping films exhibited lower percentage area coverage by water after a few seconds. The second removal mechanism, <span class="hlt">droplet</span> ejection was analyzed with respect to surface wave formation and inertia. Smaller <span class="hlt">droplets</span> (e.g., 1-mm diameter) were ejected at a higher frequency which is associated with a higher acceleration. Kinetic energy of the water was the largest contributor to energy required to flap the film, and low energy inputs (i.e., 3.3 W/m2) were possible. Additionally, self-flapping films could enable novel water collection and condensation with minimal energy input.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012AGUFM.A31I..07R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012AGUFM.A31I..07R"><span><span class="hlt">Droplet</span> Growth Kinetics in Various Environments</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Raatikainen, T. E.; Lathem, T. L.; Moore, R.; Lin, J. J.; Cerully, K. M.; Padro, L.; Lance, S.; Cozic, J.; Anderson, B. E.; Nenes, A.</p> <p>2012-12-01</p> <p>The largest uncertainties in the effects of atmospherics aerosols on the global radiation budget are related to their indirect effects on cloud properties (IPCC, the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, 2007). Cloud formation is a kinetic process where the resulting cloud properties depend on aerosol properties and meteorological parameters such as updraft velocity (e.g. McFiggans et al., Atmos. Chem. Phys., 6, 2593-2649, 2006). <span class="hlt">Droplet</span> growth rates are limited by the water vapor diffusion, but additional kinetic limitations, e.g., due to organic surface films, slow solute dissociation or highly viscous or glassy aerosol states have been hypothesized. Significant additional kinetic limitations can lead to increased cloud <span class="hlt">droplet</span> number concentration, thus the effect is similar to those of increased aerosol number concentration or changes in vertical velocity (e.g. Nenes et al., Geophys. Res. Lett., 29, 1848, 2002). There are a few studies where slow <span class="hlt">droplet</span> growth has been observed (e.g. Ruehl et al., Geophys. Res. Lett., 36, L15814, 2009), however, little is currently known about their global occurrence and magnitude. Cloud micro-physics models often describe kinetic limitations by an effective water vapor uptake coefficient or similar parameter. Typically, determining aerosol water vapor uptake coefficients requires experimental observations of <span class="hlt">droplet</span> growth which are interpreted by a numerical <span class="hlt">droplet</span> growth model where the uptake coefficient is an adjustable parameter (e.g. Kolb et al., Atmos. Chem. Phys., 10, 10561-10605, 2010). Such methods have not been practical for high time-resolution or long term field measurements, until a model was recently developed for analyzing <span class="hlt">Droplet</span> Measurement Technologies (DMT) cloud condensation nuclei (CCN) counter data (Raatikainen et al., Atmos. Chem. Phys., 12, 4227-4243, 2012). Model verification experiments showed that the calibration aerosol <span class="hlt">droplet</span> size can be predicted accurately</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1235288-atomistic-molecular-effects-electric-double-layers-high-surface-charges','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1235288-atomistic-molecular-effects-electric-double-layers-high-surface-charges"><span>Atomistic and molecular effects in electric double layers at high surface <span class="hlt">charges</span></span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Templeton, Jeremy Alan; Lee, Jonathan; Mani, Ali</p> <p>2015-06-16</p> <p>Here, the Poisson–Boltzmann theory for electrolytes near a <span class="hlt">charged</span> surface is known to be invalid due to unaccounted physics associated with high ion concentration regimes. In order to investigate this regime, fluids density functional theory (f-DFT) and molecular dynamics (MD) simulations were used to determine electric surface potential as a function of surface <span class="hlt">charge</span>. Based on these detailed computations, for electrolytes with nonpolar <span class="hlt">solvent</span>, the surface potential is shown to depend quadratically on the surface <span class="hlt">charge</span> in the high <span class="hlt">charge</span> limit. We demonstrate that modified Poisson–Boltzmann theories can model this limit if they are augmented with atomic packing densities providedmore » by MD. However, when the <span class="hlt">solvent</span> is a highly polar molecule water an intermediate regime is identified in which a constant capacitance is realized. Simulation results demonstrate the mechanism underlying this regime, and for the salt water system studied here, it persists throughout the range of physically realistic surface <span class="hlt">charge</span> densities so the potential’s quadratic surface <span class="hlt">charge</span> dependence is not obtained.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4801594','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4801594"><span>Spray <span class="hlt">Droplet</span> Characterization from a Single Nozzle by High Speed Image Analysis Using an In-Focus <span class="hlt">Droplet</span> Criterion</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Vulgarakis Minov, Sofija; Cointault, Frédéric; Vangeyte, Jürgen; Pieters, Jan G; Nuyttens, David</p> <p>2016-01-01</p> <p>Accurate spray characterization helps to better understand the pesticide spray application process. The goal of this research was to present the proof of principle of a <span class="hlt">droplet</span> size and velocity measuring technique for different types of hydraulic spray nozzles using a high speed backlight image acquisition and analysis system. As only part of the drops of an agricultural spray can be in focus at any given moment, an in-focus criterion based on the gray level gradient was proposed to decide whether a given <span class="hlt">droplet</span> is in focus or not. In a first experiment, differently sized <span class="hlt">droplets</span> were generated with a piezoelectric generator and studied to establish the relationship between size and in-focus characteristics. In a second experiment, it was demonstrated that <span class="hlt">droplet</span> sizes and velocities from a real sprayer could be measured reliably in a non-intrusive way using the newly developed image acquisition set-up and image processing. Measured <span class="hlt">droplet</span> sizes ranged from 24 μm to 543 μm, depending on the nozzle type and size. <span class="hlt">Droplet</span> velocities ranged from around 0.5 m/s to 12 m/s. The <span class="hlt">droplet</span> size and velocity results were compared and related well with the results obtained with a Phase Doppler Particle Analyzer (PDPA). PMID:26861338</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/26861338','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26861338"><span>Spray <span class="hlt">Droplet</span> Characterization from a Single Nozzle by High Speed Image Analysis Using an In-Focus <span class="hlt">Droplet</span> Criterion.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Minov, Sofija Vulgarakis; Cointault, Frédéric; Vangeyte, Jürgen; Pieters, Jan G; Nuyttens, David</p> <p>2016-02-06</p> <p>Accurate spray characterization helps to better understand the pesticide spray application process. The goal of this research was to present the proof of principle of a <span class="hlt">droplet</span> size and velocity measuring technique for different types of hydraulic spray nozzles using a high speed backlight image acquisition and analysis system. As only part of the drops of an agricultural spray can be in focus at any given moment, an in-focus criterion based on the gray level gradient was proposed to decide whether a given <span class="hlt">droplet</span> is in focus or not. In a first experiment, differently sized <span class="hlt">droplets</span> were generated with a piezoelectric generator and studied to establish the relationship between size and in-focus characteristics. In a second experiment, it was demonstrated that <span class="hlt">droplet</span> sizes and velocities from a real sprayer could be measured reliably in a non-intrusive way using the newly developed image acquisition set-up and image processing. Measured <span class="hlt">droplet</span> sizes ranged from 24 μm to 543 μm, depending on the nozzle type and size. <span class="hlt">Droplet</span> velocities ranged from around 0.5 m/s to 12 m/s. The <span class="hlt">droplet</span> size and velocity results were compared and related well with the results obtained with a Phase Doppler Particle Analyzer (PDPA).</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1248736-supported-silver-nanoparticle-near-interface-solution-dynamics-deep-eutectic-solvent','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1248736-supported-silver-nanoparticle-near-interface-solution-dynamics-deep-eutectic-solvent"><span>Supported Silver Nanoparticle and Near-Interface Solution Dynamics in a Deep Eutectic <span class="hlt">Solvent</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>Hammons, Joshua A.; Ustarroz, Jon; Muselle, Thibault</p> <p>2016-01-28</p> <p>Type III deep eutectic <span class="hlt">solvents</span> (DES) have attracted significant interest as both environmentally friendly and functional <span class="hlt">solvents</span> that are, in some ways, advantageous to traditional aqueous systems. While these <span class="hlt">solvents</span> continue to produce remarkable thin films and nanoparticle assemblies, their interactions with metallic surfaces are complex and difficult to manipulate. In this study, the near-surface region (2–600 nm) of a carbon surface is investigated immediately following silver nanoparticle nucleation and growth. This is accomplished, in situ, using a novel grazing transmission small-angle X-ray scattering approach with simultaneous voltammetry and electrochemical impedance spectroscopy. With this physical and electrochemical approach, the timemore » evolution of three distinct surface interaction phenomena is observed: aggregation and coalescence of Ag nanoparticles, multilayer perturbations induced by nonaggregated Ag nanoparticles, and a stepwise transport of dissolved Ag species from the carbon surface. The multilayer perturbations contain <span class="hlt">charge</span>-separated regions of positively <span class="hlt">charged</span> choline-ethylene and negatively <span class="hlt">charged</span> Ag and Cl species. Both aggregation-coalescence and the stepwise decrease in Ag precursor near the surface are observed to be very slow (~2 h) processes, as both ion and particle transport are significantly impeded in a DES as compared to aqueous electrolytes. Finally, altogether, this study shows how the unique chemistry of the DES changes near the surface and in the presence of nanoparticles that adsorb the constituent species.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2011TRACE..15..213S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2011TRACE..15..213S"><span>Freezing of Water <span class="hlt">Droplet</span> due to Evaporation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Satoh, Isao; Fushinobu, Kazuyoshi; Hashimoto, Yu</p> <p></p> <p>In this study, the feasibility of cooling/freezing of phase change.. materials(PCMs) due to evaporation for cold storage systems was experimentally examined. A pure water was used as the test PCM, since the latent heat due to evaporation of water is about 7 times larger than that due to freezing. A water <span class="hlt">droplet</span>, the diameter of which was 1-4 mm, was suspended in a test cell by a fine metal wire (O. D.= 100μm),and the cell was suddenly evacuated up to the pressure lower than the triple-point pressure of water, so as to enhance the evaporation from the water surface. Temperature of the <span class="hlt">droplet</span> was measured by a thermocouple, and the cooling/freezing behavior and the temperature profile of the <span class="hlt">droplet</span> surface were captured by using a video camera and an IR thermo-camera, respectively. The obtained results showed that the water <span class="hlt">droplet</span> in the evacuated cell is effectively cooled by the evaporation of water itself, and is frozen within a few seconds through remarkable supercooling state. When the initial temperature of the <span class="hlt">droplet</span> is slightly higher than the room temperature, boiling phenomena occur in the <span class="hlt">droplet</span> simultaneously with the freezing due to evaporation. Under such conditions, it was shown that the degree of supercooling of the <span class="hlt">droplet</span> is reduced by the bubbles generated in the <span class="hlt">droplet</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/biblio/1243278-lipidomic-proteomic-analysis-caenorhabditis-elegans-lipid-droplets-identification-acs-lipid-droplet-associated-protein','SCIGOV-STC'); return false;" href="https://www.osti.gov/biblio/1243278-lipidomic-proteomic-analysis-caenorhabditis-elegans-lipid-droplets-identification-acs-lipid-droplet-associated-protein"><span>Lipidomic and proteomic analysis of Caenorhabditis elegans lipid <span class="hlt">droplets</span> and identification of ACS-4 as a lipid <span class="hlt">droplet</span>-associated protein</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>Vrablik, Tracy L.; Petyuk, Vladislav A.; Larson, Emily M.</p> <p>2015-06-27</p> <p>Lipid <span class="hlt">droplets</span> are cytoplasmic organelles that store neutral lipids for membrane synthesis and energy reserves. In this study, we characterized the lipid and protein composition of purified C. elegans lipid <span class="hlt">droplets</span>. These lipid <span class="hlt">droplets</span> are composed mainly of triacylglycerols, surrounded by a phospholipid monolayer composed primarily of phosphatidylcholine and phosphatidylethanolamine. The fatty acid composition of the triacylglycerols was rich in fatty acid species obtained from the dietary E. coli, including cyclopropane fatty acids and cis-vaccenic acid. Unlike other organisms, C. elegans lipid <span class="hlt">droplets</span> contain very little cholesterol or cholesterol esters. Comparison of the lipid <span class="hlt">droplet</span> proteomes of wild type andmore » high-fat daf-2 mutant strains shows a relative decrease of MDT-28 abundance in lipid <span class="hlt">droplets</span> isolated from daf-2 mutants. Functional analysis of lipid <span class="hlt">droplet</span> proteins identified in our proteomic studies indicated an enrichment of proteins required for growth and fat homeostasis in C. elegans.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22072434','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22072434"><span>Sequential microfluidic <span class="hlt">droplet</span> processing for rapid DNA extraction.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Pan, Xiaoyan; Zeng, Shaojiang; Zhang, Qingquan; Lin, Bingcheng; Qin, Jianhua</p> <p>2011-11-01</p> <p>This work describes a novel <span class="hlt">droplet</span>-based microfluidic device, which enables sequential <span class="hlt">droplet</span> processing for rapid DNA extraction. The microdevice consists of a <span class="hlt">droplet</span> generation unit, two reagent addition units and three <span class="hlt">droplet</span> splitting units. The loading/washing/elution steps required for DNA extraction were carried out by sequential microfluidic <span class="hlt">droplet</span> processing. The movement of superparamagnetic beads, which were used as extraction supports, was controlled with magnetic field. The microdevice could generate about 100 <span class="hlt">droplets</span> per min, and it took about 1 min for each <span class="hlt">droplet</span> to perform the whole extraction process. The extraction efficiency was measured to be 46% for λ-DNA, and the extracted DNA could be used in subsequent genetic analysis such as PCR, demonstrating the potential of the device for fast DNA extraction. Copyright © 2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5666997','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=5666997"><span>Dispersions of Goethite Nanorods in Aprotic Polar <span class="hlt">Solvents</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>Coursault, Delphine; Dozov, Ivan; Nobili, Maurizio; Dupont, Laurent; Chanéac, Corinne</p> <p>2017-01-01</p> <p>Colloidal suspensions of anisotropic nanoparticles can spontaneously self-organize in liquid-crystalline phases beyond some concentration threshold. These phases often respond to electric and magnetic fields. At lower concentrations, usual isotropic liquids are observed but they can display very strong Kerr and Cotton-Mouton effects (i.e., field-induced particle orientation). For many examples of these colloidal suspensions, the <span class="hlt">solvent</span> is water, which hinders most electro-optic applications. Here, for goethite (α-FeOOH) nanorod dispersions, we show that water can be replaced by polar aprotic <span class="hlt">solvents</span>, such as N-methyl-2-pyrrolidone (NMP) and dimethylsulfoxide (DMSO), without loss of colloidal stability. By polarized-light microscopy, small-angle X-ray scattering and electro-optic measurements, we found that the nematic phase, with its field-response properties, is retained. Moreover, a strong Kerr effect was also observed with isotropic goethite suspensions in these polar aprotic <span class="hlt">solvents</span>. Furthermore, we found no significant difference in the behavior of both the nematic and isotropic phases between the aqueous and non-aqueous dispersions. Our work shows that goethite nanorod suspensions in polar aprotic <span class="hlt">solvents</span>, suitable for electro-optic applications, can easily be produced and that they keep all their outstanding properties. It also suggests that this <span class="hlt">solvent</span> replacement method could be extended to the aqueous colloidal suspensions of other kinds of <span class="hlt">charged</span> anisotropic nanoparticles. PMID:29039797</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25135067','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25135067"><span>Conformational analysis of glutamic acid: a density functional approach using implicit continuum <span class="hlt">solvent</span> model.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Turan, Başak; Selçuki, Cenk</p> <p>2014-09-01</p> <p>Amino acids are constituents of proteins and enzymes which take part almost in all metabolic reactions. Glutamic acid, with an ability to form a negatively <span class="hlt">charged</span> side chain, plays a major role in intra and intermolecular interactions of proteins, peptides, and enzymes. An exhaustive conformational analysis has been performed for all eight possible forms at B3LYP/cc-pVTZ level. All possible neutral, zwitterionic, protonated, and deprotonated forms of glutamic acid structures have been investigated in solution by using polarizable continuum model mimicking water as the <span class="hlt">solvent</span>. Nine families based on the dihedral angles have been classified for eight glutamic acid forms. The electrostatic effects included in the <span class="hlt">solvent</span> model usually stabilize the <span class="hlt">charged</span> forms more. However, the stability of the zwitterionic form has been underestimated due to the lack of hydrogen bonding between the solute and <span class="hlt">solvent</span>; therefore, it is observed that compact neutral glutamic acid structures are more stable in solution than they are in vacuum. Our calculations have shown that among all eight possible forms, some are not stable in solution and are immediately converted to other more stable forms. Comparison of isoelectronic glutamic acid forms indicated that one of the structures among possible zwitterionic and anionic forms may dominate over the other possible forms. Additional investigations using explicit <span class="hlt">solvent</span> models are necessary to determine the stability of <span class="hlt">charged</span> forms of glutamic acid in solution as our results clearly indicate that hydrogen bonding and its type have a major role in the structure and energy of conformers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhRvF...2c1601H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhRvF...2c1601H"><span><span class="hlt">Droplet</span> depinning in a wake</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hooshanginejad, Alireza; Lee, Sungyon</p> <p>2017-03-01</p> <p>Pinning and depinning of a windswept <span class="hlt">droplet</span> on a surface is familiar yet deceptively complex for it depends on the interaction of the contact line with the microscopic features of the solid substrate. This physical picture is further compounded when wind of the Reynolds number greater than 100 blows over pinned drops, leading to the boundary layer separation and wake generation. In this Rapid Communication, we incorporate the well-developed ideas of the classical boundary layer to study partially wetting <span class="hlt">droplets</span> in a wake created by a leader object. Depending on its distance from the leader, the <span class="hlt">droplet</span> is observed to exhibit drafting, upstream motion, and splitting, due to the wake-induced hydrodynamic coupling that is analogous to drafting of moving bodies. We successfully rationalize the onset of the upstream motion regime using a reduced model that computes the <span class="hlt">droplet</span> shape governed by the pressure field inside the wake.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017NatPh..13.1020D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017NatPh..13.1020D"><span>Oleoplaning <span class="hlt">droplets</span> on lubricated surfaces</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Daniel, Dan; Timonen, Jaakko V. I.; Li, Ruoping; Velling, Seneca J.; Aizenberg, Joanna</p> <p>2017-10-01</p> <p>Recently, there has been much interest in using lubricated surfaces to achieve extreme liquid repellency: a foreign <span class="hlt">droplet</span> immiscible with the underlying lubricant layer was shown to slide off at a small tilt angle <5°. This behaviour was hypothesized to arise from a thin lubricant overlayer film sandwiched between the <span class="hlt">droplet</span> and solid substrate, but this has not been observed experimentally. Here, using thin-film interference, we are able to visualize the intercalated film under both static and dynamic conditions. We further demonstrate that for a moving <span class="hlt">droplet</span>, the film thickness follows the Landau-Levich-Derjaguin law. The <span class="hlt">droplet</span> is therefore oleoplaning--akin to tyres hydroplaning on a wet road--with minimal dissipative force and no contact line pinning. The techniques and insights presented in this study will inform future work on the fundamentals of wetting for lubricated surfaces and enable their rational design.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20010004339','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20010004339"><span>Fluid Flow in An Evaporating <span class="hlt">Droplet</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Hu, H.; Larson, R.</p> <p>1999-01-01</p> <p><span class="hlt">Droplet</span> evaporation is a common phenomenon in everyday life. For example, when a <span class="hlt">droplet</span> of coffee or salt solution is dropped onto a surface and the <span class="hlt">droplet</span> dries out, a ring of coffee or salt particles is left on the surface. This phenomenon exists not only in everyday life, but also in many practical industrial processes and scientific research and could also be used to assist in DNA sequence analysis, if the flow field in the <span class="hlt">droplet</span> produced by the evaporation could be understood and predicted in detail. In order to measure the fluid flow in a <span class="hlt">droplet</span>, small particles can be suspended into the fluid as tracers. From the ratio of gravitational force to Brownian force a(exp 4)(delta rho)(g)/k(sub B)T, we find that particle's tendency to settle is proportional to a(exp 4) (a is particle radius). So, to keep the particles from settling, the <span class="hlt">droplet</span> size should be chosen to be in a range 0.1 -1.0 microns in experiments. For such small particles, the Brownian force will affect the motion of the particle preventing accurate measurement of the flow field. This problem could be overcome by using larger particles as tracers to measure fluid flow under microgravity since the gravitational acceleration g is then very small. For larger particles, Brownian force would hardly affect the motion of the particles. Therefore, accurate flow field could be determined from experiments in microgravity. In this paper, we will investigate the fluid flow in an evaporating <span class="hlt">droplet</span> under normal gravity, and compare experiments to theories. Then, we will present our ideas about the experimental measurement of fluid flow in an evaporating <span class="hlt">droplet</span> under microgravity.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017PhDT........98D','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017PhDT........98D"><span>Direct numerical simulation of <span class="hlt">droplet</span>-laden isotropic turbulence</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Dodd, Michael S.</p> <p></p> <p>Interaction of liquid <span class="hlt">droplets</span> with turbulence is important in numerous applications ranging from rain formation to oil spills to spray combustion. The physical mechanisms of <span class="hlt">droplet</span>-turbulence interaction are largely unknown, especially when compared to that of solid particles. Compared to solid particles, <span class="hlt">droplets</span> can deform, break up, coalesce and have internal fluid circulation. The main goal of this work is to investigate using direct numerical simulation (DNS) the physical mechanisms of <span class="hlt">droplet</span>-turbulence interaction, both for non-evaporating and evaporating <span class="hlt">droplets</span>. To achieve this objective, we develop and couple a new pressure-correction method with the volume-of-fluid (VoF) method for simulating incompressible two-fluid flows. The method's main advantage is that the variable coefficient Poisson equation that arises in solving the incompressible Navier-Stokes equations for two-fluid flows is reduced to a constant coefficient equation. This equation can then be solved directly using, e.g., the FFT-based parallel Poisson solver. For a 10243 mesh, our new pressure-correction method using a fast Poisson solver is ten to forty times faster than the standard pressure-correction method using multigrid. Using the coupled pressure-correction and VoF method, we perform direct numerical simulations (DNS) of 3130 finite-size, non-evaporating <span class="hlt">droplets</span> of diameter approximately equal to the Taylor lengthscale and with 5% <span class="hlt">droplet</span> volume fraction in decaying isotropic turbulence at initial Taylor-scale Reynolds number Relambda = 83. In the <span class="hlt">droplet</span>-laden cases, we vary one of the following three parameters: the <span class="hlt">droplet</span> Weber number based on the r.m.s. velocity of turbulence (0.1 ≤ Werms ≤ 5), the <span class="hlt">droplet</span>- to carrier-fluid density ratio (1 ≤ rhod/rho c ≤ 100) or the <span class="hlt">droplet</span>- to carrier-fluid viscosity ratio (1 ≤ mud/muc ≤ 100). We derive the turbulence kinetic energy (TKE) equations for the two-fluid, carrier-fluid and <span class="hlt">droplet</span>-fluid flow. These equations allow</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2012ApPhL.101d3703L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2012ApPhL.101d3703L"><span>Mapping the surface <span class="hlt">charge</span> distribution of amyloid fibril</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Lee, Gyudo; Lee, Wonseok; Lee, Hyungbeen; Woo Lee, Sang; Sung Yoon, Dae; Eom, Kilho; Kwon, Taeyun</p> <p>2012-07-01</p> <p>It is of high importance to measure and map the surface <span class="hlt">charge</span> distribution of amyloids, since electrostatic interaction between amyloidogenic proteins and biomolecules plays a vital role in amyloidogenesis. In this work, we have measured and mapped the surface <span class="hlt">charge</span> distributions of amyloids (i.e., β-lactoglobulin fibril) using Kelvin probe force microscopy. It is shown that the surface <span class="hlt">charge</span> distribution is highly dependent on the conformation of amyloids (e.g., the helical pitch of amyloid fibrils) as well as the pH of a <span class="hlt">solvent</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19990054030&hterms=Ikegami&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DIkegami','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19990054030&hterms=Ikegami&qs=N%3D0%26Ntk%3DAll%26Ntx%3Dmode%2Bmatchall%26Ntt%3DIkegami"><span>Combustion of Interacting <span class="hlt">Droplet</span> Arrays in a Microgravity Environment</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Dietrich, D. L.; Struk, P. M.; Kitano, K.; Ikegami, M.</p> <p>1999-01-01</p> <p>Investigations into <span class="hlt">droplet</span> interactions date back to Rex et al. Recently, Annamalai and Ryan and Annamalai published extensive reviews of <span class="hlt">droplet</span> array and cloud combustion studies. The authors studied the change in the burning rate constant, k, (relative to that of the single <span class="hlt">droplet</span>) that results from interactions. Under certain conditions, there exists a separation distance where the <span class="hlt">droplet</span> lifetime reaches a minimum, or average burning rate constant is a maximum . Additionally, since inter-<span class="hlt">droplet</span> separation distance, L, increases relative to the <span class="hlt">droplet</span> size, D, as the burning proceeds, the burning rate is not constant throughout the burn, but changes continuously with time. Only Law and co-workers and Mikami et al. studied interactions under conditions where buoyant forces were negligible. Comparing their results with existing theory, Law and co-workers found that theory over predicted the persistency and intensity of <span class="hlt">droplet</span> interactions. The <span class="hlt">droplet</span> interactions also depended on the initial array configuration as well as the instantaneous array configuration. They also concluded that <span class="hlt">droplet</span> heating was retarded due to interactions and that the burning process did not follow the "D-squared" law. Mikami et al. studied the combustion of a two-<span class="hlt">droplet</span> array of heptane burning in air at one atm pressure in microgravity. They showed that the instantaneous burning rate constant increases throughout the <span class="hlt">droplet</span> lifetime, even for a single <span class="hlt">droplet</span>. Also, the burn time of the array reached a minimum at a critical inter-<span class="hlt">droplet</span> spacing. In this article, we examine <span class="hlt">droplet</span> interactions in normal and microgravity environments. The microgravity experiments were in the NASA GRC 2.2 and 5.2 second drop towers, and the JAMIC (Japan Microgravity Center) 10 second drop tower. Special emphasis is directed to combustion under conditions that yield finite extinction diameters, and to determine how <span class="hlt">droplet</span> interactions affect the extinction process.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2014PhFl...26k2003C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2014PhFl...26k2003C"><span>Inertial migration of deformable <span class="hlt">droplets</span> in a microchannel</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chen, Xiaodong; Xue, Chundong; Zhang, Li; Hu, Guoqing; Jiang, Xingyu; Sun, Jiashu</p> <p>2014-11-01</p> <p>The microfluidic inertial effect is an effective way of focusing and sorting <span class="hlt">droplets</span> suspended in a carrier fluid in microchannels. To understand the flow dynamics of microscale <span class="hlt">droplet</span> migration, we conduct numerical simulations on the <span class="hlt">droplet</span> motion and deformation in a straight microchannel. The results are compared with preliminary experiments and theoretical analysis. In contrast to most existing literature, the present simulations are three-dimensional and full length in the streamwise direction and consider the confinement effects for a rectangular cross section. To thoroughly examine the effect of the velocity distribution, the release positions of single <span class="hlt">droplets</span> are varied in a quarter of the channel cross section based on the geometrical symmetries. The migration dynamics and equilibrium positions of the <span class="hlt">droplets</span> are obtained for different fluid velocities and <span class="hlt">droplet</span> sizes. <span class="hlt">Droplets</span> with diameters larger than half of the channel height migrate to the centerline in the height direction and two equilibrium positions are observed between the centerline and the wall in the width direction. In addition to the well-known Segré-Silberberg equilibrium positions, new equilibrium positions closer to the centerline are observed. This finding is validated by preliminary experiments that are designed to introduce <span class="hlt">droplets</span> at different initial lateral positions. Small <span class="hlt">droplets</span> also migrate to two equilibrium positions in the quarter of the channel cross section, but the coordinates in the width direction are between the centerline and the wall. The equilibrium positions move toward the centerlines with increasing Reynolds number due to increasing deformations of the <span class="hlt">droplets</span>. The distributions of the lift forces, angular velocities, and the deformation parameters of <span class="hlt">droplets</span> along the two confinement direction are investigated in detail. Comparisons are made with theoretical predictions to determine the fundamentals of <span class="hlt">droplet</span> migration in microchannels. In</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2017NatCo...814831T','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017NatCo...814831T"><span>Mechano-regulated surface for manipulating liquid <span class="hlt">droplets</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Tang, Xin; Zhu, Pingan; Tian, Ye; Zhou, Xuechang; Kong, Tiantian; Wang, Liqiu</p> <p>2017-04-01</p> <p>The effective transfer of tiny liquid <span class="hlt">droplets</span> is vital for a number of processes such as chemical and biological microassays. Inspired by the tarsi of meniscus-climbing insects, which can climb menisci by deforming the water/air interface, we developed a mechano-regulated surface consisting of a background mesh and a movable microfibre array with contrastive wettability. The adhesion of this mechano-regulated surface to liquid <span class="hlt">droplets</span> can be reversibly switched through mechanical reconfiguration of the microfibre array. The adhesive force can be tuned by varying the number and surface chemistry of the microfibres. The in situ adhesion of the mechano-regulated surface can be used to manoeuvre micro-/nanolitre liquid <span class="hlt">droplets</span> in a nearly loss-free manner. The mechano-regulated surface can be scaled up to handle multiple <span class="hlt">droplets</span> in parallel. Our approach offers a miniaturized mechano-device with switchable adhesion for handling micro-/nanolitre <span class="hlt">droplets</span>, either in air or in a fluid that is immiscible with the <span class="hlt">droplets</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1987snpw.proc..137M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1987snpw.proc..137M"><span>Experimental test of liquid <span class="hlt">droplet</span> radiator performance</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mattick, A. T.; Simon, M. A.</p> <p></p> <p>The liquid <span class="hlt">droplet</span> radiator (LDR) is a heat rejection system for space power systems wherein an array of heated liquid <span class="hlt">droplets</span> radiates energy directly to space. The use of submillimeter <span class="hlt">droplets</span> provides large radiating area-to-mass ratio, resulting in radiator systems which are several times lighter than conventional solid surface radiators. An experiment is described in which the power radiated by an array of 2300 streams of silicone oil <span class="hlt">droplets</span> is measured to test a previously developed theory of the LDR radiation process. This system would be capable of rejecting several kW of heat in space. Furthermore, it would be suitable as a modular unit of an LDR designed for 100-kW power levels. The experiment provided confirmation of the theoretical dependence of <span class="hlt">droplet</span> array emissivity on optical depth. It also demonstrated the ability to create an array of more than 1000 <span class="hlt">droplet</span> streams having a divergence less than 1 degree.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19880035457&hterms=use+LDR&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Duse%2BLDR','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19880035457&hterms=use+LDR&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Duse%2BLDR"><span>Experimental test of liquid <span class="hlt">droplet</span> radiator performance</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Mattick, A. T.; Simon, M. A.</p> <p>1987-01-01</p> <p>The liquid <span class="hlt">droplet</span> radiator (LDR) is a heat rejection system for space power systems wherein an array of heated liquid <span class="hlt">droplets</span> radiates energy directly to space. The use of submillimeter <span class="hlt">droplets</span> provides large radiating area-to-mass ratio, resulting in radiator systems which are several times lighter than conventional solid surface radiators. An experiment is described in which the power radiated by an array of 2300 streams of silicone oil <span class="hlt">droplets</span> is measured to test a previously developed theory of the LDR radiation process. This system would be capable of rejecting several kW of heat in space. Furthermore, it would be suitable as a modular unit of an LDR designed for 100-kW power levels. The experiment provided confirmation of the theoretical dependence of <span class="hlt">droplet</span> array emissivity on optical depth. It also demonstrated the ability to create an array of more than 1000 <span class="hlt">droplet</span> streams having a divergence less than 1 degree.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2005AcSpA..62..987X','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2005AcSpA..62..987X"><span>Spectral and photophysical properties of intramolecular <span class="hlt">charge</span> transfer fluorescence probe: 4'-Dimethylamino-2,5-dihydroxychalcone</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Xu, Zhicheng; Bai, Guan; Dong, Chuan</p> <p>2005-12-01</p> <p>The spectral and photophysical properties of a new intramolecular <span class="hlt">charge</span> transfer (ICT) probe, namely 4'-dimethylamino-2,5-dihydroxychalcone (DMADHC) were studied in different <span class="hlt">solvents</span> by using steady-state absorption and emission spectroscopy. Whereas the absorption spectrum undergoes minor change with increasing polarity of the <span class="hlt">solvents</span>, the fluorescence spectrum experiences a distinct bathochromic shift in the band position and the fluorescence quantum yield increases reaching a maximum before decrease with increasing the <span class="hlt">solvent</span> polarity. The magnitude of change in the dipole moment was calculated based on the Lippert-Mataga equation. These results give the evidence about the intramolecular <span class="hlt">charge</span> transfer character in the emitting singlet state of this compound.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/15897003','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/15897003"><span>Spectral and photophysical properties of intramolecular <span class="hlt">charge</span> transfer fluorescence probe: 4'-dimethylamino-2,5-dihydroxychalcone.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Xu, Zhicheng; Bai, Guan; Dong, Chuan</p> <p>2005-12-01</p> <p>The spectral and photophysical properties of a new intramolecular <span class="hlt">charge</span> transfer (ICT) probe, namely 4'-dimethylamino-2,5-dihydroxychalcone (DMADHC) were studied in different <span class="hlt">solvents</span> by using steady-state absorption and emission spectroscopy. Whereas the absorption spectrum undergoes minor change with increasing polarity of the <span class="hlt">solvents</span>, the fluorescence spectrum experiences a distinct bathochromic shift in the band position and the fluorescence quantum yield increases reaching a maximum before decrease with increasing the <span class="hlt">solvent</span> polarity. The magnitude of change in the dipole moment was calculated based on the Lippert-Mataga equation. These results give the evidence about the intramolecular <span class="hlt">charge</span> transfer character in the emitting singlet state of this compound.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19940027380&hterms=spray+drying&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dspray%2Bdrying','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19940027380&hterms=spray+drying&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D30%26Ntt%3Dspray%2Bdrying"><span>Thermocapillary Convection in Liquid <span class="hlt">Droplets</span></span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p></p> <p>1986-01-01</p> <p>The purpose of this video is to understand the effects of surface tension on fluid convection. The fluid system chosen is the liquid sessile <span class="hlt">droplet</span> to show the importance in single crystal growth, the spray drying and cooling of metal, and the advance <span class="hlt">droplet</span> radiators of the space stations radiators. A cross sectional representation of a hemispherical liquid <span class="hlt">droplet</span> under ideal conditions is used to show internal fluid motion. A direct simulation of buoyancy-dominant convection and surface tension-dominant convection is graphically displayed. The clear differences between two mechanisms of fluid transport, thermocapillary convection, and bouncy dominant convection is illustrated.</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('http://adsabs.harvard.edu/abs/2017APS..MARS25003R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017APS..MARS25003R"><span>Molecular-Scale Investigation of Heavy Metal Ions at a <span class="hlt">Charged</span> Langmuir Monolayer</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Rock, William; Qiao, Baofu; Uysal, Ahmet; Bu, Wei; Lin, Binhua</p> <p></p> <p><span class="hlt">Solvent</span> extraction - the surfactant-aided preferential transfer of a species from an aqueous to an organic phase - is an important technique used in heavy and precious metal refining and reprocessing. <span class="hlt">Solvent</span> extraction requires transfer through an oil/water interface, and interfacial interactions are expected to control transfer kinetics and phase stability, yet these key interactions are poorly understood. Langmuir monolayers with <span class="hlt">charged</span> headgroups atop concentrated salt solutions containing heavy metal ions act as a model of <span class="hlt">solvent</span> extraction interfaces; studies of ions at a <span class="hlt">charged</span> surface are also fundamentally important to many other phenomena including protein solvation, mineral surface chemistry, and electrochemistry. We probe these <span class="hlt">charged</span> interfaces using a variety of surface-sensitive techniques - vibrational sum frequency generation (VSFG) spectroscopy, x-ray reflectivity (XRR), x-ray fluorescence near total reflection (XFNTR), and grazing incidence diffraction (GID). We integrate experiments with Molecular Dynamics (MD) simulations to uncover the molecular-level interfacial structure. This work is supported by the U.S. DOE, BES, Contract DE-AC02-06CH11357. ChemMatCARS is supported by NSF/CHE-1346572.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1990ApPhB..51....9A','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1990ApPhB..51....9A"><span>Stimulated raman scattering of fuel <span class="hlt">droplets</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Acker, William P.; Serpengüzel, Ali; Chang, Richard K.; Hill, Steven C.</p> <p>1990-07-01</p> <p>The strong stimulated Raman scattering (SRS) from diesel fuel <span class="hlt">droplets</span> has the potential of providing the relative concentration of multicomponent fuel and the absolute size of individual <span class="hlt">droplets</span>. The morphology-dependent resonances (MDRs) of a sphere cause the <span class="hlt">droplet</span> to act as an optical resonator which greatly lowers the SRS threshold. The number density, quality factor, and frequency shift of several MDRs are calculated as a function of the ratio of the index of refraction of the liquid and the surrounding gas, which approaches unity at the thermodynamic critical condition for the fuel spray. The SRS spectra of monodispersed <span class="hlt">droplets</span> of toluene, pentane, Exxon-Aromatic-150, and Mobil D-2 are presented. The exponential growth region of the SRS intensity I 1S as a function of the input laser intensity I input is investigated for the toluene carbon ring breathing mode v 2 and the pentane C-H stretching region. The I 1S ratio of toluene and pentane is measured as a function of the ratio of the toluene and pentane concentration for monodispersed <span class="hlt">droplets</span>. The reduced fluctuation in I 1S when I input is changed from multimode to single-mode is displayed as a histogram of the I 1S of the v 2 mode of toluene <span class="hlt">droplets</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016APS..MARK53005G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016APS..MARK53005G"><span>Caustics and the growth of <span class="hlt">droplets</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Govindarajan, Rama; Ravichandran, S.; Ray, Samriddhi; Deepu, P.</p> <p></p> <p>Caustics are formed when inertial particles of very different velocities collide in a flow, and are a consequence of the dissipative nature of particle motion in a suspension. Using a model vortex-dominated flow with heavy <span class="hlt">droplets</span> in a saturated environment, we suggest that sling caustics form only within a neighbourhood around a vortex, the square of whose radius is proportional to the product of circulation and particle inertia. <span class="hlt">Droplets</span> starting close to this critical radius congregate very close together, resulting in large spikes in (Lagrangian) number density. Allowing for merger when <span class="hlt">droplets</span> collide, we show that <span class="hlt">droplets</span> starting out close to the critical radius display a much more rapid growth in size than those starting elsewhere, and a large fraction of the large <span class="hlt">droplets</span> are those that originate within the caustics-forming region. We test these predictions in a two-dimensional simulation of turbulent flow. We hope that our study will be of interest in long-standing problems of physical interest such as the mechanism of broadening of <span class="hlt">droplet</span> spectra in a turbulent flow. Support from the Ministry of Earth Sciences, Government of India for the project Coupled physical processes in the Bay of Bengal and monsoon air-sea interaction under OMM is gratefully acknowledged.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2018AIPA....8e5226L','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2018AIPA....8e5226L"><span>Morphology of supercooled <span class="hlt">droplets</span> freezing on solid surfaces</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>La, Shiren; Huang, Zhiting; Liu, Cong; Zhang, Xingyi</p> <p>2018-05-01</p> <p>Supercooled <span class="hlt">droplets</span> freezing on solid surfaces are ubiquitous in nature. This letter investigates the influences of <span class="hlt">droplet</span> viscosity on freezing velocity and frosting formation. Several experiments were conducted for three kinds of sessile <span class="hlt">droplets</span> (water, silicone oil and oil) on two types of substrates (copper and iron) with different surface roughness at various temperatures. The results show that the water <span class="hlt">droplets</span> exhibit obvious phase transition lines and their freezing speeds increase when the temperature of substrates decreases. It is found that the freezing speed is independent of the thermal conductivities of the substrates. Notably, the water <span class="hlt">droplets</span> develop prominent bulges after freezing and subsequently nucleate to frost. In contrast, the high viscosity oil and silicone oil do not manifest an obvious phase transition line. Besides, no bulges are observed in these two kinds of <span class="hlt">droplets</span>, suggesting that these frosting forms are of different mechanisms compared with water <span class="hlt">droplets</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29276391','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29276391"><span>Some Lipid <span class="hlt">Droplets</span> Are More Equal Than Others: Different Metabolic Lipid <span class="hlt">Droplet</span> Pools in Hepatic Stellate Cells.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Molenaar, Martijn R; Vaandrager, Arie B; Helms, J Bernd</p> <p>2017-01-01</p> <p>Hepatic stellate cells (HSCs) are professional lipid-storing cells and are unique in their property to store most of the retinol (vitamin A) as retinyl esters in large-sized lipid <span class="hlt">droplets</span>. Hepatic stellate cell activation is a critical step in the development of chronic liver disease, as activated HSCs cause fibrosis. During activation, HSCs lose their lipid <span class="hlt">droplets</span> containing triacylglycerols, cholesteryl esters, and retinyl esters. Lipidomic analysis revealed that the dynamics of disappearance of these different classes of neutral lipids are, however, very different from each other. Although retinyl esters steadily decrease during HSC activation, triacylglycerols have multiple pools one of which becomes transiently enriched in polyunsaturated fatty acids before disappearing. These observations are consistent with the existence of preexisting "original" lipid <span class="hlt">droplets</span> with relatively slow turnover and rapidly recycling lipid <span class="hlt">droplets</span> that transiently appear during activation of HSCs. Elucidation of the molecular machinery involved in the regulation of these distinct lipid <span class="hlt">droplet</span> pools may open new avenues for the treatment of liver fibrosis.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27087486','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27087486"><span>Improved electron transport properties of n-type naphthalenediimide polymers through refined molecular ordering and orientation induced by processing <span class="hlt">solvents</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>An, Yujin; Long, Dang Xuan; Kim, Yiho; Noh, Yong-Young; Yang, Changduk</p> <p>2016-05-14</p> <p>To determine the role played by the choice of processing <span class="hlt">solvents</span> in governing the photophysics, microstructure, and <span class="hlt">charge</span> carrier transport in naphthalenediimide (NDI)-based polymers, we have prepared two new NDI-bithiophene (T2)- and NDI-thienothiophene (TTh)-containing polymers with hybrid siloxane pentyl chains (SiC5) (P(NDI2SiC5-T2) and P(NDI2SiC5-TTh)). Among the various processing <span class="hlt">solvents</span> studied here, the films prepared using chloroform exhibited far better electron mobilities (0.16 ± 0.1-0.21 ± 0.05 cm(2) V(-1) s(-1)) than the corresponding samples prepared from different <span class="hlt">solvents</span>, exceeding one order of magnitude higher, indicating the significant influence of the processing <span class="hlt">solvent</span> on the <span class="hlt">charge</span> transport. Upon thin-film analysis using atomic force microscopy and grazing incidence X-ray diffraction, we discovered that molecular ordering and orientation are affected by the choice of the processing <span class="hlt">solvent</span>, which is responsible for the change in the transport characteristics of this class of polymers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=MSFC-0300178&hterms=Shrink&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3DShrink','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=MSFC-0300178&hterms=Shrink&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D60%26Ntt%3DShrink"><span>Series of Two <span class="hlt">Droplets</span> on Fiber Approaching Ignition</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>The Fiber-Supported <span class="hlt">Droplet</span> Combustion (FSDC) uses two <span class="hlt">droplets</span> positioned on the fiber wire, instead of the usual one. Two <span class="hlt">droplets</span> more closely simulates the environment in engines, which ignite many fuel <span class="hlt">droplets</span> at once. The behavior of the burning was also unexpected -- the <span class="hlt">droplets</span> moved together after ignition, generating quite a bit of data for understanding the interaction of fuel <span class="hlt">droplets</span> while they burn. Because FSDC is backlit (the bright glow behind the drops), you carnot see the glow of the <span class="hlt">droplets</span> while they burn -- instead, you see them shrink! The small blobs left on the wire after the burn are the beads used to center the fuel <span class="hlt">droplet</span> on the wire. This image was taken on STS-94, July 12, 1997, MET:10/19:13 (approximate). FSDC-2 studied fundamental phenomena related to liquid fuel <span class="hlt">droplet</span> combustion in air. Pure fuels and mixtures of fuels were burned as isolated single and dual <span class="hlt">droplets</span> with and without forced air convection. The FSDC guest investigator was Forman Williams, University of California, San Diego. The experiment was part of the space research investigations conducted during the Microgravity Science Laboratory-1R mission (STS-94, July 1-17 1997). Advanced combustion experiments will be a part of investigations planned for the International Space Station. (251KB JPEG, 1350 x 1523 pixels; downlinked video, higher quality not available) The MPG from which this composite was made is available at http://mix.msfc.nasa.gov/ABSTRACTS/MSFC-0300179.html</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015AIPC.1642..441H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015AIPC.1642..441H"><span>A novel percussion type <span class="hlt">droplet</span>-on-demand generator</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hussain, Taaha; Patel, Priyesh; Balachandran, Ramanarayanan; Ladommatos, Nicos</p> <p>2015-01-01</p> <p>Numerous engineering applications require generation of <span class="hlt">droplets</span> on demand which are of high uniformity and constant size. The common method to produce <span class="hlt">droplets</span> is to drive liquid at high pressure through a small orifice/nozzle. The liquid stream disintegrates into small <span class="hlt">droplets</span>. However this method normally requires large volumes of liquid and is not suitable for applications where single <span class="hlt">droplets</span> of constant size is required. Such applications require <span class="hlt">droplet</span>-on-demand generators which commonly employ piezoelectric or pneumatic actuation. It is well known that piezoelectric generators are hard to employ at high pressure and, high temperature applications, and the pneumatic generators often produce satellite (secondary) <span class="hlt">droplets</span>. This paper describes the development of a novel percussion type <span class="hlt">droplet</span>-on-demand generator, which overcomes some of the above difficulties and is capable of producing single <span class="hlt">droplets</span> on demand. The generator consists of a cylindrical liquid filled chamber with a small orifice at the bottom. The top of the chamber is covered with a thin flexible metal disc. A small metal pin is employed to hammer/impact the top metal surface to generate a pressure pulse inside the liquid chamber. The movement and the momentum of the metal pin are controlled using a solenoid device. The pressure pulse generated overcomes the surface tension of the liquid meniscus at the exit of the orifice and ejects a single <span class="hlt">droplet</span>. The work presented in this paper will demonstrate the capabilities of the <span class="hlt">droplet</span> generator.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=MSFC-0300165&hterms=movies&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dmovies','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=MSFC-0300165&hterms=movies&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D20%26Ntt%3Dmovies"><span><span class="hlt">Droplet</span> Combustion Experiment movie</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>The <span class="hlt">Droplet</span> Combustion Experiment (DCE) was designed to investigate the fundamental combustion aspects of single, isolated <span class="hlt">droplets</span> under different pressures and ambient oxygen concentrations for a range of <span class="hlt">droplet</span> sizes varying between 2 and 5 mm. The DCE principal investigator was Forman Williams, University of California, San Diego. The experiment was part of the space research investigations conducted during the Microgravity Science Laboratory-1 mission (STS-83, April 4-8 1997; the shortened mission was reflown as MSL-1R on STS-94). Advanced combustion experiments will be a part of investigations plarned for the International Space Station. (1.1 MB, 12-second MPEG, screen 320 x 240 pixels; downlinked video, higher quality not available)A still JPG composite of this movie is available at http://mix.msfc.nasa.gov/ABSTRACTS/MSFC-0300164.html.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/17503939','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/17503939"><span>Three-axis acoustic device for levitation of <span class="hlt">droplets</span> in an open gas stream and its application to examine sulfur dioxide absorption by water <span class="hlt">droplets</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Stephens, Terrance L; Budwig, Ralph S</p> <p>2007-01-01</p> <p>Two acoustic devices to stabilize a <span class="hlt">droplet</span> in an open gas stream (single-axis and three-axis levitators) have been designed and tested. The gas stream was provided by a jet apparatus with a 64 mm exit diameter and a uniform velocity profile. The acoustic source used was a Langevin vibrator with a concave reflector. The single-axis levitator relied primarily on the radial force from the acoustic field and was shown to be limited because of significant <span class="hlt">droplet</span> wandering. The three-axis levitator relied on a combination of the axial and radial forces. The three-axis levitator was applied to examine <span class="hlt">droplet</span> deformation and circulation and to investigate the uptake of SO(2) from the gas stream to the <span class="hlt">droplet</span>. <span class="hlt">Droplets</span> ranging in diameters from 2 to 5 mm were levitated in gas streams with velocities up to 9 ms. <span class="hlt">Droplet</span> wandering was on the order of a half <span class="hlt">droplet</span> diameter for a 3 mm diameter <span class="hlt">droplet</span>. <span class="hlt">Droplet</span> circulation ranged from the predicted Hadamard-Rybczynski pattern to a rotating <span class="hlt">droplet</span> pattern. <span class="hlt">Droplet</span> pH over a central volume of the <span class="hlt">droplet</span> was measured by planar laser induced fluorescence. The results for the decay of <span class="hlt">droplet</span> pH versus time are in general agreement with published theory and experiments.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007RScI...78a4901S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007RScI...78a4901S"><span>Three-axis acoustic device for levitation of <span class="hlt">droplets</span> in an open gas stream and its application to examine sulfur dioxide absorption by water <span class="hlt">droplets</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Stephens, Terrance L.; Budwig, Ralph S.</p> <p>2007-01-01</p> <p>Two acoustic devices to stabilize a <span class="hlt">droplet</span> in an open gas stream (single-axis and three-axis levitators) have been designed and tested. The gas stream was provided by a jet apparatus with a 64mm exit diameter and a uniform velocity profile. The acoustic source used was a Langevin vibrator with a concave reflector. The single-axis levitator relied primarily on the radial force from the acoustic field and was shown to be limited because of significant <span class="hlt">droplet</span> wandering. The three-axis levitator relied on a combination of the axial and radial forces. The three-axis levitator was applied to examine <span class="hlt">droplet</span> deformation and circulation and to investigate the uptake of SO2 from the gas stream to the <span class="hlt">droplet</span>. <span class="hlt">Droplets</span> ranging in diameters from 2to5mm were levitated in gas streams with velocities up to 9m /s. <span class="hlt">Droplet</span> wandering was on the order of a half <span class="hlt">droplet</span> diameter for a 3mm diameter <span class="hlt">droplet</span>. <span class="hlt">Droplet</span> circulation ranged from the predicted Hadamard-Rybczynski pattern to a rotating <span class="hlt">droplet</span> pattern. <span class="hlt">Droplet</span> pH over a central volume of the <span class="hlt">droplet</span> was measured by planar laser induced fluorescence. The results for the decay of <span class="hlt">droplet</span> pH versus time are in general agreement with published theory and experiments.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2006JAP...100l4701C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2006JAP...100l4701C"><span>Removal of biofilms by impinging water <span class="hlt">droplets</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Cense, A. W.; van Dongen, M. E. H.; Gottenbos, B.; Nuijs, A. M.; Shulepov, S. Y.</p> <p>2006-12-01</p> <p>The process of impinging water <span class="hlt">droplets</span> on Streptococcus mutans biofilms was studied experimentally and numerically. <span class="hlt">Droplets</span> were experimentally produced by natural breakup of a cylindrical liquid jet. <span class="hlt">Droplet</span> diameter and velocity were varied between 20 and 200 μm and between 20 and 100 m/s, respectively. The resulting erosion process of the biofilm was determined experimentally with high-speed recording techniques and a quantitative relationship between the removal rate, <span class="hlt">droplet</span> size, and velocity was determined. The shear stress and the pressure on the surface during <span class="hlt">droplet</span> impact were determined by numerical simulations, and a qualitative agreement between the experiment and the simulation was obtained. Furthermore, it was shown that the stresses on the surface are strongly reduced when a water film is present.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4189219','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4189219"><span><span class="hlt">Droplet</span>-based microfluidic washing module for magnetic particle-based assays</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, Hun; Xu, Linfeng; Oh, Kwang W.</p> <p>2014-01-01</p> <p>In this paper, we propose a continuous flow <span class="hlt">droplet</span>-based microfluidic platform for magnetic particle-based assays by employing in-<span class="hlt">droplet</span> washing. The <span class="hlt">droplet</span>-based washing was implemented by traversing functionalized magnetic particles across a laterally merged <span class="hlt">droplet</span> from one side (containing sample and reagent) to the other (containing buffer) by an external magnetic field. Consequently, the magnetic particles were extracted to a parallel-synchronized train of washing buffer <span class="hlt">droplets</span>, and unbound reagents were left in an original train of sample <span class="hlt">droplets</span>. To realize the <span class="hlt">droplet</span>-based washing function, the following four procedures were sequentially carried in a <span class="hlt">droplet</span>-based microfluidic device: parallel synchronization of two trains of <span class="hlt">droplets</span> by using a ladder-like channel network; lateral electrocoalescence by an electric field; magnetic particle manipulation by a magnetic field; and asymmetrical splitting of merged <span class="hlt">droplets</span>. For the stable <span class="hlt">droplet</span> synchronization and electrocoalescence, we optimized <span class="hlt">droplet</span> generation conditions by varying the flow rate ratio (or <span class="hlt">droplet</span> size). Image analysis was carried out to determine the fluorescent intensity of reagents before and after the washing step. As a result, the unbound reagents in sample <span class="hlt">droplets</span> were significantly removed by more than a factor of 25 in the single washing step, while the magnetic particles were successfully extracted into washing buffer <span class="hlt">droplets</span>. As a proof-of-principle, we demonstrate a magnetic particle-based immunoassay with streptavidin-coated magnetic particles and fluorescently labelled biotin in the proposed continuous flow <span class="hlt">droplet</span>-based microfluidic platform. PMID:25379098</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4024758','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4024758"><span>Controlled multistep synthesis in a three-phase <span class="hlt">droplet</span> reactor</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Nightingale, Adrian M.; Phillips, Thomas W.; Bannock, James H.; de Mello, John C.</p> <p>2014-01-01</p> <p>Channel-fouling is a pervasive problem in continuous flow chemistry, causing poor product control and reactor failure. <span class="hlt">Droplet</span> chemistry, in which the reaction mixture flows as discrete <span class="hlt">droplets</span> inside an immiscible carrier liquid, prevents fouling by isolating the reaction from the channel walls. Unfortunately, the difficulty of controllably adding new reagents to an existing <span class="hlt">droplet</span> stream has largely restricted <span class="hlt">droplet</span> chemistry to simple reactions in which all reagents are supplied at the time of <span class="hlt">droplet</span> formation. Here we describe an effective method for repeatedly adding controlled quantities of reagents to <span class="hlt">droplets</span>. The reagents are injected into a multiphase fluid stream, comprising the carrier liquid, <span class="hlt">droplets</span> of the reaction mixture and an inert gas that maintains a uniform <span class="hlt">droplet</span> spacing and suppresses new <span class="hlt">droplet</span> formation. The method, which is suited to many multistep reactions, is applied to a five-stage quantum dot synthesis wherein particle growth is sustained by repeatedly adding fresh feedstock. PMID:24797034</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/16091462','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/16091462"><span>The electroosmotic <span class="hlt">droplet</span> switch: countering capillarity with electrokinetics.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Vogel, Michael J; Ehrhard, Peter; Steen, Paul H</p> <p>2005-08-23</p> <p>Electroosmosis, originating in the double-layer of a small liquid-filled pore (size R) and driven by a voltage V, is shown to be effective in pumping against the capillary pressure of a larger liquid <span class="hlt">droplet</span> (size B) provided the dimensionless parameter sigmaR(2)/epsilon|zeta|VB is small enough. Here sigma is surface tension of the <span class="hlt">droplet</span> liquid/gas interface, epsilon is the liquid dielectric constant, and zeta is the zeta potential of the solid/liquid pair. As <span class="hlt">droplet</span> size diminishes, the voltage required to pump electroosmotically scales as V approximately R(2)/B. Accordingly, the voltage needed to pump against smaller higher-pressure <span class="hlt">droplets</span> can actually decrease provided the pump poresize scales down with <span class="hlt">droplet</span> size appropriately. The technological implication of this favorable scaling is that electromechanical transducers made of moving <span class="hlt">droplets</span>, so-called "<span class="hlt">droplet</span> transducers," become feasible. To illustrate, we demonstrate a switch whose bistable energy landscape derives from the surface energy of a <span class="hlt">droplet-droplet</span> system and whose triggering derives from the electroosmosis effect. The switch is an electromechanical transducer characterized by individual addressability, fast switching time with low voltage, and no moving solid parts. We report experimental results for millimeter-scale <span class="hlt">droplets</span> to verify key predictions of a mathematical model of the switch. With millimeter-size water <span class="hlt">droplets</span> and micrometer-size pores, 5 V can yield switching times of 1 s. Switching time scales as B(3)/VR(2). Two possible "grab-and-release" applications of arrays of switches are described. One mimics the controlled adhesion of an insect, the palm beetle; the other uses wettability to move a particle along a trajectory.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://eosweb.larc.nasa.gov/project/airmspi/airmspi_oracles_l2_cloud_droplet_v1','SCIGOV-ASDC'); return false;" href="https://eosweb.larc.nasa.gov/project/airmspi/airmspi_oracles_l2_cloud_droplet_v1"><span>AirMSPI ORACLES Cloud <span class="hlt">Droplet</span> Data V001</span></a></p> <p><a target="_blank" href="http://eosweb.larc.nasa.gov/">Atmospheric Science Data Center </a></p> <p></p> <p>2018-05-05</p> <p>AirMSPI_ORACLES_Cloud_<span class="hlt">Droplet</span>_Size_and_Cloud_Optical_Depth L2 Derived Geophysical Parameters ... Order: Earthdata Search Parameters:  Cloud Optical Depth Cloud <span class="hlt">Droplet</span> Effective Radius Cloud <span class="hlt">Droplet</span> ...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=20040006403&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DG%2526T','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=20040006403&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D70%26Ntt%3DG%2526T"><span>Experimental Study of Supercooled Large <span class="hlt">Droplet</span> Impingement Effects</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Papadakis, M.; Rachman, A.; Wong, S. C.; Hung, K. E.; Vu, G. T.</p> <p>2003-01-01</p> <p>Typically, ice accretion results from small supercooled <span class="hlt">droplets</span> (<span class="hlt">droplets</span> cooled below freezing), usually 5 to 50 microns in diameter, which can freeze upon impact with an aircraft surface. Recently, ice accretions resulting from supercooled large <span class="hlt">droplet</span> (SLD) conditions have become a safety concern. Current ice accretion codes have been extensively tested for Title 14 Code of Federal Regulations Part 25, Appendix C icing conditions but have not been validated for SLD icing conditions. This report presents experimental methods for investigating large <span class="hlt">droplet</span> impingement dynamics and for obtaining small and large water <span class="hlt">droplet</span> impingement data.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20010074042','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20010074042"><span>Combustion of Interacting <span class="hlt">Droplet</span> Arrays in a Microgravity Environment</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Dietrich, D. L.; Struk, P. M.; Ikegami, M.; Nagaishi, H.; Honma, S.; Ikeda, K.</p> <p>2001-01-01</p> <p>Investigations into <span class="hlt">droplet</span> interactions date back to Rex et al. Annamalai and Ryan and Annamalai published extensive reviews of <span class="hlt">droplet</span> array and cloud combustion studies. In the majority of the reviewed studies, the authors examined the change in the burning rate constant, k, (relative to that of the single <span class="hlt">droplet</span>) that results from interactions. More recently, Niioka and co-workers have examined ignition and flame propagation along arrays of interacting <span class="hlt">droplets</span> with the goal of relating these phenomena in this simplified geometry to the more practical spray configuration. Our work has focussed on <span class="hlt">droplet</span> interactions under conditions where flame extinction occurs at a finite <span class="hlt">droplet</span> diameter. In our previous work, we reported that in normal gravity, reduced pressure conditions, <span class="hlt">droplet</span> interactions improved flame stability and extended flammability limits (by inference). In our recent work, we examine <span class="hlt">droplet</span> interactions under conditions where the flame extinguishes at a finite <span class="hlt">droplet</span> diameter in microgravity. The microgravity experiments were in the NASA GRC 2.2 and 5.2 second drop towers, and the JAMIC (Japan Microgravity Center) 10 second drop tower. We also present progress on a numerical model of single <span class="hlt">droplet</span> combustion that is in the process of being extended to model a binary <span class="hlt">droplet</span> array.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27792377','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27792377"><span>Dispersion of <span class="hlt">Droplet</span> Clouds in Turbulence.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Bocanegra Evans, Humberto; Dam, Nico; Bertens, Guus; van der Voort, Dennis; van de Water, Willem</p> <p>2016-10-14</p> <p>We measure the absolute dispersion of clouds of monodisperse, phosphorescent <span class="hlt">droplets</span> in turbulent air by means of high-speed image-intensified video recordings. Laser excitation allows the initial preparation of well-defined, pencil-shaped luminous <span class="hlt">droplet</span> clouds in a completely nonintrusive way. We find that the dispersion of the clouds is faster than the dispersion of fluid elements. We speculate that preferential concentration of inertial <span class="hlt">droplet</span> clouds is responsible for the enhanced dispersion.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004PhDT........95P','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004PhDT........95P"><span>Dynamics of <span class="hlt">droplet</span> collision and flame-front motion</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Pan, Kuo-Long</p> <p></p> <p>Three physical phenomena were experimentally and computationally investigated in this research, namely the dynamics of head-on <span class="hlt">droplet-droplet</span> collision, head-on <span class="hlt">droplet</span>-film collision, and laminar premixed flames, with emphasis on the transition between bouncing and merging of the liquid surfaces for the <span class="hlt">droplet</span> collision studies, and on the susceptibility to exhibit hydrodynamic instability for the flame dynamics. All three problems share the common feature of having an active deformable interface separating two flow regions of disparate densities, and as such can be computationally described using the adopted immersed boundary technique. Experimentally, the <span class="hlt">droplets</span> (˜300 mum diameter) were generated using the ink jet printing technique, and imaged using stroboscopy for the <span class="hlt">droplet-droplet</span> collision events and high-speed cine-photography for the <span class="hlt">droplet</span>-film collision events. For the study of <span class="hlt">droplet-droplet</span> collision, the instant of merging was experimentally determined and then used as an input in the computational simulation of the entire collision event. The simulation identified the differences between collision and merging at small and large Weber numbers, and satisfactorily described the dynamics of the inter-<span class="hlt">droplet</span> gap including the role of the van der Waals force in effecting surface rupture. For the study of <span class="hlt">droplet</span>-film collision, extensive experimental mapping showed that the collision dynamics is primarily affected by the <span class="hlt">droplet</span> Weber number (We) and the film thickness scaled by the <span class="hlt">droplet</span> radius (H), that while <span class="hlt">droplet</span> absorption by the film is facilitated with increasing <span class="hlt">droplet</span> Weber number, the boundary of transition is punctuated by an absorption peninsula, in the We-H space, within which absorption is further facilitated for smaller Weber numbers. Results from computation simulation revealed the essential dependence of the collision dynamics on the restraining nature of the solid surface, the energy exchange between the <span class="hlt">droplet</span> and the</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://www.osti.gov/servlets/purl/1406529','DOE-PATENT-XML'); return false;" href="https://www.osti.gov/servlets/purl/1406529"><span>Ambient infrared laser ablation mass spectrometry (AIRLAB-MS) with plume capture by continuous flow <span class="hlt">solvent</span> probe</span></a></p> <p><a target="_blank" href="http://www.osti.gov/doepatents">DOEpatents</a></p> <p>O'Brien, Jeremy T.; Williams, Evan R.; Holman, Hoi-Ying N.</p> <p>2017-10-31</p> <p>A new experimental setup for spatially resolved ambient infrared laser ablation mass spectrometry (AIRLAB-MS) that uses an infrared microscope with an infinity-corrected reflective objective and a continuous flow <span class="hlt">solvent</span> probe coupled to a Fourier transform ion cyclotron resonance mass spectrometer is described. The efficiency of material transfer from the sample to the electrospray ionization emitter was determined using glycerol/methanol <span class="hlt">droplets</span> containing 1 mM nicotine and is .about.50%. This transfer efficiency is significantly higher than values reported for similar techniques.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1985AcAau..12..591M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1985AcAau..12..591M"><span>Liquid <span class="hlt">droplet</span> radiator performance studies</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mattick, A. T.; Hertzberg, A.</p> <p></p> <p>By making use of <span class="hlt">droplets</span> rather than solid surfaces to radiate waste heat in space, the liquid <span class="hlt">droplet</span> radiator (LDR) achieves a radiating area/mass much larger than that of conventional radiators which use fins or heat pipes. The lightweight potential of the LDR is shown to be limited primarily by the radiative properties of the <span class="hlt">droplets</span>. The requirement that the LDR heat transfer fluid have a very low vapor pressure limits the choice of fluids to relatively few—several liquid metals and Dow 705 silicone fluid are the only suitable candidates so far identified. An experimental determination of the emittance of submillimeter <span class="hlt">droplets</span> of Dow 705 fluid indicates than an LDR using this fluid at temperatures of 275-335 K would be ⋍ 10 times lighter than the lightest solid surface radiators. Although several liquid metals appear to offer excellent performance in LDR applications at temperatures between 200 K and 975 K, experimental determination of liquid metal emissivities is needed for a conclusive assessment.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19850031025&hterms=use+LDR&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Duse%2BLDR','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19850031025&hterms=use+LDR&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Duse%2BLDR"><span>Liquid <span class="hlt">droplet</span> radiator performance studies</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Mattick, A. T.; Hertzberg, A.</p> <p>1984-01-01</p> <p>By making use of <span class="hlt">droplets</span> rather than solid surfaces to radiate waste heat in space, the liquid-<span class="hlt">droplet</span> radiator (LDR) achieves a radiating area/mass much larger than that of conventional radiators which use fins or heat pipes. The light-weight potential of the LDR is shown to be limited primarily by the radiative properties of the <span class="hlt">droplets</span>. The requirement that the LDR heat-transfer fluid have a very low vapor pressure limits the choice of fluids to relatively few several liquid metals and a silicone fluid are the only suitable candidates so far identified. An experimental determination of the emittance of submillimeter <span class="hlt">droplets</span> of the silicon fluid indicates that an LDR using this fluid at temperatures of 275-335 K would be about 10 times lighter than the lightest solid-surface radiators. Although several liquid metals appear to offer excellent performance in LDR applications at temperatures between 200 and 975 K, experimental determination of liquid-metal emissivities is needed for a conclusive assessment.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1984laus.iafcQS...M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1984laus.iafcQS...M"><span>Liquid <span class="hlt">droplet</span> radiator performance studies</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Mattick, A. T.; Hertzberg, A.</p> <p>1984-10-01</p> <p>By making use of <span class="hlt">droplets</span> rather than solid surfaces to radiate waste heat in space, the liquid-<span class="hlt">droplet</span> radiator (LDR) achieves a radiating area/mass much larger than that of conventional radiators which use fins or heat pipes. The light-weight potential of the LDR is shown to be limited primarily by the radiative properties of the <span class="hlt">droplets</span>. The requirement that the LDR heat-transfer fluid have a very low vapor pressure limits the choice of fluids to relatively few several liquid metals and a silicone fluid are the only suitable candidates so far identified. An experimental determination of the emittance of submillimeter <span class="hlt">droplets</span> of the silicon fluid indicates that an LDR using this fluid at temperatures of 275-335 K would be about 10 times lighter than the lightest solid-surface radiators. Although several liquid metals appear to offer excellent performance in LDR applications at temperatures between 200 and 975 K, experimental determination of liquid-metal emissivities is needed for a conclusive assessment.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/21601213','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/21601213"><span>Dispersive liquid-liquid microextraction based on solidification of floating organic <span class="hlt">droplet</span> followed by high-performance liquid chromatography with ultraviolet detection and liquid chromatography-tandem mass spectrometry for the determination of triclosan and 2,4-dichlorophenol in water samples.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Zheng, Cao; Zhao, Jing; Bao, Peng; Gao, Jin; He, Jin</p> <p>2011-06-24</p> <p>A novel, simple and efficient dispersive liquid-liquid microextraction based on solidification of floating organic <span class="hlt">droplet</span> (DLLME-SFO) technique coupled with high-performance liquid chromatography with ultraviolet detection (HPLC-UV) and liquid chromatography-tandem mass spectrometry (LC-MS/MS) was developed for the determination of triclosan and its degradation product 2,4-dichlorophenol in real water samples. The extraction <span class="hlt">solvent</span> used in this work is of low density, low volatility, low toxicity and proper melting point around room temperature. The extractant <span class="hlt">droplets</span> can be collected easily by solidifying it at a lower temperature. Parameters that affect the extraction efficiency, including type and volume of extraction <span class="hlt">solvent</span> and dispersive <span class="hlt">solvent</span>, salt effect, pH and extraction time, were investigated and optimized in a 5 mL sample system by HPLC-UV. Under the optimum conditions (extraction <span class="hlt">solvent</span>: 12 μL of 1-dodecanol; dispersive <span class="hlt">solvent</span>: 300 of μL acetonitrile; sample pH: 6.0; extraction time: 1 min), the limits of detection (LODs) of the pretreatment method combined with LC-MS/MS were in the range of 0.002-0.02 μg L(-1) which are lower than or comparable with other reported approaches applied to the determination of the same compounds. Wide linearities, good precisions and satisfactory relative recoveries were also obtained. The proposed technique was successfully applied to determine triclosan and 2,4-dichlorophenol in real water samples. Copyright © 2011 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26263052','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26263052"><span>Molecular Dynamics Simulations of the Initial-State Predict Product Distributions of Dediazoniation of Aryldiazonium in Binary <span class="hlt">Solvents</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Cruz, Gustavo N; Lima, Filipe S; Dias, Luís G; El Seoud, Omar A; Horinek, Dominik; Chaimovich, Hernan; Cuccovia, Iolanda M</p> <p>2015-09-04</p> <p>The dediazoniation of aryldiazonium salts in mixed <span class="hlt">solvents</span> proceeds by a borderline SN1 and SN2 pathway, and product distribution should be proportional to the composition of the solvation shell of the carbon attached to the -N2 group (ipso carbon). The rates of dediazoniation of 2,4,6-trimethylbenzenediazonium in water, methanol, ethanol, propanol, and acetonitrile were similar, but measured product distributions were noticeably dependent on the nature of the water/cosolvent mixture. Here we demonstrated that <span class="hlt">solvent</span> distribution in the first solvation shell of the ipso carbon, calculated from classical molecular dynamics simulations, is equal to the measured product distribution. Furthermore, we showed that regardless of the <span class="hlt">charge</span> distribution of the initial state, i.e., whether the positive <span class="hlt">charge</span> is smeared over the molecule or localized on phenyl moiety, the <span class="hlt">solvent</span> distribution around the reaction center is nearly the same.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20100001352','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20100001352"><span>Supercapacitor Electrolyte <span class="hlt">Solvents</span> with Liquid Range Below -80 C</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Brandon, Erik; Smart, Marshall; West, William</p> <p>2010-01-01</p> <p>A previous NASA Tech Brief ["Low-Temperature Supercapacitors" (NPO-44386) NASA Tech Briefs, Vol. 32, No 7 (July 2008), page 32] detailed ongoing efforts to develop non-aqueous supercapacitor electrolytes capable of supporting operation at temperatures below commercially available cells (which are typically limited to <span class="hlt">charging</span> and discharging at > or equal to -40 C). These electrolyte systems may enable energy storage and power delivery for systems operating in extreme environments, such as those encountered in the Polar regions on Earth or in the exploration of space. Supercapacitors using these electrolytes may also offer improved power delivery performance at moderately low temperatures (e.g. -40 to 0 C) relative to currently available cells, offering improved cold-cranking and cold-weather acceleration capabilities for electrical or hybrid vehicles. Supercapacitors store <span class="hlt">charge</span> at the electrochemical double-layer, formed at the interface between a high surface area electrode material and a liquid electrolyte. The current approach to extending the low-temperature limit of the electrolyte focuses on using binary <span class="hlt">solvent</span> systems comprising a high-dielectric-constant component (such as acetonitrile) in conjunction with a low-melting-point co-<span class="hlt">solvent</span> (such as organic formates, esters, and ethers) to depress the freezing point of the system, while maintaining sufficient solubility of the salt. Recent efforts in this area have led to the identification of an electrolyte <span class="hlt">solvent</span> formulation with a freezing point of -85.7 C, which is achieved by using a 1:1 by volume ratio of acetonitrile to 1,3-dioxolane</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/28742361','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/28742361"><span>Kinetic-Dominated <span class="hlt">Charging</span> Mechanism within Representative Aqueous Electrolyte-based Electric Double-Layer Capacitors.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Yang, Huachao; Yang, Jinyuan; Bo, Zheng; Chen, Xia; Shuai, Xiaorui; Kong, Jing; Yan, Jianhua; Cen, Kefa</p> <p>2017-08-03</p> <p>The chemical nature of electrolytes has been demonstrated to play a pivotal role in the <span class="hlt">charge</span> storage of electric double-layer capacitors (EDLCs), whereas primary mechanisms are still partially resolved but controversial. In this work, a systematic exploration into EDL structures and kinetics of representative aqueous electrolytes is performed with numerical simulation and experimental research. Unusually, a novel <span class="hlt">charging</span> mechanism exclusively predominated by kinetics is recognized, going beyond traditional views of manipulating capacitances preferentially via interfacial structural variations. Specifically, strikingly distinctive EDL structures stimulated by diverse ion sizes, valences, and mixtures manifest a virtually identical EDL capacitance, where the dielectric nature of <span class="hlt">solvents</span> attenuates ionic effects on electrolyte redistributions, in stark contradiction with <span class="hlt">solvent</span>-free counterpart and traditional Helmholtz theory. Meanwhile, corresponding kinetics evolve conspicuously with ionic species, intimately correlated with ion-<span class="hlt">solvent</span> interactions. The achieved mechanisms are subsequently illuminated by electrochemical measurements, highlighting the crucial interplay between ions and <span class="hlt">solvents</span> in regulating EDLC performances.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/1994JAtS...51..397M','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/1994JAtS...51..397M"><span>Fractal Analyses of High-Resolution Cloud <span class="hlt">Droplet</span> Measurements.</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Malinowski, Szymon P.; Leclerc, Monique Y.; Baumgardner, Darrel G.</p> <p>1994-02-01</p> <p>Fractal analyses of individual cloud <span class="hlt">droplet</span> distributions using aircraft measurements along one-dimensional horizontal cross sections through clouds are performed. Box counting and cluster analyses are used to determine spatial scales of inhomogeneity of cloud <span class="hlt">droplet</span> spacing. These analyses reveal that <span class="hlt">droplet</span> spatial distributions do not exhibit a fractal behavior. A high variability in local <span class="hlt">droplet</span> concentration in cloud volumes undergoing mixing was found. In these regions, thin filaments of cloudy air with <span class="hlt">droplet</span> concentration close to those observed in cloud cores were found. Results suggest that these filaments may be anisotropic. Additional box counting analyses performed for various classes of cloud <span class="hlt">droplet</span> diameters indicate that large and small <span class="hlt">droplets</span> are similarly distributed, except for the larger characteristic spacing of large <span class="hlt">droplets</span>.A cloud-clear air interface defined by a certain threshold of total <span class="hlt">droplet</span> count (TDC) was investigated. There are indications that this interface is a convoluted surface of a fractal nature, at least in actively developing cumuliform clouds. In contrast, TDC in the cloud interior does not have fractal or multifractal properties. Finally a random Cantor set (RCS) was introduced as a model of a fractal process with an ill-defined internal scale. A uniform measure associated with the RCS after several generations was introduced to simulate the TDC records. Comparison of the model with real TDC records indicates similar properties of both types of data series.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004OptCo.234..245E','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004OptCo.234..245E"><span>Steady-state and time-resolved studies on the photophysical properties of fullerene-pyropheophorbide a complexes in polar and nonpolar <span class="hlt">solvents</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ermilov, E. A.; Al-Omari, S.; Helmreich, M.; Jux, N.; Hirsch, A.; Röder, B.</p> <p>2004-04-01</p> <p>A novel monofullerene-bis(pyropheophorbide a) dyad has been photophysically characterized by steady-state as well as time-resolved techniques. It was revealed that in this complex strong and fast quenching of the first excited singlet state of the pyropheophorbide a (pyroPheo) molecule occurs by efficient photoinduced electron transfer to the fullerene moiety in both polar (DMF) and nonpolar (toluene) <span class="hlt">solvents</span>. In DMF the energy of the <span class="hlt">charge</span>-separated state is 0.94 eV and it undergoes directly transition to the ground state resulting in a very low value of photosensitized singlet oxygen generation. In contrast to the situation in a polar <span class="hlt">solvent</span>, in toluene the <span class="hlt">charge</span>-separated state lies above the exited triplet state of pyroPheo as well as that of C 60. It has been shown that in a nonpolar <span class="hlt">solvent</span> a sufficient amount of singlet oxygen was generated by energy transfer from the excited triplet state of pyroPheo which has been populated via relaxation of the <span class="hlt">charge</span>-separated state.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29906161','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29906161"><span>Evaporation-Triggered Segregation of Sessile Binary <span class="hlt">Droplets</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Li, Yaxing; Lv, Pengyu; Diddens, Christian; Tan, Huanshu; Wijshoff, Herman; Versluis, Michel; Lohse, Detlef</p> <p>2018-06-01</p> <p><span class="hlt">Droplet</span> evaporation of multicomponent <span class="hlt">droplets</span> is essential for various physiochemical applications, e.g., in inkjet printing, spray cooling, and microfabrication. In this work, we observe and study the phase segregation of an evaporating sessile binary <span class="hlt">droplet</span>, consisting of a miscible mixture of water and a surfactantlike liquid (1,2-hexanediol). The phase segregation (i.e., demixing) leads to a reduced water evaporation rate of the <span class="hlt">droplet</span>, and eventually the evaporation process ceases due to shielding of the water by the nonvolatile 1,2-hexanediol. Visualizations of the flow field by particle image velocimetry and numerical simulations reveal that the timescale of water evaporation at the <span class="hlt">droplet</span> rim is faster than that of the Marangoni flow, which originates from the surface tension difference between water and 1,2-hexanediol, eventually leading to segregation.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19930040896&hterms=mobil&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dmobil','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19930040896&hterms=mobil&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D50%26Ntt%3Dmobil"><span>Soot agglomeration in isolated, free <span class="hlt">droplet</span> combustion</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Choi, M. Y.; Dryer, F. L.; Green, G. J.; Sangiovanni, J. J.</p> <p>1993-01-01</p> <p>Under the conditions of an isolated, free <span class="hlt">droplet</span> experiment, hollow, carbonaceous structures, called soot spheres, were observed to form during the atmospheric pressure, low Reynolds number combustion of 1-methylnaphthalene. These structures which are agglomerates composed of smaller spheroidal units result from both thermophoretic effects induced by the envelope flame surrounding each drop and aerodynamic effects caused by changes in the relative gas/drop velocities. A chemically reacting flow model was used to analyze the process of sootshell formation during microgravity <span class="hlt">droplet</span> combustion. The time-dependent temperature and gas property field surrounding the <span class="hlt">droplet</span> was determined, and the soot cloud location for microgravity combustion of n-heptane <span class="hlt">droplets</span> was predicted. Experiments showed that the sooting propensity of n-alkane fuel <span class="hlt">droplets</span> can be varied through diluent substitution, oxygen-index variations, and ambient pressure reductions.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/22841080','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/22841080"><span>Rapid analysis of chlorinated anilines in environmental water samples using ultrasound assisted emulsification microextraction with solidification of floating organic <span class="hlt">droplet</span> followed by HPLC-UV detection.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Ramkumar, Abilasha; Ponnusamy, Vinoth Kumar; Jen, Jen-Fon</p> <p>2012-08-15</p> <p>The present study demonstrates a simple, rapid and efficient method for the determination of chlorinated anilines (CAs) in environmental water samples using ultrasonication assisted emulsification microextraction technique based on solidification of floating organic <span class="hlt">droplet</span> (USAEME-SFO) coupled with high performance liquid chromatography-ultraviolet (HPLC-UV) detection. In this extraction method, 1-dodecanol was used as extraction <span class="hlt">solvent</span> which is of lower density than water, low toxicity, low volatility, and low melting point (24 °C). After the USAEME, extraction <span class="hlt">solvent</span> could be collected easily by keeping the extraction tube in ice bath for 2 min and the solidified organic <span class="hlt">droplet</span> was scooped out using a spatula and transferred to another glass vial and allowed to thaw. Then, 10 μL of extraction <span class="hlt">solvent</span> was diluted with mobile phase (1:1) and taken for HPLC-UV analysis. Parameters influencing the extraction efficiency, such as the kind and volume of extraction <span class="hlt">solvent</span>, volume of sample, ultrasonication time, pH and salt concentration were thoroughly examined and optimized. Under the optimal conditions, the method showed good linearity in the concentration range of 0.05-500 ng mL(-1) with correlation coefficients ranging from 0.9948 to 0.9957 for the three target CAs. The limit of detection based on signal to noise ratio of 3 ranged from 0.01 to 0.1 ng mL(-1). The relative standard deviations (RSDs) varied from 2.1 to 6.1% (n=3) and the enrichment factors ranged from 44 to 124. The proposed method has also been successfully applied to analyze real water samples and the relative recoveries of environmental water samples ranged from 81.1 to 116.9%. Copyright © 2012 Elsevier B.V. All rights reserved.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1185691-enhanced-droplet-based-liquid-microjunction-surface-sampling-system-coupled-hplc-esi-ms-ms-spatially-resolved-analysis','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1185691-enhanced-droplet-based-liquid-microjunction-surface-sampling-system-coupled-hplc-esi-ms-ms-spatially-resolved-analysis"><span>An enhanced <span class="hlt">droplet</span>-based liquid microjunction surface sampling system coupled with HPLC-ESI-MS/MS for spatially resolved analysis</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Van Berkel, Gary J.; Weiskittel, Taylor M.; Kertesz, Vilmos</p> <p>2014-11-07</p> <p><span class="hlt">Droplet</span>-based liquid microjunction surface sampling coupled with high-performance liquid chromatography (HPLC)-electrospray ionization (ESI)-tandem mass spectrometry (MS/MS) for spatially resolved analysis provides the possibility of effective analysis of complex matrix samples and can provide a greater degree of chemical information from a single spot sample than is typically possible with a direct analysis of an extract. Described here is the setup and enhanced capabilities of a discrete <span class="hlt">droplet</span> liquid microjunction surface sampling system employing a commercially available CTC PAL autosampler. The system enhancements include incorporation of a laser distance sensor enabling unattended analysis of samples and sample locations of dramatically disparatemore » height as well as reliably dispensing just 0.5 μL of extraction <span class="hlt">solvent</span> to make the liquid junction to the surface, wherein the extraction spot size was confined to an area about 0.7 mm in diameter; software modifications improving the spatial resolution of sampling spot selection from 1.0 to 0.1 mm; use of an open bed tray system to accommodate samples as large as whole-body rat thin tissue sections; and custom sample/<span class="hlt">solvent</span> holders that shorten sampling time to approximately 1 min per sample. Lastly, the merit of these new features was demonstrated by spatially resolved sampling, HPLC separation, and mass spectral detection of pharmaceuticals and metabolites from whole-body rat thin tissue sections and razor blade (“crude”) cut mouse tissue.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/1185691','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/1185691"><span>An enhanced <span class="hlt">droplet</span>-based liquid microjunction surface sampling system coupled with HPLC-ESI-MS/MS for spatially resolved analysis</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>Van Berkel, Gary J.; Weiskittel, Taylor M.; Kertesz, Vilmos</p> <p></p> <p><span class="hlt">Droplet</span>-based liquid microjunction surface sampling coupled with high-performance liquid chromatography (HPLC)-electrospray ionization (ESI)-tandem mass spectrometry (MS/MS) for spatially resolved analysis provides the possibility of effective analysis of complex matrix samples and can provide a greater degree of chemical information from a single spot sample than is typically possible with a direct analysis of an extract. Described here is the setup and enhanced capabilities of a discrete <span class="hlt">droplet</span> liquid microjunction surface sampling system employing a commercially available CTC PAL autosampler. The system enhancements include incorporation of a laser distance sensor enabling unattended analysis of samples and sample locations of dramatically disparatemore » height as well as reliably dispensing just 0.5 μL of extraction <span class="hlt">solvent</span> to make the liquid junction to the surface, wherein the extraction spot size was confined to an area about 0.7 mm in diameter; software modifications improving the spatial resolution of sampling spot selection from 1.0 to 0.1 mm; use of an open bed tray system to accommodate samples as large as whole-body rat thin tissue sections; and custom sample/<span class="hlt">solvent</span> holders that shorten sampling time to approximately 1 min per sample. Lastly, the merit of these new features was demonstrated by spatially resolved sampling, HPLC separation, and mass spectral detection of pharmaceuticals and metabolites from whole-body rat thin tissue sections and razor blade (“crude”) cut mouse tissue.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://hdl.handle.net/2060/20040053529','NASA-TRS'); return false;" href="http://hdl.handle.net/2060/20040053529"><span>Combustion Of Interacting <span class="hlt">Droplet</span> Arrays In Microgravity</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Dietrich, D. L.; Struk, P. M.; Ikegami, M.; Xu, G.</p> <p>2003-01-01</p> <p>Theory and experiments involving single <span class="hlt">droplet</span> combustion date back to 1953, with the first microgravity work appearing in 1956. The problem of a spherical <span class="hlt">droplet</span> burning in an infinite, quiescent microgravity environment is a classical problem in combustion research with the classical solution appearing in nearly every textbook on combustion. The microgravity environment offered by ground-based facilities such as drop towers and space-based facilities is ideal for studying the problem experimentally. A recent review by Choi and Dryer shows significant advances in <span class="hlt">droplet</span> combustion have been made by studying the problem experimentally in microgravity and comparing the results to one dimensional theoretical and numerical treatments of the problem. Studying small numbers of interacting <span class="hlt">droplets</span> in a well-controlled geometry represents a logical step in extending single <span class="hlt">droplet</span> investigations to more practical spray configurations. Studies of <span class="hlt">droplet</span> interactions date back to Rex and co-workers, and were recently summarized by Annamalai and Ryan. All previous studies determined the change in the burning rate constant, k, or the flame characteristics as a result of interactions. There exists almost no information on how <span class="hlt">droplet</span> interactions a effect extinction limits, and if the extinction limits change if the array is in the diffusive or the radiative extinction regime. Thus, this study examined experimentally the effect that <span class="hlt">droplet</span> interactions have on the extinction process by investigating the simplest array configuration, a binary <span class="hlt">droplet</span> array. The studies were both in normal gravity, reduced pressure ambients and microgravity facilities. The microgravity facilities were the 2.2 and 5.2 second drop towers at the NASA Glenn Research Center and the 10 second drop tower at the Japan Microgravity Center. The experimental apparatus and the data analysis techniques are discussed in detail elsewhere.</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 <span class="hlt">droplet</span> vaporization</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-<span class="hlt">droplets</span> can be vaporized via ultrasound, a process termed acoustic <span class="hlt">droplet</span> vaporization. Our interest is primarily motivated by a developmental gas embolotherapy technique for cancer treatment. In this methodology, infarction of tumors is induced by selectively formed vascular gas bubbles that arise from the acoustic vaporization 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 <span class="hlt">droplet</span> vaporization through experiments and theoretical/computational fluid mechanics models, and investigate the bioeffects of acoustic <span class="hlt">droplet</span> vaporization on endothelial cells and in vivo. Early timescale vaporization events, including phase change, are directly visualized using ultra-high speed imaging, and the influence of acoustic parameters on <span class="hlt">droplet</span>/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('http://adsabs.harvard.edu/abs/2017NRL....12...68K','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2017NRL....12...68K"><span><span class="hlt">Droplet</span> Epitaxy Image Contrast in Mirror Electron Microscopy</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Kennedy, S. M.; Zheng, C. X.; Jesson, D. E.</p> <p>2017-01-01</p> <p>Image simulation methods are applied to interpret mirror electron microscopy (MEM) images obtained from a movie of GaAs <span class="hlt">droplet</span> epitaxy. Cylindrical symmetry of structures grown by <span class="hlt">droplet</span> epitaxy is assumed in the simulations which reproduce the main features of the experimental MEM image contrast, demonstrating that <span class="hlt">droplet</span> epitaxy can be studied in real-time. It is therefore confirmed that an inner ring forms at the <span class="hlt">droplet</span> contact line and an outer ring (or skirt) occurs outside the <span class="hlt">droplet</span> periphery. We believe that MEM combined with image simulations will be increasingly used to study the formation and growth of quantum structures.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2010ApPhL..96s3701G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2010ApPhL..96s3701G"><span><span class="hlt">Droplet</span> electric separator microfluidic device for cell sorting</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Guo, Feng; Ji, Xing-Hu; Liu, Kan; He, Rong-Xiang; Zhao, Li-Bo; Guo, Zhi-Xiao; Liu, Wei; Guo, Shi-Shang; Zhao, Xing-Zhong</p> <p>2010-05-01</p> <p>A simple and effective <span class="hlt">droplet</span> electric separator microfluidic device was developed for cell sorting. The aqueous <span class="hlt">droplet</span> without precharging operation was influenced to move a distance in the channel along the electric field direction by applying dc voltage on the electrodes beside the channel, which made the target <span class="hlt">droplet</span> flowing to the collector. Single <span class="hlt">droplet</span> can be isolated in a sorting rate of ˜100 Hz with microelectrodes under a required pulse. Single or multiple mammalian cell (HePG2) encapsulated in the surfactant free alginate <span class="hlt">droplet</span> could be sorted out respectively. This method may be used for single cell operation or analysis.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007JTST...16..736S','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007JTST...16..736S"><span>3D Modeling of Transport Phenomena and the Injection of the Solution <span class="hlt">Droplets</span> in the Solution Precursor Plasma Spraying</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Shan, Yanguang; Coyle, Thomas W.; Mostaghimi, Javad</p> <p>2007-12-01</p> <p>Solution precursor plasma spraying has been used to produce finely structured ceramic coatings with nano- and sub-micrometric features. This process involves the injection of a solution spray of ceramic salts into a DC plasma jet under atmospheric condition. During the process, the <span class="hlt">solvent</span> vaporizes as the <span class="hlt">droplet</span> travel downstream. Solid particles are finally formed due to the precipitation of the solute, and the particle are heated up and accelerated to the substrate to generate the coating. This article describes a 3D model to simulate the transport phenomena and the trajectory and heating of the solution spray in the process. The jet-spray two-way interactions are considered. A simplified model is employed to simulate the evolution process and the formation of the solid particle from the solution <span class="hlt">droplet</span> in the plasma jet. The temperature and velocity fields of the jet are obtained and validated. The particle size, velocity, temperature, and position distribution on the substrate are predicted.</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('https://ntrs.nasa.gov/search.jsp?R=MSFC-0300173&hterms=html&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Dhtml','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=MSFC-0300173&hterms=html&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D90%26Ntt%3Dhtml"><span>Burning Heptane <span class="hlt">Droplets</span> on STS-94</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>Fuel ignites and burns in the <span class="hlt">Droplet</span> Combustion Experiment (DCE) on STS-94 on July 11, 1997. This round of experiments burned heptane <span class="hlt">droplets</span> in 1/2 atmosphere pressure consisting of oxygen and helium. During this mission, scientists have seen for the first time <span class="hlt">droplets</span> which stop burning due to heat loss by radiation. From these data, the investigators hope to understand the physical and chemical processes that take place in <span class="hlt">droplet</span> combustion in different environments, including conditions under which the flames extinguish, the chemistry of the combustion reaction, and the production of pollutants such as nitrogen oxides and soot particles. The DCE was designed to investigate the fundamental combustion aspects of single, isolated <span class="hlt">droplets</span> under different pressures and ambient oxygen concentrations for a range of <span class="hlt">droplet</span> sizes varying between 2 and 5 mm. The DCE principal investigator was Forman Williams, University of California, San Diego. The experiment was part of the space research investigations conducted during the Microgravity Science Laboratory-1R mission (STS-94, July 1-17 1997). Advanced combustion experiments will be a part of investigations plarned for the International Space Station.(983KB, 9-second MPEG, screen 320 x 240 pixels; downlinked video, higher quality not available) A still JPG composite of this movie is available at http://mix.msfc.nasa.gov/ABSTRACTS/MSFC-0300172.html.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2015NatSR...515196H','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2015NatSR...515196H"><span>One-to-one encapsulation based on alternating <span class="hlt">droplet</span> generation</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Hirama, Hirotada; Torii, Toru</p> <p>2015-10-01</p> <p>This paper reports the preparation of encapsulated particles as models of cells using an alternating <span class="hlt">droplet</span> generation encapsulation method in which the number of particles in a <span class="hlt">droplet</span> is controlled by a microchannel to achieve one-to-one encapsulation. Using a microchannel in which wettability is treated locally, the fluorescent particles used as models of cells were successfully encapsulated in uniform water-in-oil-in-water (W/O/W) emulsion <span class="hlt">droplets</span>. Furthermore, 20% of the particle-containing <span class="hlt">droplets</span> contained one particle. Additionally, when a surfactant with the appropriate properties was used, the fluorescent particles within each inner aqueous <span class="hlt">droplet</span> were enclosed in the merged <span class="hlt">droplet</span> by spontaneous <span class="hlt">droplet</span> coalescence. This one-to-one encapsulation method based on alternating <span class="hlt">droplet</span> generation could be used for a variety of applications, such as high-throughput single-cell assays, gene transfection into cells or one-to-one cell fusion.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/26487193','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/26487193"><span>One-to-one encapsulation based on alternating <span class="hlt">droplet</span> generation.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Hirama, Hirotada; Torii, Toru</p> <p>2015-10-21</p> <p>This paper reports the preparation of encapsulated particles as models of cells using an alternating <span class="hlt">droplet</span> generation encapsulation method in which the number of particles in a <span class="hlt">droplet</span> is controlled by a microchannel to achieve one-to-one encapsulation. Using a microchannel in which wettability is treated locally, the fluorescent particles used as models of cells were successfully encapsulated in uniform water-in-oil-in-water (W/O/W) emulsion <span class="hlt">droplets</span>. Furthermore, 20% of the particle-containing <span class="hlt">droplets</span> contained one particle. Additionally, when a surfactant with the appropriate properties was used, the fluorescent particles within each inner aqueous <span class="hlt">droplet</span> were enclosed in the merged <span class="hlt">droplet</span> by spontaneous <span class="hlt">droplet</span> coalescence. This one-to-one encapsulation method based on alternating <span class="hlt">droplet</span> generation could be used for a variety of applications, such as high-throughput single-cell assays, gene transfection into cells or one-to-one cell fusion.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/29256179','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/29256179"><span>Lattice Boltzmann study of chemically-driven self-propelled <span class="hlt">droplets</span>.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Fadda, F; Gonnella, G; Lamura, A; Tiribocchi, A</p> <p>2017-12-19</p> <p>We numerically study the behavior of self-propelled liquid <span class="hlt">droplets</span> whose motion is triggered by a Marangoni-like flow. This latter is generated by variations of surfactant concentration which affect the <span class="hlt">droplet</span> surface tension promoting its motion. In the present paper a model for <span class="hlt">droplets</span> with a third amphiphilic component is adopted. The dynamics is described by Navier-Stokes and convection-diffusion equations, solved by the lattice Boltzmann method coupled with finite-difference schemes. We focus on two cases. First, the study of self-propulsion of an isolated <span class="hlt">droplet</span> is carried on and, then, the interaction of two self-propelled <span class="hlt">droplets</span> is investigated. In both cases, when the surfactant migrates towards the interface, a quadrupolar vortex of the velocity field forms inside the <span class="hlt">droplet</span> and causes the motion. A weaker dipolar field emerges instead when the surfactant is mainly diluted in the bulk. The dynamics of two interacting <span class="hlt">droplets</span> is more complex and strongly depends on their reciprocal distance. If, in a head-on collision, <span class="hlt">droplets</span> are close enough, the velocity field initially attracts them until a motionless steady state is achieved. If the <span class="hlt">droplets</span> are vertically shifted, the hydrodynamic field leads to an initial reciprocal attraction followed by a scattering along opposite directions. This hydrodynamic interaction acts on a separation of some <span class="hlt">droplet</span> radii otherwise it becomes negligible and <span class="hlt">droplets</span> motion is only driven by the Marangoni effect. Finally, if one of the <span class="hlt">droplets</span> is passive, this latter is generally advected by the fluid flow generated by the active one.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3400154','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3400154"><span>Microfluidic <span class="hlt">Droplet</span> Sorting with a High Frequency Ultrasound Beam</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, Changyang; Lee, Jungwoo; Kim, Hyung Ham; Teh, Shia-Yen; Lee, Abraham; Chung, In-Young; Park, Jae Yeong; Shung, K. Kirk</p> <p>2012-01-01</p> <p>This paper presents experimental results demonstrating the feasibility of high frequency ultrasonic sensing and sorting for screening single oleic acid (lipid or oil) <span class="hlt">droplets</span> under continuous flow in a microfluidic channel. In these experiments, hydrodynamically focused lipid <span class="hlt">droplets</span> of two different diameters (50 μm and 100 μm) are centered along the middle of the channel that is filled with deionized (DI) water. A 30 MHz lithium niobate (LiNbO3) transducer, placed outside the channel, first transmits short sensing pulses to non-invasively determine acoustic scattering properties of individual <span class="hlt">droplets</span> that are passing through the beam’s focus. Integrated backscatter (IB) coefficients, utilized as a sorting criterion, are measured by analyzing received echo signals from each <span class="hlt">droplet</span>. When the IB values corresponding to 100 μm <span class="hlt">droplets</span> are obtained, a custom-built LabVIEW panel commands the transducer to emit sinusoidal burst signals to commence the sorting operation. The number of <span class="hlt">droplets</span> tested for the sorting is 139 for 50 μm <span class="hlt">droplets</span> and 95 for 100 μm <span class="hlt">droplets</span>. The sensing efficiencies are estimated to be 98.6 % and 99.0 %, respectively. The sorting is carried out by applying acoustic radiation forces to 100 μm <span class="hlt">droplets</span> to direct them towards the upper sheath flow, thus separating them from the centered <span class="hlt">droplet</span> flow. The sorting efficiencies are 99.3 % for 50 μm <span class="hlt">droplets</span> and 85.3 % for 100 μm <span class="hlt">droplets</span>. The results suggest that this proposed technique has the potential to be further developed into a cost-effective and efficient cell/microparticle sorting instrument. PMID:22643737</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/23567746','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/23567746"><span><span class="hlt">Droplet</span> morphometry and velocimetry (DMV): a video processing software for time-resolved, label-free tracking of <span class="hlt">droplet</span> parameters.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Basu, Amar S</p> <p>2013-05-21</p> <p>Emerging assays in <span class="hlt">droplet</span> microfluidics require the measurement of parameters such as drop size, velocity, trajectory, shape deformation, fluorescence intensity, and others. While micro particle image velocimetry (μPIV) and related techniques are suitable for measuring flow using tracer particles, no tool exists for tracking <span class="hlt">droplets</span> at the granularity of a single entity. This paper presents <span class="hlt">droplet</span> morphometry and velocimetry (DMV), a digital video processing software for time-resolved <span class="hlt">droplet</span> analysis. <span class="hlt">Droplets</span> are identified through a series of image processing steps which operate on transparent, translucent, fluorescent, or opaque <span class="hlt">droplets</span>. The steps include background image generation, background subtraction, edge detection, small object removal, morphological close and fill, and shape discrimination. A frame correlation step then links <span class="hlt">droplets</span> spanning multiple frames via a nearest neighbor search with user-defined matching criteria. Each step can be individually tuned for maximum compatibility. For each <span class="hlt">droplet</span> found, DMV provides a time-history of 20 different parameters, including trajectory, velocity, area, dimensions, shape deformation, orientation, nearest neighbour spacing, and pixel statistics. The data can be reported via scatter plots, histograms, and tables at the granularity of individual <span class="hlt">droplets</span> or by statistics accrued over the population. We present several case studies from industry and academic labs, including the measurement of 1) size distributions and flow perturbations in a drop generator, 2) size distributions and mixing rates in drop splitting/merging devices, 3) efficiency of single cell encapsulation devices, 4) position tracking in electrowetting operations, 5) chemical concentrations in a serial drop dilutor, 6) drop sorting efficiency of a tensiophoresis device, 7) plug length and orientation of nonspherical plugs in a serpentine channel, and 8) high throughput tracking of >250 drops in a reinjection system. Performance metrics</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27541726','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27541726"><span><span class="hlt">Solvent</span> effects on the properties of hyperbranched polythiophenes.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Torras, Juan; Zanuy, David; Aradilla, David; Alemán, Carlos</p> <p>2016-09-21</p> <p>The structural and electronic properties of all-thiophene dendrimers and dendrons in solution have been evaluated using very different theoretical approaches based on quantum mechanical (QM) and hybrid QM/molecular mechanics (MM) methodologies: (i) calculations on minimum energy conformations using an implicit solvation model in combination with density functional theory (DFT) or time-dependent DFT (TD-DFT) methods; (ii) hybrid QM/MM calculations, in which the solute and <span class="hlt">solvent</span> molecules are represented at the DFT level as point <span class="hlt">charges</span>, respectively, on snapshots extracted from classical molecular dynamics (MD) simulations using explicit <span class="hlt">solvent</span> molecules, and (iii) QM/MM-MD trajectories in which the solute is described at the DFT or TD-DFT level and the explicit <span class="hlt">solvent</span> molecules are represented using classical force-fields. Calculations have been performed in dichloromethane, tetrahydrofuran and dimethylformamide. A comparison of the results obtained using the different approaches with the available experimental data indicates that the incorporation of effects associated with both the conformational dynamics of the dendrimer and the explicit <span class="hlt">solvent</span> molecules is strictly necessary to satisfactorily reproduce the properties of the investigated systems. Accordingly, QM/MM-MD simulations are able to capture such effects providing a reliable description of electronic properties-conformational flexibility relationships in all-Th dendrimers.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2004JNuM..334..189G','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2004JNuM..334..189G"><span>Carbothermic reduction of uranium oxides into <span class="hlt">solvent</span> metallic baths</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Guisard Restivo, Thomaz A.; Capocchi, José D. T.</p> <p>2004-09-01</p> <p>The carbothermic reduction of UO 2 and U 3O 8 is studied employing tin and silicon <span class="hlt">solvent</span> metallic baths in thermal analysis equipment, under Ar inert and N 2 reactive atmospheres. The metallic <span class="hlt">solvents</span> are expected to lower the U activity by several orders of magnitude owing to strong interactions among the metals. The reduction products are composed of the <span class="hlt">solvent</span> metal matrix and intermetallic U compounds. Silicon is more effective in driving the reduction since there is no residual UO 2 after the reaction. The gaseous product detected by mass spectrometer (MS) during the reduction is CO. A kinetic study for the Si case was accomplished by the stepwise isothermal analysis (SAI) method, leading to the identification of the controlling mechanisms as chemical reaction at the surface and nucleation, for UO 2 and U 3O 8 <span class="hlt">charges</span>, respectively. One example for another system containing Al 2O 3 is also shown.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016APS..MARP37003R','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016APS..MARP37003R"><span>On the pH of Aqueous Attoliter-Volume <span class="hlt">Droplets</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Ramos, Kieran P.; Velpula, Samson S.; Demille, Trevor B.; Pajela, Ryan; Goldner, Lori S.</p> <p></p> <p><span class="hlt">Droplets</span> of water dispersed in perfluorinated liquids have widespread use including microfluidics, drug delivery and single-molecule measurements. Perfluorinated liquids are distinctly biocompatible due to their stability, low surface tension, lipophobicity, and hydrophobicity. For this reason, the effect of the perfluorinated surface on <span class="hlt">droplet</span> contents is usually ignored. However, as the <span class="hlt">droplet</span> diameter is reduced, we expect that any effect of the water/oil interface on <span class="hlt">droplet</span> contents will become more obvious. We studied the pH of attoliter-volume aqueous <span class="hlt">droplets</span> in perfluorinated liquids using pH-sensing fluorescent dyes. <span class="hlt">Droplets</span> were prepared either by sonication or extrusion from buffer and perfluorinated liquids (FC40 or FC77). A non-ionic surfactant was used to stabilize the <span class="hlt">droplets</span>. Buffer strength, ionic strength, and pH of the aqueous phase were varied and resulting <span class="hlt">droplet</span> pH compared to the pH of the buffer from which they were formed. Preliminary data are consistent with a pH in <span class="hlt">droplets</span> that depends on the concentration of non-ionic surfactant. At low surfactant concentrations, the pH in <span class="hlt">droplets</span> is distinctly lower than the stock buffer. However, as the concentration of non-ionic surfactant is increased the change in pH decreases. This work was funded by NSF/DBI-1152386.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://cfpub.epa.gov/si/si_public_record_report.cfm?dirEntryId=36008&Lab=ORD&keyword=corona&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=36008&Lab=ORD&keyword=corona&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">CHARGE</span> MEASUREMENTS ON INDIVIDUAL PARTICLES EXITING LABORATORY PRECIPITATORS WITH POSITIVE AND NEGATIVE CORONA AT VARIOUS TEMPERATURES</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 reports measurements of <span class="hlt">charge</span> values on individual particles exiting three different laboratory electrostatic precipitators (ESPs) in an experimental apparatus containing a Millikan cell. Dioctylphthalate (DOP) <span class="hlt">droplets</span> and fly ash particles were measured at temperatur...</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2007ApPhL..91l4102X','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2007ApPhL..91l4102X"><span>Marangoni flow in an evaporating water <span class="hlt">droplet</span></span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Xu, Xuefeng; Luo, Jianbin</p> <p>2007-09-01</p> <p>Marangoni effect has been observed in many liquids, but its existence in pure water is still a debated problem. In the present work, the Marangoni flow in evaporating water <span class="hlt">droplets</span> has been observed by using fluorescent nanoparticles. Flow patterns indicate that a stagnation point where the surface flow, the surface tension gradient, and the surface temperature gradient change their directions exists at the <span class="hlt">droplet</span> surface. The deduced nonmonotonic variation of the <span class="hlt">droplet</span> surface temperature, which is different from that in some previous works, is explained by a heat transfer model considering the adsorbed thin film of the evaporating liquid <span class="hlt">droplet</span>.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19950031215&hterms=How+temperature+effect+rate+evaporation&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DHow%2Btemperature%2Beffect%2Brate%2Bevaporation','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19950031215&hterms=How+temperature+effect+rate+evaporation&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D10%26Ntt%3DHow%2Btemperature%2Beffect%2Brate%2Bevaporation"><span>Theoretical and Experimental Investigations on <span class="hlt">Droplet</span> Evaporation and <span class="hlt">Droplet</span> Ignition at High Pressures</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Ristau, R.; Nagel, U.; Iglseder, H.; Koenig, J.; Rath, H. J.; Normura, H.; Kono, M.; Tanabe, M.; Sato, J.</p> <p>1993-01-01</p> <p>The evaporation of fuel <span class="hlt">droplets</span> under high ambient pressure and temperature in normal gravity and microgravity has been investigated experimentally. For subcritical ambient conditions, <span class="hlt">droplet</span> evaporation after a heat-up period follows the d(exp 2)-law. For all data the evaporation constant increases as the ambient temperature increases. At identical ambient conditions the evaporation constant under microgravity is smaller compared to normal gravity. This effect can first be observed at 1 bar and increases with ambient pressure. Preliminary experiments on ignition delay for self-igniting fuel <span class="hlt">droplets</span> have been performed. Above a 1 s delay time, at identical ambient conditions, significant differences in the results of the normal and microgravity data are observed. Self-ignition occurs within different temperature ranges due to the influence of gravity. The time dependent behavior of the <span class="hlt">droplet</span> is examined theoretically. In the calculations two different approaches for the gas phase are applied. In the first approach the conditions at the interface are given using a quasi steady theory approximation. The second approach uses a set of time dependent governing equations for the gas phase which are then evaluated. In comparison, the second model shows a better agreement with the drop tower experiments. In both cases a time dependent gasification rate is observed.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('http://adsabs.harvard.edu/abs/2016ApPhL.109n3510C','NASAADS'); return false;" href="http://adsabs.harvard.edu/abs/2016ApPhL.109n3510C"><span>Microfluidic <span class="hlt">droplet</span> sorting using integrated bilayer micro-valves</span></a></p> <p><a target="_blank" href="http://adsabs.harvard.edu/abstract_service.html">NASA Astrophysics Data System (ADS)</a></p> <p>Chen, Yuncong; Tian, Yang; Xu, Zhen; Wang, Xinran; Yu, Sicong; Dong, Liang</p> <p>2016-10-01</p> <p>This paper reports on a microfluidic device capable of sorting microfluidic <span class="hlt">droplets</span> utilizing conventional bilayer pneumatic micro-valves as sorting controllers. The device consists of two micro-valves placed symmetrically on two sides of a sorting area, each on top of a branching channel at an inclined angle with respect to the main channel. Changes in transmitted light intensity, induced by varying light absorbance by each <span class="hlt">droplet</span>, are used to divert the <span class="hlt">droplet</span> from the sorting area into one of the three outlet channels. When no valve is activated, the <span class="hlt">droplet</span> flows into the outlet channel in the direction of the main channel. When one of the valves is triggered, the flexible membrane of valve will first be deflected. Once the <span class="hlt">droplet</span> leaves the detection point, the deflected membrane will immediately return to its default flattened position, thereby exerting a drawing pressure on the <span class="hlt">droplet</span> and deviating it from its original streamline to the outlet on the same side as the valve. This sorting method will be particularly suitable for numerous large-scale integrated microfluidic systems, where pneumatic micro-valves are already used. Only few structural modifications are needed to achieve <span class="hlt">droplet</span> sorting capabilities in these systems. Due to the mechanical nature of diverting energy applied to <span class="hlt">droplets</span>, the proposed sorting method may induce only minimal interference to biological species or microorganisms encapsulated inside the <span class="hlt">droplets</span> that may accompany electrical, optical and magnetic-based techniques.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://ntrs.nasa.gov/search.jsp?R=19990054071&hterms=prescribed+burning&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dprescribed%2Bburning','NASA-TRS'); return false;" href="https://ntrs.nasa.gov/search.jsp?R=19990054071&hterms=prescribed+burning&qs=Ntx%3Dmode%2Bmatchall%26Ntk%3DAll%26N%3D0%26No%3D40%26Ntt%3Dprescribed%2Bburning"><span>Dynamics of <span class="hlt">Droplet</span> Extinction in Slow Convective Flows</span></a></p> <p><a target="_blank" href="http://ntrs.nasa.gov/search.jsp">NASA Technical Reports Server (NTRS)</a></p> <p>Nayagam, V.; Haggard, J. B., Jr.; Williams, F. A.</p> <p>1999-01-01</p> <p>The classical model for <span class="hlt">droplet</span> combustion predicts that the square of the <span class="hlt">droplet</span> diameter decreases linearly with time. It also predicts that a <span class="hlt">droplet</span> of any size will burn to completion over a period of time. However, it has been known for some time that under certain conditions flames surrounding a <span class="hlt">droplet</span>, in a quiescent environment, could extinguish because of insufficient residence time for the chemistry to proceed to completion. This type of extinction that occurs for smaller <span class="hlt">droplets</span> has been studied extensively in the past. Large <span class="hlt">droplets</span>, on the other hand, exhibit a different type of extinction where excessive radiative heat loss from the flame zone leads to extinction. This mode of "radiative extinction" was theoretically predicted for <span class="hlt">droplet</span> burning by Chao et al. and was observed in recent space experiments in a quiescent environment. Thus far, the fundamental flammability limit prescribed by radiative extinction of liquid <span class="hlt">droplets</span> has been measured only under quiescent environmental conditions. In many space platforms, however, ventilation systems produce small convective flows and understanding of the influences of this convection on the extinction process will help better define the radiative extinction flammability boundaries. Boundaries defined by experiments and captured using theoretical models could provide enhanced fire safety margin in space explor1999063d investigation of convective effects will help in interpretations of burning-rate data obtained during free-floated <span class="hlt">droplet</span> combustion experiments with small residual velocities.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/servlets/purl/5142017','SCIGOV-STC'); return false;" href="https://www.osti.gov/servlets/purl/5142017"><span>Secondary atomization of single coal-water fuel <span class="hlt">droplets</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>Hassel, G.R.; Scaroni, A.W.</p> <p>1989-03-01</p> <p>The evaporative behavior of single, well characterized <span class="hlt">droplets</span> of a lignite coal-water slurry fuel (CWSF) and a carbon black in water slurry was studied as a function of heating rate and <span class="hlt">droplet</span> composition. Induced <span class="hlt">droplet</span> heating rates were varied from 0 to 10{sup 5} K/s. <span class="hlt">Droplets</span> studied were between 97 and 170 {mu}m in diameter, with compositions ranging from 25 to 60% solids by weight. The effect of a commercially available surfactant additive package on <span class="hlt">droplet</span> evaporation rate, explosive boiling energy requirements, and agglomerate formation was assessed. Surfactant concentrations were varied from none to 2 and 4% by weight solutionmore » (1.7 and 3.6% by weight of active species on a dry coal basis). The experimental system incorporated an electrodynamic balance to hold single, free <span class="hlt">droplets</span>, a counterpropagating pulsed laser heating arrangement, and both video and high speed cinematographic recording systems. Data were obtained for ambient <span class="hlt">droplet</span> evaporation by monitoring the temporal size, weight, and solids concentration changes. 49 refs., 31 figs.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/27997205','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/27997205"><span>Interaction of <span class="hlt">Droplets</span> Separated by an Elastic Film.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Liu, Tianshu; Xu, Xuejuan; Nadermann, Nichole; He, Zhenping; Jagota, Anand; Hui, Chung-Yuen</p> <p>2017-01-10</p> <p>The Laplace pressure of a <span class="hlt">droplet</span> placed on one side of an elastic thin film can cause significant deformation in the form of a bulge on its opposite side. Here, we show that this deformation can be detected by other <span class="hlt">droplets</span> suspended on the opposite side of the film, leading to interaction between <span class="hlt">droplets</span> separated by the solid (but deformable) film. The interaction is repulsive when the drops have a large overlap and attractive when they have a small overlap. Thus, if two identical <span class="hlt">droplets</span> are placed right on top of each other (one on either side of the thin film), they tend to repel each other, eventually reaching an equilibrium configuration where there is a small overlap. This observation can be explained by analyzing the energy landscape of the <span class="hlt">droplets</span> interacting via an elastically deformed film. We further demonstrate this idea by designing a pattern comprising a big central drop with satellite <span class="hlt">droplets</span>. This phenomenon can lead to techniques for directed motion of <span class="hlt">droplets</span> confined to one side of a thin elastic membrane by manipulations on the other side.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.ncbi.nlm.nih.gov/pubmed/25501881','PUBMED'); return false;" href="https://www.ncbi.nlm.nih.gov/pubmed/25501881"><span>Rapid and continuous magnetic separation in <span class="hlt">droplet</span> microfluidic devices.</span></a></p> <p><a target="_blank" href="https://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pubmed">PubMed</a></p> <p>Brouzes, Eric; Kruse, Travis; Kimmerling, Robert; Strey, Helmut H</p> <p>2015-02-07</p> <p>We present a <span class="hlt">droplet</span> microfluidic method to extract molecules of interest from a <span class="hlt">droplet</span> in a rapid and continuous fashion. We accomplish this by first marginalizing functionalized super-paramagnetic beads within the <span class="hlt">droplet</span> using a magnetic field, and then splitting the <span class="hlt">droplet</span> into one <span class="hlt">droplet</span> containing the majority of magnetic beads and one <span class="hlt">droplet</span> containing the minority fraction. We quantitatively analysed the factors which affect the efficiency of marginalization and <span class="hlt">droplet</span> splitting to optimize the enrichment of magnetic beads. We first characterized the interplay between the <span class="hlt">droplet</span> velocity and the strength of the magnetic field and its effect on marginalization. We found that marginalization is optimal at the midline of the magnet and that marginalization is a good predictor of bead enrichment through splitting at low to moderate <span class="hlt">droplet</span> velocities. Finally, we focused our efforts on manipulating the splitting profile to improve the enrichment provided by asymmetric splitting. We designed asymmetric splitting forks that employ capillary effects to preferentially extract the bead-rich regions of the <span class="hlt">droplets</span>. Our strategy represents a framework to optimize magnetic bead enrichment methods tailored to the requirements of specific <span class="hlt">droplet</span>-based applications. We anticipate that our separation technology is well suited for applications in single-cell genomics and proteomics. In particular, our method could be used to separate mRNA bound to poly-dT functionalized magnetic microparticles from single cell lysates to prepare single-cell cDNA libraries.</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.osti.gov/pages/biblio/1346969-rapid-continuous-magnetic-separation-droplet-microfluidic-devices','SCIGOV-DOEP'); return false;" href="https://www.osti.gov/pages/biblio/1346969-rapid-continuous-magnetic-separation-droplet-microfluidic-devices"><span>Rapid and continuous magnetic separation in <span class="hlt">droplet</span> microfluidic devices</span></a></p> <p><a target="_blank" href="http://www.osti.gov/pages">DOE PAGES</a></p> <p>Brouzes, Eric; Kruse, Travis; Kimmerling, Robert; ...</p> <p>2014-12-03</p> <p>Here, we present a <span class="hlt">droplet</span> microfluidic method to extract molecules of interest from a <span class="hlt">droplet</span> in a rapid and continuous fashion. We accomplish this by first marginalizing functionalized super-paramagnetic beads within the <span class="hlt">droplet</span> using a magnetic field, and then splitting the <span class="hlt">droplet</span> into one <span class="hlt">droplet</span> containing the majority of magnetic beads and one <span class="hlt">droplet</span> containing the minority fraction. We quantitatively analysed the factors which affect the efficiency of marginalization and <span class="hlt">droplet</span> splitting to optimize the enrichment of magnetic beads. We first characterized the interplay between the <span class="hlt">droplet</span> velocity and the strength of the magnetic field and its effect on marginalization.more » We found that marginalization is optimal at the midline of the magnet and that marginalization is a good predictor of bead enrichment through splitting at low to moderate <span class="hlt">droplet</span> velocities. Finally, we focused our efforts on manipulating the splitting profile to improve the enrichment provided by asymmetric splitting. We designed asymmetric splitting forks that employ capillary effects to preferentially extract the bead-rich regions of the <span class="hlt">droplets</span>. Our strategy represents a framework to optimize magnetic bead enrichment methods tailored to the requirements of specific <span class="hlt">droplet</span>-based applications. We anticipate that our separation technology is well suited for applications in single-cell genomics and proteomics. In particular, our method could be used to separate mRNA bound to poly-dT functionalized magnetic microparticles from single cell lysates to prepare single-cell cDNA libraries.« less</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4323160','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=4323160"><span>Rapid and continuous magnetic separation in <span class="hlt">droplet</span> microfluidic devices</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Brouzes, Eric; Kruse, Travis; Kimmerling, Robert; Strey, Helmut H.</p> <p>2015-01-01</p> <p>We present a <span class="hlt">droplet</span> microfluidic method to extract molecules of interest from a <span class="hlt">droplet</span> in a rapid and continuous fashion. We accomplish this by first marginalizing functionalized super-paramagnetic beads within the <span class="hlt">droplet</span> using a magnetic field, and then splitting the <span class="hlt">droplet</span> into one <span class="hlt">droplet</span> containing the majority of magnetic beads and one <span class="hlt">droplet</span> containing the minority fraction. We quantitatively analysed the factors which affect the efficiency of marginalization and <span class="hlt">droplet</span> splitting to optimize the enrichment of magnetic beads. We first characterized the interplay between the <span class="hlt">droplet</span> velocity and the strength of the magnetic field and its effect on marginalization. We found that marginalization is optimal at the midline of the magnet and that marginalization is a good predictor of bead enrichment through splitting at low to moderate <span class="hlt">droplet</span> velocities. Finally, we focused our efforts on manipulating the splitting profile to improve the enrichment provided by asymmetric splitting. We designed asymmetric splitting forks that employ capillary effects to preferentially extract the bead-rich regions of the <span class="hlt">droplets</span>. Our strategy represents a framework to optimize magnetic bead enrichment methods tailored to the requirements of specific <span class="hlt">droplet</span>-based applications. We anticipate that our separation technology is well suited for applications in single-cell genomics and proteomics. In particular, our method could be used to separate mRNA bound to poly-dT functionalized magnetic microparticles from single cell lysates to prepare single-cell cDNA libraries. PMID:25501881</p> </li> <li> <p><a target="_blank" onclick="trackOutboundLink('https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3142295','PMC'); return false;" href="https://www.pubmedcentral.nih.gov/articlerender.fcgi?tool=pmcentrez&artid=3142295"><span>Surveying implicit <span class="hlt">solvent</span> models for estimating small molecule absolute hydration free energies</span></a></p> <p><a target="_blank" href="http://www.ncbi.nlm.nih.gov/entrez/query.fcgi?DB=pmc">PubMed Central</a></p> <p>Knight, Jennifer L.</p> <p>2011-01-01</p> <p>Implicit <span class="hlt">solvent</span> models are powerful tools in accounting for the aqueous environment at a fraction of the computational expense of explicit <span class="hlt">solvent</span> representations. Here, we compare the ability of common implicit <span class="hlt">solvent</span> models (TC, OBC, OBC2, GBMV, GBMV2, GBSW, GBSW/MS, GBSW/MS2 and FACTS) to reproduce experimental absolute hydration free energies for a series of 499 small neutral molecules that are modeled using AMBER/GAFF parameters and AM1-BCC <span class="hlt">charges</span>. Given optimized surface tension coefficients for scaling the surface area term in the nonpolar contribution, most implicit <span class="hlt">solvent</span> models demonstrate reasonable agreement with extensive explicit <span class="hlt">solvent</span> simulations (average difference 1.0-1.7 kcal/mol and R2=0.81-0.91) and with experimental hydration free energies (average unsigned errors=1.1-1.4 kcal/mol and R2=0.66-0.81). Chemical classes of compounds are identified that need further optimization of their ligand force field parameters and others that require improvement in the physical parameters of the implicit <span class="hlt">solvent</span> models themselves. More sophisticated nonpolar models are also likely necessary to more effectively represent the underlying physics of solvation and take the quality of hydration free energies estimated from implicit <span class="hlt">solvent</span> models to the next level. PMID:21735452</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>