diff --git a/ThesisThinFilmFlowsLBM.pdf b/ThesisThinFilmFlowsLBM.pdf deleted file mode 100644 index 6c33ec0..0000000 Binary files a/ThesisThinFilmFlowsLBM.pdf and /dev/null differ diff --git a/bibliography/Thesis_Swalbe.bib b/bibliography/Thesis_Swalbe.bib index 783e527..c9ccf0c 100644 --- a/bibliography/Thesis_Swalbe.bib +++ b/bibliography/Thesis_Swalbe.bib @@ -73,7 +73,6 @@ @article{alandUnifiedNumericalModel2021 volume = {122}, number = {4}, pages = {903--918}, - issn = {1097-0207}, doi = {10.1002/nme.6567}, abstract = {The wetting of deformable elastic structures has been recently shown to comprise a rich variety of new physical phenomena (stick-slip motion, durotaxis, Shuttleworth effect, etc.) whose fundamental understanding demands for numerical simulation tools. In this article, we develop a novel unified model and numerical method for this problem. As special features, the method includes exact incompressibility and a linear monolithic assembly of the Navier-Stokes and Cahn-Hilliard equations which stabilizes dominant surface tension at small interface lengths. Also, solid viscosity is included which offers the opportunity to simulate the surfing behavior of droplets on viscoelastic substrates for the first time. We show that the method is highly accurate and more robust than most previous approaches and illustrate its potential by numerical simulation examples.}, keywords = {binary fluid structure interaction,elasto-capillarity,moving contact line,phase field method,soft wetting}, @@ -106,7 +105,6 @@ @article{ayresStimuliResponsivePolyelectrolytePolymer2007 number = {1}, pages = {182--189}, publisher = {{American Chemical Society}}, - issn = {0743-7463}, doi = {10.1021/la061526l}, abstract = {We present an account of our research into polyelectrolyte polymer brushes that are capable of acting as stimuli-responsive films. We first detail the synthesis of poly(acrylic acid) polymer brushes using ATRP in a ``grafting from'' strategy. Significantly, we employed a chemical-free deprotection step that should leave the anchoring ester groups intact. We have demonstrated how these polymer assemblies respond to stimuli such as pH and electrolyte concentration. We have used poly(acrylic acid) polymer brushes for the synthesis of metallic nanoparticles and review this work. We have used XPS, ATR-FTIR, and AFM spectroscopy to show the presence of silver and palladium nanoparticles within polymer brushes. Finally, we report the synthesis of AB diblock polyampholyte polymer brushes that represent an extension of polyelectrolyte polymer brushes.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\GFMKMLH8\\Ayres et al. - 2007 - Stimuli-Responsive Polyelectrolyte Polymer Brushes.pdf;C\:\\Users\\Zitzero\\Zotero\\storage\\XMUSHNQ2\\la061526l.html} @@ -164,7 +162,6 @@ @article{beckerThinfilmEquationRecent2005 volume = {17}, number = {9}, pages = {S291}, - issn = {0953-8984}, doi = {10.1088/0953-8984/17/9/002}, urldate = {2023-07-19}, abstract = {This paper is concerned with mathematical aspects of lubrication equations. In the first part, we discuss recent analytical achievements for various types of thin-film equations. Of interest are issues like (non-)uniqueness, wetting behaviour and contact line motion, in particular optimal propagation rates and waiting time or dead core phenomena. In the second part, we shall present novel numerical results for thin-film flow on heterogeneous substrates based on entropy consistent schemes. Finally, we will be concerned with new algorithmic concepts for the simulation of thin-film flow of shear-thinning liquids.} @@ -180,7 +177,6 @@ @article{beckerComplexDewettingScenarios2003 number = {1}, pages = {59--63}, publisher = {{Nature Publishing Group}}, - issn = {1476-4660}, doi = {10.1038/nmat788}, abstract = {In the course of miniaturization of electronic and microfluidic devices, reliable predictions of the stability of ultrathin films have a strategic role for design purposes. Consequently, efficient computational techniques that allow for a direct comparison with experiment become increasingly important. Here we demonstrate, for the first time, that the full complex spatial and temporal evolution of the rupture of ultrathin films can be modelled in quantitative agreement with experiment. We accomplish this by combining highly controlled experiments on different film-rupture patterns with computer simulations using novel numerical schemes for thin-film equations. For the quantitative comparison of the pattern evolution in both experiment and simulation we introduce a novel pattern analysis method based on Minkowski measures. Our results are fundamental for the development of efficient tools capable of describing essential aspects of thin-film flow in technical systems.}, copyright = {2002 Nature Publishing Group}, @@ -227,7 +223,6 @@ @article{benetDisjoiningPressureHealing2014 number = {38}, pages = {22079--22089}, publisher = {{American Chemical Society}}, - issn = {1932-7447}, doi = {10.1021/jp506534b}, abstract = {In this work, we simulate the adsorption of wetting liquid argon films on a model substrate. We calculate the disjoining pressure isotherm and show that it is completely dominated by the long-range van der Waals interactions. Thick films exhibit the expected Hamaker power law decay, but a quantitative description of thin films requires consideration of the detailed structure of the adsorbed layer. The spectrum of film height fluctuations is calculated and shown to provide reliable estimates of the disjoining pressure for all films studied. However, it is observed that the full spectrum can only be reproduced provided that we account for a film height dependent surface tension proportional to the derivative of the disjoining pressure. A simple theory is worked out that describes well the observed film height dependence. Having at hand both the surface tension and the disjoining pressure, we calculate the healing distance of the liquid films, which differs from the classical expectation by a constant of the same order of magnitude as the bulk correlation length. We show these findings have important implications on the behavior of adsorbed liquids and determine corrections to the augmented Young\textendash Laplace equation at the subnanometer length scale.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\U44MB574\\Benet et al. - 2014 - Disjoining Pressure, Healing Distance, and Film He.pdf;C\:\\Users\\Zitzero\\Zotero\\storage\\DZGB9XPQ\\jp506534b.html} @@ -243,7 +238,6 @@ @article{benziLatticeBoltzmannEquation1992a volume = {222}, number = {3}, pages = {145--197}, - issn = {0370-1573}, doi = {10.1016/0370-1573(92)90090-M}, urldate = {2024-01-05}, abstract = {The basic elements of the theory of the lattice Boltzmann equation, a special lattice gas kinetic model for hydrodynamics, are reviewed. Applications are also presented together with some generalizations which allow one to extend the range of applicability of the method to a number of fluid dynamics related problems.}, @@ -261,7 +255,6 @@ @article{berendsenGROMACSMessagepassingParallel1995 volume = {91}, number = {1}, pages = {43--56}, - issn = {0010-4655}, doi = {10.1016/0010-4655(95)00042-E}, abstract = {A parallel message-passing implementation of a molecular dynamics (MD) program that is useful for bio(macro)molecules in aqueous environment is described. The software has been developed for a custom-designed 32-processor ring GROMACS (GROningen MAchine for Chemical Simulation) with communication to and from left and right neighbours, but can run on any parallel system onto which a a ring of processors can be mapped and which supports PVM-like block send and receive calls. The GROMACS software consists of a preprocessor, a parallel MD and energy minimization program that can use an arbitrary number of processors (including one), an optional monitor, and several analysis tools. The programs are written in ANSI C and available by ftp (information: gromacs@chem.rug.nl). The functionality is based on the GROMOS (GROningen MOlecular Simulation) package (van Gunsteren and Berendsen, 1987; BIOMOS B.V., Nijenborgh 4, 9747 AG Groningen). Conversion programs between GROMOS and GROMACS formats are included. The MD program can handle rectangular periodic boundary conditions with temperature and pressure scaling. The interactions that can be handled without modification are variable non-bonded pair interactions with Coulomb and Lennard-Jones or Buckingham potentials, using a twin-range cut-off based on charge groups, and fixed bonded interactions of either harmonic or constraint type for bonds and bond angles and either periodic or cosine power series interactions for dihedral angles. Special forces can be added to groups of particles (for non-equilibrium dynamics or for position restraining) or between particles (for distance restraints). The parallelism is based on particle decomposition. Interprocessor communication is largely limited to position and force distribution over the ring once per time step.}, keywords = {Molecular dynamics,Parallel computing}, @@ -278,7 +271,6 @@ @article{berendsenRuptureThinLiquid2012 number = {26}, pages = {9977--9985}, publisher = {{American Chemical Society}}, - issn = {0743-7463}, doi = {10.1021/la301353f}, abstract = {Thin liquid films on partially wetting substrates are subjected to laminar axisymmetric air-jets impinging at normal incidence. We measured the time at which film rupture occurs and dewetting commences as a function of diameter and Reynolds number of the air-jet. We developed numerical models for the air flow as well as the height evolution of the thin liquid film. The experimental results were compared with numerical simulations based on the lubrication approximation and a phenomenological expression for the disjoining pressure. We achieved quantitative agreement for the rupture times. We found that the film thickness profiles were highly sensitive to the presence of minute quantities of surface-active contaminants.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\DEUTPBM9\\Berendsen et al. - 2012 - Rupture of Thin Liquid Films Induced by Impinging .pdf;C\:\\Users\\Zitzero\\Zotero\\storage\\E76HK6I8\\la301353f.html} @@ -294,7 +286,6 @@ @article{bergeronControllingDropletDeposition2000 number = {6788}, pages = {772--775}, publisher = {{Nature Publishing Group}}, - issn = {1476-4687}, doi = {10.1038/35015525}, abstract = {Controlling the impact of drops onto solid surfaces is important for a wide variey of coating and deposition processes\textemdash for example, the treatment of plants with herbicides and pesticides requires precise targeting in order to meet stringent toxicological regulations. However, the outer wax-like layer of the leaves is a non-wetting substrate that causes sprayed droplets to rebound; often less than 50\% of the initial spray is retained by the plant1. Although the impact and subsequent retraction of non-wetting aqueous drops on a hydrophobic surface have been the subjects of extensive experimental and theoretical work2,3,4,5,6,7, non-newtonian rheological effects have not been considered in any detail. Here we report that, by adding very small amounts of a flexible polymer to the aqueous phase, we can inhibit droplet rebound on a hydrophobic surface and markedly improve deposition without significantly altering the shear viscosity of the solutions. Our results can be understood by taking into account the non-newtonian elongational viscosity, which provides a large resistance to drop retraction after impact, thereby suppressing droplet rebound.}, copyright = {2000 Macmillan Magazines Ltd.}, @@ -310,7 +301,6 @@ @article{bergstromHamakerConstantsInorganic1997 journal = {Advances in Colloid and Interface Science}, volume = {70}, pages = {125--169}, - issn = {0001-8686}, doi = {10.1016/S0001-8686(97)00003-1}, urldate = {2024-01-08}, abstract = {Calculations of Hamaker constants using Lifshitz theory require the availability of accurate dielectric data, especially in the ultraviolet spectral region, and the use of a convenient and appropriate mathematical representation. In this review, a multiple oscillator model {\textemdash} the so-called Ninham-Parsegian (N-P) representation {\textemdash} has been used and spectral parameters for 31 different inorganic materials (including diamond) have been generated from critically evaluated optical data or collected from the literature. For most materials, a two-oscillator model (one UV and one IR term) was used but more detailed representations were included when available. The spectral parameters presented here can be combined with previous data, mainly focused on hydrocarbon and organic systems, to yield an extensive spectral data base for both solids and liquids enabling Lifshitz calculations of Hamaker constants for many materials combinations. Non-retarded Hamaker constants for symmetric material combinations across vacuum (A1v1) and water (A1w1) have been calculated for the different materials; these calculations were performed using the full Lifshitz theory. Asymmetric combinations, A1v3 and A1w3, against four commonly used materials in atomic force microscopy studies: silica, amorphous silicon nitride, sapphire, and muscovite mica, have also been covered. The use of a new dielectric representation for water resulted in significantly lower values of A1w1 compared to previous calculations. Analytical approximations to the full Lifshitz theory were evaluated and found to give surprisingly accurate results (the Tabor-Winterton approximation) for A1v1 when the IR contribution is of minor importance. An attempt to make the TW approximation more general by establishing some scaling relationship between n0 and {$\omega$}UV was met with little success; only the UV spectral parameters of the covalent oxides, sulphides and nitrides may be fitted to a simple power law relation. The Lifshitz calculations in this study were compared with an alternative method where a more detailed dielectric representation in the visible-ultraviolet spectral range was obtained through Kramers-Kronig (K-K) transformation of reflectivity data over a broad frequency range. Despite the difference in dielectric information, the two methods generally yield non-retarded Hamaker constants which do not differ significantly. This is not true for all materials, e.g. water, where a more detailed representation using either an N-P representation with several oscillators or the K-K representation must be used. It was shown that the omission of the static and low frequency contribution in the latter method may result in a significant underestimation of the value for A1w1 when the dispersive contribution becomes very small.}, @@ -329,7 +319,6 @@ @article{besardEffectiveExtensibleProgramming2019 volume = {30}, number = {4}, pages = {827--841}, - issn = {1558-2183}, doi = {10.1109/TPDS.2018.2872064}, abstract = {GPUs and other accelerators are popular devices for accelerating compute-intensive, parallelizable applications. However, programming these devices is a difficult task. Writing efficient device code is challenging, and is typically done in a low-level programming language. High-level languages are rarely supported, or do not integrate with the rest of the high-level language ecosystem. To overcome this, we propose compiler infrastructure to efficiently add support for new hardware or environments to an existing programming language. We evaluate our approach by adding support for NVIDIA GPUs to the Julia programming language. By integrating with the existing compiler, we significantly lower the cost to implement and maintain the new compiler, and facilitate reuse of existing application code. Moreover, use of the high-level Julia programming language enables new and dynamic approaches for GPU programming. This greatly improves programmer productivity, while maintaining application performance similar to that of the official NVIDIA CUDA toolkit.}, keywords = {code generation,Graphics processing units,Graphics processors,Hardware,High level languages,Libraries,Programming,retargetable compilers,very high-level languages}, @@ -344,7 +333,6 @@ @article{besardRapidSoftwarePrototyping2019 journal = {Advances in Engineering Software}, volume = {132}, pages = {29--46}, - issn = {0965-9978}, doi = {10.1016/j.advengsoft.2019.02.002}, abstract = {The software needs of scientists and engineers are growing and their programs are becoming more compute-heavy and problem-specific. This has led to an influx of non-expert programmers, who need to use and program high-performance computing platforms. With the continued stagnation of single-threaded performance, using hardware accelerators such as GPUs or FPGAs is necessary. Adapting software to these compute platforms is a difficult task, especially for non-expert programmers, leading to applications being unable to take advantage of new hardware or requiring extensive rewrites. We propose a programming model that allows non-experts to benefit from high-performance computing, while enabling expert programmers to take full advantage of the underlying hardware. In this model, programs are generically typed, the location of the data is encoded in the type system, and multiple dispatch is used to select functionality based on the type of the data. This enables rapid prototyping, retargeting and reuse of existing software, while allowing for hardware specific optimization if required. Our approach allows development to happen in one source language enabling domain experts and performance engineers to jointly develop a program, without the overhead, friction, and challenges associated with developing in multiple programming languages for the same project. We demonstrate the viability and the core principles of this programming model in Julia using realistic examples, showing the potential of this approach for rapid prototyping, and its applicability for real-life engineering. We focus on usability for non-expert programmers and demonstrate that the potential of the underlying hardware can be fully exploited.}, keywords = {CUDA,Distributed computing,Generic programming,Heterogeneous systems,Julia}, @@ -360,7 +348,6 @@ @article{bestehorn3DLargeScale2003 volume = {33}, number = {4}, pages = {457--467}, - issn = {1434-6036}, doi = {10.1140/epjb/e2003-00186-3}, abstract = {We study large scale surface deformations of a liquid filmunstable due to the Marangoni effect caused by external heatingon a smooth and solid substrate. The work is based on the thinfilm equation which can be derived from the basic hydrodynamicequations. To prevent rupture, a repelling disjoining pressureis included which accounts for the stabilization of a thinprecursor film and so prevents the occurrence of completely dryregions. Linear stability analysis, nonlinear stationarysolutions, as well as three-dimensional time dependent numericalsolutions for horizontal and inclined substrates reveal a richscenario of possible structures for several realistic fluidparameters.}, keywords = {47.20.Dr Surface-tension-driven instability,68.15.+e Liquid thin films,68.55.-a Thin film structure and morphology}, @@ -378,7 +365,6 @@ @article{bezansonJuliaFreshApproach2017 number = {1}, pages = {65--98}, publisher = {{Society for Industrial and Applied Mathematics}}, - issn = {0036-1445}, doi = {10.1137/141000671}, abstract = {Bridging cultures that have often been distant, Julia combines expertise from the diverse fields of computer science and computational science to create a new approach to numerical computing. Julia is designed to be easy and fast and questions notions generally held to be ``laws of nature" by practitioners of numerical computing: \textbackslash beginlist \textbackslash item High-level dynamic programs have to be slow. \textbackslash item One must prototype in one language and then rewrite in another language for speed or deployment. \textbackslash item There are parts of a system appropriate for the programmer, and other parts that are best left untouched as they have been built by the experts. \textbackslash endlist We introduce the Julia programming language and its design---a dance between specialization and abstraction. Specialization allows for custom treatment. Multiple dispatch, a technique from computer science, picks the right algorithm for the right circumstance. Abstraction, which is what good computation is really about, recognizes what remains the same after differences are stripped away. Abstractions in mathematics are captured as code through another technique from computer science, generic programming. Julia shows that one can achieve machine performance without sacrificing human convenience.}, keywords = {65Y05,68N15,97P40,Julia,numerical,parallel,scientific computing}, @@ -395,7 +381,6 @@ @article{bhardwajLikelihoodSurvivalCoronavirus2020 number = {6}, pages = {061704}, publisher = {{American Institute of Physics}}, - issn = {1070-6631}, doi = {10.1063/5.0012009}, abstract = {We predict and analyze the drying time of respiratory droplets from a COVID-19 infected subject, which is a crucial time to infect another subject. Drying of the droplet is predicted by using a diffusion-limited evaporation model for a sessile droplet placed on a partially wetted surface with a pinned contact line. The variation in droplet volume, contact angle, ambient temperature, and humidity are considered. We analyze the chances of the survival of the virus present in the droplet based on the lifetime of the droplets under several conditions and find that the chances of the survival of the virus are strongly affected by each of these parameters. The magnitude of shear stress inside the droplet computed using the model is not large enough to obliterate the virus. We also explore the relationship between the drying time of a droplet and the growth rate of the spread of COVID-19 in five different cities and find that they are weakly correlated.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\8RQDHW8F\\Bhardwaj and Agrawal - 2020 - Likelihood of survival of coronavirus in a respira.pdf} @@ -491,7 +476,6 @@ @article{bouchutNewModelShallow2013 number = {08}, pages = {1479--1526}, publisher = {{World Scientific Publishing Co.}}, - issn = {0218-2025}, doi = {10.1142/S0218202513500140}, abstract = {We propose a new reduced model for gravity-driven free-surface flows of shallow viscoelastic fluids. It is obtained by an asymptotic expansion of the upper-convected Maxwell model for viscoelastic fluids. The viscosity is assumed small (of order epsilon, the aspect ratio of the thin layer of fluid), but the relaxation time is kept finite. In addition to the classical layer depth and velocity in shallow models, our system describes also the evolution of two components of the stress. It has an intrinsic energy equation. The mathematical properties of the model are established, an important feature being the non-convexity of the physically relevant energy with respect to conservative variables, but the convexity with respect to the physically relevant pseudo-conservative variables. Numerical illustrations are given, based on a suitable well-balanced finite-volume discretization involving an approximate Riemann solver.}, keywords = {Maxwell model,Oldroyd model,pseudo-conservative variables,Saint-Venant model,shallow-water,Viscoelastic fluids,well-balanced scheme}, @@ -506,7 +490,6 @@ @article{brabecPlasticSolarCells2001 volume = {11}, number = {1}, pages = {15--26}, - issn = {1616-3028}, doi = {10.1002/1616-3028(200102)11:1<15::AID-ADFM15>3.0.CO;2-A}, abstract = {Recent developments in conjugated-polymer-based photovoltaic elements are reviewed. The photophysics of such photoactive devices is based on the photo-induced charge transfer from donor-type semiconducting conjugated polymers to acceptor-type conjugated polymers or acceptor molecules such as Buckminsterfullerene, C60. This photo-induced charge transfer is reversible, ultrafast (within 100 fs) with a quantum efficiency approaching unity, and the charge-separated state is metastable (up to milliseconds at 80 K). Being similar to the first steps in natural photosynthesis, this photo-induced electron transfer leads to a number of potentially interesting applications, which include sensitization of the photoconductivity and photovoltaic phenomena. Examples of photovoltaic architectures are presented and their potential in terrestrial solar energy conversion discussed. Recent progress in the realization of improved photovoltaic elements with 3 \% power conversion efficiency is reported.}, keywords = {Conducting Polymers,Photovoltaics,Solar Cells}, @@ -524,7 +507,6 @@ @article{brasjenDewettingThinLiquid2013 volume = {14}, number = {3}, pages = {669--682}, - issn = {1613-4990}, doi = {10.1007/s10404-012-1086-4}, abstract = {Using experiments and numerical simulations, we investigate the dewetting of thin liquid films on chemically patterned substrates. The patterns consist of long and narrow hydrophobic stripes, separated by larger hydrophilic domains. We characterize the morphology and dynamics of the dewetting front starting from an initially present dry-spot. Moreover, we study the distortion of the liquid film on the adjacent hydrophilic domains as a function of film thickness, hydrophobic contact angle and pattern dimensions. Implications of our results on the solution processing of organic electronic devices on chemically patterned surfaces are discussed.}, keywords = {Chemically patterned surfaces,Dewetting,Solution processing,Thin liquid films}, @@ -558,7 +540,6 @@ @article{brenierGeneralizedSolutionsHydrostatic2008 volume = {237}, number = {14}, pages = {1982--1988}, - issn = {0167-2789}, doi = {10.1016/j.physd.2008.02.026}, abstract = {Solutions to the Euler equations on a 3D domain D3 (typically the unit cube or the periodic unit cube) can be formally obtained by minimizing the action of an incompressible fluid moving inside D3 between two given configurations. When these two configurations are very close to each other, classical solutions do exist, as shown by Ebin and Marsden. However, Shnirelman found a class of data (essentially 2D in the sense that they trivially depend on the vertical coordinate) for which there cannot be any classical minimizer. For such data, generalized solutions can be shown to exist, as a substitute for classical solutions. These generalized solutions have unusual features that look highly unphysical (in particular, different fluid parcels can cross at the same point and at the same time), but the pressure field, which does not depend on the vertical coordinate, is well and uniquely defined. In the present paper, we show that these generalized solutions are actually quite conventional in the sense they obey, up to a suitable change of variable, a well-known variant (widely used for geophysical flows) of the 3D Euler equations, for which the vertical acceleration is neglected according to the so-called hydrostatic approximation.}, keywords = {Euler equations,Generalized solutions,Hydrostatic approximation,Incompressible fluids}, @@ -624,7 +605,6 @@ @article{buttAdaptiveWettingAdaptation2018 number = {38}, pages = {11292--11304}, publisher = {{American Chemical Society}}, - issn = {0743-7463}, doi = {10.1021/acs.langmuir.8b01783}, abstract = {Many surfaces reversibly change their structure and interfacial energy upon being in contact with a liquid. Such surfaces adapt to a specific liquid. We propose the first order kinetic model to describe dynamic contact angles of such adaptive surfaces. The model is general and does not refer to a particular adaptation process. The aim of the proposed model is to provide a quantitative description of adaptive wetting and to link changes in contact angles to microscopic adaptation processes. By introducing exponentially relaxing interfacial energies and applying Young's equation locally, we predict a change of advancing and receding contact angles depending on the velocity of the contact line. Even for perfectly homogeneous and smooth surfaces, a dynamic contact angle hysteresis is obtained. As possible adaptations, we discuss changes and reconstruction of polymer surfaces or monolayers, diffusion and swelling, adsorption of surfactants, replacement of contaminants, reorientation of liquid molecules, or formation of an electric double layer.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\3LQPE3Q8\\Butt et al. - 2018 - Adaptive Wetting—Adaptation in Wetting.pdf;C\:\\Users\\Zitzero\\Zotero\\storage\\498GYLNX\\acs.langmuir.html} @@ -652,7 +632,6 @@ @article{cassieWettabilityPorousSurfaces1944 journal = {Trans. Faraday Soc.}, volume = {40}, pages = {546}, - issn = {0014-7672}, doi = {10.1039/tf9444000546}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\9BH7BDJV\\Cassie and Baxter - 1944 - Wettability of porous surfaces.pdf} } @@ -682,7 +661,6 @@ @article{cazabatDynamicsWettingEffects1986 volume = {90}, number = {22}, pages = {5845--5849}, - issn = {0022-3654, 1541-5740}, doi = {10.1021/j100280a075}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\Z4RRQDEY\\Cazabat and Stuart - 1986 - Dynamics of wetting effects of surface roughness.pdf} } @@ -695,7 +673,6 @@ @article{liLatticeBoltzmannMethod2020 journal = {IEEE Access}, volume = {8}, pages = {27649--27675}, - issn = {2169-3536}, doi = {10.1109/ACCESS.2020.2971546}, abstract = {Great advances have been made with the lattice Boltzmann (LB) method for complicated fluid phenomena and fundamental thermal processes over the past three decades. This paper presents a systematic overview of the LB method from 1990 to 2018, based on bibliometric analysis and the Science Citation Index Expanded (SCI-E) database. The results show that China took the leading position in this field, followed by the USA and UK. The Chinese Academy of Sciences had the most publications, while the Los Alamos National Laboratory was first as far as highest average citation per paper and h-index are concerned. Physical Review E was the most productive journal and ``Mechanics'' was the most frequently used subject category. Keyword analysis indicated that recent research has focused on the natural convection and heat transfer of nanofluid or multiphase flow in complex porous media. Hydrothermal treatment of nanofluid with shape factor on the conditions, such as variable magnetic fields, thermal radiation and slipping boundary, were the research hotspots. Further research perspectives mainly explore the multiscale models for coupling multiple transport phenomena, morphology optimization of porous parameters, new nanoparticles with shape factor, multicomponent LB method considering Knudsen diffusion effect, LB-based hybrid methods, radiation performance or boiling-heat transfer of nanofluid, and the active control of droplets, may continue to attract more attention. Moreover, some new applications, such as phase change of metal foam, erosion induced by nanaofluid, anode circulating, 3D modeling in thermal systems with vibration, and magnetohydrodynamics microfluid devices, could be of interest going forward.}, keywords = {bibliometrics,Bibliometrics,Computational modeling,fluid systems,Heat transfer,Lattice Boltzmann method,Lattice Boltzmann methods,Market research,Mathematical model,Microscopy,multidisciplinary,multiscale modeling,thermal processes}, @@ -769,7 +746,6 @@ @article{chenExperimentsSpreadingDrop1988 volume = {122}, number = {1}, pages = {60--72}, - issn = {0021-9797}, doi = {10.1016/0021-9797(88)90287-1}, abstract = {Experimental results are reported on the spreading of a nonvolatile liquid drop on a horizontal smooth solid surface with negligible effects of gravity and inertia. The liquid is strongly wetting to the solid with an advancing static contact angle \texttheta s less than 1\textdegree. The drop radius R, the apex height H, and the advancing dynamic contact angle \texttheta{} were measured as functions of time t. For each drop the data of R-t, H-t, and \texttheta -t follow different power laws, {$\alpha$}t{$\beta$}. The values of {$\alpha$} and {$\beta$} for R are close to those predicted by Starov (J. Colloid USSR (Engl. Trans)45(6), 1009 (1983)) for a spreading drop under the action of capillary and viscous forces. But the data for H and \texttheta{} show less satisfactory agreement with the predictions. For \texttheta{} two different quantities are measured: \texttheta 0 is the angle between the tangent to the drop edge and the solid surface and \texttheta RH is the angle calculated from R and H for a spherical drop. The measured \texttheta{} range from 1.5 to 17\textdegree, for capillary number C from 1.8 \texttimes{} 10-6 to 3.3 \texttimes{} 10-4. Correlation by a least-squares fit for the \texttheta -C data gives \texttheta 0 = 222C0.353and \texttheta RH = 242C0.353 (both in degrees). Good agreement is found between our \texttheta -C data and correlations with those presented in the literature for experiments in a capillary tube (R. L. Hoffman, J. Colloid Interface Sci.50, 228 (1975); T. S. Jiang et al., J. Colloid Interface Sci.69, 74 (1979) and in parallel plate (C. G. Ngan and E. B. Dussan V, J. Fluid Mech.118, 27 (1982)) for a wide range of \texttheta{} and C. This suggests that, when the capillary and viscous forces are the only dominant factors, \texttheta{} is independent of the experimental geometry. And hence the correlations can be used to estimate \texttheta{} when a nonvolatile liquid displaces air with \texttheta s {$<$} 1\textdegree{} under a dynamically similar condition.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\ZS9NV3GM\\0021979788902871.html} @@ -784,7 +760,6 @@ @article{chenLatticeBoltzmannMethod1998 volume = {30}, number = {1}, pages = {329--364}, - issn = {0066-4189, 1545-4479}, doi = {10.1146/annurev.fluid.30.1.329}, abstract = {We present an overview of the lattice Boltzmann method (LBM), a parallel and efficient algorithm for simulating single-phase and multiphase fluid flows and for incorporating additional physical complexities. The LBM is especially useful for modeling complicated boundary conditions and multiphase interfaces. Recent extensions of this method are described, including simulations of fluid turbulence, suspension flows, and reaction diffusion systems.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\2B4C4HAW\\Chen and Doolen - 1998 - LATTICE BOLTZMANN METHOD FOR FLUID FLOWS.pdf} @@ -800,7 +775,6 @@ @article{chenLatticeMethodsTheir1995 volume = {19}, number = {6}, pages = {617--646}, - issn = {0098-1354}, doi = {10.1016/0098-1354(94)00072-7}, abstract = {The recent development of the lattice gas automata method and its extension to the lattice Boltzmann method have provided new computational schemes for solving a variety of partial differential equations and modeling chemically reacting systems. The lattice gas method, regarded as the simplest microscopic and kinetic approach which generates meaningful macroscopic dynamics, is fully parallel and can, as a result, be easily programmed on parallel machines. In this paper, we introduce the basic principles of the lattice gas method and the lattice Boltzmann method, their numerical implementations and applications to chemically reacting systems. Comparisons of the lattice Boltzmann method with the lattice gas technique and other traditional numerical schemes, including the finite difference scheme and the pseudo-spectral method, for solving the Navier-Stokes hydrodynamic fluid flows will be discussed. Recent developments of the lattice gas and the lattice Boltzmann method and their applications to pattern formation in chemical reaction-diffusion systems, multiphase fluid flows and polymeric dynamics will be presented.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\ZHJLQQ2L\\Chen et al. - 1995 - Lattice methods and their applications to reacting.pdf;C\:\\Users\\Zitzero\\Zotero\\storage\\B8DRYGQP\\0098135494000727.html} @@ -815,7 +789,6 @@ @article{chenOpenFOAMComputationalFluid2014 volume = {61}, number = {4}, pages = {354}, - issn = {0002-9920, 1088-9477}, doi = {10.1090/noti1095}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\33Z5T6LD\\Chen et al. - 2014 - OpenFOAM for Computational Fluid Dynamics.pdf} } @@ -863,7 +836,6 @@ @article{courantUeberPartiellenDifferenzengleichungen1928 volume = {100}, number = {1}, pages = {32--74}, - issn = {1432-1807}, doi = {10.1007/BF01448839}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\V2EWHVIB\\Courant et al. - 1928 - Über die partiellen Differenzengleichungen der mat.pdf} } @@ -878,7 +850,6 @@ @article{crankPracticalMethodNumerical1947 number = {1}, pages = {50--67}, publisher = {{Cambridge University Press}}, - issn = {1469-8064, 0305-0041}, doi = {10.1017/S0305004100023197}, abstract = {This paper is concerned with methods of evaluating numerical solutions of the non-linear partial differential equation where subject to the boundary conditions A, k, q are known constants.Equation (1) is of the type which arises in problems of heat flow when there is an internal generation of heat within the medium; if the heat is due to a chemical reaction proceeding at each point at a rate depending upon the local temperature, the rate of heat generation is often defined by an equation such as (2).} } @@ -892,7 +863,6 @@ @article{cardiffThirtyYearsFinite2021 volume = {28}, number = {5}, pages = {3721--3780}, - issn = {1886-1784}, doi = {10.1007/s11831-020-09523-0}, urldate = {2023-07-25}, abstract = {Since early publications in the late 1980s and early 1990s, the finite volume method has been shown suitable for solid mechanics analyses. At present, there are several flavours of the method, which can be classified in a variety of ways, such as grid arrangement (cell-centred vs. staggered vs.~vertex-centred), solution algorithm (implicit vs. explicit), and stabilisation strategy (Rhie\textendash Chow vs. Jameson\textendash Schmidt\textendash Turkel vs. Godunov upwinding). This article gives an overview, historical perspective, comparison and critical analysis of the different approaches where a close comparison with the de facto standard for computational solid mechanics, the finite element method, is given. The article finishes with a look towards future research directions and steps required for finite volume solid mechanics to achieve more widespread acceptance.}, @@ -924,7 +894,6 @@ @article{darhuberSelectiveDipcoatingChemically2000 number = {9}, pages = {5119--5126}, publisher = {{American Institute of Physics}}, - issn = {0021-8979}, doi = {10.1063/1.1317238}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\GE4FQFZ6\\Darhuber et al. - 2000 - Selective dip-coating of chemically micropatterned.pdf} } @@ -937,7 +906,6 @@ @article{dasilvasobrinhoStudyDefectsUltrathin1999 journal = {Surf. Coat. Technol.}, volume = {116--119}, pages = {1204--1210}, - issn = {0257-8972}, doi = {10.1016/S0257-8972(99)00152-8}, abstract = {Ultra-thin layers of SiO2 and SiN prepared, for example, by plasma-enhanced chemical vapor deposition (PECVD) are increasingly used as gas barriers on flexible polymeric materials and of plastic containers. Despite the excellent barrier properties provided by these materials, all published data show some residual permeation, even when the barrier coatings are relatively thick ({$\geq$}70 nm). This residual permeation is attributed to the presence of microscopic defects in the coatings. In this article we present new techniques, based mainly on reactive ion etching in oxygen plasma, to render visible micrometer- or sub-micrometer-sized defects in transparent ceramic films on polymers. These techniques can be used to visualize and better understand the origins of defects in these coatings on a microscopic scale, as well as for mapping and counting defect density on a macroscopic scale (tens of cm2 or more).}, keywords = {Barrier coatings,Defects,Microscopies,PECVD,Permeation,Silicon compounds}, @@ -996,7 +964,6 @@ @article{degennesFluidWallSlippage2002 number = {9}, pages = {3413--3414}, publisher = {{American Chemical Society}}, - issn = {0743-7463}, doi = {10.1021/la0116342}, abstract = {Certain (nonpolymeric) fluids show an anomalously low friction when flowing against well-chosen solid walls. We discuss here one possible explanation, postulating that a gaseous film of small thickness h is present between fluid and wall. When h is smaller than the mean free path l of the gas (Knudsen regime), the Navier length b is expected to be independent of h and very large (micrometers).}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\PRYCDPKZ\\de Gennes - 2002 - On FluidWall Slippage.pdf;C\:\\Users\\Zitzero\\Zotero\\storage\\2VXNMEJ6\\la0116342.html} @@ -1043,7 +1010,6 @@ @article{dellarShallowWaterEquations2005 number = {10}, pages = {106601}, publisher = {{American Institute of Physics}}, - issn = {1070-6631}, doi = {10.1063/1.2116747}, abstract = {This paper derives a set of two-dimensional equations describing a thin inviscid fluid layer flowing over topography in a frame rotating about an arbitrary axis. These equations retain various terms involving the locally horizontal components of the angular velocity vector that are discarded in the usual shallow water equations. The obliquely rotating shallow water equations are derived both by averaging the three-dimensional equations and from an averaged Lagrangian describing columnar motion using Hamilton's principle. They share the same conservation properties as the usual shallow water equations, for the same energy and modified forms of the momentum and potential vorticity. They may also be expressed in noncanonical Hamiltonian form using the usual shallow water Hamiltonian and Poisson bracket. The conserved potential vorticity takes the standard shallow water form, but with the vertical component of the rotation vector replaced by the component locally normal to the surface midway between the upper and lower boundaries.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\E79XJMLZ\\Dellar and Salmon - 2005 - Shallow water equations with a complete Coriolis f.pdf} @@ -1059,7 +1025,6 @@ @article{delormeAzobenzeneContainingMonolayerPhotoswitchable2005 number = {26}, pages = {12278--12282}, publisher = {{American Chemical Society}}, - issn = {0743-7463}, doi = {10.1021/la051517x}, abstract = {A compact monolayer containing azobenzene has been prepared on silicon substrates. The elaboration route consisted of covalent grafting of freshly synthesized azobenzene moieties onto an isocyanate-functionalized self-assembled monolayer (SAM). The highly packed and ordered isocyanate-functionalized SAM and the azobenzene-functionalized SAM were monitored and characterized by contact angle measurements and X-ray reflectivity (XR). Photoswitching of the wettability of the film induced by the reversible cis-trans isomerization of the azobenzene chromophores is experimentally shown from water and olive oil contact angle measurements.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\TEUBA48S\\Delorme et al. - 2005 - Azobenzene-Containing Monolayer with Photoswitchab.pdf;C\:\\Users\\Zitzero\\Zotero\\storage\\2XCZVNII\\la051517x.html} @@ -1087,7 +1052,6 @@ @article{derjaguinThicknessLiquidFilm1993 volume = {43}, number = {1}, pages = {134--137}, - issn = {0079-6816}, doi = {10.1016/0079-6816(93)90022-N}, abstract = {Film thickness determines many key coating properties including gloss. However, current literature hardly addresses the development of methods to directly correlate film thickness with gloss. Thus, the lack of accurate models hinders further product design and optimisation of coating products and processes. In this contribution, a previously derived gloss-viscosity relationship is turned into a gloss-film thickness mathematical model. Experimental results of both matting and glossy agents are found to be successfully predicted by a model that is revealed as a simple and useful tool for day-to-day calculations. Polybenzoxazine coatings were elaborated by dip coating of a solution, prepared from a commercial bisphenol A benzoxazine (BA-a), on a 1050 aluminum alloy. The monomer was dissolved in acetone and the influence of the different application parameters (withdrawal speed and viscosity of the solution) on the wet coating thickness was evaluated. A heat treatment was then performed on the coating to polymerize the benzoxazine monomer by a ring opening mechanism attested for by Fourier Transform Infrared spectroscopy (FT-IR) and followed by Differential Scanning Calorimetry (DSC). Dielectric Analysis (DEA) and Thermogravimetric Analysis (TGA) showed a particular behavior related to a partial decomposition taking place at 180 \textdegree C and associated with the creation of intermediary ionic and volatile species. Finally, the barrier protection was evaluated by Electrochemical Impedance Spectroscopy (EIS) for 30 days in sodium chloride solution (0.1 M). The results showed an improvement of the impedance modulus from 104 {$\Omega$} cm2 for an uncoated aluminum to a value as high as 109 {$\Omega$} cm2 with a 10-{$\mu$}m thick polybenzoxazine coating.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\5C9ZP9NK\\007968169390022N.html} @@ -1102,7 +1066,6 @@ @article{deryckGravityInertiaEffects1998 volume = {203}, number = {2}, pages = {278--285}, - issn = {0021-9797}, doi = {10.1006/jcis.1998.5444}, abstract = {Coating processes are of technological interest. We focus here on the extraction of a vertical plate out of a wetting liquid. We first summarize the Landau-Levich-Derjaguin theory, including the gravity corrections which have been proposed in particular by White and Tallmadge. We propose a new numerical solution to the problem. Then, we discuss further developments: for liquids of low viscosity, it is shown that above a threshold in capillary number, the film is thickened because of inertia. A simple scaling argument is proposed for predicting the location of the threshold.}, keywords = {fluid coating,thin liquid films,wetting}, @@ -1119,7 +1082,6 @@ @article{diezBreakupFluidRivulets2009 number = {8}, pages = {082105}, publisher = {{American Institute of Physics}}, - issn = {1070-6631}, doi = {10.1063/1.3211248}, abstract = {We study the stability of rivulets on horizontal substrates. The implemented model includes the effects of capillarity, fluid-solid interaction, and gravity if appropriate, within the framework of the lubrication approximation. We find that the results compare favorably with those in literature, in the regime where previous analyses are valid. By isolating the effect of van der Waals interactions for nanoscale rivulets, and of gravity for macrosize rivulets, we are able to analyze the influence of these forces on the stability. We discuss in detail the scaling of the emerging wavelengths (distance between drops formed after the breakup process) with the rivulet cross-sectional area. Perhaps surprisingly, we uncover close connection between this scaling and the one for the breakup of a free-space fluid jet (Rayleigh\textendash Plateau instability). Finally, we consider rivulets of finite length and find that the finite size effects are considerably different from the ones obtained previously for semi-infinite fluid films.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\XRGARXZP\\Diez et al. - 2009 - On the breakup of fluid rivulets.pdf} @@ -1149,7 +1111,6 @@ @article{hamakerLondonVanWaals1937 volume = {4}, number = {10}, pages = {1058--1072}, - issn = {0031-8914}, doi = {10.1016/S0031-8914(37)80203-7}, urldate = {2023-07-19}, abstract = {Frequently we experience the existance of adhesive forces between small particles. It seems natural to ascribe this adhesion for a large part to London-v.d. Waals forces. To obtain general information concerning their order of magnitude the London-v. d. Waals interaction between two spherical particles is computed as a function of the diameters and the distance separating them. A table is calculated which enables numerical application of the formulae derived. Besides approximations are added, which may be used when the distance between the particles is small. In a separate section it is investigated how the results must be modified, when both particles are immersed in a liquid. Here we are led to the important conclusion that even in that case London-v. d. Waals forces generally cause an attraction.}, @@ -1180,7 +1141,6 @@ @article{diezStabilityFinitelengthRivulet2009 volume = {166}, number = {1}, pages = {012009}, - issn = {1742-6596}, doi = {10.1088/1742-6596/166/1/012009}, abstract = {We study the stability of a finite-length fluid rivulet at rest on a partially wetting surface. We consider the problem by including the intermolecular force (van der Waals interaction) within the framework of the lubrication approximation. The results are validated by comparison with numerical simulations of the full nonlinear equation. For finite length rivulets, we show that the distance between drops after breakup is very close to the wavelength of maximum growth rate predicted by the linear theory for infinite rivulets. Finally, we compare theoretical and numerical results with reported experimental data.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\7A8BHRMS\\Diez et al. - 2009 - Stability of a finite-length rivulet under partial.pdf} @@ -1195,7 +1155,6 @@ @article{ducheminInviscidCoalescenceDrops2003 volume = {487}, pages = {167--178}, publisher = {{Cambridge University Press}}, - issn = {1469-7645, 0022-1120}, doi = {10.1017/S0022112003004646}, abstract = {We study the coalescence of two drops of an ideal fluid driven by surface tension. The velocity of approach is taken to be zero and the dynamical effect of the outer fluid (usually air) is neglected. Our approximation is expected to be valid on scales larger than \$\textbackslash ell\_\{\textbackslash nu\} = \textbackslash rho\textbackslash nu\^2/\textbackslash sigma\$, which is 10 nm for water. Using a high-precision boundary integral method, we show that the walls of the thin retracting sheet of air between the drops reconnect in finite time to form a toroidal enclosure. After the initial reconnection, retraction starts again, leading to a rapid sequence of enclosures. Averaging over the discrete events, we find the minimum radius of the liquid bridge connecting the two drops to scale like \$r\_b \textbackslash propto t\^\{1/2\}\$.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\Y2IDQ3LL\\Duchemin et al. - 2003 - Inviscid coalescence of drops.pdf} @@ -1223,7 +1182,6 @@ @article{duran-olivenciaInstabilityRuptureFluctuations2019 volume = {174}, number = {3}, pages = {579--604}, - issn = {1572-9613}, doi = {10.1007/s10955-018-2200-0}, abstract = {Thin liquid films are ubiquitous in natural phenomena and technological applications. They have been extensively studied via deterministic hydrodynamic equations, but thermal fluctuations often play a crucial role that needs to be understood. An example of this is dewetting, which involves the rupture of a thin liquid film and the formation of droplets. Such a process is thermally activated and requires fluctuations to be taken into account self-consistently. In this work we present an analytical and numerical study of a stochastic thin-film equation derived from first principles. Following a brief review of the derivation, we scrutinise the behaviour of the equation in the limit of perfectly correlated noise along the wall-normal direction, as opposed to the perfectly uncorrelated limit studied by Gr\"un et al. (J. Stat. Phys. 122(6):1261\textendash 1291, 2006). We also present a numerical scheme based on a spectral collocation method, which is then utilised to simulate the stochastic thin-film equation. This scheme seems to be very convenient for numerical studies of the stochastic thin-film equation, since it makes it easier to select the frequency modes of the noise (following the spirit of the long-wave approximation). With our numerical scheme we explore the fluctuating dynamics of the thin film and the behaviour of its free energy in the vicinity of rupture. Finally, we study the effect of the noise intensity on the rupture time, using a large number of sample paths as compared to previous studies.}, keywords = {Dewetting,Fluctuating hydrodynamics,Thermal fluctuations,Thin liquid films}, @@ -1270,7 +1228,6 @@ @article{eggersCoalescenceLiquidDrops1999 volume = {401}, pages = {293--310}, publisher = {{Cambridge University Press}}, - issn = {1469-7645, 0022-1120}, doi = {10.1017/S002211209900662X}, abstract = {When two drops of radius R touch, surface tension drives an initially singular motion which joins them into a bigger drop with smaller surface area. This motion is always viscously dominated at early times. We focus on the early-time behaviour of the radius rm of the small bridge between the two drops. The flow is driven by a highly curved meniscus of length 2{$\pi$}rm and width {$\Delta$}[Lt ]rm around the bridge, from which we conclude that the leading-order problem is asymptotically equivalent to its two-dimensional counterpart. For the case of inviscid surroundings, an exact two-dimensional solution (Hopper 1990) shows that {$\Delta\propto$}r3m and rm{$\sim$}(t{$\gamma$}/{$\pi\eta$}) ln [t{$\gamma$}({$\eta$}R)]; and thus the same is true in three dimensions. We also study the case of coalescence with an external viscous fluid analytically and, for the case of equal viscosities, in detail numerically. A significantly different structure is found in which the outer-fluid forms a toroidal bubble of radius {$\Delta\propto$}r3/2m at the meniscus and rm{$\sim$}(t{$\gamma$}/4{$\pi\eta$}) ln [t{$\gamma$}/({$\eta$}R)]. This basic difference is due to the presence of the outer-fluid viscosity, however small. With lengths scaled by R a full description of the asymptotic flow for rm(t)[Lt ]1 involves matching of lengthscales of order r2m, r3/2m, rm, 1 and probably r7/4m.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\W96MUJUM\\Eggers et al. - 1999 - Coalescence of liquid drops.pdf} @@ -1327,7 +1284,6 @@ @article{fantonSpreadingInstabilitiesInduced1998 number = {9}, pages = {2554--2561}, publisher = {{American Chemical Society}}, - issn = {0743-7463}, doi = {10.1021/la971292t}, abstract = {We study the spreading of liquid films driven by surface tension gradients induced by evaporation from a two-component mixture. The films climb from a macroscopic reservoir on a plane tilted surface and their length L is found to depend linearly on the square root of time t:\, L(t) = (Dt)0.5. We develop a semiquantitative analysis that shows which parameters control the value of D for ideal mixtures and for nonideal ones. We report also experimental results about the time evolution and the spatiotemporal behavior of the interfacial instability that develops at the meniscus between the reservoir and the film. Results agree well with previous experimental and theoretical studies.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\5DRPUUAL\\Fanton and Cazabat - 1998 - Spreading and Instabilities Induced by a Solutal M.pdf;C\:\\Users\\Zitzero\\Zotero\\storage\\KZCG92AJ\\la971292t.html} @@ -1343,7 +1299,6 @@ @article{fantonThicknessShapeFilms1996 number = {24}, pages = {5875--5880}, publisher = {{American Chemical Society}}, - issn = {0743-7463}, doi = {10.1021/la960488a}, abstract = {We study an example of spreading driven by surface tension gradients, i.e. the climbing of a film on a plane wall against gravity. We show which parameters control the thickness of the film. Moreover, we present a theoretical and experimental study of the crossover between the film and the macroscopic reservoir.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\X2WDPMDU\\Fanton et al. - 1996 - Thickness and Shape of Films Driven by a Marangoni.pdf;C\:\\Users\\Zitzero\\Zotero\\storage\\22MAYH4G\\la960488a.html} @@ -1359,7 +1314,6 @@ @article{fedorchenkoFormationDrySpots2021 number = {2}, pages = {023601}, publisher = {{American Institute of Physics}}, - issn = {1070-6631}, doi = {10.1063/5.0035547}, abstract = {Here, the phenomenon of food sticking when frying in a frying pan is experimentally explained. Thermocapillary convection causes a dry spot formation in the center of the frying pan upon heating of the sunflower oil film. It is shown that the speed of formation of a dry spot is similar to the speed of receding motion of the edge of a droplet upon impact and spreading on a solid surface. This allows theoretical determination of the speed of dewetting. For the thin liquid film flowing vertically over a solid surface, the critical volumetric flow rate qcr partitions two regimes: metastable or subcritical, when small perturbation of the film free surface results in the film rupture (q {$<$} qcr) and stable or supercritical at q {$>$} qcr. For the falling thin liquid film, the critical volumetric flow rate qcr partitions two regimes: metastable or subcritical (q {$<$} qcr) and stable or supercritical at q {$>$} qcr. At q {$<$} qcr, small deformations of the film free surface result in the film rupture. For the case of the temperature distribution in the form of a unit step function, the fundamental solution G1(x) describing the deformation of the film free surface has been derived by the perturbation technique. This solution is important by itself since it describes the most ``dangerous'' film surface profile at a prescribed value of the temperature drop. For an arbitrary surface temperature distribution \texttheta{} ({$\xi$}), the convolution of G1({$\xi$}) and \texttheta{} {${'}$}({$\xi$}) yields the film thickness profile.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\LDDKMA6N\\Fedorchenko and Hruby - 2021 - On formation of dry spots in heated liquid films.pdf} @@ -1391,7 +1345,6 @@ @article{fetzerQuantifyingHydrodynamicSlip2007 number = {21}, pages = {10559--10566}, publisher = {{American Chemical Society}}, - issn = {0743-7463}, doi = {10.1021/la7010698}, abstract = {To characterize nontrivial boundary conditions of a liquid flowing past a solid, the slip length is commonly used as a measure. From the profile of a retracting liquid front (e.g., measured with atomic force microscopy), the slip length can be extracted with the help of a Stokes model for a thin liquid film dewetting from a solid substrate. Specifically, we use a lubrication model derived from the Stokes model for strong slippage and linearize the film profile around the flat, unperturbed film. For small slip lengths, we expand the linearized full Stokes model for small slopes up to third order. Using the respective model, we obtain, in addition to the slip length, the capillary number, from which we can estimate the viscosity of the fluid film. We compare numerical and experimental results, test the consistency and the validity of the models/approximations, and give an easy-to-follow guide of how they can be used to analyze experiments.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\IL4S9AJ9\\Fetzer et al. - 2007 - Quantifying Hydrodynamic Slip A Comprehensive An.pdf;C\:\\Users\\Zitzero\\Zotero\\storage\\LAG6UTBU\\la7010698.html} @@ -1422,7 +1375,6 @@ @article{fischerExistencePositiveSolutions2018 number = {1}, pages = {411--455}, publisher = {{Society for Industrial and Applied Mathematics}}, - issn = {0036-1410}, doi = {10.1137/16M1098796}, abstract = {We construct martingale solutions to stochastic thin-film equations by introducing a (spatial) semidiscretization and establishing convergence. The discrete scheme allows for variants of the energy and entropy estimates in the continuous setting as long as the discrete energy does not exceed certain threshold values depending on the spatial grid size h h . Using a stopping time argument to prolongate high-energy paths constant in time, arbitrary moments of coupled energy/entropy functionals can be controlled. Having established H\"older regularity of approximate solutions, the convergence proof is then based on compactness arguments---in particular on Jakubowski's generalization of Skorokhod's theorem---weak convergence methods, and recent tools on martingale convergence.}, keywords = {35G20,35K25,35K65,35Q35,60H15,73D03,76D08,nonnegativity-preserving scheme,stochastic partial differential equation,stochastic thin-film equation,thermal noise,thin-film flow}, @@ -1452,7 +1404,6 @@ @article{furmidgeStudiesPhaseInterfaces1962 volume = {17}, number = {4}, pages = {309--324}, - issn = {0095-8522}, doi = {10.1016/0095-8522(62)90011-9}, abstract = {The surface properties of the spray liquid/solid combination are among the most important factors controlling the retention of spray liquids on solid surfaces. The effect of these properties on the retention of pesticide sprays has been examined by a study of the sliding of drops of water and solutions of surfactants on wax and cellulose acetate surfaces. A theory, supported by experimental results, has been evolved to explain the movement of drops in terms of the size of the drop, the angle of tilt of the surface, the air/liquid surface tension, and the advancing and receding contact angles. This theory has been developed to predict the volume of spray liquid that will be retained on a solid surface. The retentions found in practice agree well with the theoretical figures for most of the spray/solid combinations examined under comparable conditions of spray droplet impaction. A slight modification of the theory permits the correction of some anomalous results obtained with easily wetted surfaces. A retention factor F for any spray liquid/solid combination may be calculated to predict the degree of retention of the liquid on the solid. This factor may be simplified, without undue loss of accuracy, to the form: F = \texttheta M[{$\gamma$}AL(cos \texttheta R - cos \texttheta A)/{$\rho$}]12, where \textbullet\texttheta A is the advancing contact angle;\textbullet\texttheta R is the receding contact angle;\textbullet\texttheta M is the arithmetic mean of \texttheta A and \texttheta R;\textbullet{$\gamma$}al is the air/liquid surface tension; and\textbullet{$\rho$} is the density of the spray liquid. The use of such a retention factor facilitates the study of the formulation of spray fluids.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\RGXIDNSC\\0095852262900119.html} @@ -1486,7 +1437,6 @@ @article{gonzalezInstabilityLiquidCu2013 number = {30}, pages = {9378--9387}, publisher = {{American Chemical Society}}, - issn = {0743-7463}, doi = {10.1021/la4009784}, abstract = {We study the instability of nanometric Cu thin films on SiO2 substrates. The metal is melted by means of laser pulses for some tens of nanoseconds, and during the liquid lifetime, the free surface destabilizes, leading to the formation of holes at first and then in later stages of the instability to metal drops on the substrate. By analyzing the Fourier transforms of the SEM (scanning electron microscope) images obtained at different stages of the metal film evolution, we determine the emerging length scales at relevant stages of the instability development. The results are then discussed within the framework of a long-wave model. We find that the results may differ whether early or final stages of the instability are considered. On the basis of the interpretation of the experimental results, we discuss the influence of the parameters describing the interaction of the liquid metal with the solid substrate. By considering both the dependence of dominant length scales on the film thickness and the measured contact angle, we isolate a model which predicts well the trends found in the experimental data.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\U7NPKA9W\\González et al. - 2013 - Instability of Liquid Cu Films on a SiO2 Substrate.pdf;C\:\\Users\\Zitzero\\Zotero\\storage\\7A5ARBL5\\la4009784.html} @@ -1501,7 +1451,6 @@ @article{grabowThinFilmGrowth1988 volume = {194}, number = {3}, pages = {333--346}, - issn = {0039-6028}, doi = {10.1016/0039-6028(88)90858-8}, abstract = {We discuss the wetting of a crystalline substrate by a deposit, and the equilibrium and metastable configurations that can be expected under various conditions. The stability of different configurations of the deposit was investigated by the use of molecular dynamics techniques. The chemical potential of a uniform film is compared with that or a deposit that aggregates into three-dimensional clusters. In the case where the uniform film is metastable, the rate of nucleation of clusters is calculated in terms of the driving force and the surface and interfacial energies of the system. The degree of misfit between the film and substrate material and the resulting strain in the film is found to have a fundamental role in determining the stability of the structures and nucleation rates. Simulations have been performed using two very different interatomic potentials and crystal structures: (1) the truncated Lennard-Jones potential with the face centered cubic structure, and (2) the Stillinger-Weber potential for silicon with the diamond cubic structure.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\DSJDXQJ4\\0039602888908588.html} @@ -1530,7 +1479,6 @@ @article{grawitterSteeringDropletsSubstrates2021 number = {9}, pages = {2454--2467}, publisher = {{The Royal Society of Chemistry}}, - issn = {1744-6848}, doi = {10.1039/D0SM02082F}, abstract = {Droplets move on substrates with a spatio-temporal wettability pattern as generated, for example, on light-switchable surfaces. To study such cases, we implement the boundary-element method to solve the governing Stokes equations for the fluid flow field inside and on the surface of a droplet and supplement it by the Cox\textendash Voinov law for the dynamics of the contact line. Our approach reproduces the relaxation of an axisymmetric droplet in experiments, which we initiate by instantaneously switching the uniform wettability of a substrate quantified by the equilibrium contact angle. In a step profile of wettability the droplet moves towards higher wettability. Using a feedback loop to keep the distance or offset between step and droplet center constant, induces a constant velocity with which the droplet surfs on the wettability step. We analyze the velocity in terms of droplet offset and step width for typical wetting parameters. Moving instead the wettability step with constant speed, we determine the maximally possible droplet velocities under various conditions. The observed droplet speeds agree with the values from the feedback study for the same positive droplet offset.}, keywords = {boundary_integral,switchable_substrate,third_paper}, @@ -1591,7 +1539,6 @@ @article{savvaViscousSheetRetraction2009 volume = {626}, pages = {211--240}, publisher = {{Cambridge University Press}}, - issn = {1469-7645, 0022-1120}, doi = {10.1017/S0022112009005795}, urldate = {2023-07-24}, abstract = {We present the results of a combined theoretical and numerical investigation of the rim-driven retraction of flat fluid sheets in both planar and circular geometries. Particular attention is given to the influence of the fluid viscosity on the evolution of the sheet and its bounding rim. In both geometries, after a transient that depends on the sheet viscosity and geometry, the film edge eventually attains the Taylor\textendash Culick speed predicted on the basis of inviscid theory. The emergence of this result in the viscous limit is rationalized by consideration of both momentum and energy arguments. We first consider the planar geometry considered by Brenner \& Gueyffier (Phys. Fluids, vol. 11, 1999, p. 737) and deduce new analytical expressions for the speed of the film edge at the onset of rupture and the evolution of the maximum film thickness for viscous films. In order to consider the expansion of a circular hole, we develop an appropriate lubrication model that predicts the form of the early stage dynamics of film rupture. Simulations of a broad range of flow parameters confirm the importance of geometry on the dynamics, verifying the exponential hole growth reported in early experimental studies. We demonstrate the sensitivity of the initial retraction speed on the film profile, and so suggest that the anomalous rate of retraction reported in these experiments may be attributed in part to geometric details of the puncture process.}, @@ -1621,7 +1568,6 @@ @article{grunThinFilmFlowInfluenced2006 volume = {122}, number = {6}, pages = {1261--1291}, - issn = {1572-9613}, doi = {10.1007/s10955-006-9028-8}, abstract = {We study the influence of thermal fluctuations on the dewetting dynamics of thin liquid films. Starting from the incompressible Navier-Stokes equations with thermal noise, we derive a fourth-order degenerate parabolic stochastic partial differential equation which includes a conservative, multiplicative noise term\textemdash the stochastic thin-film equation. Technically, we rely on a long-wave-approximation and Fokker\textendash Planck-type arguments. We formulate a discretization method and give first numerical evidence for our conjecture that thermal fluctuations are capable of accelerating film rupture and that discrepancies with respect to time-scales between physical experiments and deterministic numerical simulations can be resolved by taking noise effects into account.}, keywords = {microfluidics,stochastic hydrodynamics,thin film flow,wetting}, @@ -1689,7 +1635,6 @@ @article{guruswamyReviewNumericalFluids2002 volume = {80}, number = {1}, pages = {31--41}, - issn = {0045-7949}, doi = {10.1016/S0045-7949(01)00164-X}, abstract = {Domain decomposition approaches require efficient interface techniques when fluids and structures are solved in independent computational domains for aerospace applications. Fluid/structure interfacing techniques for solutions from equations based on low-fidelity approaches that are in the linear domain are well advanced and are incorporated in production codes NASTRAN and ASTROS. However, for computations involving high-fidelity equations such as the Navier\textendash Stokes for fluids and finite elements for structures, interface approaches are still under development. This paper provides a technical overview of methods for interfacing flow solutions from the Euler/Navier\textendash Stokes methods with structural solutions using modal/finite-element methods. Validity of the methods is supported by previously presented results.}, keywords = {Aeroelasticity,Computers,Finite elements,High fidelity,Navier–Stokes}, @@ -1721,7 +1666,6 @@ @article{haefnerInfluenceSlipPlateau2015 number = {1}, pages = {7409}, publisher = {{Nature Publishing Group}}, - issn = {2041-1723}, doi = {10.1038/ncomms8409}, abstract = {The Plateau\textendash Rayleigh instability of a liquid column underlies a variety of fascinating phenomena that can be observed in everyday life. In contrast to the case of a free liquid cylinder, describing the evolution of a liquid layer on a solid fibre requires consideration of the solid\textendash liquid interface. Here we revisit the Plateau\textendash Rayleigh instability of a liquid coating a fibre by varying the hydrodynamic boundary condition at the fibre\textendash liquid interface, from no slip to slip. Although the wavelength is not sensitive to the solid\textendash liquid interface, we find that the growth rate of the undulations strongly depends on the hydrodynamic boundary condition. The experiments are in excellent agreement with a new thin-film theory incorporating slip, thus providing an original, quantitative and robust tool to measure slip lengths.}, copyright = {2015 The Author(s)}, @@ -1738,7 +1682,6 @@ @article{harrisonFactorsInfluencingQuality1985 volume = {8}, number = {1}, pages = {1--20}, - issn = {1745-4557}, doi = {10.1111/j.1745-4557.1985.tb00828.x}, abstract = {Food emulsions are an important part of the food industry. Use of one such emulsion, mayonnaise, has grown enormously since it first was produced commercially in the early 1900's. A number of factors influence the mayonnaise emulsion. These factors that have been discussed are: the egg yolk, the relative volume of the phases, the emulsifying effect of mustard, the method of mixing, water hardness, and viscosity. Numerous research efforts have shown these factors to be of great significance in the formation and stability of a high quality mayonnaise.}, annotation = {\_eprint: https://onlinelibrary.wiley.com/doi/pdf/10.1111/j.1745-4557.1985.tb00828.x}, @@ -1747,13 +1690,12 @@ @article{harrisonFactorsInfluencingQuality1985 @article{rayleighInvestigationCharacterEquilibrium1882, title = {Investigation of the {{Character}} of the {{Equilibrium}} of an {{Incompressible Heavy Fluid}} of {{Variable Density}}}, - author = {{Rayleigh}}, + author = {Rayleigh, J. W.}, year = {1882}, journal = {Proc. Lond. Math. Soc.}, volume = {s1-14}, number = {1}, pages = {170--177}, - issn = {1460-244X}, doi = {10.1112/plms/s1-14.1.170}, urldate = {2023-08-10}, copyright = {\textcopyright{} 1882 London Mathematical Society}, @@ -1844,7 +1786,6 @@ @article{hughesIntroductionInhomogeneousLiquids2014 number = {12}, pages = {1119--1129}, publisher = {{American Association of Physics Teachers}}, - issn = {0002-9505}, doi = {10.1119/1.4890823}, abstract = {Classical density functional theory (DFT) is a statistical mechanical theory for calculating the density profiles of the molecules in a liquid. It is widely used, for example, to study the density distribution of the molecules near a confining wall, the interfacial tension, wetting behavior, and many other properties of nonuniform liquids. DFT can, however, be somewhat daunting to students entering the field because of the many connections to other areas of liquid-state science that are required and used to develop the theories. Here, we give an introduction to some of the key ideas, based on a lattice-gas (Ising) model fluid. This approach builds on knowledge covered in most undergraduate statistical mechanics and thermodynamics courses, so students can quickly get to the stage of calculating density profiles, etc., for themselves. We derive a simple DFT for the lattice gas and present some typical results that can readily be calculated using the theory.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\WD2UQLBH\\Hughes et al. - 2014 - An introduction to inhomogeneous liquids, density .pdf} @@ -1859,7 +1800,6 @@ @article{huhHydrodynamicModelSteady1971 volume = {35}, number = {1}, pages = {85--101}, - issn = {0021-9797}, doi = {10.1016/0021-9797(71)90188-3}, abstract = {Movement of the contact line over a solid surface violates the adherence, or ``no slip,'' boundary condition which is otherwise obeyed by flowing liquids. A flat fluid interface moving steadily over a flat solid is modeled with the creeping flow approximation, which turns out to be self-consistent. Adherence is required except at the contact line itself. Though the velocity field appears to be realistic, stresses and viscous dissipation are found to increase without bound at the contact line. The way the hydrodynamic model breaks down suggests that in reality there may be steep gradients, rheological anomalies, and discontinuous processes around the contact line. Slip and the role of long-range forces are explored with the aid of the lubrication flow approximation.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\MKZFVJPU\\0021979771901883.html} @@ -1883,7 +1823,7 @@ @article{ichimuraLightDrivenMotionLiquids2000 @book{IntroductionQuantumField2018, title = {An {{Introduction To Quantum Field Theory}}}, year = {2018}, -author={Peskin, M.}, + author={Peskin, M.}, month = 5, publisher = {{CRC Press}}, doi = {10.1201/9780429503559}, @@ -1902,7 +1842,6 @@ @article{jabbarzadehWallSlipMolecular1999 number = {5}, pages = {2612--2620}, publisher = {{American Institute of Physics}}, - issn = {0021-9606}, doi = {10.1063/1.477982}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\PNYJHQLZ\\Jabbarzadeh et al. - 1999 - Wall slip in the molecular dynamics simulation of .pdf} } @@ -1952,7 +1891,6 @@ @article{jalaliFabricationCharacterizationScalable2018 number = {1}, pages = {7612}, publisher = {{Nature Publishing Group}}, - issn = {2045-2322}, doi = {10.1038/s41598-018-25812-y}, abstract = {Texturing a large surface with oily micro-drops with controlled size, shape and volume provides an unprecedented capability in investigating complex interactions of bacteria, cells and interfaces. It has particular implications in understanding key microbial processes involved in remediation of environmental disasters, such as Deepwater Horizon oil spill. This work presents a development of scalable micro-transfer molding to functionalize a substrate with oily drop array to generate a microcosm mimicking bacteria encountering a rising droplet cloud. The volume of each drop within a large ``printed'' surface can be tuned by varying base geometry and area with characteristic scales from 5 to 50\,{$\mu$}m. Contrary to macroscopic counterparts, drops with non-Laplacian shapes, i.e. sharp corners, that appears to violate Young-Laplacian relationship locally, are produced. Although the drop relaxes into a spherical cap with constant mean curvature, the contact line with sharp corners remains pinned. Relaxation times from initial to asymptotic shape require extraordinarily long time ({$>$}7 days). We demonstrate that non-Laplacian drops are the direct results of self-pinning of contact line by nanoparticles in the oil. This technique has been applied to study biofilm formation at the oil-water interface and can be readily extended to other colloidal fluids.}, copyright = {2018 The Author(s)}, @@ -1969,7 +1907,6 @@ @article{jambon-puilletSpreadingDynamicsContact2018 volume = {844}, pages = {817--830}, publisher = {{Cambridge University Press}}, - issn = {0022-1120, 1469-7645}, doi = {10.1017/jfm.2018.142}, abstract = {The spreading of evaporating drops without a pinned contact line is studied experimentally and theoretically, measuring the radius R(t)R(t)R(t) of completely wetting alkane drops of different volatility on glass. Initially the drop spreads (RRR increases), then owing to evaporation reverses direction and recedes with an almost constant non-zero contact angle \texttheta{$\propto\beta$}1/3{$\mathsl{\theta}\propto\mathsl{B}$}1/3\textbackslash unicode[STIX]\{x1D703\}\textbackslash propto \textbackslash unicode[STIX]\{x1D6FD\}\^\{1/3\}, where {$\beta\mathsl{B}\backslash$}unicode[STIX]\{x1D6FD\} measures the rate of evaporation; eventually the drop vanishes at a finite-time singularity. Our theory, based on a first-principles hydrodynamic description, well reproduces the dynamics of RRR and the value of \texttheta{$\mathsl{\theta}\backslash$}unicode[STIX]\{x1D703\} during retraction.}, keywords = {condensation/evaporation,contact lines,drops}, @@ -1986,7 +1923,6 @@ @article{jamesNewFrictionModel2019 number = {1}, pages = {269--299}, publisher = {{EDP Sciences}}, - issn = {0764-583X, 1290-3841}, doi = {10.1051/m2an/2018076}, abstract = {The derivation of shallow water models from Navier\textendash Stokes equations is revisited yielding a class of two-layer shallow water models. An improved velocity profile is proposed, based on the superposition of an inviscid fluid and a viscous layer inspired by the Interactive Boundary Layer interaction used in aeronautics. This leads to a new friction law which depends not only on velocity and depth but also on the variations of velocity and thickness of the viscous layer. The resulting system is an extended shallow water model consisting of three depth-integrated equations: the first two are mass and momentum conservation in which a slight correction on hydrostatic pressure has been made; the third one, known as von K\'arm\'an equation, describes the evolution of the viscous layer. This coupled model is shown to be conditionally hyperbolic, and a Godunov-type finite volume scheme is also proposed. Several numerical examples are provided and compared to the Multi-Layer Saint-Venant model. They emphasize the ability of the model to deal with unsteady viscous effects. They illustrate also the phase-lag between friction and topography, and even recover possible reverse flows.}, copyright = {\textcopyright{} The authors. Published by EDP Sciences, SMAI 2019}, @@ -2002,7 +1938,6 @@ @article{janssenDynamicsLiquidliquidMixing1995 volume = {35}, number = {22}, pages = {1766--1780}, - issn = {1548-2634}, doi = {10.1002/pen.760352206}, abstract = {The development of multiphase liquid-liquid morphologies during mixing at small Reynolds numbers has been modeled. The mixing process is divided into (i) stretching of dispersed drops. (ii) breakup of the liquid threads formed, and (iii) coalescence of the final droplets upon collision. Rules and criteria of the distinct processes are presented and combined to a general 2-zone mixing model simplifying the flow field into a sequence of alternating ``strong and weak zones.'' In a ``strong zone,'' dispersed drops and threads are stretched unless their radius is too small; meanwhile, the stretching threads might break up into droplets. In the subsequent ``weak zone,'' the remaining threads may disintegrate while any drops present may coalesce. After passing a number of zones, stretching, breakup, and coalescence lead to a dynamic equilibrium that could be considered as the ``final'' morphology. Using the 2-zone mixing model, the influence of material parameters and processing conditions on the morphology has been studied. Interestingly, increasing either viscosity (dispersed or continuous phase) yields a finer morphology due to the delay of thread breakup, allowing for further stretching and suppression of coalescence.}, annotation = {\_eprint: https://onlinelibrary.wiley.com/doi/pdf/10.1002/pen.760352206}, @@ -2026,7 +1961,6 @@ @article{jasakOpenFOAMOpenSource2009 volume = {1}, number = {2}, pages = {89--94}, - issn = {2092-6782}, doi = {10.2478/IJNAOE-2013-0011}, abstract = {The current focus of development in industrial Computational Fluid Dynamics (CFD) is integration of CFD into Computer-Aided product development, geometrical optimisation, robust design and similar. On the other hand, in CFD research aims to extend the boundaries of practical engineering use in ``non-traditional'' areas. Requirements of computational flexibility and code integration are contradictory: a change of coding paradigm, with object orientation, library components, equation mimicking is proposed as a way forward. This paper describes OpenFOAM, a C++ object oriented library for Computational Continuum Mechanics (CCM) developed by the author. Efficient and flexible implementation of complex physical models is achieved by mimicking the form of partial differential equation in software, with code functionality provided in library form. Open Source deployment and development model allows the user to achieve desired versatility in physical modeling without the sacrifice of complex geometry support and execution efficiency.}, keywords = {C++,CFD,Equation mimicking,Finite volume,Object-Oriented,Open Source}, @@ -2051,7 +1985,6 @@ @article{joEvaluationJetPerformance2009 volume = {26}, number = {2}, pages = {339--348}, - issn = {1975-7220}, doi = {10.1007/s11814-009-0057-2}, abstract = {Inkjet printing has been widely used in many applications and has been studied for many years. However, there are not many systematic researches on the mechanism of jet formation, nor is there any reliable platform that enables us to evaluate jet performance. In this study, an approach to practically evaluate the jet stability of the dropon-demand (DOD) inkjet printing has been proposed, based on which the transient behavior of the DOD drop formation has been studied experimentally for Newtonian liquids with a range of different viscosities (1.0\textendash 11 cp) but of a comparable surface tension. For more viscous liquids, the rate of the jet retraction after a pinch-off from the nozzle was found to increase as the thread motion became more sharp and conical as a result of the shape effect. The break-up time of the jet also increased because the rate of capillary wave propagation was lower for more viscous liquids. The jet stability graph, which can be drawn in terms of jet retraction and break-up time, was employed to characterize the jetting stability, and the degree of satellite drop generation was quantitatively evaluated by two critical jet speeds. The effect of an electric pulse imposed on a piezoelectric plate inside the printhead was also studied. The single-peak electric pulse was used in this experiment for simple analysis, and the jet speed variation was measured under different operating conditions. Both the optimal dwell time and the maximum stable jetting frequency were affected by viscosity and they were explained in terms of the propagation theory.}, keywords = {Drop,Drop-on-demand,Inkjet Printing,Jetting Stability,Jetting Window,Viscosity}, @@ -2059,7 +1992,7 @@ @article{joEvaluationJetPerformance2009 } @article{xuUNSPLITTINGBGKTYPESCHEMES1999, - title = {UNSPLITTING {BGK}-TYPE SCHEMES FOR THE SHALLOW WATER EQUATIONS}, + title = {Unsplitting {BGK}-Type Schemes for the Shallow Water Equations}, author = {Xu, K.}, year = {1999}, month = 4, @@ -2103,7 +2036,6 @@ @article{junkDiscretizationsIncompressibleNavier2000 number = {1}, pages = {1--19}, publisher = {{Society for Industrial and Applied Mathematics}}, - issn = {1064-8275}, doi = {10.1137/S1064827599357188}, abstract = {A discrete velocity model with spatial and velocity discretization based on a lattice Boltzmann method is considered in the low Mach number limit. A uniform numerical scheme for this model is investigated. In the limit, the scheme reduces to a finite difference scheme for the incompressible Navier--Stokes equation, which is a projection method with a second order spatial discretization on a regular grid. The discretization is analyzed and the method is compared to Chorin's original spatial discretization. Numerical results supporting the analytical statements are presented.}, keywords = {35B25,65M06,76D05,76P05,Chorin's projection scheme,discrete velocity models,finite difference method,incompressible Navier--Stokes equations,lattice Boltzmann method,low Mach number limit,relaxation systems}, @@ -2120,7 +2052,6 @@ @article{karguptaDewettingThinFilms2002 number = {5}, pages = {1893--1903}, publisher = {{American Chemical Society}}, - issn = {0743-7463}, doi = {10.1021/la010469n}, abstract = {The instability, dynamics, and morphological transitions of patterns in thin liquid films on physically and chemically heterogeneous patterned surfaces are investigated on the basis of 3D nonlinear simulations. On a chemically striped surface (consisting of alternating less and more wettable stripes) the film breakup is suppressed on some potentially destabilizing nonwettable stripes when their spacing is below a characteristic length scale of instability ({$\lambda$}h), which is of the same order as the spinodal length scale ({$\lambda$}l) of instability on the less wettable stripes. The thin film pattern replicates the substrate surface energy pattern closely only when (a) the periodicity of the substrate pattern lies between {$\lambda$}h and 2{$\lambda$}h and (b) the less wettable stripe width is within a range bounded by a lower critical length, below which no heterogeneous rupture occurs, and an upper transition length, above which complex morphological features bearing little resemblance to the substrate pattern are formed. The thin film pattern on a periodic physically heterogeneous surface shows the loss of ideal templating when the periodicity of the surface is smaller than about 0.8 times the spinodal wavelength evaluated at the minimum film thickness. On a physicochemically patterned periodic surface, the chemical heterogeneity largely controls the thin film pattern and the effect of small to moderate physical heterogeneity is minimal.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\4XZILRLK\\Kargupta and Sharma - 2002 - Dewetting of Thin Films on Periodic Physically and.pdf;C\:\\Users\\Zitzero\\Zotero\\storage\\IK55AQWQ\\la010469n.html} @@ -2136,7 +2067,6 @@ @article{karguptaInstabilityPatternFormation2000 number = {26}, pages = {10243--10253}, publisher = {{American Chemical Society}}, - issn = {0743-7463}, doi = {10.1021/la000759o}, abstract = {The surface instability, dynamics, morphology, and spontaneous dewetting of a thin liquid film on chemically heterogeneous substrates are studied on the basis of 3D nonlinear simulations. A new mechanism of dewetting in the presence of heterogeneity is proposed where the instability is engendered by the gradient of intermolecular interactions that lead to a microscale wettability contrast. The time scale of instability, which can be several orders smaller than the spinodal dewetting time scale on homogeneous surfaces, varies inversely with the potential difference induced by the heterogeneity. Heterogeneity can even destabilize spinodally stable films, reduce the time of rupture substantially for thicker films, and decrease the dependence of rupture time on the film thickness. The presence of heterogeneity produces complex and locally ordered morphological features that are not predicted by the spinodal dewetting, for example ``ripples'' and ``castle-moat'' structures, radially symmetric structures, and a lack of undulations before the birth of a hole. The precise morphological pattern selection depends on the size of the heterogeneity, the potential difference caused by the heterogeneity, the film thickness, and also the spinodal characteristics of the substrate. The resulting morphologies can be understood on the basis of simple arguments that consider interplay among these factors.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\ZT6QCVDN\\Kargupta et al. - 2000 - Instability and Pattern Formation in Thin Liquid F.pdf;C\:\\Users\\Zitzero\\Zotero\\storage\\BGWLZ8MI\\la000759o.html} @@ -2152,7 +2082,6 @@ @article{karguptaMesopatterningThinLiquid2003 number = {12}, pages = {5153--5163}, publisher = {{American Chemical Society}}, - issn = {0743-7463}, doi = {10.1021/la026634w}, abstract = {Surface directed instability and dewetting in thin films and resulting morphologies are studied using 3D nonlinear simulations based on the equations of motion, both for the isotropic and anisotropic 2D substrate patterns. Three different substrate (wettability) patterns are considered:\, (a) arrays of more (or completely) wettable rectangular blocks on a less wettable substrate, (b) arrays of less wettable blocks on a more (or completely) wettable substrate, and (c) a checkerboard pattern of alternating more and less wettable blocks. An ideal replication of the surface energy pattern produces an ordered 2D array of liquid columns (in case 1), or a matrix of holes on a flat liquid sheet (in case 2), or a checkerboard pattern of alternating liquid columns and holes/depressions (in case 3). The effects of pattern periodicity, domain widths, anisotropy, and wettability on the morphological phase transitions are presented. Regardless of the precise geometry of the substrate pattern, templating is found to be better on a completely wettable substrate containing the less wettable blocks, rather than on a less wettable substrate having more wettable blocks. Complete wettability (zero contact angle at all thicknesses), of either the blocks or their surroundings, ensures the pinning of the liquid contact line at the block boundaries. Thus, complete wettability leads to better templating compared to partially wettable substrates. Ideal templating is found possible only when the following conditions are met:\, (a) the periodicity (Lpx and Lpy) of the pattern is more than {$\lambda$}h; where {$\lambda$}h is a characteristic length scale found to be close to the spinodal length scale of the less wettable part, {$\lambda$}m, (b) the less wettable area fraction, Af, should be less than a transition value beyond which the liquid spills over the less wettable part leading to a morphological transition from discrete columnar structure to continuous liquid ridges, (c) less wettable block/channel width should be less than a transition length scale ({$\sim$}0.5{$\lambda$}m), and (d) the aspect ratio of the periodicity intervals (Lpx/Lpy) should be close to 1. Anisotropy in the substrate periodicity leads to stripe-like liquid patterns whenever either Lpx or Lpy is less than {$\lambda$}h. Large values of periodicity lead to the formation of novel ``block mountain-rift valley'' or ``flower'' like microstructures that do not replicate the substrate energy pattern. Interestingly, the near ideal templating of more complex substrate patterns, e.g., alphabets, is also guided by the above conditions.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\TNFPHUKK\\Kargupta and Sharma - 2003 - Mesopatterning of Thin Liquid Films by Templating .pdf;C\:\\Users\\Zitzero\\Zotero\\storage\\QN4BWXVK\\la026634w.html} @@ -2168,7 +2097,6 @@ @article{karguptaMorphologicalSelforganizationDewetting2002 number = {7}, pages = {3042--3051}, publisher = {{American Institute of Physics}}, - issn = {0021-9606}, doi = {10.1063/1.1434949}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\Z3ZCAVCN\\Kargupta and Sharma - 2002 - Morphological self-organization by dewetting in th.pdf} } @@ -2199,7 +2127,6 @@ @article{karpitschkaCoalescenceNoncoalescenceSessile2014 number = {23}, pages = {6826--6830}, publisher = {{American Chemical Society}}, - issn = {0743-7463}, doi = {10.1021/la500459v}, abstract = {Due to capillarity, sessile droplets of identical liquids will instantaneously fuse when they come in contact at their three-phase lines. However, with drops of different, completely miscible liquids, instantaneous coalescence can be suppressed. Instead, the drops remain in a state of noncoalescence for some time, with the two drop bodies connected only by a thin neck. The reason for this noncoalescence is the surface tension difference, {$\Delta\gamma$}, between the liquids. If {$\Delta\gamma$} is sufficiently large, then it induces a sufficiently strong Marangoni flow, which keeps the main drop bodies temporarily separated. Studies with spreading drops have revealed that the boundary between instantaneous coalescence and noncoalescence is sharp (Karpitschka, S.; Riegler, H. J. Fluid. Mech. 2014, 743, R1). The boundary is a function of two parameters only: {$\Delta\gamma$} and {$\overline{\Theta}$}a, the arithmetic mean of the contact angles in the moment of drop\textendash drop contact. It appears plausible that surface forces (the disjoining pressure) could also influence the coalescence behavior. However, in experiments with spreading drops, surface forces always promote coalescence and their influence might be obscured. Therefore, we present here coalescence experiments with partially wetting liquids and compare the results to the spreading case. We adjust different equilibrium contact angles (i.e., different surface forces) with different substrate surface coatings. As for spreading drops, we observe a sharp boundary between regimes of coalescence and noncoalescence. The boundary follows the same power law relation for both partially and completely wetting cases. Therefore, we conclude that surface forces have no significant, explicit influence on the coalescence behavior of sessile drops from different miscible liquids.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\YVGUF26F\\Karpitschka et al. - 2014 - Coalescence and Noncoalescence of Sessile Drops I.pdf;C\:\\Users\\Zitzero\\Zotero\\storage\\FTSSG5AA\\la500459v.html} @@ -2230,7 +2157,6 @@ @article{karpitschkaSharpTransitionCoalescence2014 volume = {743}, pages = {R1}, publisher = {{Cambridge University Press}}, - issn = {0022-1120, 1469-7645}, doi = {10.1017/jfm.2014.73}, abstract = {Unexpectedly, under certain conditions, sessile drops from different but completely miscible liquids do not always coalesce instantaneously upon contact: the drop bodies remain separated in a temporary state of non-coalescence, connected through a thin liquid bridge. Here we investigate the transition between the states of instantaneous coalescence and temporary non-coalescence. Experiments reveal that it is barely influenced by viscosities and absolute surface tensions. The main system control parameters for the transition are the arithmetic means of the three-phase angles, {$\Theta$}\textasciimacron\textasciimacron\textasciimacron\textasciimacron a{$\Theta$}\textasciimacron a\textbackslash overline\{\textbackslash Theta \}\_\{a\}, and the surface tension differences {$\Delta\gamma\Delta\gamma\backslash$}Delta \textbackslash gamma between the two liquids. These relevant parameters can be combined into a single system parameter, a specific Marangoni number M\texttildelow =3{$\Delta\gamma$}/(2{$\gamma$}\textasciimacron\textasciimacron\textasciimacron{$\Theta$}\textasciimacron\textasciimacron\textasciimacron\textasciimacron 2a)M\textasciitilde =3{$\Delta\gamma$}/(2{$\gamma$}\textasciimacron{$\Theta$}\textasciimacron a2)\textbackslash widetilde\{M\}=3\textbackslash Delta \textbackslash gamma /(2\textbackslash overline\{\textbackslash gamma \}\textbackslash overline\{\textbackslash Theta \}\_\{a\}\^2). This M\texttildelow M\textasciitilde\textbackslash widetilde\{M\} universally characterizes the coalescence transition behaviour as a function of both the physicochemical liquid properties and the shape of the liquid body in the contact region. The transition occurs at a certain threshold value M\texttildelow tM\textasciitilde t\textbackslash widetilde\{M\}\_t and is sharp within the experimental resolution. The experimentally observed threshold value of M\texttildelow t{$\approx$}2M\textasciitilde t{$\approx$}2\textbackslash widetilde\{M\}\_t\textbackslash approx 2 agrees quantitatively with values obtained by simulations assuming realistic material parameters. The simulations indicate that the absolute value of M\texttildelow tM\textasciitilde t\textbackslash widetilde\{M\}\_t very weakly depends on the molecular diffusivity.}, keywords = {breakup/coalescence,convection,drops and bubbles,interfacial flows (free surface),marangoni convection,thin films}, @@ -2246,7 +2172,6 @@ @article{kellerBreakingMergingLiquid2002 volume = {42}, number = {3}, pages = {283--290}, - issn = {1573-2703}, doi = {10.1023/A:1016127027751}, abstract = {The surface of a sheet of liquid which contracts due to surface tension, breaks, and then pulls apart into two pieces, is calculated. Before breaking, the flow is the self-similar one found by Keller, Milewski and Vanden-Broeck. After breaking, it is the self-similar flow found by Keller and Miksis. A general numerical scheme, which includes the previous ones, is presented and new numerical results are discussed. There is an analogous flow of an axially symmetric liquid filament, but it is not calculated.}, keywords = {boundary-integral equations,free-surface flows,self-similar solutions,surface tension.}, @@ -2263,7 +2188,6 @@ @article{kimHighlyConductiveInk2005 number = {11}, pages = {J30}, publisher = {{IOP Publishing}}, - issn = {1944-8775}, doi = {10.1149/1.2073670}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\3MBAHHTA\\Kim and Moon - 2005 - Highly Conductive Ink Jet Printed Films of Nanosil.pdf} } @@ -2277,7 +2201,6 @@ @article{kimSlidingLiquidDrops2002 volume = {247}, number = {2}, pages = {372--380}, - issn = {0021-9797}, doi = {10.1006/jcis.2001.8156}, abstract = {A liquid drop that partially wets a solid surface will slide down the plane when it is tilted beyond a critical inclination. Here we report the study of the sliding velocity of such a drop. Experiments for measuring the steady sliding velocity of different liquids of drops are performed. We then construct a scaling law that predicts the sliding velocity given the physical properties, wetting characteristics, and size of the drop. When the sliding velocity is low and the drop distortion due to inclination is small, the scaling law is shown to correctly model the functional dependency of the measured sliding velocity.}, keywords = {capillary,liquid drop,scaling law,sliding,wetting}, @@ -2295,7 +2218,6 @@ @article{kimSprayCoolingHeat2007 volume = {28}, number = {4}, pages = {753--767}, - issn = {0142-727X}, doi = {10.1016/j.ijheatfluidflow.2006.09.003}, abstract = {Spray cooling is a technology of increasing interest for electronic cooling and other high heat flux applications, and is characterized by high heat transfer, uniformity of heat removal, small fluid inventory, low droplet impact velocity, and no temperature overshoot. The mechanisms by which heat is removed during spray cooling are poorly understood, however, due to its dependence on many parameters that are not easily varied independently, and predictive capabilities are quite limited. This paper provides an introduction to spray cooling for electronic cooling applications, reviews some proposed spray cooling heat transfer mechanisms, and summarizes the data regarding the effects of non-condensable gas, surface enhancement, spray inclination, and gravity. Some models of spray cooling are also presented.}, keywords = {Electronic cooling,Heat transfer,Spray cooling}, @@ -2311,7 +2233,6 @@ @article{kingdonReactionflowLatticeBoltzmann1992 volume = {25}, number = {14}, pages = {L907}, - issn = {0305-4470}, doi = {10.1088/0305-4470/25/14/008}, abstract = {The authors describe a development of a lattice Boltzmann fluid flow model enabling the simulation of the reaction and flow of low-concentration species in two dimensions. In principle any number of species and reactions can be implemented in this model. In making this development they derive an expression for the solution of the diffusion equation using the lattice Boltzmann approach.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\AWFS92FT\\Kingdon and Schofield - 1992 - A reaction-flow lattice Boltzmann model.pdf} @@ -2327,7 +2248,6 @@ @article{kloppDropletCoalescenceQuasiTwoDimensional2020 number = {35}, pages = {10615--10621}, publisher = {{American Chemical Society}}, - issn = {0743-7463}, doi = {10.1021/acs.langmuir.0c02139}, abstract = {Coalescence of droplets plays a crucial role in nature and modern technology. Various experimental and theoretical studies explored droplet dynamics in three-dimensional (3D) and on 2D solid or liquid substrates. In this paper, we demonstrate the complete coalescence of isotropic droplets in thin quasi-2D liquids\textemdash overheated smectic films. We observe the merging of micrometer-sized flat droplets using high-speed imaging and analyze the shape transformations of the droplets on the timescale of milliseconds. Our studies reveal the scaling laws of the coalescence time, which exhibits a different dependence on the droplet geometry from that in the case of droplets on a solid substrate. A theoretical model is proposed to explain the difference in behavior.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\TBGT2RWQ\\Klopp and Eremin - 2020 - On Droplet Coalescence in Quasi-Two-Dimensional Fl.pdf;C\:\\Users\\Zitzero\\Zotero\\storage\\DZKGB3VC\\acs.langmuir.html} @@ -2383,7 +2303,6 @@ @article{kullTheoryRayleighTaylorInstability1991 volume = {206}, number = {5}, pages = {197--325}, - issn = {0370-1573}, doi = {10.1016/0370-1573(91)90153-D}, abstract = {The theory of the Rayleigh-Taylor instability of accelerated fluid layers is systematically developed from basic fluid equations. Starting with the classical potential flow theory for moving contact surfaces, the discussion extends to various fluid systems describing inhomogeneous, viscous, compressible, and isobaric flows. Thereby an overview on the major stability issues under a broad variety of physical conditions can be given. In particular, the stability analysis is addressed to layered materials in plane and spherical geometries under various dynamical conditions, to inhomogeneous media with variable gradients and different boundary conditions, to viscous boundary layers, compressible atmospheres, and to stationary ablation fronts in laser-driven plasma experiments. The stability theory is further extended to the nonlinear stage of the Rayleigh-Taylor instability and to a discussion of bubble dynamics in two and three dimensions for closed and open bubble domains. For this purpose simple flow models are studied that can describe essential features of bubble rise and bubble growth in buoyancy-driven mixing layers.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\IKEQF6PH\\037015739190153D.html} @@ -2427,7 +2346,6 @@ @article{laugesenLinearStabilitySteady2000 volume = {154}, number = {1}, pages = {3--51}, - issn = {1432-0673}, doi = {10.1007/PL00004234}, abstract = {We study the linear stability of smooth steady states of the evolution equation\$h\_t = - (f(h)\_\{xxx\} )\_x - (g(h)h\_x )\_x - ah\$under both periodic and Neumann boundary conditions. If a{$\neq$} 0 we assume f{$\equiv$} 1. In particular we consider positive periodic steady states of thin film equations, where a=0 and f, g might have degeneracies such as f(0)=0 as well as singularities like g(0)=+{$\infty$}.}, keywords = {Linear Instability,Linear Stability,Natural Boundary Condition,Neumann Boundary Condition,Trial Function}, @@ -2456,7 +2374,6 @@ @article{lenormandLiquidsPorousMedia1990 volume = {2}, number = {S}, pages = {SA79}, - issn = {0953-8984}, doi = {10.1088/0953-8984/2/S/008}, abstract = {The basic mechanisms which take place during the displacement of immiscible fluids in porous media have been observed in micromodels and have been modelled. At the pore level, in drainage, the invading fluid chooses the largest throat. In imbibition, the displacement depends on the local geometry. For a large pore-to-throat ratio (aspect ratio), the main mechanism is the collapse of the invading fluid in the smallest channel, without entering the pore. For a small aspect ratio, the wetting fluid invades the pore first, and then the adjacent channels. From observations at the pore level, the author has modelled the displacement on a large scale in some extreme cases by using statistical theories. The different behaviours are then displayed as domains in three phase diagrams: one for drainage and two for imbibition (large and small aspect ratios). At a high rate, when viscous forces are dominant, all the diagrams show a stable domain (described by anti-DLA) and a viscous fingering domain (DLA). In drainage, low capillary numbers lead to capillary fingering represented by invasion percolation. In imbibition, the capillary domain is described either by a compact cluster growth (small aspect ratio) or percolation theory (large aspect ratio). In addition the possibility of flow by film along the roughness of the walls leads to disconnected structures.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\SASNCRE6\\Lenormand - 1990 - Liquids in porous media.pdf} @@ -2472,7 +2389,6 @@ @article{liAdaptationStyreneAcrylic2021 number = {4}, pages = {1571--1577}, publisher = {{American Chemical Society}}, - issn = {0743-7463}, doi = {10.1021/acs.langmuir.0c03226}, abstract = {Solid surfaces, in particular polymer surfaces, are able to adapt upon contact with a liquid. Adaptation results in an increase in contact angle hysteresis and influences the mobility of sliding drops on surfaces. To study adaptation and its kinetics, we synthesized a random copolymer composed of styrene and 11\textendash 25 mol\% acrylic acid (PS/PAA). We measured the dynamic advancing (\texttheta A) and receding (\texttheta R) contact angles of water drops sliding down a tilted plate coated with this polymer. We measured \texttheta A {$\approx$} 87\textdegree{} for velocities of the contact line {$<$}20 {$\mu$}m/s. At higher velocities, \texttheta A gradually increased to {$\sim$}98\textdegree. This value is similar to \texttheta A of a pure polystyrene (PS) film, which we studied for comparison. We associate the gradual increase in \texttheta A to the adaptation process to water: The presence of water leads to swelling and/or an enrichment of acid groups at the water/polymer interface. By applying the latest adaptation theory (Butt et al. Langmuir 2018, 34, 11292), we estimated the time constant of this adaptation process to be {$\ll$}1 s. For sliding water drops, \texttheta R is {$\sim$}10\textdegree{} lower compared to the reference PS surface for all tested velocities. Thus, at the receding side of a sliding drop, the surface is already enriched by acid groups. For a water drop with a width of 5 mm, the increase in contact angle hysteresis corresponds to an increase in capillary force in the range of 45\textendash 60 {$\mu$}N, depending on sliding velocity.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\LZ6HYQCW\\Li et al. - 2021 - Adaptation of a Styrene–Acrylic Acid Copolymer Sur.pdf;C\:\\Users\\Zitzero\\Zotero\\storage\\YTDT7NIX\\acs.langmuir.html} @@ -2486,7 +2402,6 @@ @article{liEvaluationForceTerms2019 volume = {90}, number = {7}, pages = {357--373}, - issn = {1097-0363}, doi = {10.1002/fld.4726}, abstract = {Proper approximation of the force terms, especially the bed slope term, is of crucial importance to simulating shallow water flows in lattice Boltzmann (LB) models. However, there is little discussion on the schemes of adding force terms to LB models for shallow water equations (SWEs). In this study, we evaluate the performance of forcing schemes coupled with different LB models (LABSWE and MLBSWE) in simulating shallow water flows over complex topography and try to find out their intrinsic characteristics and applicability. Three cases are adopted for evaluation, including a stationary case, a one-dimensional tidal wave flow over an irregular bed, and a steady flow over a two-dimensional seamount. The simulating results are compared with analytical solutions or the results produced by the finite difference method. For LABSWE, all the forcing schemes, except for the weighting factor method, fail to produce accurate solutions for the test cases; this is probably due to the mismatch between the bed slope term in source terms and the quadratic depth term of the equilibrium distribution functions in these forcing schemes. For MLBSWE, all the forcing schemes are capable of simulating flows over the complex topography accurately; furthermore, those schemes taking into account the collision effect {$\tau$} to eliminate the momentum induced by forces provide more accurate solutions with quicker convergence as the lattice size decreases. In this view, MLBSWE can bring more flexibility in treating the force terms and thus can be a better tool to simulate shallow water flows over complex topography in practical application.}, keywords = {bed slope,complex topography,forcing scheme,lattice Boltzmann,shallow water flows,source terms}, @@ -2504,7 +2419,6 @@ @article{lindahlGROMACSPackageMolecular2001 volume = {7}, number = {8}, pages = {306--317}, - issn = {0948-5023}, doi = {10.1007/s008940100045}, abstract = {GROMACS 3.0 is the latest release of a versatile and very well optimized package for molecular simulation. Much effort has been devoted to achieving extremely high performance on both workstations and parallel computers. The design includes an extraction of virial and periodic boundary conditions from the loops over pairwise interactions, and special software routines to enable rapid calculation of x\textendash 1/2. Inner loops are generated automatically in C or Fortran at compile time, with optimizations adapted to each architecture. Assembly loops using SSE and 3DNow! Multimedia instructions are provided for x86 processors, resulting in exceptional performance on inexpensive PC workstations. The interface is simple and easy to use (no scripting language), based on standard command line arguments with self-explanatory functionality and integrated documentation. All binary files are independent of hardware endian and can be read by versions of GROMACS compiled using different floating-point precision. A large collection of flexible tools for trajectory analysis is included, with output in the form of finished Xmgr/Grace graphs. A basic trajectory viewer is included, and several external visualization tools can read the GROMACS trajectory format. Starting with version 3.0, GROMACS is available under the GNU General Public License from http://www.gromacs.org.}, keywords = {Algorithmic optimization,Assembly loops,Benchmark,Keywords. Parallel molecular dynamics,Simulation}, @@ -2536,7 +2450,6 @@ @article{wangMolecularDynamicsSimulations2015 number = {27}, pages = {7457--7462}, publisher = {{American Chemical Society}}, - issn = {0743-7463}, doi = {10.1021/acs.langmuir.5b01574}, urldate = {2023-07-27}, abstract = {When an electric field with various strengths is applied to two adjacent conducting droplets, the droplets may completely coalesce, partially coalesce, or bounce off one another. To reveal an atom-scale mechanism of coalescence or non-coalescence, dynamic behaviors of two conducting nanodroplets at a homogeneous electric field are studied via molecular dynamics simulations in this work. The results show that there is a critical field strength and a critical cone angle above which the two droplets partially coalesce or bounce off. Charge transfer between the two droplets is observed when the droplets are brought into contact. The partial coalescence and the bounce-off of the two droplets at strong field strengths are found to be due to the high charge transfer rate, which leads to the breakup of the coalescing droplet at different locations.} @@ -2552,7 +2465,6 @@ @article{liuActuatingWaterDroplets2015 number = {33}, pages = {9070--9075}, publisher = {{American Chemical Society}}, - issn = {0743-7463}, doi = {10.1021/acs.langmuir.5b02335}, abstract = {A surface wettability gradient can break the equilibrium status of a liquid droplet and drives its unidirectional movement on the surface. We propose a conceptual design of the driving water droplet on a graphene surface and demonstrate that both speed and direction of the movement can be controlled via a continuous gradient of surface wettability using comprehensive molecular dynamics (MD) simulations. Controlling the water droplet toward linear and nonlinear arc paths is exemplified in one- and two-dimensional gradients of surface wettability, respectively. Unbalanced Young's equation is extended to understand the speed of the droplet movement, and the predications agree well with MD simulations.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\Z843NVRA\\Liu and Xu - 2015 - Actuating Water Droplets on Graphene via Surface W.pdf;C\:\\Users\\Zitzero\\Zotero\\storage\\TSFNRCRW\\acs.langmuir.html} @@ -2566,7 +2478,6 @@ @article{liuControlledSwitchableSurface2005 volume = {11}, number = {9}, pages = {2622--2631}, - issn = {1521-3765}, doi = {10.1002/chem.200400931}, abstract = {The macroscopic properties of a surface can be intelligently controlled by alternating the states of the modified molecules, such as polymers, metallic oxide, or self-assembled monolayers (SAMs). This article reviews various approaches to create a switchable surface and different types of external stimuli used to switch the surface properties. This area is of potential benefit for biomaterials, biosensors, information storage, microfluidic systems, adhesive materials, nanolithography, and so on.}, keywords = {monolayers,surface chemistry,wettability}, @@ -2596,7 +2507,6 @@ @article{liuMultiphaseLatticeBoltzmann2016 volume = {20}, number = {4}, pages = {777--805}, - issn = {1573-1499}, doi = {10.1007/s10596-015-9542-3}, abstract = {Over the last two decades, lattice Boltzmann methods have become an increasingly popular tool to compute the flow in complex geometries such as porous media. In addition to single phase simulations allowing, for example, a precise quantification of the permeability of a porous sample, a number of extensions to the lattice Boltzmann method are available which allow to study multiphase and multicomponent flows on a pore scale level. In this article, we give an extensive overview on a number of these diffuse interface models and discuss their advantages and disadvantages. Furthermore, we shortly report on multiphase flows containing solid particles, as well as implementation details and optimization issues.}, keywords = {47.11.-j,47.56.+r,91.60.Np,Lattice Boltzmann method,Pore scale simulation,Porous media}, @@ -2612,7 +2522,6 @@ @article{carenzaLatticeBoltzmannMethods2019 volume = {42}, number = {6}, pages = {81}, - issn = {1292-895X}, doi = {10.1140/epje/i2019-11843-6}, urldate = {2023-08-10}, abstract = {We review the state of the art of active fluids with particular attention to hydrodynamic continuous models and to the use of Lattice Boltzmann Methods (LBM) in this field. We present the thermodynamics of active fluids, in terms of liquid crystals modelling adapted to describe large-scale organization of active systems, as well as other effective phenomenological models. We discuss how LBM can be implemented to solve the hydrodynamics of active matter, starting from the case of a simple fluid, for which we explicitly recover the continuous equations by means of Chapman-Enskog expansion. Going beyond this simple case, we summarize how LBM can be used to treat complex and active fluids. We then review recent developments concerning some relevant topics in active matter that have been studied by means of LBM: spontaneous flow, self-propelled droplets, active emulsions, rheology, active turbulence, and active colloids.}, @@ -2629,7 +2538,6 @@ @article{luechingerGraphenestabilizedCopperNanoparticles2008 volume = {19}, number = {44}, pages = {445201}, - issn = {0957-4484}, doi = {10.1088/0957-4484/19/44/445201}, abstract = {Metallic copper nanoparticles were synthesized by a bottom-up approach, and in situ coated with protective shells of graphene in order to get a metal nanopowder of high air stability and chemical inertness. Using an amphiphilic surfactant, a water-based copper nanocolloid could be prepared and successfully printed onto a polymer substrate by conventional ink-jet printing using household printers. The dried printed patterns exhibited strong metallic gloss and an electrical conductivity of {$>$}1 S cm-1 without the need for a sintering or densification step. This conductivity currently limits use in electronics to low current application or shielding and decorative effects. The high stability of graphene-coated copper nanoparticles makes them economically a most attractive alternative to silver or gold nanocolloids, and will strongly facilitate the industrial use of metal nanocolloids in consumer goods.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\N4ADESE9\\Luechinger et al. - 2008 - Graphene-stabilized copper nanoparticles as an air.pdf} @@ -2646,7 +2554,6 @@ @article{mahadyComparisonNavierStokesSimulations2013 number = {11}, pages = {112103}, publisher = {{American Institute of Physics}}, - issn = {1070-6631}, doi = {10.1063/1.4828721}, abstract = {The classical long-wave theory (also known as lubrication approximation) applied to fluid spreading or retracting on a solid substrate is derived under a set of assumptions, typically including small slopes and negligible inertial effects. In this work, we compare the results obtained by using the long-wave model and by simulating directly the full two-phase Navier-Stokes equations employing a volume-of-fluid method. In order to isolate the influence of the small slope assumption inherent in the long-wave theory, we present a quantitative comparison between the two methods in the regime where inertial effects and the influence of gas phase are negligible. The flow geometries that we consider include wetting and dewetting drops within a broad range of equilibrium contact angles in planar and axisymmetric geometries, as well as liquid rings. For perfectly wetting spreading drops we find good quantitative agreement between the models, with both of them following rather closely Tanner's law. For partially wetting drops, while in general we find good agreement between the two models for small equilibrium contact angles, we also uncover differences which are particularly evident in the initial stages of evolution, for retracting drops, and when additional azimuthal curvature is considered. The contracting rings are also found to evolve differently for the two models, with the main difference being that the evolution occurs on the faster time scale when the long-wave model is considered, although the ring shapes are very similar between the two models.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\XE2R6SYW\\Mahady et al. - 2013 - Comparison of Navier-Stokes simulations with long-.pdf} @@ -2660,7 +2567,6 @@ @article{mahadyVolumeFluidMethod2015 journal = {Journal of Computational Physics}, volume = {294}, pages = {243--257}, - issn = {0021-9991}, doi = {10.1016/j.jcp.2015.03.051}, abstract = {In this paper, we present a novel approach to model the fluid/solid interaction forces in a direct solver of the Navier\textendash Stokes equations based on the volume of fluid interface tracking method. The key ingredient of the model is the explicit inclusion of the fluid/solid interaction forces into the governing equations. We show that the interaction forces lead to a partial wetting condition and in particular to a natural definition of the equilibrium contact angle. We present two numerical methods to discretize the interaction forces that enter the model; these two approaches differ in complexity and convergence. To validate the computational framework, we consider the application of these models to simulate two-dimensional drops at equilibrium, as well as drop spreading. We demonstrate that the model, by including the underlying physics, captures contact line dynamics for arbitrary contact angles. More generally, the approach permits novel means to study contact lines, as well as a diverse range of phenomena that previously could not be addressed in direct simulations.}, keywords = {Contact angle,Solid–liquid intermolecular,Thin films,Volume of fluid method}, @@ -2676,7 +2582,6 @@ @article{marcheDerivationNewTwodimensional2007 volume = {26}, number = {1}, pages = {49--63}, - issn = {0997-7546}, doi = {10.1016/j.euromechflu.2006.04.007}, abstract = {R\'esum\'e Considering the three-dimensional Navier\textendash Stokes equations with a free moving surface boundary condition and hydrostatic approximation, we study the derivation, with asymptotic analysis, of a new two-dimensional viscous shallow water model in rotating framework, with irregular topography, linear and quadratic bottom friction terms and capillary effects. A new formulation of the viscous effects, consistent with a previous one-dimensional analysis, is obtained. Finally, we propose some simple numerical experiments in order to validate the proposed model.}, keywords = {Asymptotic analysis,Capillary effects,Friction,Shallow water equations,Viscosity}, @@ -2719,7 +2624,6 @@ @article{meckeThermalFluctuationsThin2005 volume = {17}, number = {45}, pages = {S3515}, - issn = {0953-8984}, doi = {10.1088/0953-8984/17/45/042}, abstract = {In bulk fluids hydrodynamic Navier\textendash Stokes equations are proven to be valid down to the nanometre scale. However, during the dewetting process of thin liquid films of nanometre thickness the interplay of surface tension {$\gamma$}, substrate potential and thermal noise can lead to qualitatively different behaviour on laterally much larger scales up to microns. By deriving a stochastic thin film equation with a conserved noise term we show that the spectrum of capillary waves changes from an exponential decay to a power law kBT/({$\gamma$}q2) for large wavevectors q due to thermal fluctuations at temperature T. Also the time evolution of film roughness {$\sigma$}(t) and of the typical wavevector k(t) of unstable perturbations changes qualitatively. Whereas a deterministic Navier\textendash Stokes equation in the lubrication approximation predicts in the linear regime a constant k(t) = k0, one finds a coarsening due to thermal noise.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\WKP8LGA5\\Mecke and Rauscher - 2005 - On thermal fluctuations in thin film flow.pdf} @@ -2751,7 +2655,6 @@ @article{mickelShortcomingsBondOrientational2013 number = {4}, pages = {044501}, publisher = {{American Institute of Physics}}, - issn = {0021-9606}, doi = {10.1063/1.4774084}, abstract = {Local structure characterization with the bond-orientational order parameters q4, q6, \ldots{} introduced by Steinhardt et al. [Phys. Rev. B 28, 784 (1983) https://doi.org/10.1103/PhysRevB.28.784.] has become a standard tool in condensed matter physics, with applications including glass, jamming, melting or crystallization transitions, and cluster formation. Here, we discuss two fundamental flaws in the definition of these parameters that significantly affect their interpretation for studies of disordered systems, and offer a remedy. First, the definition of the bond-orientational order parameters considers the geometrical arrangement of a set of nearest neighboring (NN) spheres, NN(p), around a given central particle p; we show that the choice of neighborhood definition can have a bigger influence on both the numerical values and qualitative trend of ql than a change of the physical parameters, such as packing fraction. Second, the discrete nature of neighborhood implies that NN(p) is not a continuous function of the particle coordinates; this discontinuity, inherited by ql, leads to a lack of robustness of the ql as structure metrics. Both issues can be avoided by a morphometric approach leading to the robust Minkowski structure metrics q {${'}$} l . These q {${'}$} l are of a similar mathematical form as the conventional bond-orientational order parameters and are mathematically equivalent to the recently introduced Minkowski tensors [G. E. Schr\"oder-Turk et al., Europhys. Lett. 90, 34001 (2010) https://doi.org/10.1209/0295-5075/90/34001.; S. Kapfer et al., Phys. Rev. E 85, 030301\textendash R (2012) https://doi.org/10.1103/PhysRevE.85.030301.].}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\H9A2ICL4\\Mickel et al. - 2013 - Shortcomings of the bond orientational order param.pdf} @@ -2779,7 +2682,6 @@ @article{mitlinDewettingSolidSurface1993 volume = {156}, number = {2}, pages = {491--497}, - issn = {0021-9797}, doi = {10.1006/jcis.1993.1142}, abstract = {The formal analogy between the theory of dewetting of a solid surface covered by a fluid film and the theory of spinodal decomposition is established. It is shown that the later stage of dewetting (nucleation and growth of dry regions) can be described within the framework of the existing mathematical model of film rupture, if one uses the accurate derivation of the disjoining pressure in the range of low film thickness ({$<$}1 nm).}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\M2F6SM6X\\Mitlin - 1993 - Dewetting of Solid Surface Analogy with Spinodal .pdf;C\:\\Users\\Zitzero\\Zotero\\storage\\4UMZLAQ6\\S0021979783711422.html} @@ -2806,7 +2708,6 @@ @article{moultonEffectDisjoiningPressure2013 number = {6}, pages = {887--920}, publisher = {{Cambridge University Press}}, - issn = {0956-7925, 1469-4425}, doi = {10.1017/S0956792513000235}, abstract = {We explore the effect of disjoining pressure on a thin film equation in the presence of a non-uniform body force, motivated by a model describing the reverse draining of a magnetic film. To this end, we use a combination of numerical investigations and analytical considerations. The disjoining pressure has a regularizing influence on the evolution of the system and appears to select a single steady-state solution for fixed height boundary conditions; this is in contrast with the existence of a continuum of locally attracting solutions that exist in the absence of disjoining pressure for the same boundary conditions. We numerically implement matched asymptotic expansions to construct equilibrium solutions and also investigate how they behave as the disjoining pressure is sent to zero. Finally, we consider the effect of the competition between forcing and disjoining pressure on the coarsening dynamics of the thin film for fixed contact angle boundary conditions.}, keywords = {coarsening,evolution equation,lubrication approximation,magnetic film,matched asymptotic expansion}, @@ -2823,7 +2724,6 @@ @article{mugeleElectrowettingConvenientWay2005 volume = {17}, number = {9}, pages = {S559}, - issn = {0953-8984}, doi = {10.1088/0953-8984/17/9/016}, abstract = {Electrowetting is a versatile tool to reduce the apparent contact angle of partially wetting conductive liquids by several tens of degrees via an externally applied voltage. We studied various fundamental and applied aspects of equilibrium liquid surface morphologies both theoretically and experimentally. Our theoretical analysis showed that surface profiles on homogeneous surfaces display a diverging curvature in the vicinity of the three phase contact line. The asymptotic contact angle at the contact line is equal to Young's angle, independent of the applied voltage. With respect to the morphology of the liquid surface, contact angle variations achieved by electrowetting are equivalent to those achieved by varying the chemical nature of the substrates, except for electric field-induced distortions in a region close to the contact line. Experimentally, we studied the (global) morphology of liquid microstructure substrates with stripe-shaped electrodes. As the local contact angle is reduced by increasing the applied voltage, liquid droplets elongate along the stripe axis as expected. For droplets on a single surface with a stripe electrode, there is a discontinuous morphological transition where elongated droplets transform into translationally invariant cylinder segments with the contact line pinned along the stripe edge and vice versa. If the liquid is confined between two parallel surfaces with parallel stripe electrodes, the elongation of the droplet and its transformation into a translationally invariant morphology with pinned contact lines is continuous. Experimental results are compared to analytical and numerical models.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\8KAZGIMX\\Mugele et al. - 2005 - Electrowetting a convenient way to switchable wet.pdf} @@ -2838,7 +2738,6 @@ @article{munchDewettingRatesThin2005 volume = {17}, number = {9}, pages = {S309}, - issn = {0953-8984}, doi = {10.1088/0953-8984/17/9/003}, abstract = {We investigate the dewetting rates of thin liquid films using a lubrication model that describes the dewetting process of polymer melts on hydrophobized substrates. We study the effect of different boundary conditions at the liquid/solid interface, in particular, of the no-slip and the Navier slip boundary condition, and compare our numerical solutions for the no-slip and the slip-dominated cases to available results that originate from scaling arguments, simplified flow assumptions and energy balances. We furthermore consider these issues for an extended lubrication model that includes nonlinear curvature.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\R5YYUARV\\Münch - 2005 - Dewetting rates of thin liquid films.pdf} @@ -2853,7 +2752,6 @@ @article{munchLubricationModelsSmall2005 volume = {53}, number = {3}, pages = {359--383}, - issn = {1573-2703}, doi = {10.1007/s10665-005-9020-3}, abstract = {A set of lubrication models for the thin film flow of incompressible fluids on solid substrates is derived and studied. The models are obtained as asymptotic limits of the Navier-Stokes equations with the Navier-slip boundary condition for different orders of magnitude for the slip-length parameter. Specifically, the influence of slip on the dewetting behavior of fluids on hydrophobic substrates is investigated here. Matched asymptotics are used to describe the dynamic profiles for dewetting films and comparison is given with computational simulations. The motion of the dewetting front shows transitions from being nearly linear in time for no-slip to t2/3 as the slip is increased. For much larger slip lengths the front motion appears to become linear again. Correspondingly, the dewetting profiles undergo a transition from oscillatory to monotone decay into the uniform film layer for large slip. Increasing the slip further, to very large values, is associated with an increasing degree of asymmetry in the structure of the dewetting ridge profile}, keywords = {dewetting films,lubrication models,matched asymptotics,Navier-slip condition,stability analysis}, @@ -2887,7 +2785,7 @@ @article{Stokes @article{nagayamaIntermediateWettingState2020, - title = {Intermediate Wetting State at Nano/Microstructured Surfaces}, + title = {Intermediate Wetting State at Nano/Micro- structured Surfaces}, author = {Nagayama, G. and Zhang, D.}, year = {2020}, journal = {Soft Matter}, @@ -2934,7 +2832,6 @@ @article{nikolovSuperspreadingDrivenMarangoni2002 volume = {96}, number = {1}, pages = {325--338}, - issn = {0001-8686}, doi = {10.1016/S0001-8686(01)00087-2}, abstract = {The spontaneous spreading (called superspreading) of aqueous trisiloxane ethoxylate surfactant solutions on hydrophobic solid surfaces is a fascinating phenomenon with several practical applications. For example, the ability of trisiloxane ethoxylate surfactants to enhance the spreading of spray solutions on waxy weed leaf surfaces, such as velvetleaf (Abutilion theophrasti), makes them excellent wetting agents for herbicide applications. The superspreading ability of silicone surfactants has been known for decades, but its mechanism is still not well understood. In this paper, we suggest that the spreading of trisiloxane ethoxylates is controlled by a surface tension gradient, which forms when a drop of surfactant solution is placed on a solid surface. The proposed model suggests that, as the spreading front stretches, the surface tension increases (the surfactant concentration becomes lower) at the front relative to the top of the droplet, thereby establishing a dynamic surface tension gradient. The driving force for spreading is due to the Marangoni effect, and our experiments showed that the higher the gradient, the faster the spreading. A simple model describing the phenomenon of superspreading is presented. We also suggest that the superspreading behavior of trisiloxane ethoxylates is a consequence of the molecular configuration at the air/water surface (i.e. small and compact hydrophobic part), as shown by molecular dynamics modeling. We also found that the aggregates and vesicles formed in trisiloxane solutions do not initiate the spreading process and therefore these structures are not a requirement for the superspreading process.}, keywords = {Marangoni flow,Superspreading,Trisiloxane ethoxylate}, @@ -2951,7 +2848,6 @@ @article{nisatoExcitationSurfaceDeformation1999 number = {7}, pages = {2356--2364}, publisher = {{American Chemical Society}}, - issn = {0024-9297}, doi = {10.1021/ma981546x}, abstract = {The phase separation kinetics of ultrathin deuterated poly(styrene)/poly(butadiene) polymer blend films spun cast onto striped self-assembled monolayer (SAM) substrates is studied by atomic force microscopy (AFM). Fourier transform analysis of the AFM topographic data at various stages of the film pattern development reveals the presence of quantized surface deformation modes. These modes are excited by the phase separation process when the scale of phase separation becomes commensurate with the period of the striped surface pattern. Thus, higher frequency modes become excited at early stages of phase separation, and these excitations decay with time as the phase separation pattern further coarsens. The film ultimately self-organizes into a periodic structure in which the fundamental mode has the largest amplitude. The influence of film thickness on the film morphology in this late stage is also investigated. A decrease in the film thickness leads to surface patterns that match those of the SAM substrates with increasing resolution. However, these film patterns break up into droplet arrays along the SAM stripes if the films are made too thin. This phenomenon is attributed to a capillary wave instability.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\WUAS2JFZ\\Nisato et al. - 1999 - Excitation of Surface Deformation Modes of a Phase.pdf;C\:\\Users\\Zitzero\\Zotero\\storage\\HXH5K5KK\\ma981546x.html} @@ -2991,7 +2887,6 @@ @article{oscuratoLightDrivenWettabilityTailoring2017 number = {35}, pages = {30133--30142}, publisher = {{American Chemical Society}}, - issn = {1944-8244}, doi = {10.1021/acsami.7b08025}, abstract = {The directional light-induced mass migration phenomenon arising in the photoresponsive azobenzene-containing materials has become an increasingly used approach for the fabrication of controlled tridimensional superficial textures. In the present work we demonstrate the tailoring of the superficial wettability of an azopolymer by means of the light-driven reconfiguration of an array of imprinted micropillars. Few simple illumination parameters are controlled to induce nontrivial wetting effects. Wetting anisotropy with controlled directionality, unidirectional spreading, and even polarization-intensity driven two-dimensional paths for wetting anisotropy are obtained starting from a single pristine pillar geometry. The obtained results prove that the versatility of the light-reconfiguration process, together with the possibility of reversible reshaping at reduced costs, represents a valid approach for both applications and fundamental studies in the field of geometry-based wettability of solid surfaces.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\K65KLLPD\\Oscurato et al. - 2017 - Light-Driven Wettability Tailoring of Azopolymer S.pdf;C\:\\Users\\Zitzero\\Zotero\\storage\\E274937T\\acsami.html} @@ -3007,7 +2902,6 @@ @article{pahlavanThinFilmsPartial2018 volume = {845}, pages = {642--681}, publisher = {{Cambridge University Press}}, - issn = {0022-1120, 1469-7645}, doi = {10.1017/jfm.2018.255}, abstract = {A uniform nanometric thin liquid film on a solid substrate can become unstable due to the action of van der Waals (vdW) forces. The instability leads to dewetting of the uniform film and the formation of drops. To minimize the total free energy of the system, these drops coarsen over time until one single drop remains. Here, using a thermodynamically consistent framework, we derive a new model for thin films in partial wetting with a free energy that resembles the Cahn\textendash Hilliard form with a height-dependent surface tension that leads to a generalized disjoining pressure, and revisit the dewetting problem. Using both linear stability analysis and nonlinear simulations we show that the new model predicts a slightly smaller critical instability wavelength and a significantly (up to six-fold) faster growth rate than the classical model in the spinodal regime; this faster growth rate brings the theoretical predictions closer to published experimental observations. During coarsening at intermediate times, the dynamics become self-similar and model-independent; we therefore observe the same scalings in both the classical (with and without thermal noise) and new models. Both models also lead to a mean-field Lifshitz\textendash Slyozov\textendash Wagner (LSW)-type droplet-size distribution at intermediate times for small drop sizes. We, however, observe a skewed drop-size distribution for larger drops in the new model; while the tail of the distribution follows a Smoluchowski equation, it is not associated with a coalescence-dominated coarsening, calling into question the association made in some earlier experiments. Our observations point to the importance of the height dependence of surface tension in the early and late stages of dewetting of nanometric films and motivate new high-resolution experimental observations to guide the development of improved models of interfacial flows at the nanoscale.}, keywords = {contact lines,interfacial flows (free surface),thin films}, @@ -3039,7 +2933,6 @@ @article{pawarSymmetricAsymmetricCoalescence2019 number = {9}, pages = {092106}, publisher = {{American Institute of Physics}}, - issn = {1070-6631}, doi = {10.1063/1.5119014}, abstract = {We present an investigation of symmetric and asymmetric coalescence of two droplets of equal and unequal size on a solid surface in the inertia-dominated regime. Asymmetric coalescence can result due to the coalescence of two unequal-sized droplets or coalescence of two droplets having different contact angles with the surface due to a step gradient in wettability. Based on the solution of an analytical model and lattice Boltzmann simulations, we analyze symmetric and asymmetric coalescence of two droplets on a solid surface. The analysis of coalescence of identical droplets show that the liquid bridge height grows with time as (t*)1/2 for \texttheta{} = 90\textdegree{} and (t*)2/3 for \texttheta{} {$<$} 90\textdegree, where t* is dimensionless time. Our analysis also yields the same scaling law for the coalescence of two unequal-sized droplets on a surface with homogeneous wettability. We also discuss the coalescence of two droplets having different contact angles with the surface due to a step gradient in wettability. We show that the prediction of bridge height with time scales as (t*)2/3 irrespective of contact angles of droplet with the surface.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\8ZDRQ9KY\\Pawar et al. - 2019 - Symmetric and asymmetric coalescence of droplets o.pdf} @@ -3048,7 +2941,7 @@ @article{pawarSymmetricAsymmetricCoalescence2019 @incollection{pelinovskyHydrodynamicsTsunamiWaves2006, title = {Hydrodynamics of {{Tsunami Waves}}}, booktitle = {Waves in {{Geophysical Fluids}}: {{Tsunamis}}, {{Rogue Waves}}, {{Internal Waves}} and {{Internal Tides}}}, - author = {Pelinovsky, Efim}, + author = {Pelinovsky, E.}, editor = {Grue, J. and Trulsen, K.}, year = {2006}, series = {{{CISM International Centre}} for {{Mechanical Sciences}}}, @@ -3071,7 +2964,6 @@ @article{peschkaVariationalApproachDynamic2018 volume = {30}, number = {8}, pages = {082115}, - issn = {1070-6631}, doi = {10.1063/1.5040985}, urldate = {2023-07-24}, abstract = {This paper investigates a modeling approach for viscous flows with dynamic contact angles based on variational energy-dissipation principles. A corresponding Stokes free boundary problem is reduced to a thin-film equation and its variational structure is constructed. The usefulness of this abstract modeling approach is demonstrated by providing numerical schemes for the Stokes flow and the thin-film model and by computing numerical solutions for the problem of gravity-driven thin droplets. Some implications of the contact angle model and properties of the numerical scheme are highlighted in this setting.}, @@ -3103,7 +2995,6 @@ @article{pihler-puzovicDisplacementFlowsElastic2015 volume = {784}, pages = {487--511}, publisher = {{Cambridge University Press}}, - issn = {0022-1120, 1469-7645}, doi = {10.1017/jfm.2015.590}, abstract = {The injection of fluid into the narrow liquid-filled gap between a rigid plate and an elastic membrane drives a displacement flow that is controlled by the competition between elastic and viscous forces. We study such flows using the canonical set-up of an elastic-walled Hele-Shaw cell whose upper boundary is formed by an elastic sheet. We investigate both single- and two-phase displacement flows in which the localised injection of fluid at a constant flow rate is accommodated by the inflation of the sheet and the outward propagation of an axisymmetric front beyond which the cell remains approximately undeformed. We perform a direct comparison between quantitative experiments and numerical simulations of two theoretical models. The models couple the F\"oppl\textendash von K\'arm\'an equations, which describe the deformation of the thin elastic membrane, to the equations describing the flow, which we model by (i) the Navier\textendash Stokes equations or (ii) lubrication theory. We identify the dominant physical effects that control the behaviour of the system and critically assess modelling assumptions that were made in previous studies. The insight gained from these studies is then used in Part~2 of this work, where we formulate an improved lubrication model and develop an asymptotic description of the key phenomena.}, keywords = {flow–structure interactions,Hele-Shaw flows,low-Reynolds-number flows}, @@ -3119,7 +3010,6 @@ @article{pihler-puzovicModellingSuppressionViscous2013 volume = {731}, pages = {162--183}, publisher = {{Cambridge University Press}}, - issn = {0022-1120, 1469-7645}, doi = {10.1017/jfm.2013.375}, abstract = {Recent experiments by Pihler-Puzovic et~al.~(Phys. Rev. Lett., vol. 108, 2012, article~074502) have shown that the onset of viscous fingering in circular Hele-Shaw cells in which an air bubble displaces a viscous fluid is delayed considerably when the top boundary of the cell is replaced by an elastic membrane. Non-axisymmetric instabilities are only observed at much larger flow rates, and the large-amplitude fingers that develop are fundamentally different from the highly branched fingers in rigid-walled cells. We explain the mechanism for the suppression of the instability using a combination of linear stability analysis and direct numerical simulations, based on a theoretical model that couples a depth-averaged lubrication equation for the fluid flow to the F\"oppl\textendash von K\'arm\'an equations, which describe the deformation of the elastic membrane. We show that fluid\textendash structure interaction affects the instability primarily via two changes to the axisymmetric base flow: the axisymmetric inflation of the membrane prior to the onset of any instabilities slows down the expansion of the air bubble and forces the air\textendash liquid interface to propagate into a converging fluid-filled gap. Both of these changes reduce the destabilizing viscous effects that drive the fingering instability in a rigid-walled cell. In contrast, capillary effects only play a very minor role in the suppression of the instability.}, keywords = {fingering instability,flow–structure interaction,Hele-Shaw flows} @@ -3134,7 +3024,6 @@ @article{plimptonFastParallelAlgorithms1995 volume = {117}, number = {1}, pages = {1--19}, - issn = {0021-9991}, doi = {10.1006/jcph.1995.1039}, abstract = {Three parallel algorithms for classical molecular dynamics are presented. The first assigns each processor a fixed subset of atoms; the second assigns each a fixed subset of inter-atomic forces to compute; the third assigns each a fixed spatial region. The algorithms are suitable for molecular dynamics models which can be difficult to parallelize efficiently\textemdash those with short-range forces where the neighbors of each atom change rapidly. They can be implemented on any distributed-memory parallel machine which allows for message-passing of data between independently executing processors. The algorithms are tested on a standard Lennard-Jones benchmark problem for system sizes ranging from 500 to 100,000,000 atoms on several parallel supercomputers--the nCUBE 2, Intel iPSC/860 and Paragon, and Cray T3D. Comparing the results to the fastest reported vectorized Cray Y-MP and C90 algorithm shows that the current generation of parallel machines is competitive with conventional vector supercomputers even for small problems. For large problems, the spatial algorithm achieves parallel efficiencies of 90\% and a 1840-node Intel Paragon performs up to 165 faster than a single Cray C9O processor. Trade-offs between the three algorithms and guidelines for adapting them to more complex molecular dynamics simulations are also discussed.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\D86XCJY2\\Plimpton - 1995 - Fast Parallel Algorithms for Short-Range Molecular.pdf;C\:\\Users\\Zitzero\\Zotero\\storage\\M5KQXCFF\\S002199918571039X.html} @@ -3164,7 +3053,6 @@ @article{poissonNouvelleTheorieAction1831 number = {7}, pages = {170--175}, publisher = {{De Gruyter}}, - issn = {1435-5345}, doi = {10.1515/crll.1831.7.170}, abstract = {Article Nouvelle th\'eorie de l'action capillaire. was published on January 1, 1831 in the journal Journal f\"ur die reine und angewandte Mathematik (volume 1831, issue 7).}, chapter = {Journal f\"ur die reine und angewandte Mathematik} @@ -3193,7 +3081,6 @@ @article{prestininziEffectIntrinsicViscosity2013 number = {6}, pages = {668--680}, publisher = {{Taylor \& Francis}}, - issn = {0022-1686}, doi = {10.1080/00221686.2013.819532}, abstract = {In this work, a numerical assessment of the suitability of a Single Relaxation Time (SRT) Lattice Boltzmann Method (LBM) model to simulate axisymmetric gravity currents is carried out. The model results are compared with both experimental data and other numerical models. The particular SRT formulation employed is known to converge, in the limit of low Knudsen number, to the two-layer 2D Shallow Water Equations (SWEs) set with a viscosity term featuring a closed theoretical formulation. Even with the lowest viscosity achievable by the method, its effect is shown to become important in most of the cases analysed, thus posing some serious constraints on possible application of the single relaxation time LBM method to simulate the lock-release generated-type gravity currents analysed here. The comparison with classical numerical models shows that the the viscous effects in the LBM model can be well reproduced employing coefficients derived from the above-mentioned theoretical formulation.}, keywords = {Axisymmetric gravity currents,gravity currents,LBM,shallow water,two-layer flows,viscosity}, @@ -3211,7 +3098,6 @@ @article{qianFractionalPropagationElimination1997 number = {04}, pages = {753--761}, publisher = {{World Scientific Publishing Co.}}, - issn = {0129-1831}, doi = {10.1142/S0129183197000643}, abstract = {Lattice-based models have been attracting much interest in recent years and have been applied to many complex systems. The derivation of large scale dynamical equations of lattice-gas models as well as lattice-Boltzmann models was based on the belief that only the physically interesting quantities (mass, momentum and energy) are conserved. Staggered invariants in lattice-gas models were found in 1988 and there have been no efficient methods to eliminate these invariants. In this paper, we will first discuss the existence of staggered invariants, then we propose to use fractional propagation as an effective way of suppressing these undesired invariants. Numerical simulations will be used to confirm the theory and to show the improvement of computations.}, keywords = {BGK Models,Fractional Propagation,Lattice-Gas,Staggered Invariants} @@ -3226,7 +3112,6 @@ @article{raabeOverviewLatticeBoltzmann2004 volume = {12}, number = {6}, pages = {R13}, - issn = {0965-0393}, doi = {10.1088/0965-0393/12/6/R01}, abstract = {The article gives an overview of the lattice Boltzmann method as a powerful technique for the simulation of single and multi-phase flows in complex geometries. Owing to its excellent numerical stability and constitutive versatility it can play an essential role as a simulation tool for understanding advanced materials and processes. Unlike conventional Navier\textendash Stokes solvers, lattice Boltzmann methods consider flows to be composed of a collection of pseudo-particles that are represented by a velocity distribution function. These fluid portions reside and interact on the nodes of a grid. System dynamics and complexity emerge by the repeated application of local rules for the motion, collision and redistribution of these coarse-grained droplets. The lattice Boltzmann method, therefore, is an ideal approach for mesoscale and scale-bridging simulations. It is capable to tackling particularly those problems which are ubiquitous characteristics of flows in the world of materials science and engineering, namely, flows under complicated geometrical boundary conditions, multi-scale flow phenomena, phase transformation in flows, complex solid\textendash liquid interfaces, surface reactions in fluids, liquid\textendash solid flows of colloidal suspensions and turbulence. Since the basic structure of the method is that of a synchronous automaton it is also an ideal platform for realizing combinations with related simulation techniques such as cellular automata or Potts models for crystal growth in a fluid or gas environment. This overview consists of two parts. The first one reviews the philosophy and the formal concepts behind the lattice Boltzmann approach and presents also related pseudo-particle approaches. The second one gives concrete examples in the area of computational materials science and process engineering, such as the prediction of lubrication dynamics in metal forming, dendritic crystal growth under the influence of fluid convection, simulation of metal foam processing, flow percolation in confined geometries, liquid crystal hydrodynamics and processing of polymer blends.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\P7MFE6BZ\\Raabe - 2004 - Overview of the lattice Boltzmann method for nano-.pdf} @@ -3300,7 +3185,6 @@ @article{richardsonNoslipBoundaryCondition1973 number = {4}, pages = {707--719}, publisher = {{Cambridge University Press}}, - issn = {1469-7645, 0022-1120}, doi = {10.1017/S0022112073001801}, abstract = {It has been argued that the no-slip boundary condition, applicable when a viscous fluid flows over a solid surface, may be an inevitable consequence of the fact that all such surfaces are, in practice, rough on a microscopic scale: the energy lost through viscous dissipation as a fluid passes over and around these irregularities is sufficient to ensure that it is effectively brought to rest. The present paper analyses the flow over a particularly simple model of such a rough wall to support these physical ideas.} } @@ -3314,7 +3198,6 @@ @article{riobooTimeEvolutionLiquid2002 volume = {33}, number = {1}, pages = {112--124}, - issn = {1432-1114}, doi = {10.1007/s00348-002-0431-x}, abstract = {The normal impact of liquid drops onto solid, dry surfaces has been studied experimentally, using high-resolution digital photography. A large number of parameters were varied in a systematic manner. The focus of this paper is the quantitative determination of the influence of these parameters on the drop spreading upon impact and on the phenomenological description of the outcomes. Dimensional similarity of the spreading can only be achieved for the very early stage of the impact process. At later stages, the number of influencing factors increases, generally precluding any universal correlation. Particular emphasis is placed on the influence of the wettability and the surface roughness on spreading.}, keywords = {Influence Factor,Quantitative Determination,Surface Roughness,Time Evolution,Wettability}, @@ -3329,7 +3212,6 @@ @article{roccaDevelopmentLatticeBoltzmann2012 volume = {70}, number = {8}, pages = {1048--1072}, - issn = {1097-0363}, doi = {10.1002/fld.2742}, abstract = {In this paper the dynamics of a two-layered liquid, made of two immiscible shallow-layers of different density, has been investigated within the framework of the lattice Boltzmann method (LBM). The LBM developed in this paper for the two-layered, shallow-water flow has been obtained considering two separate sets of LBM equations, one for each layer. The coupling terms between the two sets have been defined as external forces, acted on one layer by the other. Results obtained from the LBM developed in this paper are compared with numerical results obtained solving the two-layered, shallow-water equations, with experimental and other numerical results published in literature. The results are interesting. First, the numerical results obtained by the LBM and by the shallow-water model can be considered as equivalent. Second, the LBM developed in this paper is able to simulate motion conditions on nonflat topography. Third, the agreement between the LBM (and also shallow-water model) numerical results and the experimental results is good when the evolution of the flow does not depend on the viscosity, that is, during the initial phase of the flow, dominated by gravity and inertia forces. When the viscous forces dominate the evolution of the flow the agreement between numerical and experimental results depends strongly on the viscosity; it is good if the numerical LBM viscosity has the same order of magnitude of the liquid's kinematic viscosity. Copyright \textcopyright{} 2012 John Wiley \& Sons, Ltd.}, keywords = {hydrodynamics,lattice Boltzmann,partial differential equations,particle method,shallow water,two-phase flows}, @@ -3345,7 +3227,6 @@ @article{ronsinRoleInterplaySpinodal2020 volume = {8}, number = {12}, pages = {1901468}, - issn = {2194-4296}, doi = {10.1002/ente.201901468}, abstract = {The stability of organic solar cells is strongly affected by the morphology of the photoactive layers, whose separated crystalline and/or amorphous phases are kinetically quenched far from their thermodynamic equilibrium during the production process. The evolution of these structures during the lifetime of the cell remains poorly understood. Herein, a phase-field simulation framework is proposed, handling liquid\textendash liquid demixing and polycrystalline growth at the same time to investigate the evolution of crystalline immiscible binary systems. It is found that initially, the nuclei trigger the spinodal decomposition, whereas the growing crystals quench the phase coarsening in the amorphous mixture. In contrast, the separated liquid phases guide the crystal growth along the domains of high concentration. It is also demonstrated that with a higher crystallization rate, in the final morphology, single crystals are more structured and form percolating pathways for each material with smaller lateral dimensions.}, keywords = {crystallization,liquid–liquid demixing,organic solar cells,phase-field,stability}, @@ -3395,7 +3276,6 @@ @article{rungeUeberNumerischeAufloesung1895 volume = {46}, number = {2}, pages = {167--178}, - issn = {1432-1807}, doi = {10.1007/BF01446807}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\3KJK6TNM\\Runge - 1895 - Ueber die numerische Auflösung von Differentialgle.pdf} } @@ -3409,7 +3289,6 @@ @article{ruschakLimitingFlowPremetered1976 volume = {31}, number = {11}, pages = {1057--1060}, - issn = {0009-2509}, doi = {10.1016/0009-2509(76)87026-1}, abstract = {A limiting case of flow in a pre-metered coating device is analyzed using a singular perturbation method. The ability of this device to coat a range of film thicknesses at the same coating speed is traced to the freedom of the two free surfaces to adopt any curvature within certain limits. The observed limitations on operating conditions for this coater are explained by the failure of the desired steady-state to exist for all values of the operating parameters. The parameter bounds within which the device can coat a uniform liquid layer are predicted quantitatively. Some analyses of dip coating are reviewed in the light of the present work.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\DET6JH8L\\0009250976870261.html} @@ -3424,7 +3303,6 @@ @article{ruzickaDimensionlessNumbers2008 volume = {86}, number = {8}, pages = {835--868}, - issn = {0263-8762}, doi = {10.1016/j.cherd.2008.03.007}, abstract = {The goal is to provide a little review on dimensionless numbers, commonly encountered in chemical engineering. Both their sources are considered: dimensional analysis and scaling of governing equations with boundary conditions. The numbers produced by scaling of equation are presented for transport of momentum, heat and mass. Momentum transport is considered in both single-phase and multi-phase flows. The numbers obtained are assigned the physical meaning, and their mutual relations are highlighted. Certain drawbacks of building correlations based on dimensionless numbers are pointed out.}, keywords = {Correlations,Dimensional analysis,Dimensionless numbers,Multi-phase flow,Scaling of boundary conditions,Scaling of equations,Single-phase flow}, @@ -3508,7 +3386,6 @@ @article{satinskySystematicReviewMetaanalysis2021 number = {1}, pages = {14370}, publisher = {{Nature Publishing Group}}, - issn = {2045-2322}, doi = {10.1038/s41598-021-93687-7}, abstract = {University administrators and mental health clinicians have raised concerns about depression and anxiety among Ph.D. students, yet no study has systematically synthesized the available evidence in this area. After searching the literature for studies reporting on depression, anxiety, and/or suicidal ideation among Ph.D. students, we included 32 articles. Among 16 studies reporting the prevalence of clinically significant symptoms of~depression across 23,469 Ph.D. students, the pooled estimate of the proportion of students with depression was 0.24 (95\% confidence interval [CI], 0.18\textendash 0.31; I2\,=\,98.75\%). In a meta-analysis of the nine studies reporting the prevalence of clinically significant symptoms of~anxiety across 15,626 students, the estimated proportion of students with anxiety was 0.17 (95\% CI, 0.12\textendash 0.23; I2\,=\,98.05\%). We conclude that depression and anxiety are highly prevalent among Ph.D. students. Data limitations precluded our ability to obtain a pooled estimate of suicidal ideation prevalence. Programs that systematically monitor and promote the mental health of Ph.D. students are urgently needed.}, copyright = {2021 The Author(s)}, @@ -3600,7 +3477,6 @@ @article{schallerMooreLawPresent1997 volume = {34}, number = {6}, pages = {52--59}, - issn = {1939-9340}, doi = {10.1109/6.591665}, abstract = {A simple observation, made over 30 years ago, on the growth in the number of devices per silicon die has become the central driving force of one of the most dynamic of the world's industries. Because of the accuracy with which Moore's Law has predicted past growth in IC complexity, it is viewed as a reliable method of calculating future trends as well, setting the pace of innovation, and defining the rules and the very nature of competition. And since the semiconductor portion of electronic consumer products keeps growing by leaps and bounds, the Law has aroused in users and consumers an expectation of a continuous stream of faster, better, and cheaper high-technology products. Even the policy implications of Moore's Law are significant: it is used as the baseline assumption in the industry's strategic road map for the next decade and a half.}, keywords = {Consumer products,Moore's Law,Roads,Silicon,Technological innovation}, @@ -3617,7 +3493,6 @@ @article{schallerPapaya22DIrreducible2020 volume = {5}, number = {54}, pages = {2538}, - issn = {2475-9066}, doi = {10.21105/joss.02538}, abstract = {A common challenge in scientific and technical domains is the quantitative description of geometries and shapes, e.g. in the analysis of microscope imagery or astronomical observation data. Frequently, it is desirable to go beyond scalar shape metrics such as porosity and surface to volume ratios because the samples are anisotropic or because direction-dependent quantities such as conductances or elasticity are of interest. Popular analysis software such as ImageJ and SExtractor provide only limited tooling for higher-order anisotropy characterization; usually only the tensor of inertia (rank 2) is available.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\PQNNLS6L\\Schaller et al. - 2020 - papaya2 2D Irreducible Minkowski Tensor computati.pdf} @@ -3644,7 +3519,6 @@ @article{schwartzSimulationDropletMotion1998 volume = {202}, number = {1}, pages = {173--188}, - issn = {0021-9797}, doi = {10.1006/jcis.1998.5448}, abstract = {A method of calculation is introduced that allows the simulation of the time-dependent three-dimensional motion of liquid droplets on solid substrates for systems exhibiting finite equilibrium contact angles. The contact angle is a prescribed function of position on the substrate. An evolution equation is presented, using the lubrication approximation, that includes viscous, capillary, disjoining, and gravitational forces. Motion to and from dry substrate regions is made possible by use of a thin energetically stable wetting layer. Axisymmetric spreading on a uniform substrate is calculated, and it is found, in agreement with reported experiments, that spreading rates are independent of the contact angle until the drop has almost stabilized. We simulate motion on a heterogeneous substrate composed of two different materials having widely different contact angles. Motion proceeds in an almost discontinuous fashion as the initial droplet breaks up into smaller pieces through the action of the wetting forces. Various forms of the disjoining energy functional are employed; the particular choice is found to have only a limited quantitative effect of the drop dynamics. Experimental observations confirm the basic features of the simulation, although a time-scale correction needs to be applied.}, keywords = {disjoining pressure,Droplet spreading,fluid mechanics,heterogeneous substrate,numerical simulation,surface tension}, @@ -3659,7 +3533,6 @@ @article{scrivenPhysicsApplicationsDIP1988 volume = {121}, pages = {717}, publisher = {{Cambridge University Press}}, - issn = {0272-9172, 1946-4274}, doi = {10.1557/PROC-121-717}, abstract = {Dip coating is a simple old way of depositing onto a substrate, especially small slabs and cylinders, a uniform thin film of liquid for solidification into a coating. The basic flow is steady, and in it film thickness is set by the competition among viscous force, capillary (surface tension) force and gravity. Thickness and uniformity can be sensitive to flow conditions in the liquid bath and gas overhead. The faster the substrate is withdrawn, the thicker the film deposited. This can be countered by using volatile solutes and combining rapid enough drying with the basic liquid flow. Then the physics grows more complicated, theoretical prediction of process performance more difficult, and control of the process more demanding. Outside product R\&D labs it is far less often used in precision coating manufacture than a variety of premetered coating methods.Spin coating is a more recently developed way of getting onto piecemeal substrates, especially small flat disks, a uniform thin liquid film for the same end. The basic flow is unsteady radial drainage in which centrifugal and viscous forces so compete that ordinary (Newtonian) liquid of constant viscosity tends toward a uniform film that grows ever thinner ever more slowly. Volatile solvents are commonly used because conditions can often be found that adequately separate thinning by spin-off from later thinning and solidification by drying. Thickness and uniformity, today theoretically predictable, are sensitive to speed, gas conditions, and rheology of concentrating, solidifying liquid. For the rheology of photoresist coating in microelectronics, spin coating works well. For that of suspension coatings in magnetic disk technology the process demands more careful control; actually it is often modified.} } @@ -3675,7 +3548,6 @@ @article{secchiMassiveRadiusdependentFlow2016 number = {7619}, pages = {210--213}, publisher = {{Nature Publishing Group}}, - issn = {1476-4687}, doi = {10.1038/nature19315}, abstract = {The pressure-driven flow rate through individual carbon nanotubes is precisely determined from the hydrodynamics of emerging water jets, revealing unexpectedly large and radius-dependent surface slippage.}, copyright = {2016 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.}, @@ -3706,7 +3578,6 @@ @article{sekiWideArrayPhotoinduced2018 number = {7}, pages = {1026--1057}, publisher = {{The Chemical Society of Japan}}, - issn = {0009-2673}, doi = {10.1246/bcsj.20180076}, abstract = {Molecular and macromolecular assemblies have good ``chemistry'' with light. In the research of photochromic molecules and systems, the number of papers in the most recent 10 years exceeds the half of the whole past century. Photochromic molecules have been studied for their characteristic color changes. Although the importance in this line has not dimmed, other powerful areas have emerged in the last two or three decades. Much more attention has been drawn to light induced motions triggered by photoreactions utilizing molecular and polymer assemblies. Probably, this field is one of the most developing areas in advanced soft material research. This review article mainly deals with azobenzene and diarylethene as the photochroms, and systematically introduces light triggered dynamic processes taking place at interfaces. In terms of the feature size, the systems at interfaces, mesoscopic and microscopic level materials are particularly highlighted. Molecular design and understanding of interfaces are the key to realize smart motility properties in molecular and polymer assemblies.}, keywords = {Molecular and macromolecular assemblies,Motions at interfaces,Photochromic molecules}, @@ -3722,7 +3593,6 @@ @article{shahThermalFluctuationsCapillary2019 volume = {876}, pages = {1090--1107}, publisher = {{Cambridge University Press}}, - issn = {0022-1120, 1469-7645}, doi = {10.1017/jfm.2019.595}, abstract = {Thermal fluctuations have been shown to influence the thinning dynamics of planar thin liquid films, bringing predicted rupture times closer to experiments. Most liquid films in nature and industry are, however, non-planar. Thinning of such films not just results from the interplay between stabilizing surface tension forces and destabilizing van der Waals forces, but also from drainage due to curvature differences. This work explores the influence of thermal fluctuations on the dynamics of thin non-planar films subjected to drainage, with their dynamics governed by two parameters: the strength of thermal fluctuations, {$\mathsl{\theta}\theta\backslash$}unicode[STIX]\{x1D703\}, and the strength of drainage, {$\mathsl{K}\kappa\backslash$}unicode[STIX]\{x1D705\}. For strong drainage ({$\mathsl{K}\gg\mathsl{K}$}tr{$\kappa\gg\kappa$}tr\textbackslash unicode[STIX]\{x1D705\}\textbackslash gg \textbackslash unicode[STIX]\{x1D705\}\_\{tr\}), we find that the film ruptures due to the formation of a local depression called a dimple that appears at the connection between the curved and flat parts of the film. For this dimple-dominated regime, the rupture time, trtrt\_\{r\}, solely depends on {$\mathsl{K}\kappa\backslash$}unicode[STIX]\{x1D705\}, according to the earlier reported scaling, tr{$\sim\mathsl{K}-$}10/7tr{$\sim\kappa-$}10/7t\_\{r\}\textbackslash sim \textbackslash unicode[STIX]\{x1D705\}\^\{-10/7\}. By contrast, for weak drainage ({$\mathsl{K}\ll\mathsl{K}$}tr{$\kappa\ll\kappa$}tr\textbackslash unicode[STIX]\{x1D705\}\textbackslash ll \textbackslash unicode[STIX]\{x1D705\}\_\{tr\}), the film ruptures at a random location due to the spontaneous growth of fluctuations originating from thermal fluctuations. In this fluctuations-dominated regime, the rupture time solely depends on {$\mathsl{\theta}\theta\backslash$}unicode[STIX]\{x1D703\} as tr{$\sim-$}(1/{$\mathsl{\omega}$}max)ln({$\surd$}2{$\mathsl{\theta}$}){$\mathsl{A}$}tr{$\sim-$}(1/{$\omega$}max)ln(2\texttheta --{$\surd$}){$\alpha$}t\_\{r\}\textbackslash sim -(1/\textbackslash unicode[STIX]\{x1D714\}\_\{max\})\textbackslash ln (\textbackslash sqrt\{2\textbackslash unicode[STIX]\{x1D703\}\})\^\{\textbackslash unicode[STIX]\{x1D6FC\}\}, with {$\mathsl{A}$}=1.15{$\alpha$}=1.15\textbackslash unicode[STIX]\{x1D6FC\}=1.15. This scaling is rationalized using linear stability theory, which yields {$\mathsl{\omega}$}max{$\omega$}max\textbackslash unicode[STIX]\{x1D714\}\_\{max\} as the growth rate of the fastest-growing wave and {$\mathsl{A}$}=1{$\alpha$}=1\textbackslash unicode[STIX]\{x1D6FC\}=1. These insights on if, when and how thermal fluctuations play a role are instrumental in predicting the dynamics and rupture time of non-flat draining thin films.}, keywords = {breakup/coalescence,thin films}, @@ -3754,7 +3624,6 @@ @article{shanKineticTheoryRepresentation2006 volume = {550}, pages = {413--441}, publisher = {{Cambridge University Press}}, - issn = {1469-7645, 0022-1120}, doi = {10.1017/S0022112005008153}, abstract = {We present in detail a theoretical framework for representing hydrodynamic systems through a systematic discretization of the Boltzmann kinetic equation. The work is an extension of a previously proposed formulation. Conventional lattice Boltzmann models can be shown to be directly derivable from this systematic approach. Furthermore, we provide here a clear and rigorous procedure for obtaining higher-order approximations to the continuum Boltzmann equation. The resulting macroscopic moment equations at each level of the systematic discretization give rise to the Navier\textendash Stokes hydrodynamics and those beyond. In addition, theoretical indications to the order of accuracy requirements are given for each discrete approximation, for thermohydrodynamic systems, and for fluid systems involving long-range interactions. All these are important for complex and micro-scale flows and are missing in the conventional Navier\textendash Stokes order descriptions. The resulting discrete Boltzmann models are based on a kinetic representation of the fluid dynamics, hence the drawbacks in conventional higher-order hydrodynamic formulations can be avoided.} } @@ -3783,7 +3652,6 @@ @article{sharpOverviewRayleighTaylorInstability1984 volume = {12}, number = {1-3}, pages = {3--18}, - issn = {01672789}, doi = {10.1016/0167-2789(84)90510-4}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\DNXLFCN6\\Sharp - 1984 - An overview of Rayleigh-Taylor instability.pdf} } @@ -3798,7 +3666,6 @@ @article{shiFogHarvestingHarps2018 number = {14}, pages = {11979--11986}, publisher = {{American Chemical Society}}, - issn = {1944-8244}, doi = {10.1021/acsami.7b17488}, abstract = {Fog harvesting is a useful technique for obtaining fresh water in arid climates. The wire meshes currently utilized for fog harvesting suffer from dual constraints: coarse meshes cannot efficiently capture microscopic fog droplets, whereas fine meshes suffer from clogging issues. Here, we design and fabricate fog harvesters comprising an array of vertical wires, which we call ``fog harps''. Under controlled laboratory conditions, the fog-harvesting rates for fog harps with three different wire diameters were compared to conventional meshes of equivalent dimensions. As expected for the mesh structures, the mid-sized wires exhibited the largest fog collection rate, with a drop-off in performance for the fine or coarse meshes. In contrast, the fog-harvesting rate continually increased with decreasing wire diameter for the fog harps due to efficient droplet shedding that prevented clogging. This resulted in a 3-fold enhancement in the fog-harvesting rate for the harp design compared to an equivalent mesh.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\9PAZJLIF\\Shi et al. - 2018 - Fog Harvesting with Harps.pdf;C\:\\Users\\Zitzero\\Zotero\\storage\\AA8KLZW4\\acsami.html} @@ -3812,7 +3679,6 @@ @article{singhInkjetPrintingProcess2010 volume = {22}, number = {6}, pages = {673--685}, - issn = {1521-4095}, doi = {10.1002/adma.200901141}, abstract = {In this Progress Report we provide an update on recent developments in inkjet printing technology and its applications, which include organic thin-film transistors, light-emitting diodes, solar cells, conductive structures, memory devices, sensors, and biological/pharmaceutical tasks. Various classes of materials and device types are in turn examined and an opinion is offered about the nature of the progress that has been achieved.}, keywords = {conductive structures,inkjet printing,light-emitting diodes,memory devices,organic thin-film transistors,sensors,solar cells}, @@ -3845,7 +3711,6 @@ @article{snoeijerMovingContactLines2013 volume = {45}, number = {1}, pages = {269--292}, - issn = {0066-4189, 1545-4479}, doi = {10.1146/annurev-fluid-011212-140734}, abstract = {The speed at which a liquid can move over a solid surface is strongly limited when a three-phase contact line is present, separating wet from dry regions. When enforcing large contact line speeds, this leads to the entrainment of drops, films, or air bubbles. In this review, we discuss experimental and theoretical progress revealing the physical mechanisms behind these dynamical wetting transitions. In this context, we discuss microscopic processes that have been proposed to resolve the moving\textendash contact line paradox and identify the different dynamical regimes of contact line motion.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\MLV4RHQW\\Snoeijer and Andreotti - 2013 - Moving Contact Lines Scales, Regimes, and Dynamic.pdf} @@ -3860,7 +3725,6 @@ @article{sommersCreatingMicroscaleSurface2006 volume = {16}, number = {8}, pages = {1571}, - issn = {0960-1317}, doi = {10.1088/0960-1317/16/8/018}, abstract = {A technique for fabricating micropatterned aluminum surfaces with parallel grooves 30 \textmu m wide and tens of microns in depth is described. Standard photolithographic techniques are used to obtain this precise surface-feature patterning. Positive photoresists, S1813 and AZ4620, are selected to mask the surface, and a mixture of BCl3 and Cl2 gases is used to perform the etching. Experimental data show that a droplet placed on the micro-grooved aluminum surface using a micro-syringe exhibits an increased apparent contact angle, and for droplets condensed on these etched surfaces, more than a 50\% reduction in the volume needed for the onset of droplet sliding is manifest. No chemical surface treatment is necessary to achieve this water repellency; it is accomplished solely by an anisotropic surface morphology that manipulates droplet geometry and creates and exploits discontinuities in the three-phase contact line. These micro-structured surfaces are proposed for use in a broad range of air-cooling applications, where the management of condensate and defrost liquid on the heat transfer surface is essential to the energy-efficient operation of the machine.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\JSJTQ4Q4\\Sommers and Jacobi - 2006 - Creating micro-scale surface topology to achieve a.pdf} @@ -3932,7 +3796,6 @@ @article{stuartEmergingApplicationsStimuliresponsive2010 number = {2}, pages = {101--113}, publisher = {{Nature Publishing Group}}, - issn = {1476-4660}, doi = {10.1038/nmat2614}, abstract = {Stimuli-responsive polymers can be engineered, in both film and colloid forms, to respond to a variety of inputs, from temperature to pH. The inherent flexibility in their structure and responses result in materials that lend themselves to applications ranging from drug delivery to sensing. Recent advances and future challenges in this direction are reviewed.}, copyright = {2010 Nature Publishing Group}, @@ -3963,7 +3826,6 @@ @article{suiInertialCoalescenceDroplets2013 number = {10}, pages = {101701}, publisher = {{American Institute of Physics}}, - issn = {1070-6631}, doi = {10.1063/1.4824108}, abstract = {We consider the growth rate of the height of the connecting bridge in rapid surface-tension-driven coalescence of two identical droplets attached on a partially wetting substrate. For a wide range of contact angle values, the height of the bridge grows with time following a power law with a universal exponent of 2/3, up to a threshold time, beyond which a 1/2 exponent results, that is known for coalescence of freely-suspended droplets. In a narrow range of contact angle values close to 90\textdegree, this threshold time rapidly vanishes and a 1/2 exponent results for a 90\textdegree{} contact angle. The argument is confirmed by three-dimensional numerical simulations based on a diffuse interface method with adaptive mesh refinement and a volume-of-fluid method.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\KS2WGIEU\\Sui et al. - 2013 - Inertial coalescence of droplets on a partially we.pdf} @@ -3979,7 +3841,6 @@ @article{sultanEvaporationThinFilm2005 volume = {543}, pages = {183--202}, publisher = {{Cambridge University Press}}, - issn = {1469-7645, 0022-1120}, doi = {10.1017/S0022112005006348}, abstract = {The stability of an evaporating thin liquid film on a solid substrate is investigated within lubrication theory. The heat flux due to evaporation induces thermal gradients; the generated Marangoni stresses are accounted for. Assuming the gas phase at rest, the dynamics of the vapour reduces to diffusion. The boundary condition at the interface couples transfer from the liquid to its vapour and diffusion flux. The evolution of the film is governed by a lubrication equation coupled with the Laplace problem associated with quasi-static diffusion. The linear stability of a flat film is studied in this general framework. The subsequent analysis is restricted to diffusion-limited evaporation for which the gas phase is saturated in vapour in the vicinity of the interface. The stability depends then only on two control parameters, the capillary and Marangoni numbers. The Marangoni effect is destabilizing whereas capillarity and evaporation are stabilizing processes. The results of the linear stability analysis are compared with the experiments of Poulard et al. (2003) performed in a different geometry. In order to study the resulting patterns, an amplitude equation is obtained through a systematic multiple-scale expansion. The evaporation rate is needed and is computed perturbatively by solving the Laplace problem for the diffusion of vapour. The bifurcation from the flat state is found to be a supercritical transition. Moreover, it appears that the non-local nature of the diffusion problem affects the amplitude equation unusually.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\8MM2WW2Q\\Sultan et al. - 2005 - Evaporation of a thin film diffusion of the vapou.pdf} @@ -3994,7 +3855,6 @@ @article{sumanDynamicsThinLiquid2006 volume = {304}, number = {1}, pages = {208--213}, - issn = {0021-9797}, doi = {10.1016/j.jcis.2006.08.061}, abstract = {The dynamics of thin liquid films on surfaces whose wettability changes in a time-periodic manner are examined in this work. A nonlinear evolution equation based on the lubrication approximation is used to describe the film height, and attractions due to van der Waals forces are incorporated. Film wettability is varied through an imposed sinusoidal modulation of the Hamaker constant. A linear stability analysis predicts that if the mean Hamaker constant is negative, disturbances at the film surface will eventually decay regardless of the amplitude and frequency of the oscillation. However, numerical solution of the evolution equation shows that the film can rupture at a given frequency if the amplitude is sufficiently large. The associated characteristic wavelength can be predicted from results for constant-wettability surfaces if an appropriate effective Hamaker constant is used. For positive mean Hamaker constants, film rupture can be accelerated, delayed, or prevented depending on how the Hamaker constant changes early in the oscillation cycle. The effects of spatial gradients in wettability are also considered, and it is found that oscillation can delay but not prevent rupture. Inclusion of short-range repulsive forces leads to the formation of droplet-like structures separated by ultra-thin films, but this can be prevented by sufficiently large and slow oscillations of the Hamaker constant. The results of this work may find use in applications that make use of surfaces whose wettability can be controlled by external stimuli.}, keywords = {Hamaker constant,Stability,Thin liquid films,Time-periodic,Wettability}, @@ -4036,7 +3896,6 @@ @article{suzukiSlidingBehaviorWater2008 volume = {254}, number = {6}, pages = {1797--1805}, - issn = {0169-4332}, doi = {10.1016/j.apsusc.2007.07.171}, abstract = {We prepared line-patterned hydrophobic surfaces using fluoroalkylsilane (FAS) and octadecyltrimethoxysilane (ODS) then investigated the effect of line direction on sliding behavior of water droplets by direct observation of the actual droplet motion during sliding. Water droplets slide down with a periodic large deformation of the contact line and sliding velocity fluctuation that occurred when they crossed over the 500-{$\mu$}m ODS line regions in FAS regions on a Si surface tilted at 35\textdegree. These behaviors are less marked for motion on a 100-{$\mu$}m line surface, or on lines oriented parallel to the slope direction. Smaller droplets slide down with greater displacement in the line direction on 500-{$\mu$}m line patterning when the lines were rotated at 13\textdegree{} in-plane for the slope direction. This sliding behavior depended on the droplet size and rotation angle, and is accountable by the balance between gravitational and retentive forces.}, keywords = {Coatings,Hydrophobicity,Photolithography,Silane,Surface patterning,Wetting}, @@ -4079,7 +3938,6 @@ @article{tannerSpreadingSiliconeOil1979 volume = {12}, number = {9}, pages = {1473}, - issn = {0022-3727}, doi = {10.1088/0022-3727/12/9/009}, abstract = {A study of the spreading of silicone oil drops is presented, with particular attention to the dominant effect of the conditions near the edge and the surface tension forces acting there. It is shown that the drop profiles can be predicted and that this leads to a relation between the edge velocity and the maximum slope, which occurs at an inflection point close to the edge. This relation leads to others governing the variation of drop size and slope with time. The predictions are compared with experiment, using silicone oils, and good agreement obtained. The lubrication approximation theory gives no solution for thicknesses below a value varying from 2 to 9*10-8 m at the drop edge. At thicknesses of this order, or possibly even larger, other effects become important, and some observations on edge behaviour are included.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\YP9C358L\\Tanner - 1979 - The spreading of silicone oil drops on horizontal .pdf} @@ -4107,7 +3965,6 @@ @article{tarazonaSimpleDensityFunctional1984 number = {4}, pages = {847--857}, publisher = {{Taylor \& Francis}}, - issn = {0026-8976}, doi = {10.1080/00268978400101601}, abstract = {A simple free energy functional, which incorporates both `local' thermodynamics and short ranged correlations, is formulated and applied to the calculation of the density profile of fluids near hard walls. For hard sphere fluids the calculated profiles are in reasonable agreement with Monte Carlo results. For a Lennard-Jones liquid the profiles exhibit the phenomenon of wetting by gas; the oscillations in the density profiles become much less pronounced and a layer of gas develops near the wall as the bulk density approaches its value at coexistence. Such behaviour was found earlier in Monte Carlo simulations but is not accounted for by existing integral equation theories based on closures of the wall-particle Ornstein-Zernike equation.}, annotation = {\_eprint: https://doi.org/10.1080/00268978400101601}, @@ -4124,7 +3981,6 @@ @article{teletzkeHowLiquidsSpread1987 number = {1-6}, pages = {41--82}, publisher = {{Taylor \& Francis}}, - issn = {0098-6445}, doi = {10.1080/00986448708911919}, abstract = {A theory of the wetting of solids by liquids is put forward. The theory accounts for capillary pressure gradient, gravitational potential gradient, surface tension gradient, disjoining pressure gradient driving forces of flow in thick thin-films and of surface diffusion in thin thin-films. Disjoining pressure stems from the way intermolecular forces aggregate in submicroscopically thin films. For thick thin-films of slowly varying thickness the lubrication approximation to velocity distributions is appropriate. With this approximation the spontaneous, unsteady, two-dimensional spreading of liquid is shown to be governed by a nonlinear convective-diffusion equation for the evolution of the film thickness profile. The predictions of the theory agree with Marmur and Lelah's (1980, 1981) observations of water drops spreading on glass and with Bascom, Cottington and Singleterry's (1964) and Ludviksson and Lightfoot's (1971) observations of oils spreading on high energy surfaces. The theory is used to analyze Derjaguin and co-workers' (1944, 1957, 1970) blowing-off experiments designed to measure thin-film rheology. The theory is also used to buttress the proposition that much contact angle hysteresis is due simply to slow attainment of equilibrium.}, keywords = {Capillary pressure,Gravitational potential,Surface diffusion,Surface tension}, @@ -4140,7 +3996,6 @@ @article{thampiIsotropicDiscreteLaplacian2013 journal = {Journal of Computational Physics}, volume = {234}, pages = {1--7}, - issn = {0021-9991}, doi = {10.1016/j.jcp.2012.07.037}, abstract = {We show that discrete schemes developed for lattice hydrodynamics provide an elegant and physically transparent way of deriving Laplacians with isotropic discretisation error. Isotropy is guaranteed whenever the Laplacian weights follow from the discrete Maxwell\textendash Boltzmann equilibrium since these are, by construction, isotropic on the lattice. We also point out that stencils using as few as 15 points in three dimensions, generate isotropic Laplacians. These computationally efficient Laplacians can be used in cell-dynamical and hybrid lattice Boltzmann simulations, in favor of popular anisotropic Laplacians, which make use of larger stencils. The method can be extended to provide discretisations of higher order and for other differential operators, such the gradient, divergence and curl.}, keywords = {Isotropic Laplacians,Lattice hydrodynamics}, @@ -4173,7 +4028,6 @@ @article{thieleGradientDynamicsModel2020 volume = {229}, number = {10}, pages = {1819--1832}, - issn = {1951-6401}, doi = {10.1140/epjst/e2020-900231-2}, abstract = {When a liquid drop spreads on an adaptive substrate the latter changes its properties what may result in an intricate coupled dynamics of drop and substrate. Here we present a generic mesoscale hydrodynamic model for such processes that is written as a gradient dynamics on an underlying energy functional. We specify the model details for the example of a drop spreading on a dry polymer brush. There, liquid absorption into the brush results in swelling of the brush causing changes in the brush topography and wettability. The liquid may also advance within the brush via diffusion (or wicking) resulting in coupled drop and brush dynamics. The specific model accounts for coupled spreading, absorption and wicking dynamics when the underlying energy functional incorporates capillarity, wettability and brush energy. After employing a simple version of such a model to numerically simulate a droplet spreading on a swelling brush we conclude with a discussion of possible model extensions.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\CVR3C6UI\\Thiele and Hartmann - 2020 - Gradient dynamics model for drops spreading on pol.pdf} @@ -4188,7 +4042,6 @@ @article{thielePatternedDepositionMoving2014 series = {Manuel {{G}}. {{Velarde}}}, volume = {206}, pages = {399--413}, - issn = {0001-8686}, doi = {10.1016/j.cis.2013.11.002}, abstract = {When a simple or complex liquid recedes from a smooth solid substrate it often leaves a homogeneous or structured deposit behind. In the case of a receding non-volatile pure liquid the deposit might be a liquid film or an arrangement of droplets depending on the receding speed of the meniscus and the wetting properties of the system. For complex liquids with volatile components as, e.g., polymer solutions and particle or surfactant suspensions, the deposit might be a homogeneous or structured layer of solute \textemdash{} with structures ranging from line patterns that can be orthogonal or parallel to the receding contact line via hexagonal or square arrangements of drops to complicated hierarchical structures. We review a number of recent experiments and modelling approaches with a particular focus on mesoscopic hydrodynamic long-wave models. The conclusion highlights open question and speculates about future developments.}, keywords = {Complex fluids,Depinning,Evaporation,Moving contact line,Patterned deposition,Stick-slip motion}, @@ -4203,7 +4056,6 @@ @article{thommesLatticeBoltzmannMethods2007 volume = {55}, number = {7}, pages = {673--692}, - issn = {1097-0363}, doi = {10.1002/fld.1489}, abstract = {We apply the lattice Boltzmann (LB) method for solving the shallow water equations with source terms such as the bed slope and bed friction. Our aim is to use a simple and accurate representation of the source terms in order to simulate practical shallow water flows without relying on upwind discretization or Riemann problem solvers. We validate the algorithm on problems where analytical solutions are available. The numerical results are in good agreement with analytical solutions. Furthermore, we test the method on a practical problem by simulating mean flow in the Strait of Gibraltar. The main focus is to examine the performance of the LB method for complex geometries with irregular bathymetry. The results demonstrate its ability to capture the main flow features. Copyright \textcopyright{} 2007 John Wiley \& Sons, Ltd.}, keywords = {lattice Boltzmann method,shallow water equations,Strait of Gibraltar}, @@ -4220,7 +4072,6 @@ @article{tsekovEffectThermalFluctuations1993 volume = {9}, number = {11}, pages = {3264--3269}, - issn = {0743-7463, 1520-5827}, doi = {10.1021/la00035a082}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\EMRVNZSH\\Tsekov and Ruckenstein - 1993 - Effect of thermal fluctuations on the stability of.pdf} } @@ -4234,7 +4085,6 @@ @article{tubbsMultilayerShallowWater2009 volume = {32}, number = {12}, pages = {1767--1776}, - issn = {0309-1708}, doi = {10.1016/j.advwatres.2009.09.008}, abstract = {A multilayer lattice Boltzmann (LB) model is introduced to solve three-dimensional wind-driven shallow water flow problems. The multilayer LB model avoids the expensive Navier\textendash Stokes equations and obtains stratified horizontal flow velocities as vertical velocities are relatively small and the flow is still within the shallow water regime. A single relaxation time BGK method is used to solve each layer coupled by the vertical viscosity forcing term. To increase solution stability, an implicit step is suggested to obtain flow velocities. The main advantage of using the LBM is that after selecting appropriate equilibrium distribution functions, the LB algorithm is only slightly modified for each layer and retains all the simplicities of the LBM within the high performance computing (HPC) environment. The performance of the parallel LB model for the multilayer shallow water equations is investigated on CPU-based HPC environments using OpenMP. We found that the explicit loop control with cache optimization in LBM gives better performance on execution time, speedup and efficiency than the implicit loop control as the number of processors increases. Numerical examples are presented to verify the multilayer LB model against analytical solutions. We demonstrate the model's capability of calculating lateral and vertical distributions of velocities for wind-driven circulation over non-uniform bathymetry.}, keywords = {BGK,High performance computing,Lattice Boltzmann,Three-dimensional shallow water equations,Wind-driven circulation}, @@ -4252,7 +4102,6 @@ @article{utadaDrippingJettingDrops2007 number = {9}, pages = {702--708}, publisher = {{Cambridge University Press}}, - issn = {1938-1425, 0883-7694}, doi = {10.1557/mrs2007.145}, abstract = {The following article is based on the Symposium X presentation given by David A. Weitz (Harvard University) on April 11, 2007, at the Materials Research Society Spring Meeting in San Francisco. The article describes how simple microfluidic devices can be used to control fluid flow and produce a variety of new materials. Based on the concepts of coaxial flow and hydrodynamically focused flow, used alone or in various combinations, the devices can produce precisely controlled double emulsions (droplets within droplets) and even triple emulsions (double emulsions suspended in a third droplet). These structures, which can be created in a single microfluidic device, have various applications such as encapsulants for drugs, cosmetics, or food additives.} } @@ -4275,7 +4124,6 @@ @article{vanossInterfacialLifshitzvanWaals1988 volume = {88}, number = {6}, pages = {927--941}, - issn = {0009-2665, 1520-6890}, doi = {10.1021/cr00088a006}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\6ZSFQBAD\\Van Oss et al. - 1988 - Interfacial Lifshitz-van der Waals and polar inter.pdf} } @@ -4305,7 +4153,6 @@ @article{vanthangStudy1DLattice2010 volume = {229}, number = {19}, pages = {7373--7400}, - issn = {0021-9991}, doi = {10.1016/j.jcp.2010.06.022}, abstract = {The D1Q3 lattice Boltzmann (LB) shallow water equation is analyzed in detail and compared with other numerical schemes. Analytical results are derived and used to discuss the accuracy and stability of the model. We show how such D1Q3 LB models for canal reaches may be easily coupled with various hydraulic interconnection structures to build models of complex irrigation networks.}, keywords = {1D numerical schemes,Lattice Boltzmann model,Shallow water equations}, @@ -4329,7 +4176,6 @@ @article{vieyrasalasActiveControlEvaporative2012 number = {22}, pages = {12038--12047}, publisher = {{American Chemical Society}}, - issn = {1932-7447}, doi = {10.1021/jp301092y}, abstract = {We designed and built a highly adaptive and flexible system based on modulated infrared irradiation for the active control of evaporative material deposition suitable for solution processing of organic electronic materials. We performed systematic experiments using thick layers of a high molecular weight aqueous dispersion of poly(3,4-ethylene-dioxythiophene):poly(styrenesulfonate) (PEDOT:PSS) nanoparticles as well as thin layers of light-emitting polymers in an organic solvent. We identified two complementary deposition modes, where material accumulates either in the illuminated or the nonirradiated regions, depending on the molecular weight of the solute and the solution layer thickness. We developed numerical models that account for heat transfer due to infrared illumination, solvent evaporation into the gas phase, and solute redistribution in the liquid layer. The computational results agree well with the experimental observations regarding the solute depletion and accumulation behavior.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\L569QG8H\\Vieyra Salas et al. - 2012 - Active Control of Evaporative Solution Deposition .pdf;C\:\\Users\\Zitzero\\Zotero\\storage\\BSEWAMCS\\jp301092y.html} @@ -4344,7 +4190,6 @@ @article{vrijRuptureThinLiquid1968 volume = {90}, number = {12}, pages = {3074--3078}, - issn = {0002-7863, 1520-5126}, doi = {10.1021/ja01014a015}, abstract = {The surfaces of thin liquid films are slightly corrugated because of thermal fluctuations. These corrugations show up in the light scattered by these films. It can be shown that corrugations having wavelengths larger than a critical wavelength will grow spontaneously, because of the action of van der Waals forces, and will cause the film to become rapidly thinner and break (or be stabilized as a black film, if sufficiently strong repulsive forces keep the two faces of the film separated). The critical wavelength Acrit = [ \textemdash{} l\^yJfrGjdhff-, in which y is the surface tension, G the Gibbs energy of interaction among the molecules per unit area of the film, and h the thickness of the film. The rate at which fluctuations above the critical wavelength grow depends on the viscosity of the liquid, and it can be shown that the combination of regular drainage and spontaneous growth of fluctuations leads to a critical thickness and a lifetime for the film of the same order as those found in experiments.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\A6Y8SZJR\\Vrij and Overbeek - 1968 - Rupture of thin liquid films due to spontaneous fl.pdf} @@ -4359,7 +4204,6 @@ @article{wangPhotoresponsiveSurfacesControllable2007 volume = {8}, number = {1}, pages = {18--29}, - issn = {1389-5567}, doi = {10.1016/j.jphotochemrev.2007.03.001}, abstract = {In this paper, current progress in the area of photoresponsive surfaces with controllable wettability is reviewed, including mainly surface conversion between wetting and anti-wetting, prepared from inorganic oxides (e.g., titanium dioxide, zinc oxide, and tungsten oxide) or/and photoactive organic molecules (e.g., azobenzene, and spiropyran), and movement of liquid droplets driven by molecular machines (e.g., molecular shuttles such as rotaxanes). Photoresponsive controllable wettability originates from a transition between the bistable states of photoresponsive materials. The exploration of the basic mechanisms provides a basis for the construction of novel smart responsive surfaces.}, keywords = {Controllable wettability,Hydrophilicity,Hydrophobicity,Liquid motion,Photoresponsive,Superhydrophilicity,Superhydrophobicity,Surface switch}, @@ -4377,7 +4221,6 @@ @article{wattsBurnoutUniversityTeaching2011 number = {1}, pages = {33--50}, publisher = {{Routledge}}, - issn = {0013-1881}, doi = {10.1080/00131881.2011.552235}, abstract = {Background: Teacher stress potentially impairs personal and professional competence and compromises productivity. Aversive emotional experience has been most comprehensively encapsulated by the phenomenon of burnout, which is particularly prominent for staff in human service sectors. Burnout reactions have been characterised as tripartite: the depletion of emotional reserves (emotional exhaustion), an increasingly cynical and negative approach towards others (depersonalisation) and a growing feeling of work-related dissatisfaction (personal accomplishment). Purpose: Few studies have investigated the emotional consequences of teachers' stress and even fewer have specifically focused on university educators. A systematic literature review was thus conducted to evaluate the extent of burnout for university teaching staff and specifically to reveal predictive variables, which may explain this experience in this understudied occupational group. Design and methods: Six databases including Educational Resources Information Centre (ERIC), PsychINFO and Scopus were searched using the terms burnout, university, academics, teaching staff, lecturers, research staff and faculty. Papers were limited to English language peer-reviewed empirical investigations ofburnout in full-time university teaching staff. Papers not adopting a clear operationalisation of burnout were rejected. Twelve papers met the criteria and were included in the review. A detailed data extraction form was used to reveal relevant information from each paper. Conclusions: The review revealed that staff exposure to high numbers of students, especially tuition of postgraduates, strongly predicts the experience of burnout. Other predictive variables included gender, with higher depersonalisation scores found in male teachers and female teachers typically scoring higher onthe emotional exhaustion dimension. Age also demonstrated an association, with younger staff appearing more vulnerable to emotional exhaustion. Burnout in university teachers was comparable with other service sector employees such asschoolteachers and healthcare professionals. The current review reveals a scarcity of comparative studies across different university contexts, therefore multi-site studies are required in order to control for the potential influence ofmoderating variables such as institution age when measuring burnout in university teachers.}, keywords = {burnout,postgraduates,students,support,teachers,university}, @@ -4405,7 +4248,6 @@ @article{wedershovenInfraredLaserInduced2014 number = {5}, pages = {054101}, publisher = {{American Institute of Physics}}, - issn = {0003-6951}, doi = {10.1063/1.4863318}, abstract = {We studied the deformation and destabilization of thin liquid films on stationary substrates via infrared illumination. The film thickness evolution was measured using interference microscopy. We developed numerical models for the temperature evolution and the liquid redistribution. The substrate wettability is explicitly accounted for via a phenomenological expression for the disjoining pressure. We systematically measured the film thinning- and rupture dynamics as a function of laser power, which are accurately reproduced by the simulations. While smaller laser spots generally lead to shorter rupture times, the latter can become independent of the spotsize for very narrow beams due to capillary suppression.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\LMRYFB5H\\Wedershoven et al. - 2014 - Infrared laser induced rupture of thin liquid film.pdf} @@ -4421,7 +4263,6 @@ @article{weimarNonlinearReactionsAdvected1996 volume = {224}, number = {1}, pages = {207--215}, - issn = {0378-4371}, doi = {10.1016/0378-4371(95)00355-X}, abstract = {Flows can modify the effects of nonlinear reactions. Here we use a lattice Boltzmann method to simulate an athermal flow which advects reactant species. As an example we consider a reactive system modeled by the Brusselator which exhibits Turing structures; the system is subjected to mixing as induced by the wakes behind obstacles. The resulting effect is a modification of the reaction patterns which can be as dramatic as their disappearance.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\K4XAXC7J\\Weimar and Pierre Boon - 1996 - Nonlinear reactions advected by a flow.pdf;C\:\\Users\\Zitzero\\Zotero\\storage\\QP7HY4VR\\037843719500355X.html} @@ -4453,7 +4294,6 @@ @article{bertrandDynamicsDewettingNanoscale2007 number = {7}, pages = {3774--3785}, publisher = {{American Chemical Society}}, - issn = {0743-7463}, doi = {10.1021/la062920m}, urldate = {2023-07-27}, abstract = {Large-scale molecular dynamics simulations are used to model the dewetting of solid surfaces by partially wetting thin liquid films. Two levels of solid-liquid interaction are considered that give rise to large equilibrium contact angles. The initial length and thickness of the films are varied over a wide range at the nanoscale. Spontaneous dewetting is initiated by removing a band of molecules either from each end of the film or from its center. As observed experimentally and in previous simulations, the films recede at an initially constant speed, creating a growing rim of liquid with a constant receding dynamic contact angle. Consistent with the current understanding of wetting dynamics, film recession is faster on the more poorly wetted surface to an extent that cannot be explained solely by the increase in the surface tension driving force. In addition, the rates of recession of the thinnest films are found to increase with decreasing film thickness. These new results imply not only that the mobility of the liquid molecules adjacent to the solid increases with decreasing solid-liquid interactions, but also that the mobility adjacent to the free surface of the film is higher than in the bulk, so that the effective viscosity of the film decreases with thickness.}, @@ -4470,7 +4310,6 @@ @article{koplikPearlingInstabilityNanoscale2006 volume = {18}, number = {3}, pages = {032104}, - issn = {1070-6631}, doi = {10.1063/1.2178786}, urldate = {2023-07-27}, abstract = {We investigate the flow of a nanoscale incompressible ridge of low-volatility liquid along a ``chemical channel'': a long, straight, and completely wetting stripe embedded in a planar substrate, and sandwiched between two extended less wetting solid regions. Molecular dynamics simulations, a simple long-wavelength approximation, and a full stability analysis based on the Stokes equations are used, and give qualitatively consistent results. While thin liquid ridges are stable both statically and during flow, a (linear) pearling instability develops if the thickness of the ridge exceeds half of the width of the channel. In the flowing case, periodic bulges propagate along the channel and subsequently merge due to nonlinear effects. However, the ridge does not break up even when the flow is unstable, and the qualitative behavior is unchanged even when the fluid can spill over onto a partially wetting exterior solid region.}, @@ -4488,7 +4327,6 @@ @article{wengMolecularDynamicsInvestigation2000 number = {14}, pages = {5917--5923}, publisher = {{American Institute of Physics}}, - issn = {0021-9606}, doi = {10.1063/1.1290698}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\RZ33ZFE2\\Weng et al. - 2000 - Molecular dynamics investigation of thickness effe.pdf} } @@ -4502,7 +4340,6 @@ @article{whymanRigorousDerivationYoung2008 volume = {450}, number = {4}, pages = {355--359}, - issn = {0009-2614}, doi = {10.1016/j.cplett.2007.11.033}, abstract = {The rigorous derivation of Young, Cassie\textendash Baxter and Wenzel equations carried out in the framework of the unified thermodynamic approach is presented. Wetting of rough surfaces controlled with external stimuli is treated. Areas of validity of Cassie\textendash Baxter and Wenzel approaches are discussed. General properties of the contact angle hysteresis are investigated on the same thermodynamic basis.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\BMRLHPAY\\Whyman et al. - 2008 - The rigorous derivation of Young, Cassie–Baxter an.pdf;C\:\\Users\\Zitzero\\Zotero\\storage\\S6A6KULM\\S000926140701545X.html} @@ -4517,7 +4354,6 @@ @article{wijshoffDynamicsPiezoInkjet2010 volume = {491}, number = {4}, pages = {77--177}, - issn = {0370-1573}, doi = {10.1016/j.physrep.2010.03.003}, abstract = {The operation of a piezo inkjet printhead involves a chain of processes in many physical domains at different length scales. The final goal is the formation of droplets of all kinds of fluids with any desired volume, velocity, and a reliability as high as possible. The physics behind the chain of processes comprise the two-way coupling from the electrical to the mechanical domain through the piezoelectric actuator, where an electrical signal is transformed into a mechanical deformation of the printhead structure. The next two steps are the coupling to the acoustic domain inside the ink channels, and the coupling to the fluid dynamic domain, i.e.~the drop formation process. The dynamics of the printhead structure are coupled via the acoustics to the drop formation process in the nozzle. Furthermore, wetting of the nozzle plate and air bubbles can have a negative influence on the printhead performance. The five topics (actuation, channel acoustics, drop formation, wetting, and air bubbles) are reviewed in this paper. This research connects the product developments for many emerging new industrial applications of the inkjet technology to the fundamental physical phenomena underlying the printhead operation.}, keywords = {Acoustics,Bubble dynamics,Drop formation,Inkjet,Piezoelectricity,Wetting}, @@ -4549,7 +4385,6 @@ @article{williamsNonlinearTheoryFilm1982 volume = {90}, number = {1}, pages = {220--228}, - issn = {0021-9797}, doi = {10.1016/0021-9797(82)90415-5}, abstract = {The present work aims at examining nonlinear effects on film rupture by investigating the stability of thin films to finite amplitude disturbances. The dynamics of the liquid film is formulated using the Navier-Stokes equations augmented by a body force describing the London/van der Waals attractions. The liquid film is assumed to be charge neutralized, nondraining, and laterally unbounded. A nonlinear evolution equation is derived for h(x, t), the film thickness. This strongly nonlinear partial differential equation is solved by numerical methods as part of an initial-value problem for periodic boundary conditions in x, the lateral space dimension. Given this model, one obtains true rupture in the sense that the film thickness becomes zero in a finite time. The results reveal rupture characteristics and effects of nonlinearities on the rupture properties.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\B26JYAHD\\0021979782904155.html} @@ -4580,7 +4415,6 @@ @article{williamsonStandardTestSet1992 volume = {102}, number = {1}, pages = {211--224}, - issn = {0021-9991}, doi = {10.1016/S0021-9991(05)80016-6}, abstract = {A suite of seven test cases is proposed for the evaluation of numerical methods intended for the solution of the shallow water equations in spherical geometry. The shallow water equations exhibit the major difficulties associated with the horizontal dynamical aspects of atmospheric modeling on the spherical earth. These cases are designed for use in the evaluation of numerical methods proposed for climate modeling and to identify the potential trade-offs which must always be made in numerical modeling. Before a proposed scheme is applied to a full baroclinic atmospheric model it must perform well on these problems in comparison with other currently accepted numerical methods. The cases are presented in order of complexity. They consist of advection across the poles, steady state geostrophically balanced flow of both global and local scales, forced nonlinear advection of an isolated low, zonal flow impinging on an isolated mountain, Rossby-Haurwitz waves, and observed atmospheric states. One of the cases is also identified as a computer performance/algorithm efficiency benchmark for assessing the performance of algorithms adapted to massively parallel computers.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\AZPAZSQ7\\S0021999105800166.html} @@ -4628,7 +4462,6 @@ @article{wongMicrodropletContaminantsWhen2020 number = {4}, pages = {3836--3846}, publisher = {{American Chemical Society}}, - issn = {1936-0851}, doi = {10.1021/acsnano.9b08211}, abstract = {Superamphiphobic surfaces are commonly associated with superior anticontamination and antifouling properties. Visually, this is justified by their ability to easily shed off drops and contaminants. However, on micropillar arrays, tiny droplets are known to remain on pillars' top faces while the drop advances. This raises the question of whether remnants remain even on nanostructured superamphiphobic surfaces. Are superamphiphobic surfaces really self-cleaning? Here we investigate the presence of microdroplet contaminants on three nanostructured superamphiphobic surfaces. After brief contact with liquids having different volatilities and surface tension (water, ethylene glycol, hexadecane, and an ionic liquid), confocal microscopy reveals a ``blanket-like'' layer of microdroplets remaining on the surface. It appears that the phenomenon is universal. Notably, when placing subsequent drops onto the contaminated surface, they are still able to roll off. However, adhesion forces can gradually increase by up to 3 times after repeated liquid drop contact. Therefore, we conclude that superamphiphobic surfaces do not warrant self-cleaning and anticontamination capabilities at sub-micrometric length scales.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\ERHQD93W\\Wong et al. - 2020 - Microdroplet Contaminants When and Why Superamphi.pdf;C\:\\Users\\Zitzero\\Zotero\\storage\\KFHC6GYA\\acsnano.html} @@ -4684,7 +4517,6 @@ @article{yakhotRenormalizationGroupAnalysis1986 volume = {1}, number = {1}, pages = {3--51}, - issn = {1573-7691}, doi = {10.1007/BF01061452}, abstract = {We develop the dynamic renormalization group (RNG) method for hydrodynamic turbulence. This procedure, which uses dynamic scaling and invariance together with iterated perturbation methods, allows us to evaluate transport coefficients and transport equations for the large-scale (slow) modes. The RNG theory, which does not include any experimentally adjustable parameters, gives the following numerical values for important constants of turbulent flows: Kolmogorov constant for the inertial-range spectrumCK=1.617; turbulent Prandtl number for high-Reynolds-number heat transferPt=0.7179; Batchelor constantBa=1.161; and skewness factor\textasciimacron S3=0.4878. A differentialK-\$\$\textbackslash bar \textbackslash varepsilon \$\$model is derived, which, in the high-Reynolds-number regions of the flow, gives the algebraic relationv=0.0837 K2/\$\$\textbackslash bar \textbackslash varepsilon \$\$, decay of isotropic turbulence asK=O(t-1.3307), and the von Karman constant{$\kappa$}=0.372. A differential transport model, based on differential relations betweenK,\$\$\textbackslash bar \textbackslash varepsilon \$\$, and{$\nu$}, is derived that is not divergent whenK\textrightarrow{} 0 and\$\$\textbackslash bar \textbackslash varepsilon \$\$is finite. This latter model is particularly useful near walls.}, keywords = {computational fluid dynamics,inertial range,large-eddy simulation,Renormalization group,Reynolds number,turbulence theory,turbulence transport}, @@ -4702,7 +4534,6 @@ @article{yongCononsolvencyTransitionPolymer2018 number = {6}, pages = {991}, publisher = {{Multidisciplinary Digital Publishing Institute}}, - issn = {1996-1944}, doi = {10.3390/ma11060991}, abstract = {In this study, the cononsolvency transition of poly(N-isopropylacrylamide) (PNiPAAm) brushes in aqueous ethanol mixtures was studied by using Vis-spectroscopic ellipsometry (SE) discussed in conjunction with the adsorption-attraction model. We proved that the cononsolvency transition of PNiPAAm brushes showed features of a volume phase transition, such as a sharp collapse, reaching a maximum decrease in thickness for a very narrow ethanol volume composition range of 15\% to 17\%. These observations are in agreement with the recently published preferential adsorption model of the cononsolvency effect.}, copyright = {http://creativecommons.org/licenses/by/3.0/}, @@ -4735,7 +4566,6 @@ @article{zhangInkjetPrintingDirect2015 number = {6}, pages = {2844--2852}, publisher = {{The Royal Society of Chemistry}}, - issn = {2050-7496}, doi = {10.1039/C4TA05862C}, abstract = {The preparation of biomimetic superhydrophobic surfaces with hydrophilic micro-sized patterns is highly desirable, but a one-step, mask-free method to produce such surfaces has not previously been reported. We have developed a direct method to produce superhydrophilic micropatterns on superhydrophobic surfaces based on inkjet printing technology. This work was inspired by the efficient fog-harvesting behavior of Stenocara beetles in the Namib Desert. A mussel-inspired ink consisting of an optimized solution of dopamine was applied directly by inkjet printing to superhydrophobic surfaces. Stable Wenzel's microdroplets of the dopamine solution with well-defined micropatterns were obtained on these surfaces. Superhydrophilic micropatterns with well-controlled dimensions were then readily achieved on the superhydrophobic surfaces by the formation of polydopamine via in situ polymerization. The micropatterned superhydrophobic surfaces prepared by this inkjet printing method showed enhanced water collection efficiency compared with uniform superhydrophilic and superhydrophobic surfaces. This method can be used for the facile large-scale patterning of superhydrophobic surfaces with high precision and superior pattern stability and is therefore a key step toward patterning superhydrophobic surfaces for practical applications.}, file = {C\:\\Users\\Zitzero\\Zotero\\storage\\CMI75I3E\\Zhang et al. - 2015 - Inkjet printing for direct micropatterning of a su.pdf} @@ -4781,7 +4611,6 @@ @article{zhangNonNewtonianEffectsLubricant2005 volume = {51}, number = {1}, pages = {1--13}, - issn = {1573-2703}, doi = {10.1007/s10665-004-1342-z}, abstract = {An analysis of non-Newtonian effects on lubrication flows is presented based on the upper-convected Maxwell constitutive equation, which is the simplest viscoelastic model having a constant viscosity and relaxation time. By employing characteristic lubricant relaxation times in all order of magnitude analysis, a perturbation method is developed to analyze the flow of a non-Newtonian lubricant between two surfaces. The effect of viscoelasticity on the lubricant velocity and pressure fields is examined, and the influence of minimum film thickness on lubrication characteristics is investigated. Numerical simulations show a significant enhancement in the pressure field when the minimum film thickness is sufficiently small. This mechanism suggests that viscoelasticity does indeed produce a beneficial effect on lubrication performance, which is consistent with experimental observations.}, keywords = {lubrication,non-Newtonian fluid,perturbation method,thin-film viscoelacity}, @@ -4842,7 +4671,6 @@ @article{zitzSwalbeJlLattice2022 volume = {7}, number = {77}, pages = {4312}, - issn = {2475-9066}, doi = {10.21105/joss.04312}, abstract = {Zitz et al., (2022). Swalbe.jl: A lattice Boltzmann solver for thin film hydrodynamics. Journal of Open Source Software, 7(77), 4312, https://doi.org/10.21105/joss.04312}, copyright = {Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License (CC-BY-NC-ND)}, diff --git a/chapters/Introduction.tex b/chapters/Introduction.tex index 4bee282..f3f3418 100644 --- a/chapters/Introduction.tex +++ b/chapters/Introduction.tex @@ -437,7 +437,7 @@ \section{Outline} Use cases and a derivation of the model can be found in Chap.~\ref{chapter:first_paper}. In this chapter the mandatory modelling assumptions are introduced that allow to match the shallow water system with the thin film equation. -It further highlights which modifications are made to the shallow water lattice Boltzmann algortihm. +It further highlights which modifications are made to the shallow water lattice Boltzmann algorithm. With some emphasis on the numerical implementation of e.g., the computation of gradients and the Laplacian in agreement with Ref~\cite{junkDiscretizationsIncompressibleNavier2000, thampiIsotropicDiscreteLaplacian2013}. After the derivation, numerical experiments are displayed to validate the model. Among those are the Rayleigh-Taylor instability, an instability that occurs when a heavy fluid is on top of a light fluid. diff --git a/chapters/Paper_one.tex b/chapters/Paper_one.tex index a5cd77a..07237a0 100644 --- a/chapters/Paper_one.tex +++ b/chapters/Paper_one.tex @@ -234,7 +234,7 @@ \subsection{The Rayleigh-Taylor instability} \begin{figure} \centering - \includegraphics[width=0.75\textwidth]{graphics/Fig_3_new_RTI_spectra_single_color_same_tau_capkc_correct_rescaled_x_axis_inset.png} + \includegraphics[width=0.65\textwidth]{graphics/Fig_3_new_RTI_spectra_single_color_same_tau_capkc_correct_rescaled_x_axis_inset.png} \caption{Power spectrum of the height fluctuations versus wave number. The different colors and symbols belong to different values of graviational acceleration; $g=4\cdot 10^{-5}$ is given by blue circles (\textcolor{pyblue}{$\bullet$}), $g=6\cdot 10^{-5}$ by orange triangles (\textcolor{pyorange}{$\blacktriangle$}) and $g=8\cdot 10^{-5}$ is given by green squares (\textcolor{pygreen}{$\blacksquare$}). Same colored lines are taken at different time steps. In the inset we show the growth rate $\sigma(k)$ for the largest value of $g$ (symbols) and the theoretical growth rate according to Eq.~(\ref{eq:RTgrowth}) (solid line).} \label{fig:RTI} \end{figure} @@ -339,7 +339,7 @@ \subsection{A spreading droplet} \begin{figure} \centering - \includegraphics[width=0.75\textwidth]{graphics/Fig_5_Indirect_Cox_Voinov_all_data_visually_appealing_slip_2_m_nosci.png} + \includegraphics[width=0.65\textwidth]{graphics/Fig_5_Indirect_Cox_Voinov_all_data_visually_appealing_slip_2_m_nosci.png} \caption{Difference of cubed instantaneous and equilibrium contact angles, $\theta_{num}^3-\theta_{eq}^3$, vs. capillary number $Ca$ for a spreading droplet; the dashed line shows a linear dependence (consistent with the Cox-Voinov law). The different symbols represent different viscosities, while the dashed line is a linear function of the capillary number.} \label{fig:Cox-Voinov} \end{figure} @@ -349,7 +349,7 @@ \subsection{A spreading droplet} In Fig.~\ref{fig:Cox-Voinov} we plot $\theta^3(t) - \theta_{eq}^3$ vs $Ca(t)$ from a numerical simulation of a spreading drop: A good linear scaling, in agreement with the Cox-Voinov law, is observed, as highlighted by the dashed line. \begin{figure} \centering - \includegraphics[width=0.75\textwidth]{graphics/Fig_6_Tanners_law_slip_2_paper_rescaled_t.png} + \includegraphics[width=0.65\textwidth]{graphics/Fig_6_Tanners_law_slip_2_paper_rescaled_t.png} \caption{Time evolution of the droplet base radius of a spreading droplet; the dashed red line shows a $\tilde{t}^{1/10}$ power law (consistent with Tanner's law). As in Fig.~\ref{fig:Cox-Voinov} different symbols refer to different viscosities. The radius clearly grows with the predicted power law until it saturates. On rescaling the time with $\tau_{\mbox{\tiny{cap}}}$ the curves of all three viscosities collapse into a single one.} @@ -382,7 +382,7 @@ \subsection{A sliding droplet} $Bo_c$ is the critical Bond number, defined in terms of the critical tilting angle $\alpha_c$. \begin{figure} \centering - \includegraphics[width=0.75\textwidth]{graphics/Fig_7_Ca_Bo_true_with_pic.png} + \includegraphics[width=0.65\textwidth]{graphics/Fig_7_Ca_Bo_true_with_pic.png} \caption{$Ca$ vs $Bo$ for a sliding droplet: Notice that a finite minimum forcing (corresponding to $Bo_c$) is needed to actuate the droplet. For $Bo > Bo_c$ a linear relation, $Ca \sim Bo$, is observed. In the insets we show the shape of the droplet for both, the pinned (upper left) as well as the sliding (lower right) case. diff --git a/chapters/Paper_three.tex b/chapters/Paper_three.tex index 8fef472..f477b94 100644 --- a/chapters/Paper_three.tex +++ b/chapters/Paper_three.tex @@ -88,7 +88,7 @@ \section{Results} We first investigate how the rupture times depend on the param where $U_{\Theta}$ is the retraction speed $U_{\Theta} = \frac{\gamma \Theta^3}{9\mu}$~\cite{edwardsNotSpreadingReverse2016}, with $\Theta = \max_{\mathbf{x}}\{\theta(\mathbf{x})\}$. \begin{figure} \centering - \includegraphics[width=0.4\textwidth]{graphics/Figure_2.pdf} + \includegraphics[width=0.65\textwidth]{graphics/Figure_2.pdf} \caption{Rupture times $\tau_r$ as a function of the pattern wavelength $\lambda$, for $v_{\theta}=0$ (\textcolor{jlblue}{$\bullet$}) and $v_{\theta}=20 v_0$ (\textcolor{jlorange}{$\star$}). The continuous and dashed lines indicate the linear, $\sim \lambda$, and quadratic, $\sim \lambda^2$, scaling laws, respectively. } @@ -97,7 +97,7 @@ \section{Results} We first investigate how the rupture times depend on the param We now focus on the long-time dynamics, the characterization of the dewetting morphologies, and how they are affected by the speed of the wettability wave. \begin{figure} \centering - \includegraphics[width=0.4\textwidth]{graphics/Figure_3.pdf} + \includegraphics[width=0.65\textwidth]{graphics/Figure_3.pdf} \caption{Main panel: Time evolution of the height fluctuations $\Delta h(t)$ during the dewetting process on the patterned substrate given by Eq.~(\ref{eq:sinetheta}) with $v_{\theta}= 0$ and $\lambda= 512 h_0$ (\textcolor{jlblue}{$\bullet$}), $\lambda=256 h_0$ (\textcolor{jlorange}{$\blacksquare$}) and $\lambda=170 h_0$ (\textcolor{jlgreen}{$\star$}). Inset: Number of droplets $N(t)$ as a function of time. The three horizontal dashed lines indicate the number of minima of Eq.~(\ref{eq:sinetheta}), which is $2\left(\frac{L}{\lambda}\right)^2$. @@ -113,7 +113,7 @@ \section{Results} We first investigate how the rupture times depend on the param This represents a measure of the mean droplet height $h_d$ (since droplets are essentially monodisperse), decreasing with the pattern wavelength (as expected, due to a decreasing droplet volume, $V_d = \frac{h_0 \lambda^2}{2}$). \begin{figure} \centering - \includegraphics[width=0.4\textwidth]{graphics/Figure_4.pdf} + \includegraphics[width=0.65\textwidth]{graphics/Figure_4.pdf} \caption{Time evolution of the second-order Minkowski structure metric $q_2(t)$ for different $\Gamma$ values, on a substrate with pattern wavelength $\lambda=256 h_0$. The grey-scale insets supply snapshots of the corresponding film thickness fields.} \label{fig:msm_q2} @@ -134,7 +134,7 @@ \section{Results} We first investigate how the rupture times depend on the param Notice, though, that the $q_2$ signal for any $v_{\theta} >0$ always stays above the one for the static case, suggesting that even the smallest pattern velocity introduces a sizeable deformation of the spherical cap shape. \begin{figure} \centering - \includegraphics[width=0.4\textwidth]{graphics/Figure_5.pdf} + \includegraphics[width=0.65\textwidth]{graphics/Figure_5.pdf} \caption{Main panel: Rivulet lifetimes $\tau_{\text{riv}}$ for various $\Gamma$. The dashed line is a guide to the eye to highlight the logarithmic dependence, in agreement with the theoretical prediction, Eq.~(\ref{eq:rivlt}). Inset: Height fluctuations $\Delta h(t)$ vs time, along the rivulet axis, for three different $\Gamma$. @@ -182,13 +182,12 @@ \section{Conclusions} \section{Acknowledgements} We acknowledge financial support from the German Research Foundation (DFG) (priority program SPP2171 / project HA-4382/11 and Project-ID 431791331—CRC1452), and from the Independent Research Fund Denmark (grant 9063-00018B). \newpage -\section{Supplemental material}. -\label{suppmat} +\section{Supplemental material}\label{suppmat} \subsection{Minkowski's structure metric \texorpdfstring{$q_2$}{hmm} in an extended parameter space.} \begin{figure} \centering - \includegraphics[width=0.4\textwidth]{graphics/SupMatFig_1.pdf} - \includegraphics[width=0.4\textwidth]{graphics/SupMatFig_2.pdf} + \includegraphics[width=0.45\textwidth]{graphics/SupMatFig_1.pdf} + \includegraphics[width=0.45\textwidth]{graphics/SupMatFig_2.pdf} \caption{LEFT PANEL. Minkowski's structure metric $q_2$ for three different wavelengths $\lambda=64 h_0$, $\lambda=128 h_0$ and $\lambda=256 h_0$ (as in figure 4 of the main text) and for $\Gamma=1.5$ (all other parameters are as in the main text). The time interval during which $q_2 \approx 1$ signals the emergence of rivulets, also for $\lambda = 64 h_0$ which is comparable with the spinodal wavelength $\lambda_s \approx 70 h_0$. RIGHT PANEL. Comparison of the Minkowski's structure metric $q_2$ for $\delta\theta=5^{\circ}$ and $\delta\theta=10^{\circ}$, $\Gamma = 15$ and $\lambda = 256 h_0$.} diff --git a/chapters/Paper_two.tex b/chapters/Paper_two.tex index 679da17..848b76c 100644 --- a/chapters/Paper_two.tex +++ b/chapters/Paper_two.tex @@ -241,7 +241,7 @@ \section{Results}\label{sec:results_two} \subsection{Testing the stochastic term}\label{subsec:validation} \begin{figure} \centering - \includegraphics[width=0.75\textwidth]{graphics/spectrum_theta_20_nob_fill_pattern.pdf} + \includegraphics[width=0.65\textwidth]{graphics/spectrum_theta_20_nob_fill_pattern.pdf} \caption{Height fluctuations spectra from deterministic (circles) and stochastic (triangles) simulations at $t=0.2 t_0$ (filled symbols) and $t=0.7 t_0$ (empty symbols), on a substrate with $\theta =\pi/9$. The theoretical predictions, Eq.~(\ref{eq:structure_factor_2}), are reported with solid ($\sigma=0$) and dashed ($\sigma = \sigma_0$) lines.} \label{fig:theory_simulation_structure_factor} @@ -259,7 +259,7 @@ \subsection{Testing the stochastic term}\label{subsec:validation} \subsection{Droplet size distributions}\label{subsec:morphandrup} \begin{figure} \centering - \includegraphics[width=0.75\textwidth]{graphics/Droplet_height_['00', '1e-7']_distri_new_35_nodist.pdf} + \includegraphics[width=0.65\textwidth]{graphics/Droplet_height_['00', '1e-7']_distri_new_35_nodist.pdf} \caption{Histograms of the droplet height distributions from the deterministic (blue, line-patterned) and stochastic (orange, dot-patterned) simulations with $\theta = \pi/9$, at $t \approx 20 t_0$; the corresponding Gaussian kernel density estimations are also plotted, as a guide to the eye, with solid blue (dashed orange) curve for the deterministic (stochastic) data.} \label{fig:droplet_distribution} \end{figure} @@ -282,7 +282,7 @@ \subsection{Droplet size distributions}\label{subsec:morphandrup} These findings are in agreement with what was reported by Nesic \textit{et al.}~\cite{nesicFullyNonlinearDynamics2015}. \begin{figure} \centering - \includegraphics[width = 0.75\textwidth]{graphics/Correct_t0_normed_delta_h_evo_seminar.pdf} + \includegraphics[width = 0.65\textwidth]{graphics/Correct_t0_normed_delta_h_evo_seminar.pdf} \caption{Time evolution of $\Delta h(t) = \max_x\{h(x,t)\} - \min_x\{h(x,t)\}$ for the athermal (bullets) and fluctuating (triangles) systems with $\theta=\pi/9$. The blue dashed-dotted line depicts an exponential fit of the data for the athermal case ($\sigma=0$). The vertical lines mark the rupture times for the deteministic (solid) and stochastic (dotted) simulations.} @@ -351,7 +351,7 @@ \subsection{Rupture times and role of contact angle}\label{subsec:results_contac The latter relation tells us that $\chi_{\sigma}(\theta)$, indeed, increases with the temperature (confirming that the rupture times are shorter for the fluctuating systems), but decreases with the contact angle. \begin{figure} \centering - \includegraphics[width=0.75\textwidth]{graphics/Andrea_model_t0_normed_rupture_times.pdf} + \includegraphics[width=0.65\textwidth]{graphics/Andrea_model_t0_normed_rupture_times.pdf} \caption{Ratio of rupture times from athermal and fluctuating dewetting, $\chi_{\sigma_0}(\theta)$ [Eq.~(\ref{eq:defchi}], as a function of the contact angle $\theta$. The dashed line is Eq.~(\ref{eq:chi}) for $\sigma=\sigma_0$. Inset: Rupture times $\tau_r$ normalized by $t_0$ (see Table~\ref{tab:t_0s}) vs $\theta$, for the deterministic (orange bullets \textcolor{pyorange}{$\bullet$}) and stochastic (green triangles \textcolor{pygreen}{$\blacktriangle$}) simulations. @@ -361,7 +361,7 @@ \subsection{Rupture times and role of contact angle}\label{subsec:results_contac Equation~(\ref{eq:chi}) is plotted in Fig.~\ref{fig:rupture_times_semilogy_more_theta} (dashed line), showing a nice agreement with the numerical data. \begin{figure} \centering - \includegraphics[width=0.75\textwidth]{graphics/evolution_qm_with_inset_slip0.pdf} + \includegraphics[width=0.65\textwidth]{graphics/evolution_qm_with_inset_slip0.pdf} \caption{Time dependence of the maximum wavenumber, $q_m$, at which the spectrum has the global maximum for various contact angles $\theta$ and $\sigma = \sigma_0$, using $\delta=0$. The full blue, orange and green lines correspond to theoretical curve $q_m(t)$ derived from Eq.~(\ref{eq:structure_factor_2}) for $\theta = \pi/9, 5\pi/36, \pi/6$. The dashed lines show the corresponding $q_0(\theta)$. @@ -391,7 +391,7 @@ \subsubsection{Sine wave pattern}\label{subsubsec:sine} Indeed, the deterministic dewetting leads to the formation of precisely $L/\lambda_{\theta}$ droplets. \begin{figure} \centering - \includegraphics[width=0.95\textwidth]{graphics/spacedepCA_['sine', '1e7', '10', 25, '9_3', 10000000.0]_v2.png} + \includegraphics[width=0.75\textwidth]{graphics/spacedepCA_['sine', '1e7', '10', 25, '9_3', 10000000.0]_v2.png} \caption{Space-time plot of the height field $h(x,t)$ evolution over a sinusoidally patterned substrate undergoing athermal (a) and fluctuating (b) dewetting, respectively. In panel (c) we report the contact angle profile $\theta(x)$ [Eq.~(\ref{eq:sine_angle})].} \label{fig:patterned_sine8_difference_20-30} @@ -407,7 +407,7 @@ \subsubsection{Sine wave pattern}\label{subsubsec:sine} Consequently, since $q_0$ decreases with $\theta$, we should expect the actual most unstable wavenumber to be slightly below $q_0$. \begin{figure} \centering - \includegraphics[width=0.95\textwidth]{graphics/psd_spacedepCA_sine_25_10_00_[0, 450000, 1150000]_9_3_rescaled.pdf} + \includegraphics[width=0.75\textwidth]{graphics/psd_spacedepCA_sine_25_10_00_[0, 450000, 1150000]_9_3_rescaled.pdf} \caption{Height profiles from deterministic dewetting ($\sigma=0$) over the sinusoidally patterned substrate, Eq.~(\ref{eq:sine_angle}) at $t=0.26 t_0$ (a) and $t=0.66 t_0$ (b) and corresponding spectra (c). In panel (c) also the initial spectrum, $S_0(q)$, is reported; the vertical lines indicate the pattern wave mode $q_{\theta}$ (solid) and its multiples $2q_{\theta}$ (dashed) and $3q_{\theta}$ (dashed dotted). As explained in the text, the convention $t_0 = t_0(\pi/6)$ and $q_0 = q_0(\pi/6)$ applies.} @@ -415,7 +415,7 @@ \subsubsection{Sine wave pattern}\label{subsubsec:sine} \end{figure} \begin{figure} \centering - \includegraphics[width=0.95\textwidth]{graphics/psd_spacedepCA_sine_25_10_1e-7_[0, 450000, 1150000]_9_3_rescaled.pdf} + \includegraphics[width=0.75\textwidth]{graphics/psd_spacedepCA_sine_25_10_1e-7_[0, 450000, 1150000]_9_3_rescaled.pdf} \caption{Height profiles from stochastic dewetting ($\sigma=\sigma_0$) over the sinusoidally patterned substrate, Eq.(\ref{eq:sine_angle}) at $t=0.26 t_0$ (a) and $t=0.66 t_0$ (b) and corresponding spectra (c). In panel (c) also the initial spectrum, $S_0(q)$, is reported; the vertical lines indicate the pattern wave mode $q_{\theta}$ (solid) and its multiples $2q_{\theta}$ (dashed) and $3q_{\theta}$ (dashed dotted).} \label{fig:spectral_analysis_stoch_sine8} @@ -435,7 +435,7 @@ \subsubsection{Sine wave pattern}\label{subsubsec:sine} \subsubsection{Square wave pattern}\label{subsubsec:square_wave} \begin{figure} \centering - \includegraphics[width=0.95\textwidth]{graphics/spacedepCA_['delta', '1e7', '10', 25, '9_3', 2000000.0]_v2.png} + \includegraphics[width=0.75\textwidth]{graphics/spacedepCA_['delta', '1e7', '10', 25, '9_3', 2000000.0]_v2.png} \caption{Space-time plot of the height field $h(x,t)$ evolution over a square-wave patterned substrate undergoing athermal (a) and fluctuating (b) dewetting, respectively. In panel (c) we report the contact angle profile $\theta(x)$ [Eq.~(\ref{eq:sharp_contact_angle_spatial})].} \label{fig:patterned_step8_difference_20-30} @@ -474,7 +474,7 @@ \subsubsection{Square wave pattern}\label{subsubsec:square_wave} in excellent agreement with the measured value. \begin{figure} \centering - \includegraphics[width=0.95\textwidth]{graphics/Rupture_events_with_film.pdf} + \includegraphics[width=0.75\textwidth]{graphics/Rupture_events_with_film.pdf} \caption{(a), (b) Space-time plots showing the evolution of the height field for athermal (a) and fluctuating (b) dewetting over the square-wave patterned substrate, in a neighbourhood of the instant of time at which the first rupture event (in the athermal case) occurred; for the sake of visualization we mark the rupture events with red bullets (\textcolor{red}{$\bullet$}). (c) Distribution of times of occurrence of rupture events for the athermal ($\sigma=0$, blue, line-patterned) and fluctuating ($\sigma = \sigma_0$, orange, dot-patterned) dewetting.} \label{fig:rupture_time_distri_square_wave8} @@ -488,7 +488,7 @@ \subsubsection{Square wave pattern}\label{subsubsec:square_wave} \begin{figure} \centering - \includegraphics[width=0.75\textwidth]{graphics/square_wave_det_stoch.pdf} + \includegraphics[width=0.65\textwidth]{graphics/square_wave_det_stoch.pdf} \caption{Spectra of the dewetting film on the $\theta^{(2)}(x)$ pattern, Eq.(\ref{eq:sharp_contact_angle_spatial}), at $t = 0.1t_0(\pi/6)$, with ($\sigma = \sigma_0$, \textcolor{pyorange}{$\blacktriangle$}) and without ($\sigma =0$, \textcolor{pyblue}{$\bullet$}) thermal fluctuations. The vertical lines indicate the substrate wave mode $q_{\theta}$ (solid) and its multiples $2q_{\theta}$ (dashed), $3q_{\theta}$ (thin dashed-dotted) and $7q_{\theta}$ (thick dashed-dotted).} \label{fig:square_wave_both} diff --git a/others/titlepage.tex b/others/titlepage.tex index bf76d31..d4a9267 100644 --- a/others/titlepage.tex +++ b/others/titlepage.tex @@ -17,7 +17,7 @@ \node at ($ (current page.center) + (0, 0.7) $) { \parbox{\textwidth}{% \begin{center} - \large{Der Technischen Fakul\"at \break der Friedrich-Alexander-Universit\"at \break Erlangen-N\"urnberg \break zur \break Erlangung des Doktorgrades Dr.-Ing. \break vorgelegt von} \\[2ex] + \large{Der Technischen Fakult\"at \break der Friedrich-Alexander-Universit\"at \break Erlangen-N\"urnberg \break zur \break Erlangung des Doktorgrades Dr.-Ing. \break vorgelegt von} \\[2ex] \large{Stefan Zitz \break aus Graz, \"Osterreich}