Numerical Simulation Of Electrochemical Desalination

1y ago
7 Views
2 Downloads
885.28 KB
13 Pages
Last View : 21d ago
Last Download : 3m ago
Upload by : Gideon Hoey
Transcription

Journal of Physics: Condensed MatterJ. Phys.: Condens. Matter 28 (2016) 194001 (13pp)doi:10.1088/0953-8984/28/19/194001Numerical simulation of electrochemicaldesalinationD Hlushkou1, K N Knust2, R M Crooks2 and U Tallarek1,31Department of Chemistry, Philipps-Universität Marburg, Hans-Meerwein-Strasse 4, 35032 Marburg,Germany2Department of Chemistry and the Center for Nano and Molecular Science and Technology,The University of Texas at Austin, 105 E. 24th St., Stop A5300, Austin, TX 78712, USAE-mail: ulrich.tallarek@staff.uni-marburg.deReceived 15 August 2015, revised 8 February 2016Accepted for publication 10 March 2016Published 19 April 2016AbstractWe present an effective numerical approach to simulate electrochemically mediateddesalination of seawater. This new membraneless, energy efficient desalination method relieson the oxidation of chloride ions, which generates an ion depletion zone and local electric fieldgradient near the junction of a microchannel branch to redirect sea salt into the brine stream,consequently producing desalted water. The proposed numerical model is based on resolutionof the 3D coupled Navier–Stokes, Nernst–Planck, and Poisson equations at non-uniformspatial grids. The model is implemented as a parallel code and can be employed to simulatemass–charge transport coupled with surface or volume reactions in 3D systems showing anarbitrarily complex geometrical configuration.Keywords: electrochemical desalination, membraneless desalination,parallel numerical simulation, mass–charge transport, electrochemistry(Some figures may appear in colour only in the online journal)1. Introductionresources—where the population is predicted to reach 2.4 billion people by 2050 [7].About 97.5% of the water reserves on the Earth are stored ineasily accessible reservoirs, with around 1.0% being brackishgroundwater (0.05–3.0% salts by weight) and the remaining96.5% being seawater (3.0–5.0% salts) [8]. Thus, the vastamounts of saline water could offer a seemingly unlimitedand steady fresh water ( 0.05% salts) supply. However, apractical desalination process has to meet several essentialrequirements such as technological reliability, relative simplicity, amenability to large-scale continuous operation, economic viability, minor impact on the environment, energyefficiency, etc [9]. At present, the most common desalinationtechniques employed for large-scale purification of seawaterand brackish groundwater are thermal desalination (TD) andreverse osmosis (RO), which account together for roughly90% of desalinated water production. The former techniquebecame commercially available in the 1950s. The minimumenergy required to drive TD is directly related to the heat ofwater vaporization, which is substantially higher than that ofother desalination techniques. Although TD requires minimalAt present, water scarcity is a global challenge affecting everycontinent around the world. About 11% of the world’s population currently relies on unimproved drinking water sources[1]. Nearly one-fifth of the global population lives in areaswhere water is physically scarce [2]. Regionally, the limitof ecological sustainability of water available for abstraction will be exceeded for about half of the human populationby 2030 [3]. An analysis of groundwater extraction revealedthat depletion rates doubled between 1960 and 2000 and areespecially high in parts of India, China, and the United States[4]. Even in a number of European countries (Great Britain,Poland, Spain, Germany, and the Czech Republic), the situation with total renewable water resources is characterized asvulnerable or stressed [5]. About 120 million people in theEuropean region do not have access to safe drinking water[6]. The situation is much worse in sub-Saharan Africa—the region with the most heterogeneously distributed water3Author to whom any correspondence should be addressed.0953-8984/16/194001 13 33.001 2016 IOP Publishing LtdPrinted in the UK

D Hlushkou et alJ. Phys.: Condens. Matter 28 (2016) 194001pre-treatments of the feed water, the commercial use of thisdesalination technique is limited to geographical regions,such as the Middle East, rich in energy resources. A widelyused alternative desalination technique involves RO, which isinduced if a hydrostatic pressure differential greater than theosmotic pressure of the feed water is applied across a semipermeable membrane. The minimum energy required for ROis directly related to the salt concentration of the feed water.With advances in membrane technology and energy recoverydevices, RO can nowadays operate with an energy efficiencybetter than that of any other currently commercialized desalination method [10]. However, RO requires extensive pretreatment of the feed water and product post-treatment due tofragility, contamination, and fouling of the membrane, whichgreatly increases the production cost of water purified by thistechnique [10, 11]. In recent years, a number of strategiesand approaches have been proposed as alternative routes forwater desalination. These novel techniques include microbialdesalination [12], the use of carbon nanotube membranes forfiltration and RO processes [13], shock electrodialysis [14],desalination by ion concentration polarization [15], waterpurification with biomimetic aquaporin membranes [16–18],capacitive deionization [19–23], desalination with entropybatteries [24], and gas hydrate-based desalination using newformers [25–27]. Without distinction of the diverse underlyingphenomena, a desalination process can be defined as separating a feed salt solution into deionized water and concentrated brine streams.Recently, our groups have shown that electrochemicalreactions can be employed for the enrichment and depletion[28–34], separation [35], as well as controlled delivery [36] ofcharged species from a solution flowing through microchannelsystems with embedded electrodes. Each of these applicationsis based on the principles of electrokinetic equilibrium and differences between local electromigration, diffusive, and advective fluxes of ions. The key to these techniques is the formationof a localized electric field gradient induced by electrochemical reactions at the embedded electrode surfaces. We havediscovered these same principles can be employed to desaltseawater by a technique called electrochemically mediateddesalination (EMD) [37, 38]. This novel approach providesa number of benefits relative to currently available desalination methods. Firstly, it is membraneless, thereby eliminatingmajor drawbacks of membrane-based desalination technologies: fragility and fouling of membranes and the necessity ofintensive pre-treatment of the feed seawater. In contrast, theonly pre-treatment required to perform EMD is the sedimentation of sand and debris present in seawater [37]. Becauseof this minimal pre-treatment, considerable cost savingscould be provided compared to membrane-based techniquessuch as RO and electrodialysis. Secondly, EMD is energyefficient: For 25% salt rejection in seawater at 50% recovery,the energy efficiency is 0.025 kWh m 3 [38]. This value is nearthe theoretical minimum energy calculated using the samerejection and recovery (about 0.017 kWh m 3). Thirdly, EMDrequires only a simple power supply to operate. Therefore,this desalination technique may be employed in settings witha low-voltage battery or renewable energy source without aFigure 1. Optical micrograph of the region of the PDMS/glassdesalination device near an embedded Pt electrode (anode).high-voltage converter. Finally, the EMD platform may beimplemented in a massive parallel format. The last aspectappears to be essential because our basic experimental EMDunits are currently realized in microchannel format. However,apart from its parallel implementation, the performance ofsystems employing this novel desalination technique may besubstantially enhanced by an increase in the functional efficiency of a single EMD unit. For this purpose, optimization ofthe geometrical configuration and operating conditions of theEMD unit is required. In this relation, numerical simulationsof the phenomena involved in the EMD process are able toprovide a better understanding of this desalination techniqueas well as to maximize the use of resources by replacing physical experiments with computational studies.In this contribution, we present a numerical approach tosimulate the desalination process in the EMD unit. The proposed numerical model is based on equations describingthe coupled 3D hydrodynamic, mass–charge transport, andelectrostatic problems. The model was developed based onnumerical schemes with inherent parallelism, allowing thestraightforward implementation at modern high-performancecomputational systems (supercomputers). Results obtainedwith the proposed simulation approach agree very well withthe experimental data we previously reported [37, 38]. Theanalysis of the simulated data highlights the complex interplay between the mass–charge transport, hydrodynamics, andelectrochemical reactions involved in the EMD process.2. Experimental EMD unitThe design of our current EMD unit consists of a Y-shapedpoly(dimethylsiloxane) (PDMS)/glass hybrid channel network outfitted with a 100 μm wide Pt electrode near theY-intersection that extends 50 μm toward the channel center.An optical micrograph of the region near the electrode isshown in figure 1.The Pt electrode was deposited onto the glass base of theEMD device by standard photolithographic techniques andelectron beam deposition. PDMS microchannels (feed channelinlet 2.5 mm long, 100 μm wide, and 22 μm tall; brine streamand desalted stream channel outlets 2.5 mm long, 50 µmwide, and 22 μm tall) were fabricated by a replica molding2

D Hlushkou et alJ. Phys.: Condens. Matter 28 (2016) 194001much lower than that of the sodium and chloride ions. Thisallows us to consider only Na and Cl ions as the species ofinterest in the modelled system.In seawater and an aqueous NaCl solution there are twomajor species that can be oxidized at the anode: Cl andwater. However, in the electrolysis of seawater and a concentrated NaCl solution, Cl oxidation is the preferred anodicprocess over water oxidation. The main reasons leading to thepredominance of Cl oxidation at the anode are reported asfollows [39–43]:(i) Local acidification of the electrolyte near the anodeincreases the thermodynamic potential for oxygen evol ution, favouring chloride oxidation since it is independentof the pH.(ii) The overpotential for chloride oxidation is lower than thatfor the oxidation of water.(iii) Oxygen evolution is a slow electrochemical process withlow exchange current density. The ratio of the exchangecurrent densities for chlorine and oxygen evolution atmost anode materials is in the range of 103–107. Thisratio characterizes the ease of chlorine evolution relativeto oxygen.Figure 2. Schematic of the microchannel system used forsimulation of the EMD device. The thickness (Z-dimension) of themicrochannels is 22 μm.procedure. The outlet of the brine stream channel branchedfrom the desalted stream channel at a 30 angle. At eachmicrochannel terminal, a reservoir was made using a 3 mmdiameter metal punch to remove PDMS. Finally, a PDMSmold was attached to the glass base by O2 plasma bonding.Pressure driven flow (PDF) is initiated in the EMD unitby creating a solution height differential in the reservoirs atthe ends of the channel by adding or removing seawater. Apower supply is used to apply a low-voltage bias betweenthe microfabricated Pt anode and grounded electrodes (cathodes) dipped into the reservoirs. Applied voltage is adjustedto a value (a few volts) that results in electrochemical reactionat the anode—the oxidation of chloride ions (Cl ) present inseawater: Therefore, the assumption that all of the current conductedthrough the anode results in only oxidation of Cl ions isaccurate to a very good approximation for the studied system.Taking the liquid as incompressible, the local flow velocityfield (v) can be described by the Navier–Stokes equation v v v p η 2 v,ρ (2) t 2Cl 2e Cl2. (1)where ρ and η are the mass density and dynamic viscosity ofthe liquid, and p is hydrostatic pressure. It should be mentioned that the above equation does not contain the electricbody force term which accounts for the contribution of electroconvection to charge transport. Generally, there are twomodes of electroconvective phenomena eventually observedin an electrolyte solution [44–46]. The first mode is a result ofthe action on a macroscopic scale of the electric field upon theresidual space charge. The second mode (electroosmotic flow)pertains to the electrolyte slip resulting from the action of thetangential (relative to the solid–liquid interface) component ofthe electric field upon the space charge of the electric doublelayer. Though electroconvection may change significantlythe spatiotemporal ion distributions in systems with chargeselective domains (such as metallic electrodes, ion-exchangemembranes or nanochannels), its effect under the operatingconditions in the system we modelled (with a 0.55 M electrolyte solution and a 50 nA current through the electrode)is negligible [44, 47–50]. We return to this issue in section 4when we discuss the simulation results.Bulk spatiotemporal variations in the concentrations ofNa and Cl ions are governed by balance equationsThis reaction leads to the formation of an ion depletion zone(region of high solution resistivity) near the anode, thus producing a local electric field gradient and providing a meansfor controlling the movement of ions. As a result, a seawaterfeed is separated into brine and desalted water streams at thejunction of the branched channel.3. Theoretical background and numericalapproaches to modeling EMDIn this section we present the theoretical model and the computational methods used to simulate the basic EMD process.The 3D geometry of the modelled system is exactly the sameas that used for the experiments (see figure 1) and is shown infigure 2.All microchannels are terminated by macroscopic reservoirs containing the grounded electrodes (cathodes). As modelliquid we used a 0.55 M aqueous NaCl solution (which reflectsthe Cl concentration in seawater) and furthermore assumethat Cl is oxidized at the embedded electrode (anode). Thecomputer model simulates stationary PDF initiated in themicrochannel system as well as the stationary distributions oflocal ion concentrations and local electric field strength. Infact, in an aqueous NaCl solution the ionic species includeNa , Cl , H3O , and OH . However, in a 0.55 M solution ofNaCl, the concentration of hydronium and hydroxide ions isF nNaD Na 2 φ nNa v jNa D Na 2 nNa nNa tRT (3)and3

D Hlushkou et alJ. Phys.: Condens. Matter 28 (2016) 194001where f is the acting external force and Ω is the collisionoperator. Appropriate choice of the collision operator allowsto determine accurately the macroscopic properties of a fluid,since these are not directly dependent on the details of themicroscopic behaviour but mainly defined through interactions between particles. This offers the transition toward asimplified dynamics with discrete space, time and molecularvelocities. In particular, the discrete analogy of (8) is n ClF jCl DCl 2 n Cl n Cl DCl 2 φ n Cl v, tRT (4)where n is the species volume concentration (m 3), j is theflux, D is the diffusion coefficient, and φ denotes the localelectric potential; F, R, and T represent the Faraday constant,molar gas constant, and temperature, respectively.The Poisson equation establishes the essential relationshipbetween the local ionic concentrations of the species and thelocal electric potentialn n Cl 2 φ qe Na, (5)ε0εrFα(r eαδt , t δt ) Fα(r, t ) Ωα(Fα), (9)where Fα is the distribution function for the αth discretevelocity eα at position r and time t, and δt is the time stepused in the LBE simulation. Here, we implemented the latticeBGK (Bhatnagar–Gross–Krook) model [52], a modificationof the LBE method characterized by a single-time relaxationcollision operatorwhere qe is the elementary charge; ε0 and εr are the vacuumpermittivity and dielectric constant of the liquid. In (2)–(5)it is assumed that the mass density, viscosity, diffusion coefficients, and dielectric constant are independent of the ionconcentration.Instead of a direct numerical resolution of the Navier–Stokes problem (2), the simulation of hydraulic flow throughthe microchannel system was performed using the latticeBoltzmann equation (LBE) method, a kinetic approach thatoperates with discrete space and time [51–53], where themacroscopic fluid dynamics is approximated by the motionof (and interactions between) fictitious particles on a regularlattice. This approach is based on the idea that the resultingfluid flow is determined mainly by the collective behaviourof molecules and not by their detailed interactions. Eachof the fictitious particles is then already associated with‘a large number’ of fluid molecules. Specifically, the LBEcan be regarded as a finite-difference approximation of theBoltzmann equation, which is obtained by its truncationin a Hermite velocity spectrum space [52, 54]. Among theadvantages of the LBE method are its inherent algorithmicparallelism as well as its capability to handle solid–liquidinterfaces with complex geometry, e.g. in disordered porousmedia [51].The kinetics of a fluid in terms of a statistical system canbe described with a distribution function F(r, u, t) such thatF(r, u, t)drdu is the number of fluid particles which, at time t,are located between r and (r     dr) and have velocities in therange from u to (u     du). Then, the density ρ of the fluid andits velocity v can be obtained by momentum integration of thisdistribution function1 eqF[F α (r, t ) Fα(r, t )] , (10)α(r eαδt , t δt ) Fα(r, t ) τwhere F αeq represents the equilibrium distribution function andτ is a non-dimensional relaxation time. This parameter is connected with the kinematic viscosity (κ) of the fluid by2τ 1.(11)6The values of velocities eα in (9) and (10) are chosen suchthat during one time step δt each particle moves from one lattice node to its neighbour. In this work, we adapted the D3Q19lattice, a cubic lattice that can be obtained by projecting the4D face-centred hypercubic lattice onto 3D space [56, 57].The D3Q19 lattice has 18 links at each lattice node; each lattice node is connected to its six nearest and twelve diagonalneighbours. Particles can move along the 18 links (α     1–18)or stay at the node (α     0).The equilibrium distribution function depends on the localdensity ρ(r, t) and local velocity v(r, t) κ e v(e v)2v v F αeq(r, t ) wαρ 1 α 2 α 4 2 (12)2cS2cS cS where cS is the speed of sound and wα is a weighting factorthat is governed by the length of the velocity vector eα. Withthe D3Q19 lattice, these factors are as follows: wα     1/3 forα     0, wα     1/18 for α corresponding to lattice links to thenearest neighbours, and wα     1/36 for α corresponding to lattice links to the diagonal neighbours. The lattice spacing (Δh)we used for the simulation of the PDF velocity field was setto 1.0 μm. At the solid–liquid (microchannel wall–electrolytesolution) interface, a halfway bounce-back rule was applied toimplement the required ‘liquid-stick’ velocity boundary condition [58].The Poisson and Nernst–Planck equations (3)–(5) wereresolved with conventional finite-difference methods. For thispurpose, the solution domain was split into a set of equal cubiccells with a size of Δh     1.0 μm (generating a uniform cubicgrid). The electric potential and ion concentrations are determined at the center points of the cells. The spatiotemporalfinite-difference scheme for solution of the Nernst–Planck ρ(r, t ) MmF (r,u, t )du (6)and1Mm uF (r,u, t )du,v(r, t ) (7)ρ (r, t ) where Mm is the mass of a particle. The spatiotemporalchanges in the distribution function can be described by thefollowing evolution equation fF r udt , u dt , t dt drdu F (r,u, t ) Ω(F )drdudt , Mm (8)4

D Hlushkou et alJ. Phys.: Condens. Matter 28 (2016) 194001By introducing the discrete time step Δt, the two above equations can be written as tntNa ntNa t (j j Na,X j Na,Y j Na,Y j Na,Z j Na,Z ) h Na,X (15)andntCl t ntCl t (j j Cl,X jCl,Y jCl,Y jCl,Z jCl,Z ), h Cl,X (16)where nt and nt Δt are the species concentrations at time t and(t     Δt), respectively.A finite-difference approximation of a particular (X-)component of the ion flux j X , k, l, m in a cell with discrete coordinates k, l, m (corresponding to X-, Y-, and Z-directions,respectively) isnk, l, m nk 1, l, m hFD nk, l, m nk 1, l, m φk, l, m φk 1, l, m RT2 hnk, l, m nk 1, l, m vX , k, l, m vX , k 1, l, m(17) 22j X , k, l, m D Figure 3. Denotation of the flux components in a cubic cellemployed to resolve the Nernst–Planck equation by a finitedifference scheme.where vX is the X-component of the flow velocity field. Thesign before the second term on the rhs of (17) is determinedby the valence of the species: it is ‘     ’ and ‘     ’ for Na and Cl ions, respectively. Expressions similar to (17) can bederived for all other flux components. If a cell is adjacent tothe channel wall, the corresponding terms in (15) and (16)become zero.A special treatment is used for cells adjoining theembedded electrode. The region near the electrode is characterized by steep gradients of the electric field strength andion concentrations. As a result, the spatial resolution of 1 μmcannot assure a good accuracy with the use of a finite-difference approximation. To resolve this problem, we introduce alocally non-uniform grid: Each cell adjoining the electrode(a ‘parent’cell) is split into I sub-cells (‘child’ cells or layers)with dimension of Δh     1.0 μm along X- and Y-directions(tangential to the electrode surface) and an exponentiallydecreasing thicknesses, Δhi, along the Z-direction toward theelectrode (figure 4).For the current study, we split such cells into 20 sub-cells(i     1, , I; I     20). The largest value of the index i     20corresponds to the sub-cell (layer) which is adjacent to theelectrode surface. The thickness Δhi of the ith sub-cell isdetermined asFigure 4. Illustration of the non-uniform spatial grid employed forresolving the Nernst–Planck (and Poisson) equation with finitedifference schemes. A cell adjoining the electrode (red) is split intosub-cells (layers) with exponentially decreasing thickness. For thisstudy, such cells were split into 20 layers.equations (3, 4) is based on the calculation of the total flux ina cell, thereby accounting for the flux components on each ofthe six cell surfaces (figure 3).Then, (3) and (4) can be represented, respectively, as hi 2 i h if i 1,.,I 1 (18) hi hi 1 if i I . nNa1 (j j Na,X j Na,Y j Na,Y j Na,Z j Na,Z ) h Na,X tSpecifically, with I     20 the smallest value of Δhi is about1.9     10 6 μm, which allows us to achieve an accuratefinite-difference approximation of the gradient operators forthe electric field strength and ion concentrations. Below, wedenote the discrete co-ordinates of a sub-cell with thicknessΔhi as k, l, m, i, where k, l, m are the discrete co-ordinates ofthe ‘parent’cell that is split into ‘child’ sub-cells. (13)and n Cl1 (j jCl,X j Cl,Y jCl,Y jCl,Z jCl,Z ). t h Cl,X (14)5

D Hlushkou et alJ. Phys.: Condens. Matter 28 (2016) 194001Accounting for the difference between the thicknesses ofadjoining sub-cells, the finite-difference approximations offlux components j Z , k, l, m, i and j Z , k, l, m, i (see figures 3 and 4 and(17)) can be written as1(ϕk 1, l, m, i ϕk 1, l, m, i φk, l 1, m, i φk, l 1, m, i 4ϕk, l, m, i ) h 2 φk, l, m, i 1 φk, l, m, i4 hk, l, m, i 1 hk, l, m, i 1 hi 1 hink, l, m, i nk, l, m, i 10.5( hi hi 1)FD nk, l, m, i nk, l, m, i 1 φk, l, m, i φk, l, m, i 1 RT20.5( hi hi 1)nk, l, m, i nk, l, m, i 1 vZ , k, l, m, i vZ , k, l, m, i 1(19) 22j Z , k, l, m, i D andj Z , k, l, m, i D nk, l, m, i nk, l, m, i 10.5( hi hi 1)FD nk, l, m, i nk, l, m, i 1 φk, l, m, i φk, l, m, i 1RT20.5( hi hi 1)nk, l, m, i nk, l, m, i 1 vZ , k, l, m, i vZ , k, l, m, i 1. (20) 22 In these equations, the sign before the second term on the rhsis determined by the valence of the species, as specified for(17). Expressions similar to (19) and (20) are applied also tocalculate the flux Z-component in neighbouring ‘parent’and‘child’cells by setting Δhi     Δh for a ‘parent’ cell.In the proposed numerical approach to resolve the Nernst–Planck equation, the occurrence of the electrochemical reaction (1) is accounted for through the flux component j Z , k, l, m, Ifor Cl ions. This term determines the decrement of the Cl concentration due to oxidation at the anode surface and forsub-cells with discrete co-ordinates k, l, m, I becomes4. Results and discussionThe 3D numerical approach described in section 3 was subsequently employed to simulate the EMD process in the microchannel system presented with figures 1 and 2. The simulationhas mirrored the experiment carried out for the desalinationof a 0.55 M aqueous NaCl solution with a current throughthe embedded electrode of 50 nA, a volumetric PDF rate of0.1 μl min 1, and an electrode potential of 0.9 V. The dynamicviscosity and mass density of the NaCl solution was 0.966 mPa sand 1.023     103 kg m 3, respectively. Diffusion coefficientswere adjusted to 1.334     10 9 and 2.033     10 9 m2 s 1, forNa and Cl ions, respectively. The temperature of the systemwas set to 293.16 K.Fully developed, stationary PDF in the modelled EMDsystem is required as transport mechanism for the bulk liquidphase (electrolyte solution). It can be conveniently generatedby introducing and maintaining different heights of electrolyte solution levels at the terminating ends of the respectivemicrochannels. This results in steady-state Hagen–Poiseuilleflow in the microchannel, feeding the bulk electrolyte solutiontowards the Y-intersection, i.e. the microchannel branch. Theflow is characterized by a 3D channel cross-sectional velocityprofile that closely resembles the classical parabolic velocityprofile in a cylindrical pipe. That is, flow velocity increasesfrom zero at the channel walls, where the no-slip velocityboundary condition prevails (as implemented in the simulations), to a maximum in the center of the channel cross-section. However, in contrast to the classical cylinder geometry,the four corners of the rectangular microchannel geometrypresent additional low-velocity regions (with corresp ondingIj Z , k, l, m, I el , (21)Sqewhere Iel denotes the current through the electrode and S isthe area of the electrode surface in contact with the electrolytesolution (see figures 1 and 2).A similar numerical approach is employed to resolve thePoisson equation (5). Its basic finite-difference approximationis1(φk 1, l, m φk 1, l, m φk , l 1, m φk , l 1, m φk , l, m 1 h 2nNa, k, l, m n Cl, k, l, m φk, l, m 1 6φk, l, m ) qe(22)ε0εrφk, l, m 1 qe h2(nNa, k, l, m n Cl , k, l, m) φk 1, l, m φk 1, l, m 6 ε0εr φk, l 1, m φk, l 1, m φk, l, m 1 φk, l, m 1 . (24) For sub-cells adjoining the electrode surface, the electricpotential φk,l,m,I is adjusted to φel, the potential of the anoderelative to the one in the inlet and outlet reservoirs.The program realization of the presented numericalapproach to resolve the Navier–Stokes, Nernst–Planck, andPoisson equations was implemented as parallel code in C/C language using the message passing interface (MPI) standard[59]. After discretization of the modelled system, the resultingspatial grid was composed of 107 nodes. At the first stage, the3D flow velocity field was computed using the LBE approach.Then, the coupled Nernst–Planck and Poisson equations wereresolved with the presented finite-difference schemes. Allsimulations were performed on SuperMUC, a supercomputerat the ‘Leibniz-Rechenzentrum der Bayerischen Akademieder Wissenschaften’ (LRZ, Garching, Germany). The final(productive) simulation reported and analyzed in this workrequired roughly 6 h at 1024 processor cores. orϕk , l, m, i ϕk , l, m, i 1 nNa, k , l, m, i n Cl, k , l, m, i. qe h i 1 h i ε0εr(23) For cells adjacent to the anode surface, we use the same ‘splitting’ procedure described above. A finite-difference approx imation of the Poisson equation for neighbouring cells withdistinct values of their thickness (Z-dimension) is6

D Hlushkou et alJ. Phys.: Condens. Matter 28 (2016) 194001species present in the modelled system in longitudinal andtransverse directions with respect to the feed microchannelaxis results from the superposition of advection, diffusion, andmigration. The latter is determined, in particular, by the localelectric potential gradient developing in the coupled electrochemical module (realized through the embedded anode),which in turn depends on the local ion concentrations. Sincethe electric field gradient is a key to EMD device functionalityit is described in more detail.Thus, the next step beyond the relatively simple situationillustrated by figure 5 (with pure PDF) is the onset of oxidationof Cl ions. This produces electroneutral chlorine moleculesand therefore reduces the number of ionic charge carriers nearthe embedded anode. It is a straightforward electrochemicalprocess that occurs locally but steadily and, together with thediffusion and migration of the ions in the system as well as thesuperimposed PDF, after some time results in a steady-statedepleted concentration polarization zone with an associatedelectric field gradient. We have experience with in situ formedlocalized electric field gradients for

In this contribution, we present a numerical approach to simulate the desalination process in the EMD unit. The pro-posed numerical model is based on equations describing the coupled 3D hydrodynamic, mass-charge transport, and electrostatic problems. The model was developed based on numerical schemes with inherent parallelism, allowing the

Related Documents:

thermal desalination processes around the world. Direct solar desalination uses solar energy to produce distillate directly in the solar collector whereas indirect solar desalination combines conventional desalination t

The main thermal desalination processes of great commercial use are: MSF, MED and MVC, with a market share of commercial desalination plants of 87.3%, 12.5%, and 0.2%, respectively. Other types of thermal desalination processes such as SS, HDH, and freezing are not found commercially and

Concentrating Solar Thermal Power for Seawater Desalination. 3 Slide 5 0 100 200 300 400 500 600 700 2000 2010 2020 2030 2040 2050 Year . Concentrating Solar Power for Electricity, Cooling and Desalination MED Multi-Effect-Desalination AC Absorption Chiller Example: Ayla Oasis Resort, Aqaba, Jordan MED

[2] Williams, P.M., Ahmad, A., Connolly, B.S., Oatley-Radcliffe, D.L., Technology for freeze concentration in the desalination industry, Desalination 356 (2015), 314 -327. Applicability of various desalination technologies [1] A substantial amount of research on freeze desalination was conducted during 1960's and 70's.

2007-2009: Zero Discharge Desalination was the 1 st technology piloted at the BGNDRF Recovery: 85 95% 10/28/2019 Solar Powered Desalination 3. . Recovery: 98% with ZERO liquid discharge, & produced solid gypsum byproduct for farmers 10/28/2019 Solar Powered Desalination 5. Photovoltaic power makes sense

Actual performance of the floating desalination plant 11 A floating desalination facility was introduced in 2008 until a permanent land-based facility was built in 2009. 52,000 /day USD2.27/ The Cypriot government decided to build a floating desalination as a temporary solution from 2008 to 2013. 20,000 /day EUR1.38/

In this part of the science project, you will test the performance of the desalination devices. 1. Carefully take the desalination devices outside to an area that will receive direct sunlight for at least four hours. 2. Prepare your desalination devices for testing and do a final check to make sure that everything is in place and ready.

Longair: “High Energy Astrophysics” Rohlfs and Wilson: “Tools of Radio Astronomy” Dyson and Williams: “The Physics of the Interstellar Medium” Shu: “The Physics of Astrophysics I: Radiation” Radiation Processes We can measure the following quantities: The energy in the radiation as a function of a) position on the sky b) frequency The radiation’s .