You are on page 1of 8

Journal of

Electroanalytical Chemistry
Journal of Electroanalytical Chemistry 563 (2004) 213220 www.elsevier.com/locate/jelechem

Multi-ion transport and reaction simulations in turbulent parallel plate ow


Gert Nelissen *, Achim Van Theemsche, Calin Dan, Bart Van den Bossche, Johan Deconinck
Vakgroep Electrotechniek, Vrije Universiteit Brussel, Pleinlaan 2, Brussels 1050, Belgium Received 6 May 2003; received in revised form 1 September 2003; accepted 19 September 2003

Abstract The residual distribution method is used to solve the multi-ion transport and reaction model in turbulent ow. The model describes the eects of convection, diusion, migration, chemical reactions and electrode reactions on the concentration, potential and current density distributions in an electrochemical reactor. The Reynolds averaged NavierStokes (RANS) equations are used to calculate the turbulent ow, with the turbulent viscosity obtained from a low-Reynolds number kx model. Dierent turbulence models for the turbulent mass transfer are examined and validated in a parallel plate reactor for the deposition of Cu from an acid copper-plating bath. The most accurate turbulent mass transfer models are the algebraic model for the turbulent diusion and the newly suggested model with the constant turbulent Schmidt number equal to 4.5. Furthermore, it is shown that the turbulence model for the mass transfer inuences the local concentration distribution considerably but has much less eect on the current density distribution. 2003 Elsevier B.V. All rights reserved.
Keywords: Finite element method; Turbulent mass transfer; Parallel plate reactor; Current density distribution; Concentration proles

1. Introduction An electrochemical reactor consists of at least two phases: the electrodes and the electrolyte. In many practical situations, the electrolyte solution contains a large amount of solvent (e.g., water). It is assumed that the solution is dilute, so the cross-diusion terms and dierent velocities of the individual species can be neglected when describing mass transport. In this way the general equations describing the multi-ion transport and reaction model (MITReM) for (innitely) diluted solutions are obtained, as described in detail by Newman [1]. Furthermore, it is assumed that natural convection due to concentration gradients is negligible, such that the motion of the charged species does not inuence the ow of the solvent. This means that once the uid ow is determined, the equations describing the com*

Corresponding author. Tel.: +32-2-6292811; fax: +32-2-6293620. E-mail address: gnelisse@vub.ac.be (G. Nelissen).

bined charge and mass transfer of all relevant species (charged or neutral) in a given velocity eld are solved. The mass change of all (charged and neutral) species is due to diusion, migration, convection and chemical bulk reactions. The electrode reactions rely on the local surface concentration of (some of) these species and the potential dierence between electrode and electrolyte and are introduced as boundary conditions of the transport equations. Bortels et al. [2] and Van den Bossche et al. [3,4] developed general-purpose electrochemical simulation software based on the nite volume method to solve the MITReM. They performed calculations in one-dimensional, planar and axi-symmetrical congurations for a wide variety of electrochemical systems. The model includes the eects of chemical reactions in the bulk and has been applied, among other purposes, for electrode growth on printed circuit boards. Recently timeaccurate simulations [5] and electro-osmotic calculations, where ow eects and MITReM are coupled [6], have

0022-0728/$ - see front matter 2003 Elsevier B.V. All rights reserved. doi:10.1016/j.jelechem.2003.09.012

214

G. Nelissen et al. / Journal of Electroanalytical Chemistry 563 (2004) 213220

Nomenclature ck ck;b Dk Dt E0 F j jan jca k L Ltot nel p R Rk Sct T t U concentration of species k, mol m3 bulk concentration of species k, mol m3 molecular diusion constant of species k, m2 s1 turbulent diusion coecient, m2 s1 equilibrium potential of reaction relative to MSE, V Faradays constant, 96487 C mol1 current density normal to electrode, A m2 anodic exchange current density, 25 A m2 cathodic exchange current density, 25 A m2 turbulent kinetic energy per unit mass, m2 s2 dimensionless length of the cathode length of the cathode, 0.1 m number of electrons exchanged due to copper reduction, 2 static pressure, Pa universal gas constant, 8.31 J mol1 K1 production rate of species k due to homogeneous reactions, mol m3 s1 turbulent Schmidt number absolute temperature, 298 K time, s electrical potential in the electrolyte, V uk um va xL y y zk Nk n u Greek aan aca g l m mt q swall x mobility of species k, m2 mol (J s)1 electrical potential of the metal, V average electrolyte ow inlet velocity, m s1 horizontal position of the cross-section, m normal distance to the wall, m dimensionless normal distance to the wall, as dened in [23] charge of species k mass ux of species k, mol m2 s1 unit vector normal to the electrode surface velocity vector, m s1

anodic transfer coecient, 0.5 cathodic transfer coecient, 0.25 overpotential, V dynamic viscosity, kg m1 s1 kinematic viscosity, m2 s1 turbulent viscosity, m2 s1 density of the electrolyte, kg m3 surface shear stress, kg m2 s2 dissipation of turbulent kinetic energy per volume and per time unit, s1

also been presented. Their approach was the basis for the work presented in this text. Georgiadou and co-workers [79] presented results calculated for two-dimensional geometries. The solution technique is based on the nite dierence method for general curvilinear coordinates, with an upwind scheme for the convection. This model also takes into account homogeneous reaction equilibriums and non-linear electrode kinetics. Also Volgin et al. [10,11] have performed multi-ion calculations using the nite dierence method. Most of the examples presented were rather theoretical, but some interesting comments on the stability and solution speed of the multi-ion equations were made. However, all the results presented by the authors mentioned above are limited to laminar ow conditions. Numerical treatment of turbulent mass transfer is almost always supposed to be analogous to turbulent heat transfer [1,1214]. But the Schmidt number of the reaction ions, characterizing mass transfer, is at least a thousand times higher than the equivalent Prandtl number for heat transfer. Therefore, the concentration boundary layer is much thinner than the hydrodynamic or thermal boundary layer, leaving the numerical treatment of turbulent mass transfer in arbitrary geometries far from easy, well established and understood. Several authors [1518] have presented results of mass transfer at a high Schmidt number in turbulent

ow. These contributions are either based on a boundary integral method to obtain the results in the boundary layer or solve only the convectiondiusion equation, including the turbulent diusion, to obtain the limiting current situation. Chalupa et al. [19,20], for example, calculated turbulent mass transfer at a high Schmidt number in electrochemical reactors using the k x model. In these calculations the convectiondiusion model is used to describe the mass transfer. Furthermore, turbulent diusion is neglected as only short electrodes are treated. To the authors knowledge, none of the authors mentioned above uses the full MITReM including the eects of turbulence on the mass transfer. Gurniki et al. [21,22] calculated turbulent mass transfer in nonlimiting current situations. They studied the use of large eddy simulations (LES) for predicting turbulent mass transfer in a parallel plate reactor, taking into account the turbulent uctuations of the concentration. Only a binary electrolyte is considered which allowed the solution to be obtained in terms of one equivalent concentration c z1 c1 z2 c2 . The electrode reaction rate is imposed as a gradient of the concentration while the potential gradients, including migration, are neglected. The primary objective was to obtain information about mass transfer in the nearwall region between the solid boundary and a turbulent uid ow at high Schmidt numbers.

G. Nelissen et al. / Journal of Electroanalytical Chemistry 563 (2004) 213220

215

In order to model turbulent mass transfer, the turbulent uid ow needs to be calculated. By far the most common level of turbulence modeling for industrial ows is based on the Reynolds averaged NavierStokes (RANS) equations [23]. In the simplest case turbulent uctuations are modeled through the introduction of a turbulent or eddy viscosity (EVM), which allows the Reynolds stresses (higher-order moment terms introduced by the averaging) in the RANS equations to be computed. In the literature, several turbulent uid ow models are proposed to calculate the turbulent viscosity, ranging from algebraic models (e.g., BaldwinLomax) over one equation (e.g., SpalartAlmaras) and two equation (ke and kx among others) models to Reynolds stress models, large eddy simulations and direct numerical simulations. The kx model proposed by Wilcox [23] is used in this work. It has a very good trade-o between good predictive capabilities, due to the absence of wall functions, limited need for computer resources and stability [24]. A detailed description of the governing equations, the boundary conditions and the numerical solution techniques can be found in [25,26].

is used to close the system of equations. Eq. (3) states that the charged particles cannot be separated. This condition is valid everywhere in the solution, except in the diuse double-layer, a region very close to the electrodes. The boundary conditions to be imposed are [3,25]: at insulators and symmetry planes: oU ock 0; 0; on on at electrodes for non-reacting species N k n 0; at inlets: ck ck;b ; at outlets: rck 0; rU 0: 7 At the electrodes, the ux of the active species is coupled to the speed of the electrode reactions. The driving force for these reactions is the electrode overpotential: g Um U E0 ; 8 i.e., the dierence between the imposed potential on the electrode (Um ) and the solution potential at the other side of the double layer (U ), this with respect to the equilibrium potential E0 . In general, the values of Um and U can vary along the electrode surface, respectively, due to an ohmic voltage drop in the electrode and to non-uniform current densities. The overpotential depends on the local current density and concentration(s). Often a ButlerVolmer equation j f g; ck is used to describe this dependence. MITReM allows the computation of current density distributions that range from purely secondary situations, where no mass transport problems exist (the reacting species concentrations near the electrode remain close to their bulk values), to highly tertiary current distributions (one or more reacting species concentrations may drop to zero at the electrode surface). 2.1. Turbulent mass transfer Similarly to modeling turbulent uid ow, the eects of turbulence on mass transfer are modeled by adding the turbulent diusion Dt to the molecular diusion as presented in Eq. (2). In the literature, several models are proposed to calculate this turbulent diusion based on the turbulent ow and electrochemical parameters. Most of these models are straightforward extrapolations of the models for turbulent heat transfer [17]. However, because the mass transfer boundary layer is at least one order of magnitude smaller than the hydrodynamic (or temperature) boundary layer, this similarity does not always hold. rU 0; 6 4

2. Mathematical model In order to ensure conservation of mass for each species in a diluted solution, the change of concentration in an innitesimally small control volume must be equal to the net input and the local production (or consumption) due to chemical reactions [1,2,4]: ock r N k Rk ; 1 ot with r N k the divergence of the ux and Rk the production rate of a species due to the homogeneous reaction in the solution. In this paper, no homogeneous reactions are considered, in which case the term Rk is zero. The total ux of a species is given by [1]: N k Dk Dt rck ck u zk Fuk ck rU ; 2

with Dk Dt rck the contribution of the molecular and turbulent diusion [25]; ck u the contribution of the convection and zk Fuk ck rU the contribution of the migration. The ow vector u is considered known at each position in the reactor from previous laminar or turbulent ow computations. The unknowns in the mass conservation equations are the species concentrations and the potential in each point of the solution. The electroneutrality condition,
N X k1

zk ck 0;

216

G. Nelissen et al. / Journal of Electroanalytical Chemistry 563 (2004) 213220

The turbulent Schmidt number, equivalent to the turbulent Prandtl number in heat transfer, is dened as: Sct mt =Dt : 9 As described by many authors [1214,1719,2729], mainly the following models for turbulent diusion are used for turbulent high Schmidt number mass transfer calculations: Dt 0. The turbulent diusion is neglected because it is assumed that the turbulent boundary layer is so small that all the eects happen in the laminar sub-layer of the turbulent uid ow. This assumption is valid if the length of the electrode is small [30] (Ltot p swall =lm L < 700). Dt / mt or Sct const: This is the straightforward extrapolation of what is generally done in turbulent heat transfer [15,29]. A typical value for the constant is 0.71. More elaborate models consider the dependence of Sct on some global quantities of the ow (Re, Sc, boundary layer thickness, etc.) [18]. Dt is given by an algebraic turbulence model. This is the equivalent of an algebraic turbulence model for ow. Several dierent models have been developed [12,17,31,32], most of them based on one set of measurements. Most models start from the fact, both theoretical and experimental, that the turbulent diusion varies with y 3 close to the wall. A typical example is the model proposed by Aravinth [17]: Dt 0:0007y ; 0:5 m 1 0:00405y 2
3

the mass transport equations are discretized using the Nscheme. Again standard nite element discretization is applied to the diusion, migration and homogeneous reaction terms. All the numerical schemes provide at least second-order accuracy [26]. The resulting non-linear systems of equations are solved by linearization using the NewtonRaphson method, with explicit calculation of the jacobian matrices. An incomplete LU preconditioned gmres (generalized minimal residual algorithm) is used to approximate the solution of the resulting linear system [26].

3. Results and discussion Because both the turbulent ow and mass transfer are well dened and documented, the parallel plate geometry is a very suitable test case to study the validity and accuracy of turbulent ow and turbulent mass transfer calculations. The geometry under consideration is shown in Fig. 1. The validity of the kx model to calculate the uid ow in a parallel plate reactor and the inuence of the dierent turbulence models on the limiting current density distribution has been studied extensively [25]. It was demonstrated that the calculated limiting current density distribution was strongly inuenced by the turbulence model. The algebraic model and the newly suggested model of Sct 4:5 exhibited the best agreement with the correlations for a large range of Re and Sc numbers. It has also been conrmed that turbulent mass transfer becomes only important when long electrodes are considered (L > 700). 3.1. Turbulent mass transfer using MITReM for long electrodes The turbulent diusion coecient is assumed to be the same for all species involved in the multi-ion system. This is based on the observation [25], that the turbulent diusion is apparently independent of the molecular diusion coecient. Furthermore, considering dierent turbulence models for dierent species would increase the complexity of the calculations, without any indica-

0 < y < 30; Sc P 1:

10

Also some attempts exist to use an explicit algebraic turbulent stress transport model for passive scalar variables [33,34]. Although these models seem very promising, they are much more complicated then the models presented above, and so far give no guarantee for more accurate results. Therefore, they are not considered in this work. Recently, some authors [16,35] used DNS or LES calculations for the turbulent ow and scalar transport in a channel to validate the dierent turbulence models described above. They found that the turbulent Schmidt number is close to one for y > 5. Close to the wall, Dt was found to vary with y 3 . 2.2. Numerical method All partial dierential equations presented above are solved in two dimensions using the residual distribution method [36]. The discretization is carried out on grids of triangles representing the geometrical domain of interest. The LaxWendro scheme is applied to the convection terms of the NavierStokes equations, while the viscous terms are treated in a standard nite element manner. For the kx equations, to ensure positive values for both k and x, the scalar N-scheme is applied to the convective terms. For the same reason, also the convection terms in

0.2 m
Inlet Outlet

0.01 m

0.2 m

0.1 m

Fig. 1. Schematic representation of the geometry.

G. Nelissen et al. / Journal of Electroanalytical Chemistry 563 (2004) 213220

217

tion that this would increase the accuracy. In this paper, only one turbulent ow condition at Re 8333, based on the half height of the cell and the average inlet velocity, is used for all calculations. A study on the inuence of the Re number on the turbulent mass transfer for MITReM will be the subject of future work. The electrolyte consists of 0.3 M CuSO4 + 1 M H2 SO4 , which is quite within the industrial concentration range of acid copper-plating baths. This electrolyte has a kinematic viscosity of 106 m2 s1 . The ionic properties of this electrolyte, containing three species (Cu2 , HSO and H ) are shown in Table 1, obtained 4 from [9]. Assuming that only the reduction of Cu2 takes place: Cu
2

-500

-1000

J [Am-2]

-1500

-2000

-2500

-3000 0 0.2 0.4 0.6 0.8 1

2e $ Cu #;

11
Fig. 2. Current density distribution along the cathode for dierent , 10%; , 25%; , 50%; , fractions ( 75%; , 90%; , 97%) of the limiting current,Re 8333.

the ux for ion Cu2 is written as: j : N Cu2 n nel F

12

The ButlerVolmer relation describes the electrode kinetics for the copper deposition reaction, where cCu2 is the local concentration of Cu2 at the cathode surface and cCu2 ;b the bulk concentration of Cu2 : !c   aan nel F cCu2 g jca j jan exp RT cCu2 ;b   aca nel F g : 13 exp RT For the cathodic kinetic parameters in this reaction, the values suggested by Van Den Bossche et al. [4] are used: jca 25 A m2 , aca 0:25, c 1:0. The anodic parameters of the ButlerVolmer relation are chosen to ensure a smooth transition from anodic to cathodic behavior enhancing the numerical stability: jan 25 A m2 , aan 0:5. No overpotential relation is considered (primary distribution, innitely fast reaction) at the anode, because the electrochemical processes occurring there are of no interest for this study. The current density distributions along the cathode for the deposition reaction of Cu2 for dierent fractions of the limiting current density are shown in Fig. 2. The corresponding dimensionless concentration proles of the reacting ion Cu2 at the electrode surface are given in Fig. 3. The turbulent Schmidt number is assumed to be equal to one. The limiting current is obtained by solving the convectiondiusion equation
Table 1 Ionic properties of the acid copper deposition electrolyte z Cu2 HSO 4 H +2 )1 +1 cb /mol m3 300 1300 700 109 D/m2 s1 0.61 1.065 9.312

0.8

C/Cb

0.6

0.4

0.2

0 0 0.2 0.4 0.6 0.8 1

Fig. 3. Concentration distribution along the cathode for dierent , 10%; , 25%; , 50%; , fractions ( 75%; , 90%; , 97%) of the limiting current, Re 8333.

(neglecting migration) for the Cu2 ion, with the concentration equal to zero at the cathode, as explained in detail in [25]. At low current densities no mass transfer eects are seen and the typical secondary current density distributions, including the edge eects, are obtained. As expected, the downstream part of the cathode reaches the limiting current density regime rst, because the boundary layer increases in thickness along the cathode. It is important to note that even at 75% of the total limiting current, no part of the cathode is under limiting current conditions (cCu2 0). At the leading edge of the cathode, the edge eect (involving a very high current density) decreases the concentration of the reacting ion

218

G. Nelissen et al. / Journal of Electroanalytical Chemistry 563 (2004) 213220

sharply. At the downstream edge, for the lower current densities, the expected secondary edge eect is visible. However, when the total current is higher than 90% of the limiting current, the mass transfer limits this edge eect. From Figs. 2 and 3, it is clear that the current density starts to drop towards the end of the electrode, at the moment that the concentration reaches zero. Note also the bump in the concentration prole for a total current between 50% and 97% of the limiting current, similar to that observed for laminar ow conditions [37]. This is due to a competition between an increase in the concentration by diusion of ions through the boundary layer and a decrease in concentration due to the deposition reaction. As long as the boundary layer is quite thin, the diusive transport will be dominant and the concentration will rise, but at a certain point along the cathode, the boundary layer becomes too thick and then the concentration starts decreasing. For the lower currents, the edge eect does not create a concentration gradient large enough to cause an increase in concentration due to diusion. It is clear from Fig. 3 that mass transfer starts playing an important role in the current density distribution only at fractions higher than 50% of the limiting current. Of course, depending on the electrochemical plating system under consideration, morphological dierences of the deposited layer might already have occurred at higher local concentrations. 3.2. Inuence of the turbulence model on the current density and concentration distributions In what follows, calculations using the MITReM are performed using the dierent turbulent diusion models presented above. The aim is to study the inuence of these models on the resulting current density and concentration proles. First, calculations with a constant imposed potential dierence between anode and cathode equal to 600 mV are performed. Both anode and cathode are supposed to have a constant metal potential, due to their high electrical conductivity. The resulting concentration and current density proles for four dierent turbulence models are shown in Fig. 4. It is clear that the variation in the turbulent diusion coecients according to the turbulence models leads to diering current density distributions. As the local current density is dierent for the different turbulence models, it is evident that also the concentration proles vary from one model to the other. In general the model that neglects the turbulent diusion Dt 0 yields the lowest current density and the highest concentration variation along the cathode. The model Sct 1 that overestimates the turbulent diusion, results in the highest current density and the smallest concentration variation. Similarly as for the limiting current [25], both the algebraic model and the model Sct 4:5

-250

-260

-270

-280

-290

-300

0.2

0.4

0.6

0.8

L/Ltot
1

0.95

C/C bulk

0.9

0.85

0.8
0 0.2 0.4 0.6 0.8 1

Fig. 4. Current density and concentration distribution of the Cu2 ion for an imposed potential dierence of 600 mV, Re 8333, using different turbulence models: , Dt 0; , Sct 1; , Sct 4:5; Alg. model.

exhibit almost the same behavior in between the two extremes of the other models. At the moment, to verify these results and to draw decisive conclusions on the inuence of turbulent mass transfer on the MITReM, no measurements of local concentrations at the electrode surface are available. The analogy with the limiting current case points, however, in the direction that the algebraic model and the model Sct 4:5 give the best results. To study the inuence of the turbulent mass transfer models on the concentration for equal current densities, calculations are performed where the potential dierence between the anode and cathode was varied in such a way that the same total current was obtained for the dierent turbulent models.

G. Nelissen et al. / Journal of Electroanalytical Chemistry 563 (2004) 213220

219

The resulting concentration and current density distribution for two dierent mean current densities ()500 and )1200 A m2 ) are shown in Figs. 5 and 6. From this it is clear that, as the total current is the same, the turbulence model has apparently a limited eect on the overall current density distribution. This means that measurements of the local current density distribution will not suce to compare the dierent turbulence models. However, although the current density distribution is the same, the concentration varies dramatically when the dierent turbulence models are applied. Again the analogy with the results obtained from the limiting current situation leads to the conclusion that the algebraic turbulence model and the model Sct 4:5 give the best results. The dierence in concentration depends also on the electrochemical system, as the additional

-1000

-1100

-1200

-1300

-1400

-1500 -1600

0.2

0.4

0.6

0.8

L/Ltot
1

-460
0.8

-480

-500

C/C bulk

0.6

0.4

-520

0.2

-540 0 0.2 0.4 0.6 0.8 1

0 0 0.2 0.4 0.6 0.8 1

L/Ltot
1 0.9 5 0.9

Fig. 6. Current density and concentration distribution of the Cu2 ion for an average current density of 1200 A m2 , Re 8333, using different turbulence models: , Dt 0; , Sct 1; , Sct 4:5; , Alg. model.

C/C bulk

0.8 5 0.8 0.7 5 0.7 0.6 5 0.6 0 0.2 0.4 0.6 0.8 1

Fig. 5. Current density and concentration distribution of the Cu2 ion for an average current density of 500 A m2 , Re 8333, using dierent turbulence models: , Dt 0; , Sct 1; , Sct 4:5; , Alg. model.

concentration drop needs to be compensated by a higher overpotential to yield the same current density. Fig. 7 shows the concentration proles developed normal to the cathode near the end of the electrode (xL 0:9Ltot ) for an average current density of 1200 Am2 . Note rst that the current density, mainly determined by the gradient of the concentration at the electrode surface, is the same for all proles. However, the resulting concentrations at the surface and in the diusion boundary layer dier by up to 50%. As a result, also the boundary layer, dened typically as the zone where the concentration is less than 90% of the bulk value, varies quite strongly from model to model. The smallest diusion layer thickness is about 10 lm for the Dt 0 model up to almost 30 lm for the Sct 1 model. To make further progress in the modeling of turbulent mass transfer in combination with the MITReM system, a detailed measurement of the local concentration

220

G. Nelissen et al. / Journal of Electroanalytical Chemistry 563 (2004) 213220

It would also be very interesting to verify whether these conclusions can be extended to other geometrical congurations and more complex electrochemical systems.
1

References
0.8

0.6

0.4

0.2

0 0

1 10

-5

2 10

-5

3 10

-5

4 10

-5

5 10

-5

Fig. 7. Inuence of the turbulent mass transfer model on the Cu2 ion concentration proles normal to the cathode for an average current density of 1200 A m2 , Re 8333: , Dt 0; , Sct 1; , Sct 4:5; , Alg. model.

prole throughout the diusion boundary layer would provide important information. This will allow validation and improvement of the results obtained from simulations. Furthermore, if the proles of all concentrations are measured, this would also verify that the assumption that all ions have the same turbulent diusion coecient is correct.

4. Conclusions It has been demonstrated for the rst time, that the approach of modeling turbulent mass transfer can be extrapolated from the limiting current case to the much more relevant case of mixed kinetic-transport control using the MITReM. Because evidence from the limiting current case suggests that the turbulent concentration uctuations are independent of the molecular diusion coecients, the same turbulence model is used to calculate the turbulent diusion for all ions. From the calculations, it is clear that the turbulent mass transfer model has a very strong inuence on the local concentration distribution, but that the current density distribution is much less sensitive. To make further progress, a detailed measurement of the concentration proles, both along the electrode and perpendicular to the electrode throughout the boundary layer, is necessary.

[1] J.S. Newman (Ed.), Electrochemical Systems, second ed., Prentice-Hall, Englewood Clis, NJ, 1991. [2] L. Bortels, J. Deconinck, B. Van den Bossche, J. Electroanal. Chem. 404 (1996) 15. [3] B. Van den Bossche, L. Bortels, J. Deconinck, S. Vandeputte, A. Hubin, J. Electroanal. Chem. 411 (1996) 129. [4] B. Van den Bossche, L. Bortels, J. Deconinck, S. Vandeputte, A. Hubin, J. Electroanal. Chem. 397 (1995) 35. [5] C. Dan, B. Van den Bossche, L. Bortels, G. Nelissen, J. Deconinck, J. Electroanal. Chem. 505 (2001) 12. [6] A. Van Theemsche, J. Deconinck, B. Van den Bossche, L. Bortels, Anal. Chem. 74 (2002) 4919. [7] M. Georgiadou, R.J. Alkire, J. Electrochem. Soc. 141 (1994) 679. [8] M. Georgiadou, J. Electrochem. Soc. 144 (1997) 2732. [9] M. Georgiadou, D.J. Veyret, J. Electrochem. Soc. 149 (2002) C324. [10] V.M. Volgin, O.V. Volgina, A.D. Davydov, Comput. Chem. 411 (2000) 129. [11] V.M. Volgin, A.D. Davydov, Russ J. Electrochem. 31 (2001) 1197. [12] R.H. Notter, A. Sleicher, Chem. Eng. Sci. 26 (1971) 161. [13] A.J. Reynolds, Int. J. Heat Mass Trans. 18 (1975) 1055. [14] S.W. Churchill, Ind. Eng. Chem. Res. 36 (1997) 3866. [15] Y. Wang, J. Postlethwaite, Corros. Sci. 39 (1997) 1265. [16] I. Calmet, J. Magnaudet, Phys. Fluids 9 (1997) 438. [17] S. Aravinth, Int. J. Heat Mass Trans. 43 (2000) 1399. [18] C. Rosen, C. Traegaardh, Chem. Eng. J. 59 (1995) 153. [19] R. Chalupa, M. Chen, V. Modi, A.C. West, Int. J. Heat Mass Trans. 44 (2001) 3775. [20] R. Chalupa, M. Chen, V. Modi, A.C. West, J. Electrochem. Soc. 148 (2001) E92. [21] F. Gurniki, K. Fukagata, S. Zahrai, F.H. Bark, J. Appl. Electrochem. 29 (1999) 27. [22] F. Gurniki, S. Zahrai, F.H. Bark, J. Appl. Electrochem. 30 (2000) 1335. [23] D.C. Wilcox (Ed.), Turbulence Modeling for CFD, second ed., DCW Industries, La Canada, CA, 1998. [24] K. Heyerichs, A. Pollard, Int. J. Heat Mass Transfer 369 (1996) 2385. [25] G. Nelissen, B. Van den Bossche, J. Deconinck, A. Van Theemsche, C. Dan, J. Appl. Electrochem. 33 (2003) 863. [26] N. Waterson, PhD thesis, TU Delft, Delft, 2003. [27] Q. Chen, V. Modi, Int. J. Heat Mass Transfer 42 (1999) 873. [28] V. Yakhot, S.A. Orszag, A. Yakot, Int. J. Heat Mass Transfer 30 (1987) 15. [29] S. Nesic, J. Postlethwaite, D.J. Bergstrom, Int. J. Heat Mass Transfer 35 (1992) 1977. [30] D.A. Shaw, T.J. Hanratty, AIChE J. 23 (1977) 160. [31] S. Martemyanov, E. Skurygin, J. Legrand, Int. J. Heat Mass Transfer 42 (1999) 2357. [32] T. Mizushina, F. Ogino, Y. Oko, H. Fuduka, Int. J. Heat Mass Transfer 14 (1971) 1705. [33] Y. Shabany, P.A. Durbin, AIAA J. 35 (1997) 985. [34] P.M. Wikstrom, S. Wallin, A.V. Johansson, Phys. Fluids 12 (2000) 688. [35] S.L. Lyons, T.J. Hanratty, Int. J. Numer. Methods Fluid 13 (1991) 999. [36] R. Abgrall, J. Comput. Phys 167 (2001) 277. [37] W.R. Parrish, J. Newman, J. Electrochem. Soc. 117 (1970) 43.

C/C bulk

You might also like