You are on page 1of 12

chemical engineering research and design 1 0 0 ( 2 0 1 5 ) 422–433

Contents lists available at ScienceDirect

Chemical Engineering Research and Design

journal homepage: www.elsevier.com/locate/cherd

Modeling the effect of non-ideal flow pattern on


tertiary current distribution in a filter-press-type
electrochemical reactor for copper recovery

Eligio P. Rivero a,∗ , Martín R. Cruz-Díaz a , Francisco J. Almazán-Ruiz b ,


Ignacio González b
a Departamento de Ingeniería y Tecnología, Universidad Nacional Autónoma de México, Facultad de Estudios
Superiores Cuautitlán, Av. 1o de Mayo, Cuautitlán Izcalli 54740, Estado de México, Mexico
b Departamento de Química, Universidad Autónoma Metropolitana-Iztapalapa, San Rafael Atlixco 186, México D.F.

09340, Mexico

a r t i c l e i n f o a b s t r a c t

Article history: This work presents the numerical modeling of the effect of hydrodynamics on mass trans-
Received 28 September 2014 port and tertiary current and potential distribution in a filter press type electrochemical
Received in revised form 29 April reactor used to study the copper recovery process. The operating conditions of the reactor
2015 were in turbulent regime and under charge and mass transfer mixed control.
Accepted 30 April 2015 For hydrodynamics, the Reynolds averaged Navier–Stokes equations and the stan-
Available online 11 May 2015 dard k–ε turbulence model were used. The mass transfer model was a combination of
the convection–diffusion equation and a wall function adapted for mass transfer. The
Keywords: Butler–Volmer kinetics for copper reduction, simplified Tafel equations for water oxidation
Electrochemical reactors and ohmic potential drop through the electrolyte were also incorporated into the model.
Tertiary current distribution The strategic part of the proposed numerical modeling is the concentration wall function
Turbulent mass transfer that allows linking the transport equations with Cu2+ concentration at the interface in order
Copper recovery to obtain, along with interfacial potential, the electrode kinetics. Using this approach it was
Butler–Volmer kinetics possible to model a very complex interrelation between physical phenomena and the elec-
trochemical reaction taking place in a reactor under a turbulent flow regime using moderate
computer resources. The numerical results obtained are in agreement with experimental
data of mass transfer coefficient and current–potential behavior.
© 2015 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.

1. Introduction FM01-LC, like other kinds of flat plate electrochemical reactor,


exhibits a non-homogeneous flow pattern inside the reactor
The filter press type FM01-LC electrochemical reactor is a channel caused by the reactor design, especially the inlet and
common example of flat plate laboratory scale reactor used outlet liquid distributors (Frías-Ferrer et al., 2008; Wragg and
to study electrochemical processes, such as synthesis, metal Leontaritis, 1997). Stagnant zones, channelling, bypass, recir-
recovery, recycling of industrial effluents, and water treat- culation and other observed flow phenomena (Bengoa et al.,
ment. To analyse experimental data from this kind of study 2000) give rise to variations in the mass transfer coefficient
it is important to understand the effect of flow pattern on along the electrodes.
the mass transfer from or toward the electrodes and then In the FM01-LC reactor for Cu recovery, as in other types
on the electrode reaction kinetics and electric potential. The of electrochemical reactor, these mass transfer coefficient


Corresponding author. Tel.: +52 55 56232071.
E-mail address: priveromtz@msn.com (E.P. Rivero).
http://dx.doi.org/10.1016/j.cherd.2015.04.036
0263-8762/© 2015 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.
chemical engineering research and design 1 0 0 ( 2 0 1 5 ) 422–433 423

Nomenclature
ε turbulent energy dissipation rate
A empirical constant a electrolyte potential at the anode interface
a,b kinetic constants for water oxidation  electrolyte potential
B reactor channel width c electrolyte potential at the cathode interface
C constant of the turbulence model  Karman constant
Cε1 constant of the k–ε model  viscosity
Cε2 constant of the k–ε model T turbulent viscosity
C+ dimensionless concentration  density
Ci concentration of the electroactive species electrolyte conductivity
Ci,0 inlet concentration of the electroactive species k constant of k–ε model
Ci,w wall concentration of the electroactive species ε constant of k–ε model
de equivalent diameter = 2BS/(B + S)  shear stress
Di diffusion coefficient w wall shear stress
DiT turbulent diffusion coefficient
E potential
Ea anodic potential variations are very important, not only because of the amount
Ec cathodic potential of electroactive species which can reach the electrode sur-
Eeq equilibrium potential of copper reduction face, but also because of their effect on the overpotential in
Eeq,ox equilibrium potential of water oxidation each zone of the reactor. Mass transfer rate and overpoten-
E0 equilibrium cell potential difference tial are factors involved in the type of metallic copper deposit
E˝ ohmic drop formed (compactness, grain size, dendrite and powder pro-
F Faraday constant duction) as well as the flow characteristics are involved in
F external force the shear stress over the electrode surface for carrying the
J current density metallic copper out in the liquid stream. Rough deposit is
JL limiting current density formed at local laminar hydrodynamic conditions (current
Ja current density at the anode electrodeposition lower than the limiting current density con-
Jc current density at the cathode ditions), whereas a turbulent local regime provokes dendritric
J current density vector deposits (Bazan and Bisang, 2004; Robinson and Walsh, 1991a,
J0 exchange current density 1991b). In addition, high overpotentials may lead to unde-
k turbulence kinetic energy sired side reactions that may diminish the charge efficiency of
km mass transfer coefficient the reactor and lead to uncontrollable changes (pH, conduc-
L length of the reactor channel tivity, excessive gas formation, mass transfer rate, formation
n vector normal to the wall of solid byproducts, etc.). Therefore, coupling hydrodynam-
Ni mass transfer flux ics, mass transport, electrochemical kinetics and potential
Ni,w mass transfer flux at the wall and current distribution is necessary to obtain a comprehen-
P pressure sive understanding of the electrochemical reactor behavior.
Pc factor of integration of resistances in near-wall The importance of the fluid dynamics on mass transport and
layers electrochemical kinetics has been shown by modeling elec-
R gas constant trodeposition process on rotating electrodes (Mandin et al.,
Re Reynolds number = de v/ 2007a, 2007b, 2006a, 2006b). Also, in processes with inten-
S electrode spacing sive gas evolution reactions, as in water electrolysis, there is
Sc Schmidt number = /(Di ) a strong coupling of these phenomena since the bubble for-
ScT turbulent Schmidt number = T /(Di,T ) mation diminishes the electrolyte conductivity and modifies
ScT∞ constant of the Kays–Crawford model the hydrodynamic behavior affecting the current distribution
Sh Sherwood number = de km /Di over the electrode (Mandin et al., 2014, 2005). However, model-
t time ing tertiary current distribution in an electrochemical process
T temperature may be very complex, especially for reactors operating under
u velocity vector turbulent flow. Thus, several authors have studied the prob-
u velocity lem using some simplifications: the 2D current distribution
U cell potential = Ea − Ec over flat plate electrodes and other simple geometries where
u+ dimensionless velocity = u/( w /)1/2 electrolyte flows in the laminar regimen have been reported
Vin inlet velocity (Georgiadou, 2003; Boovaragavan and Basha, 2006). Also, the
x,y,z Cartesian coordinates relationships among mass transfer, ohmic drop and overpo-
y+ dimensionless distance to the tential have been studied for copper electrodeposition in a cell
wall = y( w /)1/2 / without any type of convection (Pohjoranta et al., 2010). How-
zi number of electrons exchanged in electro- ever, very few works have studied electrochemical reactions in
chemical reaction reactors under turbulent flow, which is a very important case
from a practical point of view. Even though the characteristic
Greek letters lengths of the laboratory scale FM01-LC electrochemical reac-
˛ transfer coefficient tor are small, the design of the inlet distributor produces high
ı+ dimensionless thickness selected for wall velocity streams or jets not parallel to reactor walls, which
function = ı( w /)1/2 / are discharged inside the reactor channel causing 3D flow
424 chemical engineering research and design 1 0 0 ( 2 0 1 5 ) 422–433

instabilities and eddy formation. The mixing of these jets with


the liquid inside the reactor and the changes of velocity direc-
tion generate additional turbulences avoiding the formation
of laminar zones (Rivero et al., 2012). The solution of Navier
Stokes equations for developed turbulent flow in a flat plate
electrochemical reactor was attained by Direct Numeric Sim-
ulation (DNS) coupled with mass transport equations and
electrochemical reaction kinetics through a first order Taylor
series approach of the Butler–Volmer equation (Doche et al.,
2013, 2012). However, in the simulation performed by these
authors a Schmidt number Sc = 1 was assumed owing to the
unrealistic computational requirements under typical Sc val-
ues.
Alternatively, turbulence has been often simulated by solv-
ing the Reynolds Averaged Navier Stokes (RANS) equations
using a closure condition known as a turbulence model. The
concentration and current density over the electrode have
been calculated by 2D RANS equations, the Low Reynolds (LR)
k–w turbulence model and Butler–Volmer kinetics in a flat Fig. 2 – Hydraulic and electrical setup of FM01-LC
plate electrochemical reactor used for Cu electrodeposition electrochemical reactor.
(Nelissen et al., 2004, 2003); also, it was applied for solving 2D
RANS equations on a wire (Vande Vyver et al., 2008) and for and electrode overpotential in the copper recovery in a typi-
studying the effect of anode position (Nelissen et al., 2007). The cal flat plate electrochemical reactor working under mass and
low Reynolds turbulence models use damping functions for charge transfer mixed control. To our knowledge there are not
simulation of the near-wall regions, but requires an extremely reported works dealing with the use of mass transfer wall
fine mesh size for describing the fluid movement. Moreover, functions to obtain a complete 3D tertiary current distribu-
considering typical Sc values the mass transport in such near- tion in electrochemical reactors. Thus the present work aims
wall region involves concentration changes in a thinner layer to fulfill a gap that is relevant in electrochemical engineering
of fluid that an even finer mesh is needed. In addition, it is very and can be useful for evaluating reactor performance.
often necessary to perform simulations in 3D to capture the
effect of non-ideal flow pattern on the local mass transport 2. Experimental
rate toward or from the electrode in real reactor, making the
calculations of RANS equations and LR models highly com- 2.1. Equipment and setup
puter demanding (Santos et al., 2010).
On the other hand, it is well known that the combination The configuration of the FM01-LC cell has been described else-
of RANS equations and analytical wall functions is a very effi- where (Bengoa et al., 2000; Griffiths et al., 2005). An exploded
cient way to find a velocity field description. Furthermore, we view of the cell is represented in Fig. 1 along with spacer
have shown (Rivero et al., 2012) that a Launder–Spalding wall geometry showing the inlet/outlet manifolds. The channel is
function adapted for mass transfer, using a parameter that 0.16 m long, 0.04 m wide and 0.011 m thick (gaskets included).
takes into account the change of mass transfer resistance from The electrodes consisted of a 316 stainless steel plate and
the viscous sublayer to the turbulent layer, can predict the a Pb–Ag plate as cathode and anode, respectively. The reac-
mass transfer coefficient in the reduction of ferricyanide ion tor was installed in a recirculation system (Fig. 2) formed by
in the FM01-LC electrochemical reactor. Hence, the present a 4 L reservoir made of polycarbonate, a 0.25 HP centrifu-
work propose to use this kind of mass transfer wall function gal pump and rotameter-type flow meters. The electrodes
to solve the interfacial electrode kinetics and the tertiary cur- were connected to a Bio-logic Science Instruments VMP3TM
rent distribution. The goal is to examine by simulation the potentiostat–galvanostat coupled with a 20A–20 V VMP3B
effect of non-ideal flow pattern on the local mass transfer booster. The Cu(II) concentrations were determined employ-
ing a VarianTM Atomic Absorption Spectrophotometer Model
220FS.

2.2. Experimental procedure

Experiments were performed in an undivided continuous-


flow FM01-LC cell by using the Cu2+ reduction reaction on
the flat plate electrode. Aqueous solutions of 1 M H2 SO4 , con-
taining 18.8 mM CuSO4 , were prepared with Milli-Q deionized
water and analytical reagents. Recently prepared solutions
were loaded into the electrolyte recirculation system reser-
voir and were bubbled with nitrogen for 10 min and nitrogen
atmosphere was preserved during the experiments. All elec-
trolyses were conducted in potentiostatic mode at T = 298 K. A
saturated sulfate reference electrode (SSE) was connected to
Fig. 1 – Exploded view of the FM01-LC reactor and details of a Luggin capillary, which was in contact with the electrolytic
the spacer. solution by means of a supporting device placed in the middle
chemical engineering research and design 1 0 0 ( 2 0 1 5 ) 422–433 425

of the hydraulic channel. Such a configuration ensures a (5) only Cu2+ reduction at the cathode and water oxidation at
more uniform flow and electric potential distribution so the anode are considered.
that mass transport measurements are only associated with
the flow pattern in the hydraulic channel. For experimental 3.1. Fluid flow
determination of the limiting current, polarization curves
were constructed using sampled current density vs. poten- For the purposes of the present work, the bulk-flow dynamics
tial. It is important to mention that the authors decided to can be approximated by the Reynolds averaged Navier Stokes
construct the polarization curves in this way instead of by equation (Eq. (3)) using turbulence models as closure condition
using the linear sweep voltammetry technique, because a since it is still an effective and practical tool to approach this
voltammetric study always involves competition between the kind of problem (Bernard and Wallace, 2002; Eςa and Hoekstra,
rate of electrochemical reaction at the interface and the rate 2010; Lacasse et al., 2004; Trujillo et al., 2009). The standard
of potential change. This competition is eliminated in the k–ε turbulence model and the balances of k and ε given by
potential pulse technique. Eqs. (4)–(6) are used in this work. Since this approach is not
The polarization curves have been constructed as follows: applicable in the proximity of solid surfaces, log-law based
a series of potential double pulses was applied to the elec- wall functions are commonly used as boundary conditions
trochemical reactor. The direct pulse was started from open (Bernard and Wallace, 2002; Eςa and Hoekstra, 2010; Lacasse
circuit potential OCP (−0.34 versus SSE) in the potential range et al., 2004), i.e. Eq. (3) is solved only in the fluid bulk sepa-
of −1.0 ≤ E ≤ −0.4 V versus SSE during 30 s, with increments of rated by a distance ı from walls where the velocity is linked
0.05 V. The inverse pulse was 0.58 V to ensure the oxidation to the wall function (Eq. (7)). This wall function based on the
of the previously deposited copper. The corresponding cur- universal velocity distribution in the viscous, buffer and turbu-
rent time curves were obtained at different volumetric flows lent layers defines the parallel-to-wall velocity in the turbulent
between 2 and 7 LPM (1384 ≤ Re ≤ 4845). From these current layer as a function of its distance to the wall, where molecu-
time curves obtained for direct pulses at the applied poten- lar viscosity is considered negligible compared to turbulent
tial, the current density was sampled at 15 s and polarization viscosity.
curves were constructed (J–E) for a specific flow rate. The
T
 
sampled time ensures both a negligible contribution of the (u · ∇) u = −∇P + ∇ · ( + T ) ∇u + (∇u) + F (3)
capacitive current to the measured current and a pseudo-
steady state (the current is time-independent). k2
T = C (4)
ε
3. Mathematical model and simulation  T
  1  2
u · ∇k − ∇ · + ∇k = T ∇u + (∇u)T − ε (5)
k 2
Copper recovery in a continuous bench scale electrochemical
reactor may be carried out under either potentiostatic or gal-
 T
 
vanostatic conditions where copper reduction (Eq. (1)) takes
u · ∇ε − ∇ · + ∇ε
place with a minimum effect of side cathodic reactions (hydro- ε

gen evolution reaction, HER). Pb–Ag (95.5–05 mol%) alloy plate 1 ε  2 ε2


previously anodized in sulfuric medium (to form a PbO2 film) = Cε1 T ∇u + (∇u)T − Cε2 (6)
2 k k
was used as anode in the FM01LC reactor. In this way, the reac-
tion taking place at the anode concerns only water oxidation
that forms oxygen gas bubbles (Eq. (2)). 1
u+ = 5.5 + ln y+ (7)

Cu2+ + 2e− → Cu (1)
where u is the average velocity, P is the pressure,  is the
viscosity, T is the turbulent viscosity, F is any external force, 
H2 O → 1/2O2 + 2H+ + 2e− (2) is the fluid density, C is a turbulence model constant (C = 0.09
to guarantee consistency with the log-law of the wall (Bernard
The electrolyte flows into the reactor in the turbulent and Wallace, 2002)), k is the eddy kinetic energy, ε is the tur-
regime from the inlet manifold, spreading across the entire bulent energy dissipation rate, Cε1 , Cε2 , k and ε are empirical
volume of the reactor channel and finally leaving through constants, u+ is a dimensionless velocity (=u/( w /)1/2 ), u is the
the outlet. Once the electrolysis starts, cupric ions are trans- velocity,  w is the wall shear stress and y+ is a dimensionless
ported from the bulk to the electrode surface at a local distance to the wall (=y( w /)1/2 /) and  is the wall-law pro-
hydrodynamics-dependant rate. Thus, non-homogenous portionality constant known as Karman constant. Constants
interfacial cupric ion concentration and overpotential over of the model are shown in Table 1.
the electrode surface are expected leading to a different The boundary conditions used in Eq. (3) are as follows:
electrochemical reaction rate or even different electrochem-
ical reactions. To keep the model as simple as possible, the • u = −Vin n, at the reactor inlet, where n is the normal vector
following assumptions are made: (1) the reaction extent of at the boundary.
 
metallic Cu deposition is small enough that changes on reac- • ( + T ) ∇u + ∇uT n = 0, P = 100 kPa, at the reactor outlet.
tor channel dimensions and bulk velocity field are negligible, • Logarithmic function (Eq. (7)) for y+ = ı+ at all walls.
(2) hydrodynamic and conductivity changes caused by gas
evolution reaction at the anode are not considered (a reason- This last boundary condition means that the thin layer next
able approximation when low current density and high flow to all walls in the reactor is not included in the simulation
rates are used), (3) the process is in a pseudo-steady state, (4) domain where Eq. (3) will be solved. The dimensionless thick-
there are no spatial variations of electrolyte properties, and ness of this thin layer ı+ should be greater than 30 in order
426 chemical engineering research and design 1 0 0 ( 2 0 1 5 ) 422–433

Table 1 – Electrolyte properties and parameters used in the model.


Parameter Value References
−3
Density () 1059 kg m Hotlos and Jaskula (1988)
Viscosity () 0.001 kg m−1 s−1 Hotlos and Jaskula (1988)
Diffusion coefficient of Cu2+ (Di ) 7.2 × 10−10 m2 s−1 Newman (1991)
Conductivity ( ) (this work) 39 S m−1 This work
Constants of k–ε turbulence model
C 0.09 Bernard and Wallace (2002)
Cε1 1.44 Bernard and Wallace (2002)
Cε2 1.92 Bernard and Wallace (2002)
k 1 Bernard and Wallace (2002)
ε 1.3 Bernard and Wallace (2002)
Karman constant () 0.41 Bernard and Wallace (2002)
Constant for turbulent Schmidt (ScT∞ ) 0.85 Weigand et al. (1997)
Equilibrium potential of copper reduction (Eeq ) 0.288 V This work
Exchange current density (J0 ) 1 A m−2 Nelissen et al. (2003)
Kinetic constant of water oxidation (a) 0.17 A m−2 Zeng and Zhang (2010)
Kinetic constant of water oxidation (b) 12.03 V−1 Zeng and Zhang (2010)
Transfer coefficient (˛) 0.25 Nelissen et al. (2003)

to locate the boundary in the turbulent layer and the logarith- where u is the averaged velocity vector, Ci the averaged con-
mic velocity distribution given by Eq. (7) could be applied. Fig. 3 centration field, Di the diffusion coefficient and Di,T the eddy
illustrates this wall function along with the location of viscous diffusivity or turbulent diffusivity. This last term can be deter-
(y+ ≤ 5), buffer (5 ≤ y+ ≤ 30) and turbulent (y+ ≥ 30) layers, the mined from the turbulent Schmidt number (ScT = T /(Di,T ))
mass transport wall conditions and other model conditions given by the following Kays Crawford model as described pre-
for the cathode surface. viously (Rivero et al., 2012).

   −1
1 0.3 T
T
2 −Di
ScT = +√ − 0.3 1 − exp  √  (9)
2ScT∞ ScT∞ Di Di 0.3T ScT∞

where ScT∞ = 0.85 (Weigand et al., 1997).


3.2. Mass transport The model proposed by Eq. (8) has to be solved in the
same fluid region where RANS equations are solved, i.e. sepa-
The mass transport model is derived considering the migra- rated y+ = ı+ dimensionless distance from the wall (see Fig. 3)
tion term of the Nernst–Planck equation to be negligible (Lacasse et al., 2004), because turbulent viscosity and average
because of the high concentration of supporting electrolyte. velocity field need to be known. In the unsolved near-wall
In addition, under turbulent flow conditions, the concentra- region of the electrode, a distribution of dimensionless con-
tion and velocity fluctuation terms can be grouped into a centration C+ can be considered.
so-called turbulent diffusion term (·Di,T Ci ) of an averaged The following Launder–Spalding distribution was adapted
diffusion–convection equation analogous to the RANS equa- to the mass transfer case (Trujillo et al., 2009) and is used in the
tion, as follows: present work, where Eq. (10) corresponds to the distribution
in the viscous sublayer and Eq. (11) to the distribution in the
turbulent layer.
−u · ∇Ci + ∇ · (Di + Di,T ) ∇Ci = 0 (8)

Ci + = Sc y+ (10)

1 
Ci + = ScT ln y+ + 5.5 + Pc (11)


where Ci + is defined by:

(Ci,w − Ci ) C 1/4 k1/2


Ci + = (12)
Ni,w

where Ci,w is the fluid concentration at the electrode surface,


Ci is the concentration at a distance y+ = ı+ from the wall, Ni,w
is the diffusion flux in the near-wall region and the term Pc
is given by the following empirical equation, which takes into
account the change of resistance to mass transfer from the
viscous sublayer to the turbulent layer:

Fig. 3 – Schematic of near-cathode region showing the  Sc 3/4


variables at the interphase and their relationship with Pc = A −1 (13)
velocity and copper ion concentration at y+ = ı+ . ScT
chemical engineering research and design 1 0 0 ( 2 0 1 5 ) 422–433 427

where A is a constant whose value (A = 9.24) has been deter- The ohmic potential drop can be obtained from the con-
mined from heat transfer experimental data in smooth tubes tinuity equation for the electrolyte between the electrodes,
(Jayatillake, 1969). according to
Considering the transport of Cu2+ ion toward the cathode
(i.e. subscript i is equal to Cu2+ ) where a reduction reaction ∇ · J = −∇ · ( ∇) = 0 (17)
takes place, the boundary conditions of the mass transport
model are: where J is the current density vector and the electrolyte
conductivity.
As the overpotentials are given by the Butler–Volmer equa-
• Ci = Ci,0 , at the reactor inlet.
tion for copper reduction (Eq. (14)) and Tafel equation for water
• n·(−Di Ci ) = 0, at the reactor outlet (convective flux)
oxidation (Eq. (15)), Eq. (17) applies for the potential between
• Wall function (Eqs. (11) and (12)) at the electrode, with
c and a . Assuming constant cathodic and anodic potentials,
Ni,w = Jc /zi F.
Ec and Ea respectively, the interface electrolyte potentials, c
• n·(−Di Ci + uCi ) = 0, in all other (impermeable) walls.
and a , vary over the electrode and are unknown. Therefore,
Dirichlet type boundary conditions of Eq. (17) for the electrode
where Jc is the local current density at the cathode, zi the num- interface are not possible, instead Neumann type conditions
ber of electrons in the reduction reaction (two electrons in this are to be used as follows.
case) and F the Faraday constant.
• −n · J = Jc , at the cathode.
3.3. Electrode kinetics • −n · J = Ja , at the anode.
• n · J = 0, at all other boundaries.
The most accepted mechanism for Cu2+ reduction involves
two successive monoelectronic transfer reactions, adsorp- where Jc and Ja are the local current density given by the kinetic
tion processes and ion migration over the electrode surface. equations (Eqs. (14) and (15)), which depend on overpotential.
Even though other mechanisms have been reported and there With these equations the mathematical development to
has not been a clear discrimination among several reaction obtain the tertiary current distribution in an electrochemical
sequences to elucidate the real mechanism, the experimental reactor for copper recovery under turbulent flow is complete.
electrodeposition rate data might be fitted to Butler–Volmer
kinetics, according to: 3.5. Simulation

zi (1 − ˛)F (Ec − c − Eeq ) The equations for hydrodynamic, mass transport, electrical
Jc = J0 exp
RT continuity, electrode kinetics and concentration wall func-
tion form a coupled set of partial differential equations and
CCu2+ ,w −zi ˛F (Ec − c − Eeq ) algebraic equations. However, since hydrodynamic changes
− exp (14)
CCu2+ RT caused by anodic and cathodic reactions are not considered
(assumptions 1 and 2), the hydrodynamic model was solved
where J0 is the exchange current density, Ec the cathodic first. It was made by use of finite elements using commercial
potential, c the electrolyte potential at the cathode inter- software (COMSOL Multiphysics® ) as follows.
face, Eeq the equilibrium potential, CCu2+ ,w the interface Cu2+ The hydrodynamic model was solved with the option of the
concentration, CCu2+ the bulk concentration, R the gas con- k–ε turbulence model, momentum transport, using calculation
stant and T the temperature (parameters are given in Table 1). options by default: stationary segregated solver with BICGStab
Eq. (14) corresponds to a mixed control (kinetics and diffu- iterative methods for u and P, and GMRES for logarithms
sion). The potential range for the copper electrodeposition in a of the turbulence model variables, log(k) and log(ε). Galerkin
mixed control was previously determined in a microeletrolysis Least Squares (GLS) was used as a stabilization method; tol-
study. In this potential range, HER is negligible. erance was fixed at 10−3 . Upon verifying numerical solutions
Assuming that oxidation of the medium (water) follows a reached at different values of ı+ and mesh size, the ı+ value
Tafel expression: was fixed at 100 (typical value used within the validity of
the logarithmic velocity distribution law, assuming no effects
Ja = a exp (b (Ea − a − Eeq,ox )) (15) of roughness caused by copper deposits) and the maximum
mesh size in the reactor channel at 2 × 10−3 m. Fig. 4 shows
the mesh in the reactor inlet used in simulation (37,042 total
where Ea is the anodic potential, a the electrolyte potential
mesh points and 185,250 tetrahedral elements). The results
at the anode interphase, Eeq,ox the equilibrium potential of
obtained were stored through the option available in the pro-
the water oxidation, a and b are empirical constants shown in
gram, in order to be used in the mass transfer model. The
Table 1.
diffusion–convection equation (Eq. (8)) along with mass trans-
port wall functions given by Eqs. (11)–(13) were then added to
3.4. Electric continuity
the model, linking the results of the hydrodynamic calcula-
tion: (i) the time-averaged velocity to evaluate the convection
The cell potential defined by the voltage difference (Ea –Ec ) is
term in Eq. (8), (ii) the turbulent viscosity to obtain turbulent
given by the sum of absolute values of overpotentials at both
diffusivity through the turbulent Schmidt number, ScT , and (iii)
electrodes plus the ohmic drop E˝ and the equilibrium cell
the eddy kinetic energy necessary for the concentration wall
potential difference E0 , as follows:
functions (Eqs. (11) and (12)).
The mass transport model requires the current density Jc
U = (Ea − a − Eeq,ox ) + E˝ − (Ec − ϕc − Eeq ) + E0 (16) at the boundary conditions, which is set by Cu2+ reduction
428 chemical engineering research and design 1 0 0 ( 2 0 1 5 ) 422–433

Fig. 4 – Mesh configuration of the simulation domain in the


Fig. 5 – Simulation of the velocity field inside the FM01-LC
FM01-LC reactor.
reactor channel at Re = 4845 showing the effect of inlet
distributor on the flow pattern and jet formation.
kinetics given by the Butler–Volmer equation. But, this last
equation is a function of interfacial variables: Cu2+ concentra-
tion, Ci,w , and potential, c . The first variable, Ci,w , is related
to the concentration at the boundary of the mass-transport
1.4
model domain Ci (y+ = ı+ ) by the concentration wall function 1.2
(see Fig. 3 and Eqs. (11) and (12)). The second variable, c , is
related to the interfacial potential at the anode, a , by the 1.0 z1
Velocity/m s-1

ohmic potential drop through the electrolyte given by the elec- 0.8
trical continuity equation (Eq. (17)). However, as the potential
drop is a function of the current density it is necessary to eval- 0.6
uate the boundary conditions of Eq. (17), Jc and Ja , from the z2
0.4
kinetic equations, Eqs. (14) and (15), respectively. The kinet-
ics of anodic reaction given by Eq. (15) are only related to one 0.2
interfacial variable: a , and must satisfy the current continuity
criterion i.e. anodic and cathodic currents are equal, although
0.0
z3
the current density distribution over the anode, Ja , could be -0.2
different from that of the cathode Jc . Taking into account all 0 0.01 0.02 0.03 0.04
these relationships the complete model can be solved for given
Cell width, y/m
electrode potentials, Ea , Ec and inlet Cu2+ concentration. The
mass transport and electrical continuity models were solved Fig. 6 – Axial velocity profile as a function of channel width
simultaneously by finite elements along with kinetics equa- at 0.005 m (continuous line) and 0.01 m (broken line) from
tions for Jc and Ja and the concentration wall function, using the channel inlet and different distances from the cathode,
COMSOL Multiphysics® . The solver employed was iterative, z1 = 0.005 m, z2 = 0.008 m and z3 = 0.01 m; Re = 4835.
FGMRES, and a relative tolerance of 1 × 10−6 was established.
The values of ı+ and mesh size were as those of the hydro- profiles at different distance from the cathode, z1 = 0.005 m,
dynamic model. Simulations were done using a 4-core Intel i7 z2 = 0.008 m and z3 = 0.01 m, and from the inlet, 0.005 m (solid
920 processor at 2.67 GHz, 12 GB RAM. The values of electrolyte line) and 0.01 m (broken line) are shown in Fig. 6, where
properties and model parameters used to solve the model are negative velocity indicates a recirculation pattern. These flow
shown in Table 1. instabilities caused by the inlet distribution and the absence
of calming zone make it impossible for the flow to be laminar
4. Results and discussion even at low Re. Also, it is clear that a fully developed flow
is not reached in the short axial length of the FM01-LC. All
4.1. Hydrodynamics these characteristics of the Reynolds averaged velocity field
reveal complex velocity variations in all directions making
There has been a number of works dealing with hydrody- impossible to find symmetries for performing 2D simulations
namics inside flat plate electrochemical reactors; some of so that simulations in 3D are necessary.
them have specifically treated the FM01-LC reactor (Rivero
et al., 2012; Vázquez et al., 2010). As they commonly used 4.2. Mass transfer assessment
CFD to describe the reactor behavior we only show the
main characteristics investigated in this work. The manifold Before the model for tertiary current distribution can be used
effects, mainly those of the inlet manifold, on the velocity for simulating Cu reduction in a FM01-LC reactor, the mass
field are shown in Fig. 5, where jets of different velocity are transfer rate should be adjusted by experimental results. The
clearly formed causing preferred flow paths, recirculation mass transport model given by Eqs. (8)–(13) uses an empirical
and low velocity regions. Two high velocity streams, one parameter (A) for correcting the concentration wall function.
higher than the other, close to lateral walls and a low veloc- The A-parameter value has been reported for heat transfer
ity region in between can be identified. The axial velocity over smooth surfaces (A = 9.24). This value has allowed the
chemical engineering research and design 1 0 0 ( 2 0 1 5 ) 422–433 429

2.4 of mass transfer rate caused by roughness of the copper elec-


trodeposits. This is in agreement with our previous results
for copper electrodeposition at the surface of a rotating cylin-
2.3 der electrode (Rivero et al., 2010). Although, there is a slight
difference in the A value from this last case (A = 2.9), this
log(Sh/Sc0.33)

2.2 is explained on the basis of different geometry and rela-


tive movement of fluid and solid surface making different
proportionality between the dimensionless concentration dis-
2.1 tribution Ci + and the mass transport resistances of the viscous,
buffer and turbulent layers. It is important to note that the
2 possible variations of roughness caused by different rates of
copper electrodeposition have no appreciable effect on the
values of parameters A and those of Eq. (19) (coefficient and
1.9
exponent of the Re). This is in agreement with the typical cor-
3 3.2 3.4 3.6 3.8 relations for mass transfer reported in the literature where no
log(Re) variation of parameters of these correlations in copper elec-
trodeposition is obtained upon varying the electrodeposition
Fig. 7 – Mass transfer correlation in copper reduction at the
rate.
FM01-LC reactor determined by simulation (continuous
The exponent of the Re in the empirical correlation (Eq.
line) and comparison with experimental results.
(19)) for the mass transfer coefficient is expected to be 0.8 for
fully developed turbulent flow on flat plates (Bird et al., 2007).
description of experimental mass transfer coefficients for fer-
The lower value found in this study (0.785) can be related to
ricyanide ion reduction over smooth electrodes (Rivero et al.,
the fact that in the small axial distance of the FM01-LC reac-
2012). In the case of copper reduction, a rough surface is
tor (0.16 m), fully developed flow is not reached as shown in
formed over the electrode owing to the appearance of the cop-
Fig. 5. On the other hand, this exponent of Re is slightly higher
per metallic phase, so a different A-value is expected which
than those reported by Griffiths et al. (2005) (0.73), Brown et al.
has yet to be determined experimentally.
(1993) (0.71) and Hammond et al. (1991) (0.68) for the FM01-
Under mass-transfer controlled conditions, the current
LC reactor using the ferrocyanide/ferricyanide system. This
density value is related to the maximum copper ion flux
difference can be associated with the higher Re used in the
toward the cathode surface. This limiting current density is
present study that prevents the formation of laminar zones in
not dependent on applied potential at each Re, and the mass
the reactor, and the presence of micro-turbulences in the near-
transfer coefficient can be determined from its value by the
electrode region induced by the copper deposits; both factors
following equation:
promote turbulences and, consequently, higher values of the
JL Reynolds exponent.
km = (18)
zi FCi
4.3. Potential and current distribution
where JL is the limiting current density at the cathode, zi the
number of electrons involved in the electrochemical reaction, Using the results of hydrodynamics, Section 4.1, and the value
F is the Faraday constant and Ci the bulk copper ion concen- of A-parameter determined from experimental mass trans-
tration. fer coefficient, Section 4.2, the copper reduction process was
Simulations were performed under the same experimental simulated under mass and charge transfer mixed control by
conditions searching for the A parameter value of the mass simultaneous solution of the mass transport, electrical con-
transport model. At the limiting current conditions the cop- tinuity and anode and cathode kinetic equations. The results
per ion concentration at the interface approaches zero, thus obtained are illustrated in Figs. 8–11 for FM01-LC operation
the mass transfer model might be solved independently of the under the same experimental conditions: Re = 4845, 18.8 mM
electrode kinetics and current continuity model. Therefore, inlet copper ion concentration and potential (−0.3 V vs. SHE in
the simulations were performed using only hydrodynamic Figs. 10 and 11) between the cathode and the reference elec-
and mass transport models to obtain the mass transfer flux trode placed in the middle of the reactor channel as described
distribution over the electrode and the global mass transfer in Section 2.
coefficient according to a previously reported methodology The concentration Ci,w at the cathode interface (Fig. 8),
(Rivero et al., 2012). Using a trial and error procedure, the A given by the concentration wall function, reveals the zones
parameter value was determined at 2.3 by matching experi- where the limiting current is reached (Ci,w approaches zero)
mental and calculated mass transfer coefficients. as the potential becomes more negative. The effect of the flow
Fig. 7 shows the experimentally determined mass transfer pattern produced by the inlet distributor on the interface con-
coefficient values along with the simulation results (contin- centrations is clearly observed.
uous line) in dimensionless form. The empirical correlation The ohmic potential drop across the electrolyte
found is (E˝ = a − c ) calculated from Eq. (17) along the x-axis
(reactor flow direction) at different cathodic potentials (−0.06,
Sh = 0.304Re0.785 Sc0.33 (19) −0.13, −0.21 and −0.3 V vs. SHE) is shown in Fig. 9. This nor-
mal to the electrode surface potential drop shows variations
The A parameter value found (A = 2.3), lower than those through the reactor: it is large at the entrance, but diminishes
reported for heat transfer in smooth pipes (Jayatillake, 1969) in the flow direction until reaching a minimum and, finally,
and mass transfer in a smooth flat plate electrode (Rivero et al., increase slightly near the channel exit. This potential drop
2012) (A = 9.24 in both cases), accounts for the enhancement distribution is in agreement with hydrodynamic behavior in
430 chemical engineering research and design 1 0 0 ( 2 0 1 5 ) 422–433

Fig. 10 – Total overpotential over cathode surface of the


FM01-LC in Cu2+ reduction. Inlet Cu2+
concentration = 18.8 mM, Re = 4845, potential between the
cathode and reference electrode = −0.3 V vs. SHE (see Fig. 2).

The contour lines of the total cathodic overpotential (con-


centration and activation overpotential), calculated as the
difference between the potential of the electrode Ec and the
potential of the electrolyte at the interface c , are depicted in
Fig. 10. Such overpotential distribution shows the strong effect
of hydrodynamics on the mass transport of copper ions toward
the cathode coupled to the electrochemical reaction rate and
the electrical potential. The higher overpotential lines corre-
spond to lower turbulence regions and vice versa, following
the flow pattern caused by the geometry of the reactor, mainly
the inlet and outlet distributor.
Fig. 11 shows the current density distribution at both
Fig. 8 – Simulation of Cu2+ concentration at the cathode
electrodes. The cathodic current distribution exhibits a strong
interface when different potentials between the cathode
influence of the inlet and outlet distributor; the five jets
and reference electrode placed in the middle of the reactor
formed (see Figs. 5 and 6) produce regions of low mass trans-
channel (see Fig. 2) are applied (potential values indicated
fer resistance decreasing the concentration polarization, and
in the figure vs. SHE). Inlet Cu2+ concentration = 18.8 mM,
allowing for an increase in current density, but at the same
Re = 4845.

the reactor where higher turbulence regions produce higher


mass transfer rate and higher current density. Moreover, the
change of potential drop with cathodic potential depicted
in Fig. 9 is consistent with results of concentration at the
interface shown in Fig. 8, i.e. in zones where concentration
at the interface approaches zero a minor change of potential
drop with cathodic potential is observed. It is clear that a
simple primary distribution is insufficient to describe the
behavior of the potential drop in the reactor.

0.16 -0.3 V vs. SHE


-0.21
-0.13
Potential drop/V

0.12 -0.06

0.08

0.04

0 Fig. 11 – Current density distribution over the cathode


0 0.04 0.08 0.12 0.16 surface for copper reduction and its effect on the anodic
Distance, x/m current distribution for water oxidation given by coupling
the hydrodynamic-dependent mass transport, electrode
Fig. 9 – Axial variation of the ohmic potential drop along kinetics and electrical continuity (conditions given in
the reactor channel (conditions given in Fig. 8). Fig. 10).
chemical engineering research and design 1 0 0 ( 2 0 1 5 ) 422–433 431

0
-10

Current density, Jc/A m-2


-100
-25
-200
-40
-300
-55

-400 -70

-500 -85
Fig. 12 – Patterns of copper electrodeposition over the -0.5 -0.25 0 0.25
electrode of the FM01-LC reactor at (a) Re = 2769 and 5 min Potential, Ec vs SHE/ V
electrolysis time, and (b) Re = 1384 and 10 min electrolysis
time. Fig. 13 – Average current density obtained from copper
electrodeposition simulations at FM01-LC reactor under
different cathodic potentials (solid line) and experimental
sampled current (dots). Results of simulation (broken line)
time this high current density demands a higher overpo-
of a reactor channel with homogenous inlet velocity and
tential according to the Butler–Volmer equation. In addition,
constant sectional area are shown for comparison.
high current density regions provoke high ohmic potential
Conditions: 18.8 mM CuSO4 in 1 M H2 SO4 , Re = 4845.
drop diminishing the available overpotential in potentiostatic
experiments so that a dynamic balance is reached at the
steady state. The final results are a low overpotential and Fig. 12 shows the patterns of the copper deposits on the
high current density (absolute values) regions at the electrode electrode surface obtained after 5 and 10 min of electrolysis.
surface at the entrance of the reactor (Figs. 10 and 11). There The regions covered by copper follow the current density dis-
is then a calm region, where a high overpotential and low tribution pattern at the cathode shown in Fig. 11. At short
current density are evident. It is worth noting that the mass times the figure reveals the effect of jets on copper deposition.
transfer resistance is due not only to the low local velocity Also, the effect of the two high velocity streams near the chan-
but also to the low turbulence intensity. nel wall (Fig. 5) and the consequently high current density at
At the anode, the electrochemical reaction is not the cathode edges (parallel to the flow direction in Fig. 11), is
dependent on mass transport of reactants, however, the non- displayed in Fig. 12a. At longer electrolysis times (10 min), the
uniformity of current density at the cathode surface produces copper deposit shape follows the high and low current den-
an anodic current density distribution due to the electrical sity zones (Fig. 12b) depicted by simulation in the cathode in
continuity requirement. Nevertheless, the variations of cur- Fig. 11.
rent density at the anode are smoother than those at the With regard to simulations at different cathodic potentials,
cathode showing a steady decrease of current density from Fig. 13 shows the average current density obtained by surface
the inlet in the axial direction that reaches a minimum at integration of the current density over the electrode divided
13.8 cm from the inlet and then a slight increase close to the by the electrode area at each steady state simulation (solid
outlet. Since the overpotential depends on the current density line) and the comparison with experimental data (dots).
distribution, an ovepotential distribution at the anode is also Even though the current distributions are not uniform on the
evident. It is worthy of note that the oxygen evolution at the electrode, the average current densities calculated are close to
anode was not taken into account in the model. Although gas the experimental data. This fact proves the capability of the
bubble production may lead to several changes like decreas- approach used in the model. On the other hand, the very wide
ing the effective conductivity and modifying the flow pattern, sigmoidal curve (solid line) depicted in Fig. 13 represents the
their effects were considered minor as the high flow used can wide current density distribution on the electrode, i.e. at not
carry the bubbles away rapidly. very negative cathode potential some areas of the electrode
The general behavior obtained by simulation and depicted surface reach the limiting current density while others (with
in Figs. 8–11 for copper recovery in the FM01-LC reactor was lower mass transfer resistance) close to the entrance of the
found to be qualitatively similar to the results of simulations reactor are far from limiting current density; and the more
at different Re ranging from 1384 to 4845 at the same cathodic negative is the cathode potential, the larger is the area of
potential (with respect to the reference electrode). The jet for- the electrode surface reached by the limiting current density.
mation and the flow pattern developed determine the mass For comparative purposes Fig. 13 shows the current potential
transport rate in each zone of the reactor leading to equivalent curve (broken line) predicted by the model when ideal inlet
effects on current density and overpotential on the electrode and outlet reactor distributors are used. The simulations at
surface but with faster rate as Re increases. the same Re (Re = 4845) and inlet copper ion concentration
Experimentally, the type of copper deposits is different (18.8 mM) were performed only in the reactor channel (formed
depending on Re. At high Re, keeping the same cathodic poten- by the space between the electrodes) with a homogenous
tial, the high copper deposition rate leads to loosely adherent inlet velocity. Under these circumstances, the sigmoidal curve
copper deposits making the removal of copper deposits easier is more sharply so that the mass-transfer-controlled plateau
because of the high shear stress of the fluid flowing at high is reached at less negative potential. In this case, as expected
velocity. the current density is lower than that obtained for the actual
432 chemical engineering research and design 1 0 0 ( 2 0 1 5 ) 422–433

FM01-LC reactor since the additional turbulence generated up. The results obtained from the simulation are in agree-
by distributors in the last case leads to higher mass transfer ment with potential current experimental data before the HER
coefficients. takes place. Moreover, they allow explanation of the difference
Furthermore, as the cathode potential in the FM01-LC between the potential necessary to start the HER in the FM01-
becomes more negative, the overpotential of higher mass LC and that expected from microelectrolysis experiments. It
transfer resistance zones increases more (in absolute value) is important to note the advantages in approach presented
than the average overpotential value, which increases the risk here which allowed a complete description of the tertiary cur-
of side reactions in these zones. The experimental points at rent distribution in an electrochemical reactor under the most
more negative potential in Fig. 13 show that the current begins practical type of flow regime (turbulent flow) requiring mod-
to increase (absolute value) due to the appearance of the erate computer resources.
hydrogen evolution reaction (HER) which shows that higher
overpotential regions started to develop this side reaction. It References
is important to note that at this potential (0.386 V vs. SHE) the
HER is not expected to take place based on micro electroly- Bazan, J.C., Bisang, J.M., 2004. Electrochemical removal of tin from
sis experiments on an RDE (Rivera and Nava, 2007). However, dilute aqueous sulfate solutions using a rotating cylinder
in the FM01-LC reactor, higher overpotential regions might electrode of expanded metal. J. Appl. Electrochem. 37, 501–506.
be developing before the reference point can reach such high Bengoa, C., Montillet, A., Legentilhomme, P., Legrand, J., 2000.
overpotential owing to the non uniform overpotential distribu- Characterization and modeling of the hydrodynamic behavior
tion. Clearly, under these conditions the model deviates from in the filter-press-type FM01-LC electrochemical cell by direct
flow visualization and residence time distribution. Ind. Eng.
the experimental data as the side reactions are not taken into
Chem. Res. 39, 2199–2206.
account in the model, but the implications of the model calcu- Bernard, P.S., Wallace, J.M., 2002. Turbulent Flow: Analysis,
lations remain valid since regions of higher overpotential than Measurement, and Prediction. John Wiley & Sons, New Jersey.
that where the potential is measured exhibit an overpoten- Bird, R.B., Stewart, W.E., Lightfoot, E.N., 2007. Transport
tial high enough to drive the reduction of protons. Moreover, Phenomena, second ed. John Wiley & Sons, USA.
the overpotential and current density distribution explain the Boovaragavan, V., Basha, C.A., 2006. A novel approach for
computing tertiary current distributions based on simplifying
experimental results. The picture described for the overpo-
assumptions. J. Appl. Electrochem. 36, 745–757.
tential and current density distribution where the fraction of
Brown, C.J., Pletcher, D., Walsh, F.C., 1993. Studies of
surface area of the electrode reaches the limiting current con- space-averaged mass transport in the FM01-LC laboratory
ditions and the overpotential is high enough to develop side electrolyser. J. Appl. Electrochem. 23, 38–43.
reactions as the cathode potential becomes more negative, Doche, O., Bauer, F., Tardu, S., 2012. Direct numerical simulation
produces current–potential curves with ill-defined plateaux of an electrolyte deposition under a turbulent flow—a first
which is in agreement with experimental results. approach. J. Electroanal. Chem. 664, 1–6.
Doche, O., Bauer, F., Tardu, S., 2013. Direct numerical simulations
of electrochemical reactions in turbulent flow. Electrochim.
5. Conclusions Acta 88, 365–372.
Eςa, L., Hoekstra, M., 2010. Near-wall profiles of mean flow and
A comprehensive mathematical model was developed to sim- turbulence quantities predicted by eddy-viscosity turbulence
ulate the complex relationship between hydrodynamics and models. Int. J. Numer. Methods Fluids 63, 953–988.
distribution of tertiary current and potential for copper recov- Frías-Ferrer, Á., González-García, J., Sáez, V., Ponce de León, C.,
Walsh, F.C., 2008. The effects of manifold flow on mass
ery in a continuous filter press type electrochemical reactor,
transport in electrochemical filter-press reactors. AIChE J. 54,
where non-ideal flow patterns, like backmixing, preferen-
811–823.
tial flow, inlet effects, dead zones, etc., are present. The Georgiadou, M., 2003. Modeling current density distribution in
model comprises the following equations: Reynolds averaged electrochemical systems. Electrochim. Acta 48, 4089–4095.
Navier Stokes, convection–diffusion, Butler–Volmer kinetics Griffiths, M., Ponce de León, C., Walsh, F.C., 2005. Mass transport
for copper reduction, Tafel for water oxidation and electri- in the rectangular channel of a filter-press electrolyzer (the
cal continuity. The model was able to provide information FM01-LC reactor). AIChE J. 51, 682–687.
Hammond, J.K., Robinson, D., Walsh, F.C., 1991. Mass transport
on velocity, copper ion concentration and electrical potential
studies in filterpress monopolar (FM-type) electrolysers. In:
fields inside the reactor, from which important distributions
Kreysa, G (Ed.), Electrochemical Cell Design and Optimisation
could be calculated: overpotential, mass transfer flux and cur- Procedures. Dechema Monograph, vol. 123. VCH, Weinheim, p.
rent density over the electrode surface. As part of the model 279.
development the mass transfer coefficients were determined Hotlos, J., Jaskula, M., 1988. Densities and viscosities of
experimentally and a fitting parameter (A) of the concentra- CuSO4 –H2 SO4 –H2 O solutions. Hydrometallurgy 21, 1–7.
tion wall function was tuned to the mass transfer model. Jayatillake, C.L.V., 1969. The influence of Prandtl number and
surface roughness on the resistance of the laminar sub-layer
The A-value and mass transfer coefficients obtained account
to momentum and heat transfer. Prog. Heat Mass Transfer 1,
for the enhancement of mass transfer due to the roughness 193–329.
caused by copper deposition. Simulation results show the Lacasse, D., Turgeon, E., Pelletier, D., 2004. On the judicious use of
importance of the reactor geometry, especially of the inlet and the k–ε model, wall functions and adaptivity. Int. J. Therm. Sci.
outlet distributors in the FM01-LC reactor where some flow 43, 925–938.
characteristics such as back mixing, preferential flow and low Mandin, Ph, Hamburger, J., Bessou, S., Picard, G., 2005. Modelling
and calculation of the current density distribution evolution
velocity zones cause strong variations of the mass transfer
at vertical gas-evolving electrodes. Electrochim. Acta 51,
rate in different parts of the electrode and, consequently, a
1140–1156.
nonuniform overpotential distribution over the electrode sur- Mandin, Ph, Cense, J.M., Picard, G., Lincot, D., 2006a. Simplified
face. These characteristics have profound implications on the kinetic modelling and numerical simulation of a metal oxide
reactor performance that require assessment in order to ana- chemical bath electro deposition process at a rotating
lyze experimental data of copper recovery before process scale electrode. Electrochim. Acta 52, 1296–1308.
chemical engineering research and design 1 0 0 ( 2 0 1 5 ) 422–433 433

Mandin, Ph, Fabian, C., Lincot, D., 2006b. Importance of the electrode (RCE) reactor under turbulent flow for copper
density gradient effects in modelling electro deposition recovery. Chem. Eng. Sci. 65, 3042–3049.
process at a rotating cylinder electrode. Electrochim. Acta 51, Rivero, E.P., Rivera, F.F., Cruz-Díaz, M.R., Mayen, E., González, I.,
4067–4079. 2012. Numerical simulation of mass transport in a filter press
Mandin, Ph, Cense, J.M., Fabian, C., Gbado, C., Lincot, D., 2007a. type electrochemical reactor FM01-LC: comparison of
Electrodeposition process modelling using continuous and predicted and experimental mass transfer coefficient. Chem.
discrete scales. Comput. Chem. Eng. 31, 980–992. Eng. Res. Des. 90, 1969–1978.
Mandin, Ph, Cense, J.M., Georges, B., Favre, V., Pauporté Th, Robinson, D., Walsh, F.C., 1991a. The performance of a 500 Amp
Fukunaka, Y., Lincot, D., 2007b. Prediction of the rotating cylinder electrode reactor. Part 1: Current–potential
electrodeposition process behavior with the gravity or data and single pass. Hydrometallurgy 26, 93–114.
acceleration value at continuous and discrete scale. Robinson, D., Walsh, F.C., 1991b. The performance of a 500 Amp
Electrochim. Acta 53, 233–244. rotating cylinder electrode reactor. Part 2: Batch recirculation
Mandin, Ph, Derhoumi, Z., Roustan, H., Rolf, W., 2014. Bubble studies and overall mass transport. Hydrometallurgy 26,
over-potential during two-phase alkaline water electrolysis. 115–133.
Electrochim. Acta 128, 248–258. Santos, J.L.C., Geraldes, V., Velizarov, S., Crespo, J.G., 2010.
Nelissen, G., Van den Bossche, B., Deconinck, J., Van Theemsche, Characterization of fluid dynamics and mass-transfer in an
A., Dan, C., 2003. Laminar and turbulent mass transfer electrochemical oxidation cell by experimental and CFD
simulations in a parallel plate reactor. J. Appl. Electrochem. studies. Chem. Eng. J. 157, 379–392.
33, 863–873. Trujillo, F.J., Safinski, T., Adesina, A.A., 2009. Solid–liquid mass
Nelissen, G., Van Theemsche, A., Dan, C., Van den Bossche, B., transfer analysis in a multi-phase tank reactor containing
Deconinck, J., 2004. Multi-ion transport and reaction submerged coated inclined plates: a computational fluid
simulations in turbulent parallel plate flow. J. Electroanal. dynamics approach. Chem. Eng. Sci. 42, 1143–1153.
Chem. 563, 213–2020. Vande Vyver, O., Nelissen, G., Weyns, G., Deconinck, J., Degrez, M.,
Nelissen, G., Weyns, G., Maciel, P., Deconinck, J., Vande Vyver, O., Godet, S., 2008. Mass transfer and current distribution on a
Deconinck, H., 2007. Numerical study of the influence of the metallic wire. Electrochim. Acta 53, 6452–6459.
anode position and the electrolyte flow on the deposition of Vázquez, L., Alvarez-Gallegos, A., Sierra, F.Z., Ponce de León, C.,
copper on a wire. Electrochim. Acta 52, 6584–6591. Walsh, F.C., 2010. Simulation of velocity profiles in a
Newman, J.S., 1991. Electrochemical Systems, second ed. laboratory electrolyser using computational fluid dynamics.
Prentice-Hall, New Jersey. Electrochim. Acta 55, 3437–3445.
Pohjoranta, A., Mendelson, A., Tenno, R., 2010. A copper Weigand, B., Ferguson, J.R., Crawford, M.E., 1997. An extended
electrolysis cell model including effects of the ohmic potential Kays and Crawford turbulent Prandtl number model. Int. J.
loss in the cell. Electrochim. Acta 55, 1001–1012. Heat Mass Transfer 40, 4191–4196.
Rivera, F.F., Nava, J.L., 2007. Mass transport studies at rotating Wragg, A.A., Leontaritis, A.A., 1997. Local mass transfer and
cylinder electrode (RCE) Influence of using plates and current distribution in baffled and unbaffled parallel plate
concentric cylinder as counter electrodes. Electrochim. Acta electrochemical reactors. Chem. Eng. J. 66, 1–10.
52, 5868–5872. Zeng, K., Zhang, D., 2010. Recent progress in alkaline water
Rivero, E.P., Granados, P., Rivera, F.F., Cruz, M., González, I., 2010. electrolysis for hydrogen production and applications. Prog.
Mass transfer modeling and simulation at a rotating cylinder Energy Combust. Sci. 36, 307–325.

You might also like