Pqper
Pqper
Pqper
reverse-osmosis desalination
∗1
Nicodemo Di Pasquale , Mayo Akele2 , Federico Municchi3 , John King2 , and Matteo
Icardi2
arXiv:2301.13160v1 [math.NA] 16 Jan 2023
1
Department of Chemical Engineering, Brunel University London, Uxbridge, UB8 3PH,
United Kingdom
2
Department of Mathematics, University of Nottingham, Nottingham, NG7 2RD, United
Kingdom
3
Department of Mechanical Engineering, Colorado School of Mines, Golden, CO 80401, US
Abstract
The reverse osmosis membrane module is an integral element of a desalination system as it determines the
overall performance of the desalination plant. The fraction of clean water that can be recovered via this
process is often limited by salt precipitation which plays a critical role in its sustainability. In this work, we
present a model to study the complex interplay between flow, transport and precipitation processes in reverse
osmosis membranes, which together influence recovery and in turn process sustainability. A reactive porous
interface model describes the membrane with a dynamic evolving porosity and permeability to capture the
scaling and clogging of the membrane. An open-source finite-volume numerical solver is implemented within
the OpenFOAM® library and numerical tests are presented here showing the effect of the various parameters
of the model and the robustness of the model to describe a wide range of operating conditions.
1 Introduction
The demand of freshwater has steadily increased over the last forty decades at a global level, mainly because
of the increase in population and improving living standards which are leading to an expansion of irrigated
agriculture and its human consumption. In turn, the increase in consumption is straining the freshwater
sources in their ability to supply the growing demand of water worldwide with almost two third of the
total world population experiencing severe water scarcity during at least a part of the year (Mekonnen and
Hoekstra, 2016; Jones et al., 2019). The current freshwater sources are already overexploited, and even if
better management is still needed to reduce the misuse of such resources (e.g., wastewater treatments or
waste reduction) (Najid et al., 2022), these solutions cannot still be enough to meet the future demand of
freshwater. The ongoing climate change is expected to reduce the availability of freshwater because of the
receding of glaciers with a subsequent important reduction of the flow in important rivers such as the Mekong
Yellow, or Gange (Shannon et al., 2008).
∗ Corresponding author: nicodemo.dipasquale@brunel.ac.uk
1
Almost 98% of the total liquid water on the Earth is not available for the direct use or consumption, as
it form the total water present in seas and oceans. However, this last fact also means that if the exceedingly
high saline content in seawater can be reduced or removed, we will have access to the largest source of
freshwater sources, with which the required amount of freshwater could be delivered without straining the
natural occurring resources (Jones et al., 2019). Therefore, there is a strong push into developing more effi-
cient technologies for the desalination of seawater. Among the currently available technologies are thermal
desalination and membrane processes (Fritzmann et al., 2007; Subramani and Jacangelo, 2015). In thermal
desalination, seawater is brought to evaporation through multi-effect distillation or multi-stage flash distilla-
tion (Al-hotmani et al., 2021), and the resulting vapor is subsequently condensed. In membrane technologies,
a semi-permeable membrane is employed to separate (or filtrate) the solution of salt and water.
Reverse Osmosis (RO), is a widely employed membrane technology for treating seawater and wastewater
with salinity up to 70 g/l (Hickenbottom and Cath, 2014) which, due to its relative simplicity and widespread
diffusion has been among the main topics of research in membrane filtration (Wardeh and Morvan, 2008;
Luo et al., 2019). One of the main challenges in RO is concentration polarisation (Kim and Hoek, 2005),
which is the presence, over the membrane, of a solute-rich boundary layer. Concentration polarisation can
lead to solute precipitation and fouling, significantly reducing the local permeability of the membrane with
adverse effects on the permeation flux. When the solution contains inorganic salts (such as sodium chloride
NaCl or calcium sulfate CaSO4 ), the resulting accumulation of crystals on the membrane is also called
scaling. There is extensive experimental evidence indicating that scaling reduces the membrane performance
over time (Hu et al., 2014) and therefore, scaling phenomena should play a major role in the mathematical
modelling of RO systems.
Computational Fluid Dynamics (CFD) represents a powerful to analyse concentration polarisation and
scaling, with the earliest attempts dating back to more than two decades ago (Hansen et al., 1998). The
membrane is usually included as a boundary condition where the flux of the solute is assumed equal to zero
(complete retention). Previous studies have considered dependence on the solute concentration of properties
such as the osmotic pressure (Hansen et al., 1998), viscosity, density and diffusion coefficient (Geraldes et al.,
2001; Wiley and Fletcher, 2002) to include effects of concentration polarisation on the membrane(Wiley and
Fletcher, 2002; Johnston et al., 2022). In particular, CFD simulations for membranes have been employed to
analise different geometrical configurations, such as the spacer-filled channels. These include solid elements
in the feeding flow to increase the shear stress on the surface of the membrane, which in turn increases
the local mixing and mass transfer across the membrane (Shakaib et al., 2007; Fletcher and Wiley, 2004;
Fimbres-Weihs and Wiley, 2010; Ghidossi et al., 2006; Lau et al., 2009; Santos et al., 2007; Koutsou et al.,
2009; Ranade and Kumar, 2006).
However, the effects of scaling have not been directly included in CFD simulations. In this work, we
propose a mathematical model and a CFD solver for the analysis of the performances of a RO membrane
in which we included the possibility for the solute in the feed flow to react (precipitate) on the membrane
and therefore affecting the membrane permeability. The paper is organised as follows. We describe the
mathematical framework for the analysis of the membrane, highlighting how the chemical reactions can be
accounted for in the model. We then discuss the implementation of our model into the widely used open
source finite volume library OpenFOAM® and we show some results for some typical situations, showing the
applicability of the whole framework. We then draw some conclusions and we outline the possible extension
of the model.
2 Model
In this work, we approximate a rectangular 3D membrane module as a 2D channel illustrated in fig. 1. This
is a common practice employed in other recent CFD studies (Johnston et al., 2022). While the configuration
we considered can be easily extended to more complex geometries, the focus here is to develop a complete
mathematical model for the polarization and scaling of the membrane by giving a proof of concept for a
general-purpose numerical solver. Our main goal is to show the mechanisms governing the evolution of the
solute at the membrane interface, and a 2D channel geometry allows us to focus on this task.
2
Figure 1: 2-dimensional domain considered in this work
Let us therefore assume a rectangular domain Ω ≡ (0, L) × (0, H) with boundary Γ = ∂Ω subdivided in
three different regions:
where Γm represents the membrane boundary, Γin is the inlet boundary and Γout is the outlet boundary.
With Γw we represent the remaining part of the boundary constituted by solid boundaries so that Γw =
Γ \ (Γin ∪ Γout ∪ Γm ), as shown in fig. 1. In this study, the permeate flow is not explicitly modelled, therefore
the membrane represents a boundary condition for the problem.
The analysis of the filtration process requires the simultaneous solution of the flow field coupled with the
transport of dissolved chemical species, which can involve one or more chemical reactions. Such chemical
reactions can lead to solute precipitation, modifying the permeability and porosity of the membrane, thus
causing a variation in the osmotic pressure and the final yield of the permeate. We summarised the different
mechanism and their interdependence in fig. 2.
Our mathematical model is composed of three different and intertwined parts that need to be simulta-
neously considered to obtain the overall behaviour of the membrane which we are now going to describe in
detail.
3
Figure 2: Sketch of the different mechanisms and models considered in this work.
where ρ is the density of the fluid u is the velocity vector with components (u, v)T , p is the fluid pressure
and µ is the dynamic viscosity, ∇ is the gradient operator. We impose the following boundary and initial
conditions on the system shown in eq. (2) for the velocity and pressure:
u=0 in Ω, t = 0 (3a)
u = 0, p = Pin on Γin (3b)
∇u · n = 0, p = Pout on Γout (3c)
u=0 on Γw (3d)
Finally, the membrane is modelled as a dynamic Dirichlet condition for the velocity v orthogonal to the
membrane is obtained from the Darcy law as:
k(∆p − ∆π)
u = 0, v = − on Γm (4)
`µ
where µ is the fluid viscosity, ` is the membrane thickness, k is the membrane permeability, ∆p = pm − pp
is the pressure difference between the feed side of the membrane, pm , calculated at the membrane boundary
Γm and the permeate side of the membrane, pp , whereas ∆π is the osmotic pressure gradient. This last
equation shows that the flux through the membrane is proportional to the difference between the applied
and osmotic pressure differentials.
The pressure gradient ∆p = pm − pp is the difference between the pressure on the feed side of the
membrane pm calculated at the membrane boundary Γm , and the permeate side of the membrane pp . The
osmotic pressure difference is defined as follows:
where the feed osmotic pressure π is a function of the N ions concentrations, φi , i = 1, . . . , n, whereas the
permeate osmotic pressure φp is function of the N ions concentrations in the permeate, φpi , i = 1, . . . , n.
As common in membrane applications (Linares et al., 2014) the osmotic pressure is computed using the
Van’t Hoff model (Van’t Hoff, 1888):
N
X
π(φ1 , . . . , φn ) = RT ϕ φi (6)
i
where ϕ is the osmotic coefficient, R is the gas constant and T is the temperature we can rewrite eq. (5) as:
N
X N
X
∆π = RT ϕ (φ − φpi ) = RT ϕ riφ φi (7)
i i
4
where we used the definition of membrane rejection of the species i,
φpi
riφ = 1 − , i = 1, . . . , N . (8)
φi
The Van’t Hoff equation assumes a linear relation between concentration and osmotic pressure and is more
accurate for low concentration of solutes. Different formulations were proposed to take into account the non-
linear behaviour of solutions, which consider the activity of the solvent (Khraisheh et al., 2019), calculated
using the Pitzer equation for the electrolyte solutions (Pitzer, 1973). As a proof of concept, we will consider
here the simplified model since the most complex behaviour can be straightforwardly added to the model
and the CFD code we will present in the next section. The membrane rejection of the species i expresses
the amount of solute rejected by the membrane (and therefore not present in the permeate) as a fraction of
the initial quantity. In this work, we assume riφ = 1 for every ion species, which corresponds to complete
rejections of the ions at the membrane. In this case, the concentration on the permeate side and the permeate
osmotic pressure, πp , are both equal to 0 and therefore:
N
X
∆π = RT ϕ φi . (9)
i
In the literature, the two most popular models proposed to describe the solute-solvent solute transport
through the membrane are the solution-diffusion model (Merten, 1963; Lonsdale et al., 1965; Wijmans and
Baker, 1995) and the Spiegler-Kedem model Spiegler and Kedem. The former expresses the flow through
the membrane J˙v as:
J˙v = A(∆p − ∆π) (10)
J˙v
noting that in our notation v = A(Γm ) where Sm is the membrane surface; while for the latter we have
1
J˙v = (∆p − σ∆π) (11)
Rm µ
J˙v
with the same identification for J˙v (i.e., v = A(Γ m)
) where Rm is the membrane resistance and σ is the
reflection coefficients which measure the impermeability of the membrane to the solutes. σ = 1 indicates a
membrane completely impermeable to solutes and will be the one considered for this work.
Notice that in eq. (4) we use the Darcy law to rewrite the so-called water permeability of the membrane
A which appear in the equation in terms of the permeability and thickness of the membrane and the viscosity
of water as A = S`µ mk
. Using the same reasoning for eq. (11) we obtain that Rm = Sm` k , that is to say, the
membrane resistance is inversely proportional to the permeability. The identification of the Darcy-related
terms with the water permeability A or membrane resistance has two main advantages. Firstly, we can
connect our analysis with the membranes available commercially which are described in terms of water
permeability or membrane resistance. This last fact allows us to choose the range of the parameters we are
considering which are appropriate for the description of real membranes. Secondly, using the Darcy derived
version of the equation for the flow through the membrane allows us to include more detailed mechanisms in
the model which can take into account more complex phenomena, such as chemical reactions and depositions
of solids on the membrane as we will show in more details in the next sections. One of the ways considered
in the literature to include fouling and polarisation of the membrane is to define such contribution as
additional resistance terms to be added to Rm in eq. (11) (see e.g., (Silva et al., 2011; Lee and Clark, 1998;
Yeh, 2002)). These additional terms must be derived from experiments or empirical correlations. In contrast,
our formulation leverages the well-established theory of porous media to include such effects directly in the
determination of the permeability k.
The performance of the membrane can be evaluated by using the recovery r, defined as the ratio between
the permeate flow rate, Q̇p , and the feed flow rate, Q̇f :
Q̇p vAp vL
r= = = (12)
Q̇f U A
in in uH
5
where we used the definition of flow rate through a surface, AU/`t, to calculate the flux through the in-
let (subscript in) and the membrane (subscript m) and then we replaced the area in eq. (12) with their
geometrical expression related to the domain we are considering: Ain = H × Z, Ap = L × Z.
6
where φi (t), i = 1, . . . , n denotes the volume molar concentration of chemical species Xi at time t. The
dynamics of the reaction network can be conveniently written in matrix form using the formalism developed
in (Chellaboina et al., 2009):
dφi i
= ξR = (A − B)T KφA (t), φi (0) = φ0,i , t ≥ 0 (16)
dt
where K = diag(K1 , . . . , Km ) is the diagonal matrix which contains as elements the reaction kinetics Kj ,
j = 1, . . . , m and φ0 is the initial concentration. A and B are the m × n matrices having in each entry the
stochiometric coefficients of the reactants and products respectively and φA (t) is the matrix obtained by
a
replacing each element of A with φi lp where akj is the element of A in the l-th row and p-th line. The first
term in the boundary conditions in eq. (13e) takes the form
i
= (A − B)T KφA (t) i `
ξR (17)
where the notation [·]|i stands for the i-th component of its (n × 1 vector) argument.
For a generic first order reaction X1 + X2 → X3 with kinetic constant K, we have therefore
dφ1
= −Kφ1 φ2
dt
dφ2
= −Kφ1 φ2
dt
dφ3
= Kφ1 φ2
dt
(18)
and
1
ξR = −Kφ1 φ2 `
2
ξR = −Kφ1 φ2 `
3
ξR = Kφ1 φ2 `
(19)
An example of a reaction of this type important in membrane operation is the formation of calcium carbonate,
through the reaction (Warsinger et al., 2015):
k f ()
= . (21)
k0 f (0 )
7
Thus we can describe the permeability evolution using the following power law (Hommel et al., 2018)
3
(1 − 0 )2
k
= , (22)
k0 (1 − )2 0
where k is the current permeability, is the current porosity, k0 is the initial permeability and 0 is the initial
porosity. The rate at which porosity reduction occurs is given by (Huo et al., 2019; Noiriel et al., 2004):
Zt X
Vs j
= o − ξR dt, (23)
` j
t0
where o denotes the initial porosity at t0 , Vs is the molar volume of solid precipitate in m3 /mol and r(t) is
the rate of precipitation in mol·m−2 · s−1 and the index j runs over all the possible reactions in the system.
For the first order reaction we are considering here the eq. (23) simplify as:
Zt
Vs
= o − (Kφ1 φ2 ) dt . (24)
`
t0
We should again observe the inter-dependencies and feedback mechanisms between fluid flow, transport
and reaction. Namely, in eq. (24) we see the precipitation reaction leads to a change in porosity which in
turn affects the permeability in equation eq. (22). The change in permeability impacts the flow via the fluid
velocity in eq. (4). This, in turn, alters the solute concentration distribution via eq. (13a), which ultimately
impacts the rate of precipitation again via eq. (13e). Moreover, from eq. (24) we can observe that the
variation of the porosity is proportional to the kinetic reaction constant. This last fact, in turn, simplifies
the predictions for the clogging of the membrane based on the reactions in the systems. We can expect that
if there are two chemical reactions in our systems, the first one 10 times slower than the second, then the
clogging caused by the products of the second reaction will take 10 times the time needed by the products of
the first reaction to produce the same amount of clogging. One of the strengths of our models is that allow
these kinds of qualitative analyses even without actually solving the equations.
3 Numerical discretisation
The equations presented in section 2 are solved using the open-source finite volume OpenFOAM 8.0 library
and the code is available open-source (Icardi, 2022). In order to solve the equation of motion alongside the
reaction at the membrane we developed a new solver for OpenFOAM called binaryReactionFoam which is
based on two widely used solvers, pimpleFoam and scalarTransportFoam· The former is a transient solver for
incompressible flows based on the PIMPLE algorithm while the latter is a concentration transport solver using
a user-specified velocity field. The solver also includes the possibility of modelling solid precipitation in the
fluid and a multiphase flow model for the solid particles. The most important element in the computational
framework, however is represented by the new boundary conditions implemented to model the membrane.
The membraneVelocity boundary conditions impose the fluid velocity based on the fluid pressure, the permeate
pressure, and the membrane properties. These are updated in time by linking this boundary condition to
the one for the scalar concentration, named binaryReaction, which solves for the solid precipitation at the
boundary and therefore updates the membrane permeability. The equations and boundary conditions are
coupled iteratively through Picard (fixed point) iteration (through the PIMPLE iterations) until convergence,
making the whole model fully implicit.
We simulate a two-dimensional rectangular channel with height h = 0.003 m and length of L = 0.02 m
discretised on a mesh composed by 600×200 cells. The following discretisation schemes (we direct the reader
to the OpenFOAM user guide (Ope, 2019) for a detailed description of each scheme) are used to discretise
the equations:
8
• advective fluxes (divSchemes Gauss vanLeer) are computed at the faces and the variables interpolated
with a Total Variation Diminishing scheme;
• gradient terms (gradSchemes Gauss linear) are approximated with central differencing;
• surface normal gradients for diffusive fluxes (snGradSchemes orthogonal) are approximated with central
differencing (the grid is in fact orthogonal and does not need any correction to ensure second order
accuracy);
• time derivatives (ddTSchemes backward) are approximated with third order implicit backward Euler
scheme,
We specified a fully developed velocity profile at the inlet:
y y
u(0) = 6uav 1− (25)
h h
where uav is the average velocity along the channel. By specifying the velocity profile at the inlet we need
only to specify the pressure at the outlet (the value of which is given in table 1). The pressure of the permeate
through the membrane is assumed constant along the length of the membrane and put equal to zero. For
longer membranes, this assumption is no longer valid and the permeate flow needs to be modeled explicitly
(with 1D or 2D models). This will be the subject of future extensions of our framework.
The initial value of the permeability we consider in the calculations is k = 10−16 m2 . Using the expression
for the water permeability obtained from eq. (4) and eq. (10): A = S`µ mk
, and the viscosity of water at room
temperature µ = 10 Pa s, and the surface of the membrane, Sm = 2·10−6 m2 for the channel configuration
−3
and ` = 10−7 m , we obtain a value of A = 2 · 10−12 m (Pa s)−1 . This value is in line with the values
reported for commercial membranes, which are in the range 10−14 to 10−10 (Pa s)−1 (Ruiz-Garcı́a and de la
Nuez Pestana, 2019; Lee et al., 1981; Dražević et al., 2014).
Our model is able to include the scaling of the membrane given by the chemical reaction, which can
modify the membrane permeability through the precipitation of a solid phase obtained as a product of the
reaction. In this work we considered a range of kinetic reactions, going from very slow to fast reactions,
i.e. with a value of the kinetic constant spanning four orders of magnitude, from 10−10 to 10−1 m3 /mol.
We considered such a large range of kinetics since our main goal is not to focus on a specific system (and
reaction) but to give a general description that can be applied to different specific situations.
4 Results
In this work, we employ a fixed flow profile at the inlet, which when considering the properties in table 1,
gives a Reynolds number equal to 300. Therefore, the system operates in a fully developed laminar flow
regime. The first property that can be derived from this model is the polarisation of the membrane, which
represents the accumulation of the solute at the interface of the membrane on the feed side. This is an
undesired effect since it increases the osmotic pressure reducing the extraction of the permeate per unit of
energy consumed in the process. We report the variation of the concentration profile in the domain in fig. 3
for the lowest and highest chemical kinetic rate considered. We can observe in fig. 3 that the concentration
at the membrane is different from the one in the bulk region in both cases. However, while the case with
K = 10−15 m3 /mol shows a higher concentration with respect to the bulk (see fig. 3a), the case with the
highest value of the kinetic rate (K = 10−1 m3 /mol, see fig. 3b) shows a concentration smaller than the one
in the bulk.
The latter observations show that the possible behaviour of the solution near the membrane strongly
depends on the reaction kinetics. In the first case (the one represented in fig. 3a corresponding to the lowest
kinetic rate considered K = 10−15 m3 /mol), we can observe the “standard” effect of the polarisation of
the membrane. During the filtration process, there is an accumulation of the solutes molecule on the feed
side of the membrane, which results in a higher concentration of the solution at the membrane itself. On
9
Symbol definition/value units
k 10−16 m2
0.7 -
φa,0 35 g/m3
D 0.003 m
` 0.0001 m
Vs 27 · 10−6 m3 /mol
ρ 1000 kg/m3
pperm 0 kPa
µ 10−3 Pa s
Table 1: Summary of the numerical inputs for the physical quantities used in the simulations. Note that
the units of the kinetic constant K depend on the fact that we considered a binary reaction, whereas for the
permeability k and porosity we are considering the initial value (i.e., the value at t = 0 h.
10
(a) K = 10−10 m3 /mol (b) K = 10−1 m3 /mol
Figure 3: Contour plot of the concentration within the channel given in units of the initial concentration
φ̃a = φa φin,0 . On the left: results reported for the lowest kinetic constant. On the right, results are reported
for the highest kinetic constant. Note that the starting point of the legend is not zero and is different between
the pictures to make the results more clear.
the opposite side, when the reaction rate is almost negligible (as for the case of K = 10−15 m3 /mol), the
solutes are now consumed by the reaction and they accumulate at the membrane interface, leading to the
concentration profile observed in fig. 3a. In the latter case, the solute is now consumed almost instantly at the
membrane interface. This result in a transport (convection and diffusion) limited profile of the concentration
near the interface.
The two opposite effects just described for the profile of the concentration at the membrane interface give
an interesting effect on the evolution of the porosity and permeability profiles across the membrane. As the
reaction proceeds, a new solid phase is formed which precipitates on the membrane modifying its structure
and therefore its fluid dynamical behaviour. In particular, the solid phase generated during the reaction
clogs the pores of the membrane, resulting in a variation of the porosity of the membrane with time. On top
of this, since the concentration along the membrane (in the x-direction) decreases, we can expect a decrease
in the overall reaction rate (which is proportional to the concentration) and therefore a difference in the
permeability and porosity over the membrane. When we instead observe polarisation (i.e., in the case of the
lowest reaction rate) the concentration near the membrane increases with the distance along the membrane
direction.
Therefore, we can expect that the reaction velocity (which depend on K and the concentration), at a
fixed time, will increase along the membrane interface for the case with polarisation (low kinetic reaction)
and decrease along the membrane for high value of the kinetic reaction rate. We will show quantitatively
these effects in the next section.
The description of the properties of the membrane (i.e., porosity, permeability, velocity through the
membrane) can be given in terms of global quantities, that is to say quantities averaged over all the membrane
length, which therefore becomes a function of time only. We will start our analysis by giving an account of
these global properties.
In fig. 4 we report the variation of the average of the porosity across the membrane as function of time.
For the lowest kinetic reaction time considered there is no appreciable variation of the porosity after more
than one day of operations. By increasing the kinetic reaction rate we can start to observe some deviations.
In particular, for the highest value of the reaction rate the porosity decays to 60% of its original value after
only one day of operation. This latter kind of results can be useful in determining the operation time that
we can expect from a membrane given a certain composition of the feed.
The law of variation of the permeability with the reaction depends on the variation of the porosity, and
in fact we can expect a similar behaviour. We reported the results for k in fig. 5, where we can see that there
is an order of magnitude difference between the initial value of the permeability at time t = 0 and after 28
h of operations.
The average velocity through the membrane obtained with the conditions specified is 1.8 µs, which
decreases with the decrease of the permeability of the membrane up to 0.13 µs for the lowest value of k
and shown in figs. 4 and 5. In order to maintain the flow across the membrane in the given conditions of
cross-flow in the channel and for the permeability and porosity given, we have to apply a pressure of approx
1800 kPa, which is needed to overcome an osmotic pressure of 1000 kPa, which reduces to 978 kPa in the
11
Figure 4: Plot of the porosity versus time for all the systems considered, K is in mol/m3 .
Figure 5: Plot of the permeability as a function of time for all the systems considered K is in mol/m3 .
12
case of the highest reaction rate. The difference in the osmotic pressure for the case K = 10−1 depends on
the fact that for this case the concentration at the membrane is lower than the bulk (and lower than the
case with polarisation) because of the very fast reaction (see fig. 3b).
Despite the fact the osmotic pressure is smaller for the fastest reaction case, this case remains the worse
in terms of permeate extraction, because the fast scaling of the membrane reduces quickly the porosity until
the flux stops completely, i.e., we reach a value of v = 0.13 µs when we consider a fast reaction rate, against
a value one order of magnitude higher for the case of the lowest reaction rate where we do not observe the
scaling in the simulated time.
13
(a) v (b) porosity
Figure 6: Velocity through the membrane and porosity profile along the membrane at different times for
different value of the kinetic reaction values. The continuous lines are the results at 6 h, the dashed lines at
12 h and the dot-dashed lines at 28 h. K is in mol/m3 .
5 Conclusions
In this work, we presented a comprehensive computational model to describe the solute dynamics ed evolution
near a membrane for desalination processes. In particular, we included a model to treat the scaling of the
membrane as solid precipitated following a (general) chemical reaction. We connected the accumulation of
solids at the membrane with porosity and permeability as described by the Darcy theory of porous media.
Following this approach, we were able to give a full explicit model to derive the dynamical evolution of the
filtration process by specifying a few initial parameters (e.g., the property of the solution and the kinetics
of the reaction).
The membrane is described as a dynamic boundary condition for the fluid mechanics and solute transport
equations, which are coupled together through the osmotic pressure term, and therefore the flow through
the membrane. We implemented our model in the widely used software package for CFD calculations
OpenFOAM® , and performed simulations for a selected range of operating conditions. Results show how
this model can be used to predict the decay in the flux through the membrane due to the accumulation of
the precipitated solid originating from the chemical reaction.
The formulation presented here has two main advantages which make it flexible and powerful in treat-
ing polarization. First, the proposed formulation can address all the interconnections between the different
mechanisms (fluid dynamics, solute evolution, chemical reaction, scaling, and fouling) which affect the mem-
brane performance. The second advantage is that the model can be easily extended to include more complex
geometries, or models for the osmotic pressure (such as the Pitzer model (Pitzer, 1973; Khraisheh et al.,
2019)), fluid flow conditions in the system, as well as more complex reactions paths.
References
M. M. Mekonnen and A. Y. Hoekstra, Science advances 2, e1500323 (2016).
E. Jones, M. Qadir, M. T. van Vliet, V. Smakhtin, and S.-m. Kang, Science of the Total Environment 657,
1343 (2019).
N. Najid, J. N. Hakizimana, S. Kouzbour, B. Gourich, A. Ruiz-Garcı́a, C. Vial, Y. Stiriba, and R. Semiat,
Computers & Chemical Engineering p. 107794 (2022).
14
M. A. Shannon, P. W. Bohn, M. Elimelech, J. G. Georgiadis, B. J. Mariñas, and A. M. Mayes, Nature 452,
301 (2008).
15
K. Spiegler and O. Kedem, Desalination 1, 311 (1966).
J. M. Silva, V. G. Ferreira, and S. R. Fontes, Applied mathematics and computation 217, 7955 (2011).
Y. Lee and M. M. Clark, Journal of Membrane Science 149, 181 (1998).
H. Yeh, Desalination 145, 153 (2002).
V. Chellaboina, S. P. Bhat, W. M. Haddad, and D. S. Bernstein, IEEE Control Systems Magazine 29, 60
(2009).
D. M. Warsinger, J. Swaminathan, E. Guillen-Burrieza, H. A. Arafat, et al., Desalination 356, 294 (2015).
T. Waly, M. D. Kennedy, G. J. Witkamp, G. Amy, and J. C. Schippers, Desalination and water treatment
5, 146 (2009).
C. I. Steefel, D. J. DePaolo, and P. C. Lichtner, Earth and Planetary Science Letters 240, 539 (2005).
J. Hommel, E. Coltman, and H. Class, Transport in Porous Media 124, 589 (2018).
J.-x. Huo, F.-h. Ma, and X.-l. Ji, Water Science and Engineering 12, 155 (2019).
C. Noiriel, P. Gouze, and D. Bernard, Geophysical research letters 31 (2004).
M. Icardi, membraneFoam v1.0 (2022), URL https://doi.org/10.5281/zenodo.7477180.
OpenFOAM: The Open Source CFD Toolbox, The OpenFOAM Foundation, https://openfoam.org/, v1906
ed. (2019).
A. Ruiz-Garcı́a and I. de la Nuez Pestana, Water 11, 152 (2019).
K. Lee, R. Baker, and H. Lonsdale, Journal of membrane science 8, 141 (1981).
E. Dražević, K. Košutić, and V. Freger, Water research 49, 444 (2014).
H. Krawczyk and A.-S. Jönsson, Chemical Engineering Research and Design 92, 174 (2014).
16