1. Introduction
Osmosis can be described as the spontaneous flow of water from a solution of high– water potential (low salinity) to a solution of low water potential (high salinity) across a semipermeable membrane [
1]. Osmotic–driven processes play a significant role in modern engineering systems and industries, from separation and purification units to clean water and energy production plants [
2,
3,
4]. There are three basic osmotic processes based on the water flow direction that depend on both the osmotic and hydraulic pressure difference: Reverse Osmosis (RO), Forward Osmosis (FO), and Pressure–Retarded Osmosis (PRO). In the FO process, there is no hydraulic pressure difference and hence the water flux is generated by the osmotic pressure difference only from the low-salinity (feed) solution to the high–salinity (draw) solution. In PRO, the hydraulic pressure difference applied to the draw solution is lower than the osmotic pressure difference, and, hence, water will still flow from the low–salinity to the high–salinity solution, so it is osmotic–driven like the FO process.
This class of membrane–based processes, i.e., PRO, has recently been investigated by several researchers due to its potential as a clean and sustainable energy source. However, there are limitations that are currently slowing down its scale–up and worldwide spreading [
4]. Despite the technical challenges ahead, recent investments and promising innovations by start-up companies, e.g., SaltPower in Denmark, are paving the way to the realization of osmotic energy harvesting on an industrial scale [
5]. Thus, research and development are still needed to improve osmotic power efficiency, and its commercial viability.
The fast increase in the experimental study of PRO has been accompanied by an increase in CFD modeling studies since 2010 [
6]. Even though the pioneering CFD works on osmotic-driven flows focused only on FO process simulation [
7,
8], later works included the two–dimensional (2D) PRO channel [
9], and both FO and PRO modes [
10,
11]. Nevertheless, only a few researchers took advantage of CFD to analyze local hydrodynamics and mass transfer inside three–dimensional (3D) PRO modules [
12,
13]. This is particularly due to challenges and difficulties to model the interaction of flow and mass transfer in all membrane compartments (feed, draw, and porous support layer) in a PRO module [
6].
One of the main advantages of CFD is that geometry modifications can be implemented and tested more easily than physical prototypes [
14,
15]. Also, using CFD, one can model and study 3D complex geometries and multi–physics processes [
16,
17], which is impossible via the analytical approaches. Thus, a CFD modeling framework for PRO simulation can be considered as an efficient virtual lab for the simulation and analysis of different osmotic–driven processes. Despite this usefulness of CFD, there are some considerations that should be taken into account for the modeling of an osmotically driven membrane process.
The major challenge in using CFD to describe an osmotic-driven system lies in the implementation of the membrane boundary conditions: here, unlike in pressure–driven systems, internal concentration polarization (ICP) inside the membrane porous support layer reduces the effective osmotic driving force significantly, and this effect therefore has to be explicitly accounted for in the CFD model. Fundamentally, two approaches can be categorized to capture the ICP phenomena: (1) solving the governing differential equations over the porous zone, e.g., using Brinkman porous medium model, and (2) modeling the effect of the porous layer via the analytical representation of the osmotic flux–ICP relation.
The first approach resolves hydrodynamics and mass transfer in the porous layer [
18]. Thus, it requires a fine computational mesh in the membrane zone and also determination of the permeability of the porous layer (κ). Due to such requirements, and because the characteristic length scale of the membrane support structure,
O (10
−4 m), is much smaller than the length scale of the membrane module, it is not viable to resolve the porous layer in a 3D CFD model. Alternatively, the solute profile within the support can be calculated from integrating the convection–diffusion equation between the active layer and feed channel [
12]. This is the second approach that involves adoption of an analytical but complete model that includes both hydrodynamics and mass transfer. Previously, CFD validations [
6,
8,
13], performed by comparing simulation outputs against experimental results, verified that the second approach is also a reasonable approximation due to the weak interaction between external flow and concentration polarization in the porous layer. Hence, in this work we adopt the second approach using an analytical ICP model for definition of the membrane boundary conditions (BCs) in CFD simulation, i.e., fully coupling the model to the CFD solver in such a way that the osmotic flux (as a BC for CFD solver) is calculated based on the local concentrations at the external membrane surfaces, which are resolved in return by the CFD solution of feed/draw channel flows.
On the other hand, the effect of external concentration polarization (ECP) on the draw-side (dilutive ECP or ECP
d) and feed-side (concentrative ECP or ECP
c) of the membrane are important in osmotic-driven processes, especially in the case of high–concentration solutions. Some earlier studies [
8,
19] used Sherwood number correlations for ECP modeling, while others ignored the local concentration variation or estimated it using simplified assumptions [
12,
13]. Such approaches are limited to special module geometries as they are based on the boundary layer similarity or other simplifications for calculation of local concentrations on the membrane boundaries with the feed and draw solution streams. In the current research, we adopted a more general approach, that is, ECP effects are resolved directly by CFD via implementing boundary layer (BL) mesh grids fine enough to capture thin concentration boundary layers. This approach can be executed without any limitation on geometry complexity and flow mode/regime and, as aforementioned, all CP effects are included in the simulation via two-way coupling between the membrane osmotic flux–ICP model, and the simultaneous CFD solution of feed/draw channel flows (ECPs) [
20].
The methodology presented in this paper, was implemented by means of a powerful commercial CFD solver (Ansys® Fluent, ANSYS, Inc., Canonsburg, PA, USA), which can be linked to a variety of modeling and meshing tools on Ansys Workbench Platform. For the efficient adjustment of membrane BCs, user–defined functions (UDFs) were programmed and coupled with the CFD solver. To the authors’ knowledge, there is no previously published work on the CFD simulation of 3D PRO membrane chambers (including all CP effects), or on the implementation and validation of the Ansys Fluent solver for osmotic–driven membrane process simulations, including a comparison between FO and PRO modes.
In this paper, after a description of the physics of the PRO process in the next section, the details of the modeling approach, including the PRO flux and RSF calculations as BCs for the CFD solver, are presented. Then, the results of CFD model validation against the reference experimental data are discussed for different scenarios and membrane parameters in both FO and PRO mode. Moreover, flow visualization and a comprehensive comparison of ECP/ICP contributions in local osmotic pressure drop along the membrane chamber are presented for both PRO and FO modes. Finally, the capability of the presented CFD model for the simulation of the lab–scale mixed–mode PRO module is demonstrated on a range of Re numbers from laminar to turbulent flow regimes.
2. Materials and Methods
Osmotically driven membrane processes involve the movement of water through a semipermeable membrane due to an osmotic pressure difference. There are different key phenomena involved in transmembrane mass transfer in these processes. The driving force in such processes is the osmotic pressure difference between two solutions separated by a membrane. This pressure difference causes water to move from the side with lower concentration (feed solution) to the side with higher solute concentration (draw solution). The water flux across the membrane is influenced by the osmotic pressure gradient, the membrane properties, and the concentration polarization effect. So, in an osmotic process, there are different types of physics that have effect on the transmembrane flux and the system performance, including fluid dynamics, solute transfer, membrane separation, and membrane fouling, all of which are nonlinear and interdependent (see
Figure 1).
There are also diverse scales of the involved phenomena in a membrane module. While the membrane active layer thickness is around 0.1 to 0.5 μm, the porous sublayer is in the range of 50 to 150 μm, whereas the module length can be from a few cm up to meters. Also, due to the fact that the concentration boundary layer is generally much thinner than the velocity boundary layer, high–fidelity modeling methods are required to capture the local concentration variations as the most important parameter in osmotically driven processes. As water moves through the membrane layers, solutes accumulate near the membrane surface on the feed side, creating a concentration polarization layer. This layer reduces the effective osmotic pressure difference and thus the osmotic water flux. As shown in
Figure 1, there are both external and internal concentration polarization (ICP) phenomena involved in PRO. The external concentration polarization (ECP) stems from the diffusion boundary layer development near the dense layer of the membrane, but ICP occurs in the internal porous layer of the membrane. Managing concentration polarization is crucial for maintaining high efficiency in osmotic membrane systems. While ECP can be suppressed by increasing velocity or mixing, ICP is very difficult to mitigate.
In addition to osmotic water flux, there is also a reverse flux of solutes from the draw solution to the feed solution. This reverse solute flux can affect the overall efficiency and performance of the process. The extent of reverse solute flux depends on the membrane selectivity and the nature of the solutes. Therefore, for technical and economical feasibility studies of osmotic–driven systems, many laboratory and modeling works are required. Also, for efficient computational simulation of such multi–physics multi–scale processes, several factors should be taken into consideration in order to compromise between modeling accuracy and cost. In the following part of this section, we present details of the integrated framework developed in this study for simulation of osmotic–driven processes.
2.1. Governing Equations
The membrane module is simulated in an isothermal and steady–state condition. The governing equations for the conservation of mass, momentum, and solute mass fraction in a multi–species Newtonian fluid flow, neglecting the gravity effects, are
where
u is the fluid velocity vector,
Sm,
Sv, and
SA are mass, momentum and solute source terms,
p is the static pressure,
mA is the local mass fraction of solute, and
DAB is the solute diffusivity coefficient, which together with density (
ρ) and dynamic viscosity (
μ) denote the solution properties. Assuming a dilute aqueous solution of sodium chloride (NaCl), there are empirical correlations for the mixture properties [
7]:
which are valid for NaCl solutions at 25 °C (
ρw = 997.1 kg/m
3) for concentrations up to solute mass fractions of 0.09, corresponding to molar concentrations around 1.6 M [
21]. All the case studies simulated in this work are iso–thermal (
T = 25 °C), but, for different operational temperatures or non–isothermal conditions, we must implement more general empirical correlations, which also take into account the effect of temperature on the properties [
10],
where
c (Mol/L) is molarity and
T (K) is temperature. Also, the effect of pressure on the mixture properties is negligible for the operational pressures studied in this work.
The above system of equations can be solved numerically using a pressure–based CFD solver to determine the variations of velocity and concentration (C = ρ mA), in both the feed and draw channels, provided that we have laminar flow ( < 2100) in both the feed and draw channels of the membrane chamber. In fact, for high–Reynolds–number flows, it is necessary to implement a turbulence model in the CFD simulation, where further parameters will be introduced depending on the turbulence modeling approach.
The simplest “complete models” of turbulence are the two–equation models in which the solution of two separate transport equations allows the turbulent velocity and length scales to be independently determined. One of the most popular two–equation models in practical engineering simulations is the
model that has robustness, economy, and reasonable accuracy for a wide range of turbulent flows. The standard
model is a semi–empirical model based on transport equations for the turbulence kinetic energy (
k) and its dissipation rate (
). As the strengths/weaknesses of the standard
model have become known, improvements have been made to the model to improve its performance. One of these variants is the RNG
model that we implemented for the turbulent case studies in this work. It is derived from the instantaneous Navier–Stokes equations, using a mathematical technique called “renormalization group” (RNG) methods [
22].
The RNG
model is similar in form to the standard
, but includes some refinements and features that make it more accurate and reliable for a wider class of flows. Namely, in the derivation of the standard model, and the assumption is that the flow is fully turbulent, and the effects of molecular viscosity are negligible; it is therefore valid only for high–Reynolds–number flows, while the RNG
model provides an analytically–derived differential formula for effective viscosity that accounts for low–Reynolds–number effects. For more details on the transport equations, the methods of calculating turbulent viscosity, and model constants, it is recommended to refer to the documentation of the CFD software (ANSYS Fluent) [
22]. Specifically, here we just present a short discussion on the turbulent diffusivity calculation in the following.
If we are in the turbulent flow regime, and a turbulence model is on in the CFD model, then we need to take into account the turbulent diffusivity. The effective diffusivity is then equal to the sum of laminar diffusivity (
DAB) and turbulent diffusivity (
Dturb). We know the laminar diffusivity from Equation (6), while the turbulent diffusivity can be computed from the turbulent Schmidt number (
Scturb = μturb/
rDturb), where
μturb (the turbulent viscosity) is a function of turbulence parameters (
). Hence, in turbulent flows, the mass diffusion is computed by replacing the mass diffusive flux term, i.e.,
term in Equation (3), with
. In ANSYS Fluent, the turbulent Schmidt number is typically set to a user–specified constant (default value is 0.7, which is generally considered appropriate for most turbulent flows, as it reflects the empirical relationship between the turbulent transport of momentum and mass). However, in the RNG
model, the RNG theory provides an analytical formula for turbulent Prandtl numbers, which is analogously used for turbulent Schmidt numbers, relating
Scturb and
Sc (molecular Schmidt number) [
22].
It must be noted that the specific turbulence model implemented in this work is just an example of several available turbulence models, and is not essentially the best candidate for this kind of simulation. It may be an interesting topic for future works to investigate and compare different turbulence models for the simulation of osmotic membrane modules. It is also noteworthy that, for both laminar and turbulent simulations, similar settings are used for the CFD solver, as shown in
Table 1, except for the pressure–velocity coupling scheme, where the coupled solver is generally preferred for turbulent simulations due to its robustness and faster convergence, especially in cases involving complex geometries.
2.2. Boundary Conditions
In a PRO module, a thin semipermeable membrane structure separates the feed and draw channels with a dense selective (active) layer facing to the draw–side, and a porous support layer next to the feed–side of the membrane, as illustrated in
Figure 1. Due to the concentration difference (Δ
Cm) between the two sides of the membrane, an osmotic–driven flow can be generated by the osmotic pressure difference (Δπ
eff), which is always higher than the hydraulic pressure difference in the PRO case. So, the osmotic pressure, the osmotic water flux (
Jw), and reverse salt flux (
Js) can be calculated via,
where the proportionalities are the van’t Hoff correlation coefficient (805.1 bar), water permeability (
A), and salt permeability (
B), which are the intrinsic membrane properties, typically determined experimentally via RO tests [
23].
The most important parameters in osmotic membrane modeling are
Jw and
Js, which should be calculated via Equations (7)–(9) as BCs of the CFD solution. The flux equation can be solved either by having wall flux conditions that explicitly allow the matching flux pair to be set on either side of the membrane or by using a source–sink term pairing [
18]. For the latter approach, it is crucial to note that the source at the membrane wall on the feed–side must have the opposite sign but same value as on the draw–side. Here, we are following the second approach that is considering the wall BC for the active layer and pairing source–sink terms for the adjacent cells on both sides of the membrane. A similar approach has been recently used for CFD simulation of RO membranes by Bae et al. [
16], but here we modified and extended its formulation to the simulation of PRO membranes. Initially, all the source terms in the computational domain are zero. Next, in each CFD solver iteration, the PRO flux will be accounted for through mass/solute source–sink terms (
Sm and
SA in Equations (1) and (3)), as they become non–zero in the mesh cells neighbored with the membrane walls, according to the following equations:
where the water and solute source terms (
Sw,
SA) have dimensions of kg/(m
3·s), Δ
Vm is the volume of a mesh cell directly in contact with the membrane wall, and Δ
Am denotes the corresponding face area. It is notable that source terms can be defined for the momentum equation as well, i.e.,
Sv in Equation (2); however, they are usually negligible owing to the small velocity in the boundary layer [
18].
All the BCs for different boundaries in feed and draw channels are listed in
Table 2. It is noteworthy that a complete 3D model is essential for capturing the full complexity of fluid flows in practical applications, but it is possible to reduce the computational domain using symmetry BC, if the geometry and fluid flow are symmetrical, which significantly decreases the computational resources and time required for simulations. However, while symmetry BCs can simplify 3D problems and make them more computationally feasible, the results obtained from 3D simulations with symmetry can still differ from purely 2D simulations, due to the inherent differences in dimensionality, 3D boundary layer effects, secondary flows, and turbulence modeling.
2.3. ICP Model
Recalling Equations (8) and (9), it can be seen that for the calculation of
Jw and
Js we need the values of local concentrations (or salt mass fractions) on both sides of the membrane active layer, i.e.,
Cdm and
Csm (see
Figure 1).
Cdm is known from the CFD solution of the draw channel flow, but
Csm is determined via an analytical ICP model, which is a more efficient approach compared with the numerical solution of Brinkman equations governing the incompressible flow of the solution in the homogenous and isotropic porous support. Assuming
Jw as the dominant fluid velocity across the support layer in the PRO membranes, and neglecting the gradients in all directions except normal to the membrane surface, the Brinkman equations can be reduced to an ODE for solute mass balance in the porous layer:
where
C(
y) is the concentration at position
y, and
is the effective diffusivity of the solute at the support layer, whose porosity (
ε) and tortuosity (
τ) are known. By their definitions,
Jw and
Js are defined as variables only at the position of the selective layer surface, i.e.,
y =
ts. However, it has been shown that the fluxes derived at the selective layer surface are accurate quantitative approximations of the flux and flow in the three–dimensional geometry throughout the membrane structure [
24]. Thus, at a steady state, the conservation of mass governing the solute transport in the support layer can be described as Equation (13), and can be integrated over the porous thickness with the relevant boundary conditions [
C(
y = 0) =
Cfm,
C(
y =
ts) =
Csm], and the solute concentration on the porous–side of the active layer (
Csm) comes out as,
where
K represents the solute resistivity in the support layer and is defined as
K =
S/
D, where
is the structural parameter of the membrane, which, together with A and B (water and salt permeability), form the three main parameters displaying the properties of a membrane. Replacing Equation (9), i.e.,
, in Equation (14) results in,
where both relations depend on the local concentrations on the feed–side (
Cfm) and draw–side (
Cdm) of the membrane, which are known from the CFD solution of channel flows. Now, the PRO water flux equation can be derived using Equations (7), (8) and (15)
where the osmotic pressure on the draw–side of the membrane (
) is a function of the local draw–side mass fraction of salt, which is known from the CFD solution, and
fos (=805.1 × 10
5/
ρ) at the porous–side of the active layer.
fos can be determined having
Csm from Equation (15) and the local density (
ρsm) from a concentration–based version of Equation (4), as derived in
Appendix A.
It is also described in
Appendix A that a linear approximation of the relation between osmotic pressure and salt concentration can be formulated with a constant proportionality of
, i.e.,
, which results in
. Hence, via Equations (8) and (16), an alternative equation for
Jw can be obtained without the need to evaluate the local density in the porous layer (i.e.,
ρsm), i.e., Equation (18), which is the water flux equation of the PRO process used by almost all the previously published modeling works. We present the derivation in details here to clarify the assumptions and simplifications that were not mentioned in the literature. So, now the PRO water flux equation (Equation (17) can be rewritten as,
As will be discussed later, there is no remarkable difference between the CFD results coupling the solver either with Equation (17) or Equation (18) for the PRO cases studied in this work, where the concentrations are in the range of van’t Hoff assumption applicability (<100 g/L). The advantage of Equation (17) is that it can be used for the PRO membranes with higher draw solution concentrations (with a nonlinear osmotic pressure–concentration relation) as well. Also, for FO case, it can be shown that in a similar manner the famous FO flux equation, i.e.,
[
25], can be written in an analogous form as (see
Appendix B),
Similarly, knowing the local values of
and
from the channel flow solution, or FO or PRO water flux equation, Equation (19) or Equation (18), will have only one unknown variable, i.e.,
Jw. However, both equations are nonlinear and should be solved numerically (
Figure 2) at each point on the membrane using a robust root–finding technique like Ridders’ solver, for which the details of its algorithm can be found in [
26].
4. Conclusions
In this research, a unified CFD framework was developed and implemented for the 3D simulation of osmotically driven membrane processes, incorporating all concentration polarization effects (ECPd, ECPc, ICP) and reverse salt flux. The modeling methodology was presented in detail, based on a two–way coupling between the CFD solver and an analytical ICP model in FO or PRO mode. The model was validated against experimental data from a rectangular flow chamber under various operational scenarios, showing good agreement for the PRO fluxes, with a maximum deviation of less than 10%. Additionally, comparisons between different scenarios indicated that, for a constant inlet concentration difference, increasing feed/draw concentrations reduces the PRO and FO water fluxes by approximately 70% and 50%, respectively, in alignment with experimental observations.
CFD visualization of concentration profiles within the membrane chamber revealed that the highest osmotic fluxes are concentrated near the draw inlet section, whereas the minimum osmotic flux, corresponding to the greatest osmotic pressure drop, is observed near the membrane–wall intersections. Also, simulations of pressurized PRO modules demonstrate that the overall trends in local concentration profiles (and local CP effects) are similar with unpressurized cases (Δp = 0), even though the total osmotic pressure drop is lower in pressurized cases. A comparative analysis of the two membrane types indicates that, for membranes with higher water permeability (e.g., TFC), the CP effects contribute more substantially to the total osmotic pressure drop than for membranes with lower permeability (e.g., CTA), under analogous conditions. This disparity is more noticeable in PRO mode compared with FO mode. Consequently, it can be inferred that optimizing the PRO flux is more effectively achieved through the use of low–concentration feed solutions, rather than solely relying on membranes with high permeability.
CFD simulations of a lab–scale PRO chamber across a range of Reynolds numbers indicate that, under high–Re flow conditions, internal concentration polarization (ICP) is the dominant CP effect. Conversely, in low–velocity laminar flow regimes, the increase in external concentration polarization on the draw side (ECPd) can diminish the effective osmotic pressure and PRO flux to the same extent as ICP. In addition, analysis of various membrane parameter combinations indicates that the structural parameter (S) exerts a significant influence on PRO flux. However, under low–Re laminar conditions, its effect is nearly equivalent to that of water permeability (A). The influence of solute permeability coefficient (B) is several orders of magnitude lower.
In conclusion, the presented modeling framework exhibits sufficient flexibility for the simulation of 3D osmotically driven membrane modules with complex geometries, while also providing the capability to visualize local variations in various flow regimes. The proposed CFD model can be integrated with other solvers to conduct multi–physics simulations. Future research may explore dynamic membrane modules, fluid–structure interactions (FSI) in high–pressure membrane systems, hybrid models such as MD–CFD for including interphase interactions and microscopic inhomogeneities, and development of machine learning (ML) models to optimize osmotically driven membrane modules.