Next Article in Journal
Physical Properties and Reactivity of Microdomains in Phosphatidylinositol-Containing Supported Lipid Bilayer
Next Article in Special Issue
Water as a Blood Model for Determination of CO2 Removal Performance of Membrane Oxygenators
Previous Article in Journal
Amorphous NdIZO Thin Film Transistors with Contact-Resistance-Adjustable Cu S/D Electrodes
Previous Article in Special Issue
Theoretical Analysis of a Mathematical Relation between Driving Pressures in Membrane-Based Desalting Processes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Correlations for Concentration Polarization and Pressure Drop in Spacer-Filled RO Membrane Modules Based on CFD Simulations

1
Department of Chemical Engineering, Imperial College London, London SW7 2AZ, UK
2
School of Chemical Engineering, Chonnam National University, 77 Yongbong-ro, Buk-gu, Gwangju 61186, Korea
*
Authors to whom correspondence should be addressed.
Membranes 2021, 11(5), 338; https://doi.org/10.3390/membranes11050338
Submission received: 12 April 2021 / Revised: 28 April 2021 / Accepted: 29 April 2021 / Published: 1 May 2021
(This article belongs to the Special Issue Numerical Modelling in Membrane Processes)

Abstract

:
Empirical correlations for mass transfer coefficient and friction factor are often used in process models for reverse osmosis (RO) membrane systems. These usually involve four dimensionless groups, namely Reynolds number (Re), Sherwood number (Sh), friction factor (f), and Schmidt number (Sc), with the associated coefficients and exponents being obtained by fitting to experimental data. However, the range of geometric and operating conditions covered by the experiments is often limited. In this study, new dimensionless correlations for concentration polarization (CP) modulus and friction factor are presented, which are obtained by dimensional analysis and using simulation data from computational fluid dynamics (CFD). Two-dimensional CFD simulations are performed on three configurations of spacer-filled channels with 76 combinations of operating and geometric conditions for each configuration, covering a broad range of conditions encountered in RO membrane systems. Results obtained with the new correlations are compared with those from existing correlations in the literature. There is good consistency in the predicted CP with mean discrepancies less than 6%, but larger discrepancies for pressure gradient are found among the various friction factor correlations. Furthermore, the new correlations are implemented in a process model with six spiral wound modules in series and the predicted recovery, pressure drop, and specific energy consumption are compared with a reference case obtained by ROSA (Reverse Osmosis System Analysis, The Dow Chemical Company). Differences in predicted recovery and pressure drop are up to 5% and 83%, respectively, highlighting the need for careful selection of correlations when using predictive models in process design. Compared to existing mass transfer correlations, a distinct advantage of our correlations for CP modulus is that they can be directly used to estimate the impact of permeate flux on CP at a membrane surface without having to resort to the film theory.

1. Introduction

Spiral wound modules are the most widely used membrane modules in a reverse osmosis (RO) process owing to a number of desirable features, such as their higher packing density than plate-and-frame modules and lower fouling propensity than hollow fiber modules [1]. As multiple membrane sheets are rolled around a central tube in a spiral wound module, spacers are inserted to keep membrane sheets separate. The presence of spacers affects the performance and energy consumption of an RO process, due to their influences on mass and momentum transfer. Concentration polarization (CP) is mitigated by enhanced mixing induced by spacers while pressure drop is increased due to raised frictional resistance. In order to predict membrane performance and pumping energy requirements, it is essential to gain a quantitative understanding of the mass and momentum transfer processes in a membrane module. This has motivated numerous experimental and computational studies on the effects of spacers on CP and pressure drop [2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30].
Considerable progress has been made on developing correlations for mass transfer coefficient and pressure drop in a membrane channel, with a number of empirical correlations available in the literature [3,5,6,7,8,10,11,12,13,26,27,31,32,33]. Mass transfer correlations are commonly used for quantitative description of CP at a membrane surface by using the film theory [2,34,35]. More recently, computational fluid dynamics (CFD) simulations have been performed to investigate the effects of spacers on mass transfer coefficient and pressure drop [17,18,19,20,21,22,23,24,25,26,27,28,29], and furthermore, to establish mathematical correlations incorporating geometric features of spacers [26,27]. However, the range of geometric and operating conditions covered by existing correlations is limited. Therefore, it is important to evaluate the performance of existing correlations and to modify these to suit for a wide range of RO conditions. An ideal correlation should be derived in the presence of membrane permeation, incorporate geometric information of spacers, and be valid for a wide range of RO operating conditions.
Correlations for mass transfer coefficient and pressure drop typically involve the following dimensionless groups: Sherwood number (Sh), Reynolds (Re), and Schmidt (Sc) number for mass transfer coefficient and friction factor (f) and Reynolds (Re) for pressure drop. It should be noted that different definitions of f and Sh exist due to different characteristic lengths used. Geometric parameters representing spacer filaments are taken into account either in the dimensionless groups as a characteristic length or as an additional term.
A number of empirical correlations were derived from mass transfer experiments in an electrodialysis process using an electrochemical cell without a membrane [5,7,8,27]. As a result, these correlations do not directly account for permeation velocity and should be coupled with the film theory and 1D convection-diffusion model to correlate permeation velocity and CP. Schock and Miquel [30] introduced a mathematical expression for the hydraulic diameter of a spacer-filled channel so that the dimensionless variables (i.e., Re and Sh) contain geometric features of spacers. Their experiments were performed on an RO membrane module, making their correlation well suited for prediction of CP and pressure drop in an RO process. The number of spacers tested in their work, however, is limited to five types. Da Costa et al. [11,12,13] carried out a series of experimental studies on ultrafiltration (UF) membranes and spacers. By examining a variety of spacer geometries, they developed a single correlation for Sh that incorporates a number of geometric characteristics of spacers [13]. Nevertheless, their correlation was derived at velocity conditions more appropriate for a UF rather than an RO process [26]. More recently, Koutsou et al. [26,27] presented correlations for Sh and pressure drop in 3D spacer geometries based on CFD simulations and experimental measurements in an electrochemical cell with prototype spacers. Although various spacer geometries were considered under typical operating conditions for RO processes, the correlations were derived for each spacer geometry separately, without inclusion of geometric parameters.
The present study focuses on deriving dimensionless correlations for CP modulus and friction factor for a wide range of geometric and operating conditions relevant to a spacer-filled channel of an RO membrane module. This is undertaken by means of CFD simulations of two-dimensional (2D) models of spacers with varying geometry and operating conditions. By means of dimensional analysis, a correlation for CP modulus is formulated instead of a mass transfer coefficient (or Sh), so that CP can be calculated directly without having to couple the film theory with a process model. In order to simulate a spacer-filled feed channel in an RO process as realistically as possible, the CFD model accounts for the effect of permeation through RO membranes by coupling the solution–diffusion model with the governing equations for fluid flow and solute transport. It is expected that the derived correlations are able to capture mass and momentum transfer phenomena in the presence of membrane and feed spacer for a wide range of RO operating conditions.
In the following section, the CFD modelling approach is described, which is followed by the derivation of dimensionless correlations. Next, the key performance indicators obtained from CFD simulations are presented, together with the obtained correlations for CP modulus and friction factor. Derived correlations are also compared with existing correlations in the literature. Finally, the derived and existing correlations are implemented in a process model with multiple spiral wound modules, where both feed and permeate flows are accounted for in a real size module. In doing so, the impact of using different correlations on predicted module performance, such as recovery, pressure drop, and specific energy consumption in an RO process, can be assessed.

2. Methods

Details of the CFD models are described in this section including model geometry, governing equations, and boundary conditions. This is followed by the derivation of correlations for CP and pressure drop in spacer-filled channels.

2.1. CFD Simulations

Two-dimensional (2D) models of different types of spacers are created by taking the cross-sections of spacer filaments in a feed channel, which represents the main flow path. Figure 1 illustrates three spacer types: nonwoven, partially woven and fully woven spacers [36,37]. When transverse filaments are normal to the inflow direction and are also at 90º with the axial filaments, three types of cross-sections can be created: cavity, submerged, and zigzag configurations as shown in Figure 1 [18,20,21,22,24,25,38]. In this study, 2D models of the three cross-sectional configurations are constructed as simplified representations of spacer-filled channels with upper and lower permeable membrane walls. Since each configuration has a repeating pattern of filament arrangement along the flow path, a unit cell is built to simulate a spacer-filled channel. For the cavity and submerged configurations, a single elemental unit consisting of two filaments can present the repeating pattern in a channel, while the zigzag configuration requires two elemental units consisting of three filaments. For the sake of consistency in geometry, two elemental units are included in each model, which consists of one circular filament in the middle and two half filaments at the entrance and exit of each spacer-filled channel.
In the fluid domain, governing equations for steady-state laminar flow and solute transport are solved for an incompressible and Newtonian fluid. The continuity, Navier–Stokes and convection-diffusion equations are given in Equations (1) to (3), respectively.
· u = 0
ρ u · u = p + μ 2 u
· ( D s w u ) · ( u c ) = 0
where u is the fluid velocity vector, ρ the density of the fluid, μ the viscosity of the fluid, p the pressure, c the solute concentration and Dsw the diffusivity of solute in water. A dilute aqueous solution with a single solute is assumed and the fluid properties are approximated by those of pure water.
A fully developed velocity profile is imposed at the inlet, while a constant pressure is specified at the outlet. Spacer filaments are treated as no-slip walls, where zero velocities are prescribed. For solute transport, a flat concentration profile is applied at the inlet, and a no flux condition is imposed at the outlet. At membrane walls, the solution–diffusion model is applied, which allows a realistic description of the transport phenomena through RO membranes [39,40,41]. By coupling with the van’t Hoff equation to calculate osmotic pressure, water, and solute fluxes in the solution–diffusion model can be expressed as:
J w = A [ ( p p p ) κ ( c c p ) ]
J s = B ( c c p )
where Jw is the water flux, Js the solute flux, A the hydraulic conductivity, B the solute permeability, pp the permeate pressure, cp the permeate concentration, and κ the osmotic factor, which is assumed constant under an isothermal condition.
By introducing two assumptions: (i) permeate pressure is constant (here, pp = 0) and (ii) permeate concentration is determined by the solute to water flux ratio (cp = Js/Jw), water and solute fluxes can be expressed as a function of feed concentration and pressure [37].
J w = A ( p κ c ) B + { A ( p κ c ) + B } 2 + 4 κ A B c 2
J s = B [ A ( p + κ c ) + B { A ( p κ c ) + B } 2 + 4 κ A B c 2 κ A ]
Equations (6) and (7) are used to provide boundary conditions at the upper and lower membrane walls, where the following conditions are applied:
u l , u p = [ 0 J w ]
u l , l o w = [ 0 J w ]
n · ( D s w c u c ) = J s
where n is the normal vector and ul,up and ul,low are the leaking velocities at the upper and lower membrane walls, respectively.
Model geometries are created for 45 combinations of filament diameter Df and filament spacing Lf (details are provided later). The fluid domains are discretized into quadrilateral elements of varying sizes with denser mesh near the upper and lower membrane walls in order to ensure adequate resolution in the mass transfer boundary layer. Local mesh refinement is also performed in regions where circular filaments are in contact with membrane walls in the cavity and zigzag configurations. The number of mesh elements for the baseline cases (Df = 0.5 mm and Lf = 4.5 mm) are 14,400, 16,200, and 9000 for the cavity, submerged and zigzag configurations, respectively. Mesh independence study is performed by increasing the mesh density until changes in averaged water and solute fluxes are less than 3 × 10−5 %. Moreover, the local concentration in the vicinity of center filament and membrane wall, at x = Lf + 0.1 mm and y = 0.01 mm for the cavity and submerged configuration and y = Hc – 0.01 mm for the zigzag configuration, is monitored and changes between the coarsest and finest meshes are below 0.02% for three configurations.
CFD simulations are carried out using COMSOL Multiphysics 5.1, which is based on the finite element method. Linear formulation is chosen for both velocity and pressure and the concentration field. The stationary solutions are obtained by the nonlinear solver using the fully coupled and automatic damped Newton method. The MUltifrontal Massively Parallel Spacer direct solver (MUMPS) is used to solve the system of linear algebraic equations with a convergence criterion of 10−6 for the relative error in the solution.

2.2. Selection of Geometric and Operating Conditions

Six operating and geometric parameters are varied in CFD simulations: average inlet velocity u0, solute diffusivity Dsw, inlet concentration c0 and outlet pressure p0, filament spacing Lf, and filament diameter Df. Each operating parameter is varied one at a time while the other parameters are kept the same as in the corresponding baseline case. As shown in Figure 2, geometric variables Lf and Df are simultaneously varied with the operating variables being fixed. For each configuration, a total of 76 cases are simulated as listed in Table 1. Spacers’ geometric parameters Lf and Df are varied by selecting different feed channel height Hc and Lf to Hc ratio, with Df and Hc being related as: Df = 0.5Hc. The baseline inlet concentration 603.45 mol/m3 is equivalent to 35 kg/m3 of an aqueous sodium chloride solution, which can approximate seawater. The hydraulic conductivity A and solute permeability B are selected within reported ranges [37,42,43,44]. The osmotic factor κ is obtained at temperature of 298.15 K.

2.3. Dimensional Analysis and Derivation of Correlations

Dimensional analysis using the Buckingham Pi theorem [45] is performed to identify the dimensionless groups involved in the correlations for CP modulus and pressure drop. A detailed procedure to formulate the correlations can also be found in Section A in the Supporting Information.

2.3.1. Correlation for CP Modulus

There are eight parameters involved in determining CP: CP modulus MCP, fluid density ρ, dynamic viscosity μ, solute diffusivity in fluid Dsw, cross-velocity uc, transmembrane velocity ut (or water flux at membrane walls Jw), filament diameter Df, and filament spacing Lf. In this study, a slightly different definition of CP modulus is used to facilitate its implementation in an RO process model.
M C P = c m c p c 0 c p
where cm is the feed concentration at membrane walls and c0 the inlet concentration. According to the Buckingham Pi theorem, the above dimensional parameters can be grouped into five independent dimensionless groups (8 − 3 = 5). By choosing ρ, μ, Df as repeating variables, five dimensionless groups are formed, including MCP and 4 dimensionless groups commonly used in momentum and mass transfer, such as Re and Sc. The dimensionless groups Rec, Ret, Sc, and GR are defined as:
R e c = ρ u c D f μ
R e t = ρ u t D f μ   or   ρ J w D f μ
S c = μ ρ D s w
G R = L f D f
Here the filament diameter Df is used as a characteristic length in Rec and Ret. GR, the filament spacing to diameter ratio, allows the inclusion of an important geometric feature of a spacer-filled channel. Their functional relationship can be expressed as follows:
M C P 1 = C R e c α ( m · R e t ) β S c γ G R δ
where C, α, β, γ, and δ are coefficient and exponents that are dependent on the spacer configuration (cavity, submerged, and zigzag). Although Ret could be considered negligible compared to Rec due to transmembrane velocity being usually several order of magnitude smaller than cross-velocity, the term is kept in Equation (16) for the time being with an adjustment factor m. This is discussed later in Section 3.1.1 in relation with CFD results for the effects of transmembrane velocity on CP moduli.

2.3.2. Correlation for Pressure Drop

Pressure drop in a spacer-filled channel is influenced by the following parameters: pressure gradient dp/dx, uc, ρ, μ, Df, and Lf. The transmembrane velocity ut is not included as its effect on axial pressure drop is negligible due to its small magnitude compared to the cross-velocity. Analysis using the Buckingham Pi theorem yields three dimensionless groups, which can be related as follows:
f = C R e c α G R δ + ζ
with
f = D f ρ u c 2 d p d x
where f is the derived friction factor, C′, α′, δ′, and ζ are coefficients and exponents to be determined by experimental or numerical simulation data. The local pressure gradient dp/dx in Equation (18) can be replaced by Δp/(2Lf), which is the pressure drop per unit length in the fluid domain.

2.3.3. Parameter Estimation

The coefficients and exponents in Equations (16) and (17) are estimated by data fitting, using results obtained from 2D CFD simulations. For each case, the simulation results for MCP, uc, ut, and Δp are averaged over the entire fluid domain, and spatially averaged values are used for data fitting. A wide range of values for the dimensionless parameters are considered by varying operating and geometric conditions (Sc = 111–4475, Rec = 1.12–274, Ret = 7.82 × 10−4–6.70 × 10−3, and GR = 4–12). Since transmembrane velocity, ut, is a key parameter that determines CP, a range of ut (3–12 × 10−6 m/s) are obtained by varying feed concentration and pressure in the simulations. The coefficient and exponents in Equation (16) are estimated by solving the following:
min n = 1 N = 76 [ M C P , n C F D M C P , n ( C , α , β , γ , δ ) ] 2
where MCPCFD is the CP modulus obtained from CFD simulations, MCP (C, α, β, γ, δ) the CP modulus calculated by Equation (16) with a set of parameters, n the index of data and N the total number of data. To find the values of C, α, β, γ, and δ that minimize the sum of the squared errors of CP moduli, the built-in simplex algorithm in MATLAB 2015b is utilized. Since there might exist multiple local minima, it is critical to provide a good initial guess, which is obtained by differentiating Equation (16) together with the corresponding CFD data.

2.4. Implementation of Correlations in a Process Model

The new correlations obtained in this study and existing correlations available in the literature are implemented in a predictive model for a spiral wound module [42], where feed and permeate streams are independently accounted for in a real size module, and the predicted performances using different correlations are compared. The correlations are implemented as follows:
  • Correlations for CP (obtained in this study) are coupled with the solution–diffusion model described in Equations (4) and (5).
  • Existing correlations for mass transfer coefficient are implemented using the film theory.
  • All these are incorporated in a spiral wound module model, where mass balances for the feed and permeate streams are performed separately.
Among the existing correlations in the literature, choices are made from those derived under the conditions similar to RO processes. Two correlations for mass transfer coefficient are chosen: Schock and Miquel [30] and Koutsou et al. [27] for Sherwood number, which are expressed in Equations (20) and (21), respectively.
S h = 0.065 R e c 0.875 S c 0.25
S h = 0.2 R e c 0.57 S c 0.40
Definition of Sc is the same as given in Equation (14), but Sh and Rec have different characteristic lengths; dh and Df for Equations (20) and (21), respectively. The derived expression for dh of spacer-filled channels in [30] is expressed as:
d h = 4 ε 2 / H c + ( 1 ε ) 4 / L f
where ε is the voidage of the spacer-filled channel. The selected correlations for pressure drop are those of Schock and Miquel [30] and Koutsou et al. [26], as given in Equations (23) and (24), respectively.
d p d x = f S M ρ u c 2 2 d h where   f S M = 6.23 R e c 0.3  
d p d x = f K R e c 2 ρ ν 2 D f 3   where   f K = 2.3 R e c 0.31   at   L f D f = 6   and   f K = 0.8 R e c 0.19   at   L f D f = 8
where fSM and fK are the friction factors in the respective correlations, and ν the kinematic viscosity. Two correlations, as in Equation (24), were derived using 3D CFD simulations with a constant Df (=1 mm). The main reason for selecting the aforementioned correlations is that they were derived under RO operating conditions, which usually yield a feed flow velocity below 0.4 m/s [23]. Two additional correlations for mass transfer coefficient [30,42,46,47] and pressure drop [42,48] are also included in the comparison of module performance, which are frequently used in membrane process models when spacer specifications are not available.
S h = 1.85 R e c 1 / 3 S c 1 / 3 ( d h , e m p t y L ) 1 / 3
d p d x = 12 k s p μ ( 0.5 H c ) 2 u c
where ksp represents wall roughness in a channel due to spacer filaments. The negative sign for pressure drop correlations in Equations (18), (23), (24) and (26) is omitted in order to make friction factors positive, but it is added when dp/dx is implemented in a process model. Equation (25) is for a flat channel without spacers, in which the hydraulic diameter, dh,empty, correspond to that of an empty channel.
In the model for a spiral wound module, mass balance equations are expressed in the form of differential equations so that spatial variations along the feed and permeate flow directions can be taken into account [42,43,44]. For process simulations a typical arrangement is considered, which has six spiral wound modules arrayed in series in a pressure vessel [49,50,51], and only one pressure vessel is included. The number of leaves and membrane width are estimated using the model presented in [42] based on the given module diameter in the product data sheet [52]. Permeability constants (A and B in Equations (4) and (5)) and frictional coefficients (ksp in Equation (26)) are also estimated by comparing predicted recovery, feed pressure drop and permeate concentration with those from ROSA (Reverse Osmosis System Analysis, The Dow Chemical Company), the latter serve as reference data for comparison purposes. The estimated permeabilities and friction coefficients values are listed in Table 2, as well as module and process specifications used to simulate RO processes.

3. Results and Discussions

3.1. Analysis of CFD Results

In this section, CFD results are presented and analyzed with a focus on predicted performance indicators, including water flux, CP modulus, and pressure drop. Velocity vectors and concentration profiles obtained from the CFD simulations are included in Section B in the Supporting Information.

3.1.1. Water Flux and CP Modulus

Average water fluxes and CP moduli in three configurations under various operating conditions (the baseline condition and conditions 1–30) are shown in Figure 3. It is clear that feed concentration and pressure have a direct impact on water flux-water flux increases with a decrease in feed concentration or an increase in feed pressure. This is because water permeation is driven by the net pressure defined as (ΔpκΔc), according to the solution–diffusion model in Equation (4). As can be seen in Figure 3a,c,e and g, there appears to be a direct relationship between average water flux and CP modulus when only feed concentration or pressure is varied, suggesting that CP is strongly influenced by transmembrane velocity. This finding justifies the inclusion of Ret in the CP modulus correlation in Equation (16); if Ret was neglected, the effects of transmembrane velocity caused by different feed concentrations and pressures on CP moduli would not be captured. On the other hand, increasing solute diffusivity and inlet flow velocity causes water flux to rise until it plateaus, as shown in Figure 3b,d. The increased water flux could be due to mitigated CP by accelerated diffusion of solutes back to the bulk flow at high solute diffusivity and by enhanced mixing near membrane walls at high flow velocity. This is supported by Figure 3f,h, where CP moduli are reduced as solute diffusivity or inlet flow velocity increases. The trend of water flux plateauing at high inlet velocity in Figure 3d also indicates that there is a limit to the impact of enhanced mixing.
Comparisons among the different configurations suggest that the submerged configuration has a slightly higher water flux within the ranges studied. CP moduli are lower in the submerged configuration than in the cavity and zigzag configurations, due to the absence of direct contact between spacer filaments and membrane walls. The maximum difference between the cavity and submerged configurations are 7.8% for water flux and 7.0% for CP modulus. The response of CP modulus to increasing inlet velocity appears to be biphasic, with exponential decays for inlet velocity up to 0.1 m/s and almost linear reduction thereafter for the cavity and zigzag configurations. This behavior is not observed in the submerged configuration. It is possible that the direct contact between filaments and membrane walls in the cavity and zigzag configurations is responsible for the biphasic behavior, which is influenced by the extent of flow recirculation and stagnation.
Average water fluxes and CP moduli are displayed with respect to channel height at different filament lengths to diameter (Lf/Df) ratios in Figure 4. The effect of channel height is obvious: water flux is reduced and CP modulus is elevated with an increase in channel height. Compared to the cavity configuration, the submerged configuration offers 2.1–13.5% higher water flux and 1.8–9.9% lower CP modulus by varying the geometric conditions. The results also indicate that denser spacer mesh (i.e., at small Lf/Df ratios) achieves a slightly better performance.
Similar biphasic behavior is observed in the simulation results with varying geometric conditions, most notably at small Lf/Df ratio in Figure 4d, where there is a logarithmic increase in CP modulus up to a channel height of around 0.8 mm and an almost linear increase thereafter for the cavity and zigzag configurations. Combined with the observation described previously in relation to Figure 3h, it appears that in the cavity and zigzag configurations CP moduli behave differently in two distinguished regimes, which can be separated using a critical Reynolds number, Recrit, based on the cross-flow velocity and characteristic length, Df (= 0.5Hf). Therefore, a critical Reynolds number is introduced in this study to more accurately correlate CP moduli with operating and geometric conditions for the cavity and zigzag configurations.

3.1.2. Pressure Drop and Friction Factor

Pressure drop per unit length is calculated using the predicted axial pressure drop from CFD and the channel length, while the corresponding friction factors are calculated by using Equation (18). Results for varying inlet velocities and geometric conditions are presented in Figure 5. As expected, pressure drop increases as inlet velocity increases. Regardless of the operating and geometric conditions, the submerged configuration appears to have the highest pressure drop among the three configurations. This is because the spacer filaments are located in the middle of the channel, which forces the flow to split into two streams when passing around the filament, causing increased local velocity. On the other hand, little variation in pressure drop is found with varying inlet concentration, solute diffusivity and outlet pressure, justifying the formulation of the correlation for pressure drop, as expressed in Equation (17).
The effects of geometric variables on pressure drop can also be deduced from the results for different Lf/Df ratios (Figure 5b–d). As the channel height increases, pressure drop per unit length decreases rapidly. At the same channel height, higher Lf/Df ratios result in less pronounced pressure drops per unit length. Since Lf/Df ratio affects the void volume in a channel, it has a strong influence on the mean velocity and pressure drop. Friction factors show a similar trend to pressure drop per unit length except for varying inlet velocities. The definition of friction factor in Equation (18) dictates that it is inversely proportional to the square of cross-flow velocity, which results in the trend shown in Figure 5e.

3.2. Correlations for CP Modulus and Friction Factor

Values for the parameters in the CP modulus (Equation (16)) and friction factor (Equation (17)) correlations are obtained by using the CFD simulation results, and these are shown in Table 3, together with the corresponding absolute percentage errors (APE) from data fitting. The adjustment factor m in Equation (16) is 104 in order to make the exponents of Ret and Rec be of a similar order of magnitude. Both CP modulus (MCP) and friction factor (f) correlations have a good fit to the CFD data with a mean APE less than 1.5% for all configurations. There are two sets of parameters for the MCP correlation for the cavity and zigzag configurations, depending on the critical Reynolds number, Recrit, which is found to be 24.4–31.0 for varying inlet velocities (Figure 3h) and geometric conditions (Figure 4d). As a result, Recrit is assumed to be 25.5 for both the cavity and zigzag configurations.
For all configurations, values for the exponents of Rec and Ret (i.e., α and β) reveal that CP modulus is more sensitive to the transmembrane velocity than cross-flow velocity. The value for β indicates that the relationship between Ret and MCP is almost linear, whereas the value for α suggests that Rec has a stronger influence on MCP at low Rec than at high Rec. It is also noted that the cavity and zigzag configurations have similar parameter values for the given regime, but there are differences between the two regimes. The value for γ, exponent of Sc, is assumed to be constant regardless of the flow regimes but to vary with the filament configuration. For the cavity and zigzag configurations, there is only a slight difference in α and β between the two regimes. However, values for δ are markedly different, leading to different effects: in the low Rec regime, MCP decreases slightly with an increase in Lf/Df ratio, as displayed in Figure 4d–f. In the high Rec regime, δ tends to zero for both configurations, suggesting a negligible effect of Lf/Df on MCP. For the submerged configuration, δ is very different from that for the other configurations; the positive sign of δ means that MCP increases with an increase in Lf/Df, as shown in Figure 4d–f. The coefficient C varies with hydrodynamics and geometry, hence giving different values depending on configurations and flow regimes.
Estimated parameter values for the friction factor correlations are comparable for the cavity and zigzag configurations. The value for α′, the exponent of Rec, implies an inverse relation between f and Rec as can be seen in Figure 5e. For the exponent of Lf/Df ratio, δ′, its values capture the fact that f decreases more rapidly with an increase in Lf/Df in the submerged configuration than in the cavity and zigzag configurations, as displayed in Figure 5f–h. The parameters C′ and ζ for the submerged configuration are found to be approximately 3.4–4.5 times larger than those for the cavity and zigzag configurations. This suggests the submerged configuration has a larger friction factor in most situations. Further comparisons between the CFD results and derived correlations can be found in Section C in the Supporting Information.

3.3. Comparison with Existing Correlations

Before being implemented in a process model, the new correlations are compared with exiting correlations described in Section 2.4. Comparisons are made for a range of flow velocities (0.095–0.35 m/s) using spacer specifications listed in Table 2. The chosen velocity range corresponds to the range of hydraulic Reynolds numbers examined by Schock and Miquel [30]. Koutsou et al. derived the mass transfer correlation in Equation (21) from their experimental results on prototype spacers with a similar Lf/Df ratio to commercial spacers (Lf/Df = 8 and 12 and mesh angle = 90°) [26,27]. However, the filament diameter (1 mm) adopted in their work far exceeded that of commercial spacers (0.27–0.385 mm) based on the study by Schock and Miquel on four feed spacers in spiral wound RO modules [30].

3.3.1. CP Modulus and Water Flux

Results for CP modulus and the corresponding water flux obtained using our new correlations are compared with those of Schock and Miquel [30] and Koutsou et al [26,27]. Since the latter are based on Sherwood number (or mass transfer coefficient), it is necessary to involve water flux to relate mass transfer coefficient to CP modulus [30]. In addition, it should be borne in mind that Koutsou et al. [27] measured the mass transfer coefficient in an electrochemical cell in the absence of membrane.
Figure 6 shows that the predicted behavior of CP modulus with respect to varying flow velocity is similar between our new correlations and those of Koutsou et al. with discrepancies of less than 6%, which corresponds to approximately 6% difference in water flux. On the other hand, CP moduli and water fluxes obtained with the correlation by Schock and Miquel are more sensitive to changes in flow velocity in that the CP modulus declines more rapidly with increasing flow velocity, leading to a more pronounced increase in water flux. In addition, the Schock and Miquel correlation tends to predict higher CP modulus and consequently lower water flux than others, particularly in the low velocity range. Similar findings were also reported by Koutsou et al. [27], and they attributed this discrepancy to uncertainties related to the indirect determination of mass transfer coefficient (or Sherwood number) using the solution–diffusion model and film theory. However, it should also be noted that these correlations were derived for different ranges of geometric parameters.
Results from our new correlations for the three configurations lie between those of Schock and Miquel and Koutsou et al., as can be seen in Figure 6a, although our correlations for the cavity and zigzag configurations predict a slightly higher CP modulus when flow velocities exceed 0.25 m/s. Mean and maximum differences compared to the correlations of Schock and Miquel and Koutsou et al. are summarized in Table 4. It is difficult to ascertain which correlation is more accurate for RO applications. This will be discussed further in Section 3.5 in relation to limitations of the CFD model.

3.3.2. Pressure Gradient

Values for pressure gradient dp/dx obtained with our correlations are compared with those of Schock and Miquel and Koutsou et al. in Equations (23) and (24). As shown in Figure 7, the correlations of Koutsou et al. for Lf/Df = 6 and Lf/Df = 8 predict larger pressure drops than that of Schock and Miquel and our new correlations. For the particular spacer simulated here (Table 2), Lf/Df = 7.3, which lies between the Lf/Df ratios covered by the correlations of Koutsou et al.; the latter appear to over-predict dp/dx by 2.3 times compared to the Schock and Miquel correlation for the range of flow velocities examined. This is consistent with the findings reported in Koutsou et al. [26]. The mean and maximum percentage differences between the new and existing correlations are summarized in Table 5.
The submerged correlation is closer to that of Schock and Miquel compared to the correlations for the cavity and zigzag configurations. Although Koutsou et al. also derived their correlations based on CFD results, there are a number of differences between their work and the present study: (i) their CFD models are 3D; (ii) a periodic condition with a constant mean-pressure gradient between inlet and outlet faces is imposed with the assumption of impermeable membrane walls; and (iii) Koutsou et al. simulated only four data points at flow velocities of 0.02, 0.08, 0.1, and 0.15 m/s, with a fixed channel height of 2 mm, whereas our simulations cover a wider range of operating and geometric conditions.

3.4. Simulation Results from Process Model

Correlations for CP and friction factor are implemented in a process model, and nine simulations are performed as listed in Table 6, under three operating conditions that lead to a water recovery of approximately 50%. Case 1 corresponds to the reference case, which is set to reproduce the simulation results obtained with ROSA. Cases 2 to 5 are designed to test the influence of different friction factor correlations while using the same CP correlation as in the reference case, whereas Cases 6 to 9 allow the influence of CP correlations to be determined when the same pressure drop correlation (Equation (26)) is used. Comparisons are made for the predicted water recovery, specific energy consumption, and pressure drop along six modules. Results for all simulation cases are compared with those of Case 1, and the relative differences are reported in Table 6.
Table 6 shows that changing the friction factor correlations (Cases 1 to 5) has an almost negligible effect on the predicted recovery, with differences of up to 1.5% at the highest feed flowrate, although the predicted pressure drop varies considerably. Nevertheless, the magnitude of pressure drop is too small to make a notable effect on the predicted specific energy consumption, which has a relative difference of less than 2% for all conditions examined. It is worth noting that the largest differences are observed in Case 5 where the correlation of Koutsou et al. for Lf/Df = 8 is used. Since this correlation was derived over a narrow flow velocity range (up to 0.15 m/s), it is not surprising that larger differences are found at higher feed flow velocities, although the differences are not significant.
Using different CP correlations (Case 6 to 9), the largest differences are −4.7% in recovery and 3.4% in pressure drop (Case 8 at Condition 1). Results obtained with our new correlation for the zigzag configuration has the smallest relative difference in recovery up to a feed velocity of 0.2 m/s. At higher flow velocities, the correlation for the submerged configuration appears to have the smallest relative difference.
Although using different mass transfer and pressure drop correlations do not appear to have a strong influence on the predicted performance of a single pressure vessel, the effect could be more significant in a larger system that uses multiple pressure vessels in parallel. Accurate predictions are even more important for process design when the number of membrane modules and pressure vessels in a plant are determined based on a predictive model. For example, if a single stage RO plant with a capacity of 3 million/day is to be built, using the correlations adopted in Case 1 would predict the required number of pressure vessels to be approximately 34,705 whereas those at for Case 8 would predict 36,413. This discrepancy will lead to different estimates of capital and maintenance cost. Therefore, the selection of correlations implemented in an overall process model is important.

3.5. Limitations of the CFD Model and Derived Correlations

Although the CFD model adopted in this study accounts for the presence of spacers and membrane permeation and the simulations cover a wide range of operating and geometric conditions, the present study has two main limitations. Firstly, the boundary condition for permeable walls implemented in the CFD model needs to be experimentally validated, although the solution–diffusion model is widely accepted for RO membranes. Owing to the lack of experimental data obtained in the presence of spacers and RO membranes, indirect comparisons of CP moduli are made with existing correlations in the literature. Secondly, the spacer-filled fluid domain is simplified to allow for a large number of simulations. In particular, 2D representations of spacer-filled channels do not account for mesh and flow attack angles. The degree of flow recirculation and stagnation in 2D geometries might differ from that in 3D geometries when angles are varied, which would affect CP and pressure drop. Thus, it is necessary to compare 2D and 3D CFD models with various spacer geometries in order to determine whether 2D approximations are sufficient for the purpose of performance and energy prediction.

4. Conclusions

CFD models for spacer-filled feed channels in an RO membrane module are built for three commonly adopted configurations. By incorporating the solution–diffusion model, permeation phenomena through RO membranes can be captured in CFD simulations, allowing us to predict membrane performance under a wide range of operating and geometric conditions. Using results obtained from 228 CFD simulations, new correlations for CP modulus and friction factor are obtained, which fit the CFD data very well with a maximum absolute percentage error of 1.82% for CP modulus and 6.13% for pressure drop. A distinct advantage of our correlations for CP modulus is that they can direct predict the impact of permeate flux on the polarized concentration at membrane walls, bypassing the additional steps needed when using existing correlations based on Sherwood number.
Comparisons of our new correlations for CP modulus and the selected mass transfer correlations show that predicted water fluxes are comparable with a mean difference of less than 7%. The correlations for pressure drop, on the other hand, have a larger discrepancy, up to 84%. When implemented in a process model, the predicted performance and energy consumption for a single pressure vessel with six membrane modules differ by 4.7% for recovery and 1.6% for specific energy consumption. It is, therefore, concluded that this approach to predict a process-level system by exploiting new correlations derived from CFD results demonstrates its versatile applicability to various types of spacers and modules unlike the membrane manufacturer’s projection software, which can only predict their own products.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/membranes11050338/s1, Section A: Derivation of correlations: dimensional analysis, Section B: B. Supplementary CFD simulation results, Section C: C. Data fitting to estimate parameters in derived correlations.

Author Contributions

Conceptualization, B.G., C.S.A. and X.Y.X.; methodology, B.G.; software, B.G.; validation, B.G., C.S.A. and X.Y.X.; formal analysis and investigation, B.G.; writing—original draft preparation, B.G.; writing—review and editing, B.G., C.S.A. and X.Y.X.; visualization, B.G.; supervision, C.S.A. and X.Y.X.; project administration, C.S.A. and X.Y.X.; funding acquisition, C.S.A. and X.Y.X. All authors have read and agreed to the published version of the manuscript.

Funding

The authors would like to acknowledge the funding and technical support from BP through the BP International Centre for Advanced Materials (BP-ICAM) which made this research possible.

Data Availability Statement

The data presented in this study are available upon request from the corresponding author.

Acknowledgments

This article is based on Chapters 5 and 6 of the author’s PhD thesis [54].

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Schwinge, J.; Neal, P.R.; Wiley, D.E.; Fletcher, D.F.; Fane, A.G. Spiral wound modules and spacers: Review and analysis. J. Memb. Sci. 2004. [Google Scholar] [CrossRef]
  2. Sablani, S.; Goosen, M.; Al-Belushi, R.; Wilf, M. Concentration polarization in ultrafiltration and reverse osmosis: A critical review. Desalination 2001. [Google Scholar] [CrossRef]
  3. Gekas, V.; Hallström, B. Mass transfer in the membrane concentration polarization layer under turbulent cross flow: I. Critical literature review and adaptation of existing sherwood correlations to membrane operations. J. Memb. Sci. 1987. [Google Scholar] [CrossRef]
  4. Gekas, V.; Ölund, K. Mass transfer in the membrane concentration polarization layer under turbulent cross flow: II. Application to the characterization of ultrafiltration membranes. J. Memb. Sci. 1988. [Google Scholar] [CrossRef]
  5. Winograd, Y.; Solan, A.; Toren, M. Mass transfer in narrow channels in the presence of turbulence promoters. Desalination 1973. [Google Scholar] [CrossRef]
  6. Chiolle, A.; Gianotti, G.; Gramondo, M.; Parrini, G. Mathematical model of reverse osmosis in parallel-wall channels with turbulence promoting nets. Desalination 1978. [Google Scholar] [CrossRef]
  7. Schwager, F.; Robertson, P.M.; Ibl, N. The use of eddy promoters for the enhancement of mass transport in electrolytic cells. Electrochim. Acta 1980. [Google Scholar] [CrossRef]
  8. Kuroda, O.; Takahashi, S.; Nomura, M. Characteristics of flow and mass transfer rate in an electrodialyzer compartment including spacer. Desalination 1983. [Google Scholar] [CrossRef]
  9. Zimmerer, C.C.; Kottke, V. Effects of spacer geometry on pressure drop, mass transfer, mixing behavior, and residence time distribution. Desalination 1996. [Google Scholar] [CrossRef]
  10. Kim, W.S.; Park, J.K.; Chang, H.N. Mass transfer in a three-dimensional net-type turbulence promoter. Int. J. Heat Mass Transf. 1987. [Google Scholar] [CrossRef]
  11. Da Costa, A.R.; Fane, A.G.; Fell, C.J.D.; Franken, A.C.M. Optimal channel spacer design for ultrafiltration. J. Memb. Sci. 1991. [Google Scholar] [CrossRef]
  12. Da Costa, A.R.; Fane, A.G. Net-Type spacers: Effect of configuration on fluid flow path and ultrafiltration flux. Ind. Eng. Chem. Res. 1994. [Google Scholar] [CrossRef]
  13. Da Costa, A.R.; Fane, A.G.; Wiley, D.E. Spacer characterization and pressure drop modelling in spacer-filled channels for ultrafiltration. J. Memb. Sci. 1994. [Google Scholar] [CrossRef]
  14. Dražević, E.; Košutić, K.; Dananić, V. Mass transfer of differently sized organic solutes at spacer covered and permeable nanofiltration wall. Chem. Eng. J. 2014. [Google Scholar] [CrossRef]
  15. Van den Berg, G.B.; Rácz, I.G.; Smolders, C.A. Mass transfer coefficients in cross-flow ultrafiltration. J. Memb. Sci. 1989. [Google Scholar] [CrossRef] [Green Version]
  16. Prabhakar, S.; Ramani, M.P.S. A new concept of mass transfer coefficient in reverse osmosis—practical applications. J. Memb. Sci. 1994. [Google Scholar] [CrossRef]
  17. Geraldes, V.; Semiao, V.; Pinho, M.N. Hydrodynamics and concentration polarization in NF/RO spiral-wound modules with ladder-type spacers. Desalination 2003. [Google Scholar] [CrossRef]
  18. Song, L.; Ma, S. Numerical studies of the impact of spacer geometry on concentration polarization in spiral wound membrane modules. Ind. Eng. Chem. Res. 2005. [Google Scholar] [CrossRef]
  19. Ahmad, A.L.; Lau, K.K. Impact of different spacer filaments geometries on 2D unsteady hydrodynamics and concentration polarization in spiral wound membrane channel. J. Memb. Sci. 2006. [Google Scholar] [CrossRef]
  20. Ma, S.; Song, L.; Ong, S.L.; Ng, W.J. A 2-D streamline upwind Petrov/Galerkin finite element model for concentration polarization in spiral wound reverse osmosis modules. J. Memb. Sci. 2004. [Google Scholar] [CrossRef]
  21. Wardeh, S.; Morvan, H.P. CFD simulations of flow and concentration polarization in spacer-filled channels for application to water desalination. Chem. Eng. Res. Des. 2008. [Google Scholar] [CrossRef]
  22. Subramani, A.; Kim, S.; Hoek, E.M.V. Pressure, flow, and concentration profiles in open and spacer-filled membrane channels. J. Memb. Sci. 2006. [Google Scholar] [CrossRef]
  23. Ahmad, A.L.; Lau, K.K.; Abu Bakar, M.Z. Impact of different spacer filament geometries on concentration polarization control in narrow membrane channel. J. Memb. Sci. 2005. [Google Scholar] [CrossRef]
  24. Radu, A.I.; Vrouwenvelder, J.S.; van Loosdrecht, M.C.M.; Picioreanu, C. Effect of flow velocity, substrate concentration and hydraulic cleaning on biofouling of reverse osmosis feed channels. Chem. Eng. J. 2012. [Google Scholar] [CrossRef]
  25. Radu, A.I.; Bergwerff, L.; van Loosdrecht, M.C.M.; Picioreanu, C. A two-dimensional mechanistic model for scaling in spiral wound membrane systems. Chem. Eng. J. 2014. [Google Scholar] [CrossRef]
  26. Koutsou, C.P.; Yiantsios, S.G.; Karabelas, A.J. Direct numerical simulation of flow in spacer-filled channels: Effect of spacer geometrical characteristics. J. Memb. Sci. 2007. [Google Scholar] [CrossRef]
  27. Koutsou, C.P.; Yiantsios, S.G.; Karabelas, A.J. A numerical and experimental study of mass transfer in spacer-filled channels: Effects of spacer geometrical characteristics and Schmidt number. J. Memb. Sci. 2009. [Google Scholar] [CrossRef]
  28. Fimbres-Weihs, G.A.; Wiley, D.E. Numerical study of mass transfer in three-dimensional spacer-filled narrow channels with steady flow. J. Memb. Sci. 2007. [Google Scholar] [CrossRef]
  29. Shakaib, M.; Hasani, S.M.F.; Mahmood, M. CFD modeling for flow and mass transfer in spacer-obstructed membrane feed channels. J. Memb. Sci. 2009. [Google Scholar] [CrossRef]
  30. Schock, G.; Miquel, A. Mass transfer and pressure loss in spiral wound modules. Desalination 1987. [Google Scholar] [CrossRef]
  31. Kimura, S.; Sourirajan, S. Concentration polarization effects in reverse osmosis using porous cellulose acetate membranes. Ind. Eng. Chem. Process Des. Dev. 1968. [Google Scholar] [CrossRef]
  32. Kimura, S.; Sourirajan, S. Mass transfer coefficients for use in reverse osmosis process design. Ind. Eng. Chem. Process Des. Dev. 1968. [Google Scholar] [CrossRef]
  33. Lyster, E.; Cohen, Y. Numerical study of concentration polarization in a rectangular reverse osmosis membrane channel: Permeate flux variation and hydrodynamic end effects. J. Memb. Sci. 2007. [Google Scholar] [CrossRef]
  34. Rosén, C.; Trägårdh, C. Computer simulations of mass transfer in the concentration boundary layer over ultrafiltration membranes. J. Memb. Sci. 1993. [Google Scholar] [CrossRef]
  35. Salcedo-Díaz, R.; García-Algado, P.; García-Rodríguez, M.; Fernández-Sempere, J.; Ruiz-Beviá, F. Visualization and modeling of the polarization layer in crossflow reverse osmosis in a slit-type channel. J. Memb. Sci. 2014. [Google Scholar] [CrossRef] [Green Version]
  36. Li, F.; Meindersma, W.; De Haan, A.B.; Reith, T. Optimization of commercial net spacers in spiral wound membrane modules. J. Memb. Sci. 2002. [Google Scholar] [CrossRef] [Green Version]
  37. Gu, B.; Adjiman, C.S.; Xu, X.Y. The effect of feed spacer geometry on membrane performance and concentration polarisation based on 3D CFD simulations. J. Memb. Sci. 2017, 527, 78–91. [Google Scholar] [CrossRef]
  38. Schwinge, J.; Wiley, D.E.; Fletcher, D.F. A CFD study of unsteady flow in narrow spacer-filled channels for spiral-wound membrane modules. Desalination 2002. [Google Scholar] [CrossRef]
  39. Wijmans, J.G.; Baker, R.W. The solution-diffusion model: A review. J. Memb. Sci. 1995, 107, 1–21. [Google Scholar] [CrossRef]
  40. Soltanieh, M.; Gill, W.N. Review of reverse osmosis membranes and transport models. Chem. Eng. Commun. 1981. [Google Scholar] [CrossRef]
  41. Paul, D.R. Reformulation of the solution-diffusion theory of reverse osmosis. J. Memb. Sci. 2004. [Google Scholar] [CrossRef]
  42. Gu, B.; Xu, X.Y.; Adjiman, C.S. A predictive model for spiral wound reverse osmosis membrane modules: The effect of winding geometry and accurate geometric details. Comput. Chem. Eng. 2017. [Google Scholar] [CrossRef] [Green Version]
  43. Boudinar, M.B.; Hanbury, W.T.; Avlonitis, S. Numerical simulation and optimisation of spiral-wound modules. Desalination 1992. [Google Scholar] [CrossRef]
  44. Senthilmurugan, S.; Ahluwalia, A.; Gupta, S.K. Modeling of a spiral-wound module and estimation of model parameters using numerical techniques. Desalination 2005. [Google Scholar] [CrossRef]
  45. Misic, T.; Najdanovic-Lukic, M.; Nesic, L. Dimensional analysis in physics and the Buckingham theorem. Eur. J. Phys. 2010. [Google Scholar] [CrossRef]
  46. McCutcheon, J.R.; Elimelech, M. Influence of concentrative and dilutive internal concentration polarization on flux behavior in forward osmosis. J. Memb. Sci. 2006. [Google Scholar] [CrossRef]
  47. Gu, B.; Kim, D.Y.; Kim, J.H.; Yang, D.R. Mathematical model of flat sheet membrane modules for FO process: Plate-and-frame module and spiral-wound module. J. Memb. Sci. 2011, 379, 403–415. [Google Scholar] [CrossRef]
  48. Zhou, W.; Song, L.; Guan, T.K. A numerical study on concentration polarization and system performance of spiral wound RO membrane modules. J. Memb. Sci. 2006. [Google Scholar] [CrossRef]
  49. Greenlee, L.F.; Lawler, D.F.; Freeman, B.D.; Marrot, B.; Moulin, P. Reverse osmosis desalination: Water sources, technology, and today’s challenges. Water Res. 2009, 43, 2317–2348. [Google Scholar] [CrossRef] [PubMed]
  50. Kim, Y.M.; Kim, S.J.; Kim, Y.S.; Lee, S.; Kim, I.S.; Kim, J.H. Overview of systems engineering approaches for a large-scale seawater desalination plant with a reverse osmosis network. Desalination 2009. [Google Scholar] [CrossRef]
  51. Semiat, R. Energy issues in desalination processes. Environ. Sci. Technol. 2008, 42, 8193–8201. [Google Scholar] [CrossRef] [PubMed]
  52. Product Data Sheet for DOW. FilmtecTM SW30XHR-440i Element. Available online: https://www.dupont.com/content/dam/dupont/amer/us/en/water-solutions/public/documents/en/45-D00968-en.pdf (accessed on 30 April 2021).
  53. Avlonitis, S.; Hanbury, W.T.; Boudinar, M.B. Spiral wound modules performance. An analytical solution, part I. Desalination 1991. [Google Scholar] [CrossRef]
  54. Boram, G. Modelling of Reverse Osmosis Membrane Process and Transport Phenomena: From Feed Spacer to Large-Scale Plants. Ph.D. Thesis, Imperial College London, London, UK, 2017. [Google Scholar]
Figure 1. Three-dimensional spacer geometries (ac) and their cross-sections along the main flow path (df). (a) Nonwoven spacer, (b) fully woven spacer, (c) partially woven spacer. 2D cross-sections: (d) cavity, (e) submerged, and (f) zigzag configuration.
Figure 1. Three-dimensional spacer geometries (ac) and their cross-sections along the main flow path (df). (a) Nonwoven spacer, (b) fully woven spacer, (c) partially woven spacer. 2D cross-sections: (d) cavity, (e) submerged, and (f) zigzag configuration.
Membranes 11 00338 g001
Figure 2. Schematic illustration of strategy for varying operating and geometric conditions. (a) at a fixed geometry (b) at a fixed operating condition.
Figure 2. Schematic illustration of strategy for varying operating and geometric conditions. (a) at a fixed geometry (b) at a fixed operating condition.
Membranes 11 00338 g002
Figure 3. Average water fluxes (top) and CP moduli (bottom) in the cavity, submerged and zigzag configurations for varying inlet concentration c0 (a,e), diffusivity of solute in water Dsw (b,f), outlet pressure p0 (c,g), and inlet velocity u0 (d,h).
Figure 3. Average water fluxes (top) and CP moduli (bottom) in the cavity, submerged and zigzag configurations for varying inlet concentration c0 (a,e), diffusivity of solute in water Dsw (b,f), outlet pressure p0 (c,g), and inlet velocity u0 (d,h).
Membranes 11 00338 g003
Figure 4. Average water fluxes (top) and CP moduli (bottom) in the cavity, submerged and zigzag configurations for varying channel height at selected filament length to diameter (Lf/Df) ratios. (a,d): Lf/Df =4; (b,e): Lf/Df =8; (c,f): Lf/Df =12.
Figure 4. Average water fluxes (top) and CP moduli (bottom) in the cavity, submerged and zigzag configurations for varying channel height at selected filament length to diameter (Lf/Df) ratios. (a,d): Lf/Df =4; (b,e): Lf/Df =8; (c,f): Lf/Df =12.
Membranes 11 00338 g004
Figure 5. Pressure drop per unit length (top) and friction factor (bottom) in the cavity, submerged and zigzag configurations with respect to inlet velocity (a,e) and channel height with Lf/Df =4 (b,f), Lf/Df =/8 (c,g), Lf/Df =12 (d,h).
Figure 5. Pressure drop per unit length (top) and friction factor (bottom) in the cavity, submerged and zigzag configurations with respect to inlet velocity (a,e) and channel height with Lf/Df =4 (b,f), Lf/Df =/8 (c,g), Lf/Df =12 (d,h).
Membranes 11 00338 g005
Figure 6. Comparisons of CP modulus and water flux calculated by different correlations: Sherwood number correlations of Schock and Miquel [30] and Koutsou et al. [27] expressed in Equations (20) and (21), respectively, and our new correlations. (a) CP modulus and (b) water flux.
Figure 6. Comparisons of CP modulus and water flux calculated by different correlations: Sherwood number correlations of Schock and Miquel [30] and Koutsou et al. [27] expressed in Equations (20) and (21), respectively, and our new correlations. (a) CP modulus and (b) water flux.
Membranes 11 00338 g006
Figure 7. Comparison of pressure gradient calculated by different correlations: Schock and Miquel [30] and Koutsout et al. [26] expressed in Equations (23) and (24), respectively, and our new correlations for three spacer configurations.
Figure 7. Comparison of pressure gradient calculated by different correlations: Schock and Miquel [30] and Koutsout et al. [26] expressed in Equations (23) and (24), respectively, and our new correlations for three spacer configurations.
Membranes 11 00338 g007
Table 1. Model parameters and simulation conditions where inlet concentration c0, solute diffusivity in water Dsw, outlet pressure p0, inlet velocity u0, channel height Hc, and filament spacing to channel height ratio Lf-to-Hc are varied.
Table 1. Model parameters and simulation conditions where inlet concentration c0, solute diffusivity in water Dsw, outlet pressure p0, inlet velocity u0, channel height Hc, and filament spacing to channel height ratio Lf-to-Hc are varied.
Parameters and ConditionsCase Number
Base1–78–1516–1920–3031–75
Hydraulic conductivity, A (m/(s·Pa))2.50 × 10−12
Solute permeability, B (m/s)2.50 × 10−8
Osmotic factor, κ (m3·Pa/mol)4958
Density, ρ (kg/m3)998.20
Viscosity, μ (Pa·s)8.93 × 10−4
Inlet concentration, c0 (mol/m3)603.45362.07–905.17603.45603.45603.45603.45
Solute diffusivity, Dsw (10−9 m2/s)110.2–8111
Outlet pressure, p0 (105 Pa)60606050–906060
Mean inlet velocity, u0 (m/s)0.10.10.10.10.01–0.350.1
Channel height, Hc (10−3 m)111110.2–1.4
Filament spacing to channel height ratio, Lf/Hc (-)4.54.54.54.54.52–6
Table 2. List of parameters for RO process simulation.
Table 2. List of parameters for RO process simulation.
Symbol (Unit)NameValueSource
Nele (-)Number of modules arrayed in series6[49,50,51]
Nleaf (-)Number of leaves16[42,52]
Li (m)Membrane length1.016[52]
Wi (m)Membrane width1.505[42,52]
Hc (m)Feed spacer thickness7.1 × 10−4[52]
Hc,p (m)Permeate spacer thickness4.0 × 10−4[52]
Lg (m)Glue line width in length direction4 × 10−2[53]
Wg (m)Glue line width in width direction17 × 10−2[53]
Df (m)Filament diameter3.85 × 10−4[30]
Lf (m)Filament spacing2.8 × 10−3[30]
ε (-)Voidage of spacer-filled channel0.89[30]
dh (m)Hydraulic diameter of spacer-filled channel0.95 × 10−3[30]
A (m3/m2 s Pa)Pure water permeability constant3.26 × 10−12Estimated
B (m3/m2 s)Salt permeability constant1.15 × 10−8Estimated
ksp,f (-)Frictional coefficient in feed channel1.58Estimated
ksp,p (-)Frictional coefficient in permeate channel30Estimated
ηHPP (-)Efficiency of high pressure pump0.85[42,49,50,51]
ηPX (-)Efficiency of pressure exchanger0.98[42,49,50,51]
Table 3. Estimated parameter values for CP modulus and pressure drop correlations in three configurations and APEs of the correlations compared to the corresponding CFD results.
Table 3. Estimated parameter values for CP modulus and pressure drop correlations in three configurations and APEs of the correlations compared to the corresponding CFD results.
RegimeCavitySubmergedZigzag
RecRecritRec < RecritRecRecritRec < Recrit
CP modulus, MCP
in Equation (16)
Estimated parameter valuesC2.55 × 10−47.06 × 10−45.55 × 10−33.24 × 10−47.63 × 10−4
α−0.350−0.382−0.422−0.394−0.365
β1.110.9311.091.120.909
γ0.6110.6110.6720.5970.597
δ5.40 × 10−4−0.1430.5362.18 × 10−4−0.112
Mean APE [%]0.3750.1470.516
Friction factor, f
in Equation (17)
Estimated parameter valuesC14.863.314.0
α−0.910−0.994−0.898
δ−0.525−0.810−0.510
ζ0.02560.08680.0200
Mean APE [%]1.151.480.974
Table 4. Average and maximum absolute percentage difference between the new correlations and existing correlations for CP moduli and water fluxes.
Table 4. Average and maximum absolute percentage difference between the new correlations and existing correlations for CP moduli and water fluxes.
CorrelationAbsolute Percentage Difference, |YYex| /Yex (%)
VariableExistingSchock and Miquel [30]Koutsou et al. [27]
NewCavitySubmergedZigzagCavitySubmergedZigzag
CP modulusMean1.94.51.85.51.85.4
Max.6.710.36.55.61.95.6
Water fluxMean2.76.42.65.61.65.5
Max.11.016.510.75.81.76.1
Table 5. Average and maximum absolute percentage differences between our new correlations and existing correlations for pressure gradient.
Table 5. Average and maximum absolute percentage differences between our new correlations and existing correlations for pressure gradient.
CorrelationAbsolute Percentage Difference (%)
ExistingSchock and Miquel [30]Koutsou et al. [26] (Lf/Df = 8)Koutsou et al. [26] (Lf/Df = 6)
NewCavitySubmergedZigzagCavitySubmergedZigzagCavitySubmergedZigzag
Mean541255663067805880
Max.632864744475846484
Table 6. Simulation results for water recovery Rw, specific energy consumption Esp and feed pressure drop Pd, with different combinations of CP and pressure drop correlations under three simulation conditions (with feed concentration cf = 35kg/m3).
Table 6. Simulation results for water recovery Rw, specific energy consumption Esp and feed pressure drop Pd, with different combinations of CP and pressure drop correlations under three simulation conditions (with feed concentration cf = 35kg/m3).
Case
No.
CorrelationCondition 1Condition 2Condition 3
CPdp/dxQf = 2 × 10−3 m3/s (uf = 0.13 m/s), Pf = 60 barQf = 3 × 10−3 m3/s (uf = 0.20 m/s), Pf = 70 barQf = 4.5 × 10−3 m3/s (uf = 0.30 m/s), Pf = 80 bar
Rw
(%)
Esp
(kWh/m3)
Pd
(×105 Pa)
Rw
(%)
Esp
(kWh/m3)
Pd
(×105 Pa)
Rw
(%)
Esp
(kWh/m3)
Pd
(×105 Pa)
1Equation (25)Equation (26)50.032.060.7253.432.391.0852.922.741.67
Relative difference [%]Condition 1Condition 2Condition 3
ΔRwΔEspΔPdΔRwΔEspΔPdΔRwΔEspΔPd
2Equation (25)Zig *0.53−0.64−530.56−0.66−490.65−0.81−44
3Equation (25)Sub0.066−0.096−8.0−0.0560.0342.3−0.360.3719
4Equation (25)S&M +0.31−0.40−340.075−0.15−12−0.430.3920
5Equation (25)KT8 $0.031−0.097−8.4−0.420.3928−1.51.683
6Zig *Equation (26)0.12−0.00760.34−1.00.101.1−2.60.281.6
7SubEquation (26)1.6−0.15−1.01.1−0.10-0.490.25−0.0230.081
8S&M ++Equation (26)-4.70.493.4−4.70.472.9−4.30.472.1
9KT $$Equation (26)1.9−0.18−1.51.9−0.18−1.21.7−0.17−0.78
* The correlation for the zigzag configuration is chosen due to negligible difference between the cavity and zigzag configurations. + The correlation of Schock and Miquel for pressure gradient [30] in Equation (23). ++ The correlation of Schock and Miquel [30] for mass transfer coefficient in Equation (20). $ The correlation of Koutsou et al. [26] for pressure gradient when Lf/Df = 8 in Equation (24); $$ The correlation of Koutsou et al. [27] for mass transfer coefficient in Equation (21).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gu, B.; Adjiman, C.S.; Xu, X.Y. Correlations for Concentration Polarization and Pressure Drop in Spacer-Filled RO Membrane Modules Based on CFD Simulations. Membranes 2021, 11, 338. https://doi.org/10.3390/membranes11050338

AMA Style

Gu B, Adjiman CS, Xu XY. Correlations for Concentration Polarization and Pressure Drop in Spacer-Filled RO Membrane Modules Based on CFD Simulations. Membranes. 2021; 11(5):338. https://doi.org/10.3390/membranes11050338

Chicago/Turabian Style

Gu, Boram, Claire S. Adjiman, and Xiao Yun Xu. 2021. "Correlations for Concentration Polarization and Pressure Drop in Spacer-Filled RO Membrane Modules Based on CFD Simulations" Membranes 11, no. 5: 338. https://doi.org/10.3390/membranes11050338

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop