Next Article in Journal
Energy Saving Evaluation with Low Liquid to Gas Ratio Operation in HVAC&R System
Next Article in Special Issue
CFD-Based Sensitivity-Analysis and Performance Investigation of a Hydronic Road-Heating System
Previous Article in Journal
A Coal Seam Thickness Prediction Model Based on CPSAC and WOA–LS-SVM: A Case Study on the ZJ Mine in the Huainan Coalfield
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Challenges Facing Pressure Retarded Osmosis Commercialization: A Short Review

by
Bassel A. Abdelkader
and
Mostafa H. Sharqawy
*
School of Engineering, University of Guelph, Guelph, ON N1G 2W1, Canada
*
Author to whom correspondence should be addressed.
Energies 2022, 15(19), 7325; https://doi.org/10.3390/en15197325
Submission received: 29 August 2022 / Revised: 27 September 2022 / Accepted: 29 September 2022 / Published: 5 October 2022

Abstract

:
Pressure-retarded osmosis (PRO) is a promising technology that harvests salinity gradient energy. Even though PRO has great power-generating potential, its commercialization is currently facing many challenges. In this regard, this review highlights the discrepancies between the reported power density obtained by lab-scale PRO systems, as well as numerical investigations, and the significantly low power density values obtained by PRO pilot plants. This difference in performance is mainly due to the effect of a pressure drop and the draw pressure effect on the feed channel hydrodynamics, which have significant impacts on large-scale modules; however, it has a minor or no effect on small-scale ones. Therefore, this review outlines the underlying causes of the high power density values obtained by lab-scale PRO systems and numerical studies. Moreover, other challenges impeding PRO commercialization are discussed, including the effect of concentration polarization, the solution temperature, the pressure drop, and the draw pressure effect on the feed channel hydrodynamics. In conclusion, this review sheds valuable insights on the issues facing PRO commercialization and suggests recommendations that can facilitate the successful development of PRO power plants.

1. Introduction

Salinity gradient energy (SGE) is a promising form of renewable energy that provides clean sustainable energy and recently emerged as an alternative to fossil fuel power plants. SGE is generated from the salinity difference between two solutions, such as freshwater and seawater. SGE was first introduced by Pattle [1] as the energy available through the mixing of seawater and freshwater, where the energy generated depends on the osmotic pressure difference between the two solutions. There are several processes that utilize SGE to produce power, such as reverse electrodialysis (RED), pressure-retarded osmosis (PRO), capacitive mixing (CAPMIX), and reverse vapor compression (RVC) [2]. RED utilizes the electrochemical potential difference between the solutions through the directional flow of ions across ion-exchange membranes, generating power that is captured by electrodes [3]. PRO utilizes the osmotic pressure difference between two different salinity solutions across a semipermeable membrane to generate power [4]. CAPMIX, on the other hand, does not involve the use of membranes but uses porous electrodes submerged in an electrolyte to generate power from the chemical potential energy [5]. Similar to CAPMIX, RVC does not involve the use of membranes but generates power by utilizing the difference in vapor pressure between freshwater and seawater [6].
The performance of any SGE process can be assessed by determining its efficiency and power density. The efficiency indicates the amount of useful work produced relative to the available energy, which is the Gibbs free energy of mixing the two solutions. The power density is defined as the power produced per membrane surface area for PRO and RED, or per unit electrode mass in the case of CAPMIX. PRO has a higher power density when compared with RED and CAPMIX [7,8]. In addition, PRO has a higher efficiency range (54–56%) compared with RED (18–38%), as reported by Yip and Elimelech [9]. Thus, PRO has a greater potential for power generation compared with other SGE technologies.
Compared with other forms of renewable energy technologies, PRO has a higher availability factor, as it is able to generate a constant power supply for base loads [10,11,12,13,14,15,16]. In addition, PRO has a higher energy conversion efficiency compared with other renewable energy technologies, such as those generating solar, geothermal, and wind energies [17]. The energy conversion efficiency is the ratio between the energy generated and the energy that is input or available. In the case of PRO, this efficiency is the ratio between the energy produced and the Gibbs free energy of mixing. The global power generation potential of PRO is estimated to be around 157 GW after considering losses and inefficiencies when using only 10% of the global rivers’ water discharge [18]. This estimate can cover the energy demand of around half a billion people according to the average worldwide electricity consumption [18].
PRO is an osmotically driven process that utilizes the osmotic pressure difference between low- and high-salinity solutions, which was first described and patented by Leob [19] in 1975. Figure 1 shows a schematic of the basic components and the flow of solutions in the PRO process. In PRO, a semi-permeable membrane fixed in a membrane module separates the two solutions of different salt concentrations. The high-salinity solution (draw solution) is pressurized to a high pressure that is still lower than the osmotic pressure, while the low-salinity solution (feed solution) flows at a low pressure that overcomes the pressure drop in the module flow channel. The osmotic pressure difference between the two solutions drives water from the feed solution to the draw solution across the membrane [20,21]. The flow of water through the membrane dilutes the draw solution and increases its volume flow rate. A hydro turbine is then utilized to depressurize the high-pressure draw solution to generate energy that feeds the high-pressure pump and provides net positive power. The power generated by PRO is directly related to the amount of water permeated through the membrane from the feed solution to the draw solution, which depends on the osmotic pressure difference, as well as the membrane characteristics and operating conditions.
The high- and low-salinity solutions for PRO plants are available from various natural and industrial sources. The feed and draw solutions could be from freshwater and seawater present at river mouths, seawater, and brine rejected from desalination plants, or freshwater and wastewater rejected from water treatment plants [21,22]. As the concentration difference between the feed and draw solutions increases, the power generated increases significantly [23,24]. It was reported that the PRO power density increased by 6.5 times as the draw solution concentration increased from 0.5 M NaCl (i.e., 0.5 mol/liter NaCl solution, which is close to seawater salinity) to 1 M NaCl (similar to the salinity of rejected brine from a seawater reverse osmosis (RO) desalination plant), and 17.5 times as the draw concentration increased to 2 M NaCl (similar to brine rejected from a seawater FO desalination plant) with a feed concentration of 0.05 M NaCl [25]. However, at a higher feed solution concentration, the power density is significantly lower at the same concentration difference between the feed and draw solutions due to the severe internal concentration polarization [25], as discussed later.
Despite its great potential, the commercialization of PRO faces several issues that limit its use for large-scale power production. One of the main challenges is the inefficiency of the commercially available membrane module and spacers for use in PRO [26,27,28,29,30]. The current spacers were reported to cause channel blockage and severe membrane deformation, increasing the pressure drop, and hence, reducing the PRO performance [29,30]. Moreover, the PRO performance is greatly reduced by the effect of the concentration polarization, which can either be internal, occurring inside the membrane support layer, or external, occurring near the membrane surface [21,31]. Another challenge limiting PRO performance is the occurrence of fouling, which is more significant when high-salinity solutions, such as seawater or wastewater, are used as the feed solution [32,33,34]. Moreover, the temperatures of the draw and feed solutions can significantly influence the PRO performance [35]. The solution temperatures depend on the geographical location and seasonal temperature fluctuations, as seawater temperatures can fluctuate from −2 to 35 °C throughout the year [36].
This review aimed to enhance the understanding of the challenges facing PRO commercialization. It starts with a discussion of the solution diffusion model, along with a comparison between the water flux predictions estimated using the different modified versions of the model. This is followed by a comprehensive comparison of the performance of plate-and-frame and spiral wound modules using commercially available membranes. This comparison surveys the experimental and numerical data of the performance of both modules available in the literature. A detailed discussion of the challenges facing PRO commercialization is then included, some of which were encountered and reported by the first PRO pilot plant. Finally, this review offers recommendations and insights on the future research required to achieve successful PRO commercialization.

2. Modeling of PRO Process

2.1. Osmotic Processes

Osmosis can be described as the spontaneous flow of water (or solvent) from a solution of high water potential (low salinity) to a solution of low water potential (high salinity) across a semipermeable membrane. There are three basic osmotic processes with different water flow directions that depend on both the osmotic and applied hydraulic pressure difference, namely, forward osmosis (FO), reverse osmosis (RO), and PRO. In an FO process, there is no hydraulic pressure difference applied and hence the water flux is generated by the osmotic pressure difference only from the low-salinity solution to the high-salinity solution. In PRO, the hydraulic pressure difference applied to the high-salinity solution is lower than the osmotic pressure difference, and hence, water will still flow from the low-salinity to the high-salinity solution, similar to the FO process. In RO, the hydraulic pressure difference applied is higher than the osmotic pressure difference and water flows in the opposite direction (i.e., from the high-salinity to the low-salinity solution). The flux reversal point is theoretically defined as the point at which the hydraulic pressure difference is equal to the osmotic pressure difference. In reality, the flux reversal point occurs at a hydraulic pressure difference that is lower than the osmotic pressure, as shown in Figure 2, due to the combined effects of the concentration polarization, draw solution dilution, and pressure drop. The PRO performance is assessed in terms of the power density, which is defined as the power generated per unit membrane area. Theoretically, the maximum power density occurs at the point at which the hydraulic pressure difference is equal to half the osmotic pressure difference. However, similar to the flux reversal point, the maximum power density is achieved at a lower hydraulic pressure difference, as shown in Figure 2.

2.2. Solution Diffusion Model

2.2.1. Neglecting the Effect of Concentration Polarization

The water flux across a semipermeable membrane is defined by the following equation:
J w = A Δ π Δ P
where A is the water permeability coefficient, Δπ is the osmotic pressure difference, and ΔP is the hydraulic pressure difference. The PRO performance is assessed through the power density W, which is the power generated per membrane surface area and is given by
W = J w Δ P = A Δ π Δ P Δ P
To find the optimal ΔP that achieves a maximum power density, Equation (2) should be differentiated with respect to ΔP. The maximum power density occurs at ΔP = Δπ/2, yielding the following expression for the theoretical maximum power density:
W m a x = A Δ π 2 4
In this theoretical maximum power density expression, the effects of the concentration polarization and salt flux across the membrane are neglected. In real membranes, the concentration polarization reduces the osmotic pressure difference and subsequently lowers the water flux reversal point. In addition, the salt flux across the membrane changes both feed and draw solution concentrations along the membrane length, reducing the osmotic pressure difference. Therefore, the combined effects of the concentration polarization and salt flux in real membranes will result in the maximum power density occurring at a lower pressure difference compared with the theoretical maximum power density, i.e., ΔP < Δπ/2.

2.2.2. Lee’s Model

In 1981, Lee et al. [37] considered the effect of internal concentration polarization (ICP) in the solution diffusion model but ignored the effect of the external concentration polarization (ECP). Referring to Figure 3, they neglected the ECP by assuming that C D , b = C D , m and C F , b = C F , m . This assumption is valid if there is a high degree of turbulence near the membrane surface. Therefore, they modified the osmotic pressure difference in Equation (1), yielding the following expression:
J w = A Δ π e f f Δ P = A π D , m π i c p Δ P
where Δ π e f f is the effective osmotic pressure difference between the active layer membrane surface ( π D , m ) and the active–support layer interface ( π i c p ), as shown in Figure 3. Using van’t Hoff’s model, the osmotic pressure as a function of the concentration is given as
π = β R T C
where β is the van’t Hoff coefficient, Ru is the universal gas constant, T is the absolute temperature, and C is the salt concentration.
The salt (or solute) flux across the membrane active layer is defined by Equation (6):
J s = B C D , m C i c p
where B is the salt permeability coefficient, C D , m is the concentration at the active layer, and C i c p is the concentration at the active–support layer interface.
The mass transport of the salt across the membrane support layer is caused by diffusion due to the presence of the salt gradient and convection due to water flux across the membrane. The effect of diffusion is represented by the first term in Equation (7), while the second term represents the convection effect:
J s = D ε d C x d x J w C x
where D is the diffusion coefficient in the membrane support layer and ε is the porosity of the membrane support layer. Rearrangement of Equation (7) gives
1 C x + J s J w d C x = J w D ε d x
Integrating Equation (8) using the boundary conditions C x = C F , m   a t   x = 0 and C x = C i c p   a t   x = τ t , where x is the distance measured from the membrane support layer toward the membrane active layer, yields the following expression:
C i c p = C F , m exp J w τ t D ε + J s J w exp J w τ t D ε 1
where t represents the thickness of the membrane support layer and τ is the tortuosity of the membrane support layer. Substituting Equation (6) into Equation (9) yields the following equation:
C i c p = B exp J w K 1 C D , m C i c p J w + C F , m exp J w K
where K is the solute resistivity for diffusion inside the support layer, as given by Equation (11):
K = τ t D ε = S D
where S is the membrane structure parameter.
Rearranging Equation (10) and Equation (4) gives
C i c p C D , m = B exp J w K 1 + J w C F , m C D , m exp J w K B exp J w K 1 + J w
J w = A π D , m 1 π i c p π D , m Δ P
Using van’t Hoff’s equation (Equation (5)), the ratio of the salt concentrations can be assumed to be equal to the ratio of the osmotic pressure (i.e., π i c p π D , m = C i c p C D , m ). Using this assumption, Equations (13) and (12) can be combined to yield the following expression for the water flux:
J w = A π D , m 1 C F , m C D , m exp J w K 1 + B J w exp J w K 1 Δ P
As the effect of the external concentration polarization is neglected, the concentrations at the membrane surface can be replaced with the bulk low concentrations, i.e., C D , b = C D , m and C F , b = C F , m . Therefore, Equation (14) yields
J w = A π D , b 1 C F , b C D , b exp J w K 1 + B J w exp J w K 1 Δ P

2.2.3. Solution Diffusion Model Development

In the model proposed by Lee et al. [37], the effect of the ECP was neglected. Therefore, later studies proposed several modifications to include the ECP effect in the model.

Achilli’s Model

The first modification to the solution diffusion model was proposed by Achilli et al. [38], where the effect of the ECP was considered. The dilutive ECP (at the draw side) was obtained based on film theory, where the osmotic pressure ratio in the draw side boundary layer is given by Equation (16) [39,40]:
π D , m π D , b = exp J w k D
where k D is the mass transfer coefficient of the draw side.
Substituting Equation (16) into Equation (14) (which already considers the ICP effect) and assuming the ratio of the osmotic pressures to be equal to the ratio of the salt concentrations, i.e., C F , b C D , m = π F , b π D , m , gives the following equation for the water flux:
J w = A π D , b exp J w k D π F , b exp J w K 1 + B J w exp J w K 1 Δ P

Yip’s Model

Another model by Yip et al. [41] considered the effect of the reverse salt flux and ECP by assuming the osmotic pressure to be linearly proportional to the salt concentration. Similar to the ICP effect, the salt flux across the boundary layer (on the draw side) is caused by diffusion and convection, as given by Equation (18):
J s = D d C x d x J w C x
Rearrangement of Equation (18) gives
1 C + J s J w d C = J w D d x
Integrating Equation (19) using the boundary conditions C x = C D , m   a t   x = 0 and C x = C D , b   a t   x = δ D , where x is the distance measured from the membrane active layer and δ D is the draw boundary layer thickness, gives the following expression for the draw salt concentration at the membrane surface:
C D , m = C D , b exp J w δ D D + J s J w exp J w δ D D 1
In Equation (20), the first term on the RHS represents the draw concentration at the membrane surface, while the second term represents the reduction in salt concentration due to the salt flux across the membrane active layer. Subtracting Equation (9) from Equation (20) and substituting J s from Equation (6) and rearranging yields the following expression:
C D , m C i c p = C D , b e x p J w k D C F , b e x p J w K 1 + B J w e x p J w K e x p J w k D
where k D is the mass transfer coefficient of the draw solution ( k D = D / δ D ). Assuming the osmotic pressure difference to be linearly proportional to the salt concentration difference, i.e., Δ π m = Δ C m and substituting Equation (21) into Equation (1) yields the following expression for the water flux:
J w = A π D , b exp J w k D π F , b exp J w K 1 + B J w exp J w K exp J w k D Δ P

Touati’s Model

The previously discussed modifications only considered the effect of the ECP on the draw side, neglecting its effect on the feed side. Touati et al. [42] derived a new expression to account for the effect of the external concentration polarization on both the draw and feed sides, as well as the ICP. Similar to the salt flux calculation on the draw side carried out by Yip et al. [41], the salt flux across the boundary layer on the feed side is given by Equation (18). By integrating Equation (18) using the boundary conditions C x = C F , m   a t   x = 0 and C x = C F , b   a t   x = δ F , Equation (23) is obtained, where x is the distance from the membrane support layer and δ F is the feed boundary layer thickness:
C F , m = C F , b exp J w δ F D + J s J w exp J w δ F D 1
Substituting Equations (9), (20), and (23) into Equation (5) yields
π i c p = β R u T C F , b exp J w k F + J s J w exp J w k F 1 exp J w K + J s J w exp J w K 1
π D , m = β R u T C D , b exp J w k D + J s J w exp J w k D 1
The ratio of the salt flux to the water flux is influenced by the water and salt permeability coefficients of the membranes, as shown by Equation (26) [27]. The ratio of the salt flux to the water flux can be solved by combining Equations (4)–(6), yielding the following expression:
J s J w = B A β R u T 1 + A Δ P J w
By substituting the equations for osmotic pressure (i.e., Equations (24) and (25)) and Equation (26) into Equation (4), the following expression for water flux is obtained:
J w = A π D , b + B A 1 + A Δ P J w exp J w k D π F , b + B A 1 + A Δ P J w exp J w K exp J w k F Δ P

Comparison between Different Models

A summary of the several modifications to the solution diffusion model is presented in Table 1. Generally, the modified versions of the solution diffusion model predicted results that were closer to those obtained experimentally when the effect of the ECP was considered. The predicted water flux using the different modified versions of the solution diffusion model were compared with the experimental results conducted using a draw solution of 20 g/L NaCl at 20 °C [43]. The comparison is shown in Figure 4. When the effects of both the internal and external concentration polarization were neglected, the water flux was calculated using Equation (1). The average percentage error in the water flux results obtained in this case was around 40%, while the average percentage error for Lee et al. [37] was around 17% when compared with the experimental results. However, for all the modified versions of the solution diffusion model that considered the effect of the ECP, the percentage error obtained was less than 1% when compared with the experimental data. Thus, the modified versions predict water flux values that were very close and displayed a good agreement with experimental data, as shown in Figure 4.

3. PRO Membranes and Modules

The PRO process uses semi-permeable membranes that allow for water flow and prevent salt flow. Membranes suitable for use in PRO should meet several criteria including their ability to withstand relatively high pressure, which optimally should be half of the osmotic pressure [44,45]. The membranes should also have low salt permeability and high water permeability to enhance the water flux [44]. In addition, they should have a low structure parameter to minimize the internal concentration polarization. In most PRO systems, two types of membranes are commonly used, namely, (1) membranes synthesized using cellulose acetate (CA) or cellulose triacetate (CTA) and (2) thin-film composite (TFC) membranes, which include an active nonporous layer on a porous support layer [46,47]. These membranes are arranged in different module designs to maximize the area-to-volume ratio, such as plate-and-frame, hollow fiber, tubular, and spiral wound modules [48,49], as shown in Figure 5. The following section discusses two of these modules, namely, plate-and-frame and spiral wound modules, as plate-and-frame modules are most commonly used in laboratory investigations of PRO, while spiral wound modules were used in pilot PRO power plants.

3.1. Plate-and-Frame Module

A plate-and-frame module (Figure 5a) is the simplest design for packing flat sheet membranes to increase the surface-area-to-volume ratio. In a plate-and-frame module, a membrane is sandwiched between a feed spacer and a draw spacer, and all of these are layered between two end plates. Feed and draw solutions are forced to flow parallel to the membrane surface in their corresponding channels, as shown in Figure 5a. Despite its simple design and easy manufacturing, plate-and-frame modules are not preferable for use in large-scale PRO power plants, as they have a lower surface-area-to-volume ratio compared with spiral wound and hollow fiber modules [29,30]. The surface-area-to-volume ratio for a plate-and-frame module ranges from 100 to 400 m2/m3 [50].
A single flat sheet membrane module (or membrane cell) is typically used in lab-scale studies to assess the PRO performance and measure membrane properties. Table 2 summarizes the performance and properties of several commercial membranes tested (experimental investigation) and simulated (numerical investigation) in flat sheet membrane modules. The water and salt permeability coefficients (A and B) and the membrane structure parameter (S) were determined by carrying out RO and FO tests using bench-top flat sheet modules. The experimentally determined membrane properties were used in numerical investigations to assess the PRO performance under different operating conditions using the solution diffusion model. The results obtained by the numerical and experimental investigations of flat sheet modules were commonly found to be in good agreement, as the measured and estimated power density are in the same range, as shown in Figure 6. This figure shows that the power densities from a few investigations of a membrane cell that used hypersaline draw solutions were higher than 56.4 W/m2, which is considered the minimum power density for PRO to be as cost-effective as solar photovoltaic (PV) panels [51,52].

3.2. Spiral Wound Module (SWM)

The other common module used in PRO systems is the spiral wound module (SWM), which is more suitable for use in large-scale plants since it has a large membrane-area-to-volume ratio (300–1000 m2/m3) [50]. An SWM consists of several membrane envelopes and spacers wrapped around a perforated tube, as shown in Figure 5c. The commercially available reverse osmosis SWM was modified by Foreman and Worsley [61] to be suitable initially for FO and later for the PRO process. In their design, a central glue line was added to the membrane facing the feed side, causing feed water to flow back to the central tube through a 180-degree turn, as shown in Figure 7. This causes a significant increase in the pressure drop on the feed water side due to the long flow path. In addition, this design was reported to be inadequate for PRO due to the presence of areas of low flow velocity and poor flow distribution [30].
Several studies experimentally investigated the performance of an SWM working under PRO conditions [29,62,63,64,65]. They reported many performance issues, such as the draw solution’s high-pressure effect, the deviation between experimental measurements and the model prediction, and differing performance between large and small module areas. Kim et al. [29] experimentally investigated the performance of a PRO-SWM with a membrane area of 29 m2. They reported a large increase in the feed solution pressure drop as water flowed through the draw channel. Xu et al. [65] reported a similar increase in the feed solution pressure drop, which was caused by the draw solution’s high pressure, when using a PRO-SWM with a membrane area of 0.94 m2. The PRO performance of a small SWM module with a membrane area of 0.5 m2 was investigated both numerically and experimentally by Attarde et al. [62]. Using numerical analysis, they achieved a maximum power density of 3.5 W/m2; however, they were unable to obtain this power density experimentally due to the hydraulic pressure limit of the membrane (5 bars).
Several studies numerically investigated the performance of an SWM [62,66,67,68,69]. The reported power density values of these studies were noticed to be higher compared with the experimental results. Matta et al. [66] numerically investigated the performance of an SWM with a membrane area of 30 m2, assuming a linear pressure drop across the length of the membrane. Although the results of their numerical analysis were validated against experimental data of a flat sheet module with an area of 20 cm2, their obtained power density was overestimated compared with the experimental results obtained by Kim et al. [29] for the same SWM. Another study by Altaee and Cippolina [67] also overestimated the power density values of a large-scale SWM, which might have been due to the fact that they ignored the effect of the pressure drop along the module, yielding a higher water flux. The overestimation of power density values reported by these numerical investigations was also reported by Binger and Achilli [68]. They compared the power density obtained by their numerical model to the experimental results of Achilli et al. [63] for a PRO-SWM with an area of 4.18 m2. The overestimation of the power density by some numerical investigations of SWM compared with experimental data is clearly shown in Table 3 and Figure 8, where the numerically calculated power density values are higher than the experimentally measured power density for the same membrane properties. Although several numerical SWM investigations predicted a high power density of up to 16 W/m2, the highest power density achieved through experimental investigations of the SWM was 3.5 W/m2. From Table 3, the results obtained via numerical and experimental investigations of small-scale modules were comparable. However, in large-scale modules, the power density obtained using experimental investigations was considerably lower than those obtained using numerical investigations. This could have been due to the performance of large-scale modules being significantly influenced by the effect of the draw solution’s high hydraulic pressure on the feed side pressure drop [29,65]. However, this effect is not significant in the case of small SWM and flat sheet modules. Thus, the effect of the draw hydraulic pressure on feed side pressure drop should be considered in numerical investigations when assessing the performance of a large SWM.

4. First PRO Pilot Plant

The first attempt at commercial power generation using PRO was a pilot plant in Norway; it was built by Statkraft in 2009 and later terminated in 2014 [70,71]. They used spiral wound modules with different membrane types and generated power with a power density of 0.9 W/m2, which was significantly lower than the expected power density. The inadequate production of energy could have been due to the use of inefficient spacers, as the commercially available spacers are not suitable for PRO application [30]. These spacers lead to severe membrane deformation as the hydraulic draw pressure increases, which blocks the feed channel, increasing the pressure drop on the feed side and reducing the performance [26,27,28,53]. Moreover, the module design used was reported to be unsuitable, as it resulted in the presence of areas of low flow velocity and poor flow distribution [30]. Another important factor that influenced the power density generated was the effect of the draw solution dilution [17,59]. Throughout the length of the membrane, the effective osmotic pressure difference decreases from the inlet to the outlet. This effect is more pronounced in PRO systems with a large membrane area due to the large changes in salt concentrations at both the draw and feed sides. Additionally, the presence of both internal and external concentration polarization, as well as membrane fouling, which could be reduced but not eliminated by regular cleaning, reduces the performance of PRO systems [32,33,72]. The inefficiency of the first PRO power plant prototype indicated the need to optimize the design of PRO membranes, modules, spacers, and operating conditions to reach the power-generating potential of PRO. This encouraged several efforts to enhance the design of modules and spacers to enhance PRO performance [73,74]. Some of these efforts included studies that improved the design of SWMs by modifying the central tubes [75] and modifying the arrangement of the membrane and spacers [76,77,78]. New module designs were also proposed, such as those consisting of a stack of membrane plate assemblies [79]. Moreover, to reduce membrane damage and increase its mechanical strength, a woven or non-woven mesh was added as membrane support [80], or a base layer with mechanical reinforcement was added [81]. However, despite these notable efforts, PRO commercialization is yet to be achieved.

5. PRO Challenges

5.1. Concentration Polarization

Concentration polarization is the accumulation or depletion of solutes near the membrane interface due to water flow from the feed side to the draw side of the membrane. Thus, creating a smaller concentration difference across the membrane active layer when compared with the bulk concentration difference across the membrane, which negatively affects the overall performance [82,83]. Concentration polarization can either be internal, which occurs inside the membrane support layer, or external, which occurs near the membrane surface, as shown in Figure 3. The presence of internal concentration polarization in the support layer occurs due to the accumulation of salts in the active–support layer interface, and this concentration is represented by Cicp. This reduces the effective osmotic pressure difference across the active membrane layer and consequently reduces the water flux across the membrane. The internal concentration polarization is controlled by the membrane structure parameter (S), which is the characteristic length that an ion diffuses through in the membrane support layer to reach the feed solution side [43,84]. With a lower structure parameter, the internal concentration polarization is reduced. On the other hand, external concentration polarization can be either concentrative or dilutive. As water flows from the feed to the draw side, an increase in the concentration at the support layer surface occurs (CF,b to CF,m), i.e., concentrative external concentration polarization, and a decrease in the concentration at the active layer surface occurs (CD,b to CD,m), i.e., dilutive external concentration polarization. External concentration polarization can be controlled by introducing turbulence near the membrane surface through membrane spacers. The combined effect of both the internal and external concentration polarization results in the reduction of the effective driving force from the difference in bulk concentrations (CD,b–CF,b) to the difference in concentrations across the membrane active layer (CD,m–Cicp).

5.2. Pressure Drop

Membrane spacers play an important role in modules, as they introduce turbulence at the membrane surface, which decreases the external concentration polarization [85,86,87]. However, these spacers can have a negative effect on the performance, as they cause a significant increase in the pressure drop [88,89,90]. The pressure drop has a significant effect on the water flux, as it affects the hydraulic pressure difference across the membrane [91,92,93,94,95,96,97]. Therefore, the effect of the pressure drop along the flow channels should be considered in water flux calculations when modeling the PRO process. The pressure drop along the flow channels was measured experimentally for several spacer geometries at different velocities [86,98,99]. The measurements were conducted using a rectangular channel filled with a spacer or using an RO membrane module. Therefore, the effect of the spacer elastic deformation due to high pressure and a membrane indentation depth was not considered in these measurements. The pressure drop measurements were then used to develop empirical correlations for the friction factor as a function of the Reynolds number. The correlations were used by several studies to estimate the pressure drop in actual membrane modules with both small and large membrane areas to assess the PRO performance [100,101,102,103]. It is important to emphasize that these empirical correlations can only be used to estimate the pressure drop across a channel with the same spacer used in the experimental measurement and without the effect of elastic deformation of the spacer and membrane indentation depth under high operating pressure. Even though there are numerous correlations for spacers with different characteristics [104,105], a general correlation is needed that can be used with currently available PRO spacers.
An alternative approach for estimating the pressure drop across a channel is using computational fluid dynamics (CFD) in either a simplified 2D model or a detailed 3D model [103,106,107,108,109,110]. The simplified 2D model was utilized by several studies to investigate the concentration, pressure drop, and flow distribution [111,112,113,114]. Spacers with a triangular filament cross-section geometry were reported to be more efficient at reducing the concentration polarization when compared with spacers with different geometries [115]. Although 2D models can allow for the quick estimation of the pressure drop for different spacers, a drawback to 2D models is their lack of ability to accurately estimate the pressure drop for spacers with complex geometrical features. On the other hand, 3D models allow for an accurate estimation of the pressure drop, as complex geometrical features, such as the filament intersection, angle of attack, and mesh angle, are taken into consideration [103,109,116,117,118,119,120]. Three-dimensional CFD simulations were conducted by drawing cylindrical filaments in a crossing arrangement or using captured microscopic images of the spacer as the geometry of the computational domain. This latter approach was used by Picioreanu et al. [118], who reported the pressure drop after considering the variation in filament diameters at the intersections to be twice that of the geometry based on a crossing arrangement of cylindrical filaments. However, it is worth mentioning that 3D simulations can be time-consuming and may require high computational power. Moreover, the geometries of commercially available spacers tend to be more complex than spacers with a crossing arrangement of filaments. This could be explained by the change in diameter at the filament intersection [118,121]. An alternative approach for pressure drop estimation across a channel is by treating spacer-filled channels as porous media and using Darcy–Forchheimer equations and CFD simulations [90]. This approach allows for the incorporation of the complex geometrical features of spacers without the need for high computational power. However, the main drawback to this approach is the necessity to estimate both the permeability and inertia coefficients.

5.3. Draw Pressure Effect on Feed Channel Hydrodynamics

High draw pressure affects feed channel hydrodynamics, as it reduces the feed channel gap, thus increasing the pressure drop significantly, which reduces the PRO performance. Karabelas et al. [122] investigated the effect of the draw pressure on the feed channel gap and reported that it is influenced by the indentation depth and reduction in spacer thickness, as shown in Figure 9. The indentation depth can be defined as the depth of the membrane that is deformed due to the applied pressure. Moreover, the reduction in spacer thickness is a result of the elastic deformation of the spacer due to the applied pressure. These effects caused a reduction of around 16% in the effective feed channel gap as draw pressure increased from 1 to 3 bar. Another approach to study the effect of draw hydraulic pressure on the feed channel hydrodynamics was used by Jeon et al. [123], where they changed the draw pressure and measured the feed pressure drop experimentally. Using these results, they developed a correlation to estimate the feed pressure drop using draw solution pressure. However, the draw pressure used by both studies was lower (by up to 3 bars) when compared with the usual operating draw pressure used in PRO [122,123]. She et al. [27] investigated the effect of the feed spacer geometry on the PRO performance. It was reported that feed spacers with large openings resulted in severe membrane deformation, increasing the pressure drop in the feed channel and reducing the power density significantly. Thus, the effect of the draw solution pressure on the effective feed channel gap, which significantly increases the feed pressure drop, should be considered when optimizing the PRO performance, as it is noticeable in large-scale modules.

5.4. Temperature Variation Effect

The feed and draw solution temperatures can greatly influence the PRO performance. Water temperatures can fluctuate with changes in seasonal temperature and geographical location, affecting the temperature of the water used as the draw or feed in a PRO power plant [124,125]. A solution’s temperature can influence its properties, such as the density, diffusion coefficient, viscosity, and osmotic pressure. The solution temperature also affects a membrane’s geometrical characteristics, such as its thickness, porosity, and pore size. Moreover, membrane properties, such as water and salt permeabilities, were reported to increase with increasing temperature [125,126]. However, some studies reported a decrease in salt permeability with increasing temperature [127]. The membrane structure parameter, on the other hand, was reported to decrease with increasing temperature by some studies [125], while others reported no change in the structure parameter with temperature [126]. Thus, due to the disagreement in the literature on the effect of temperature on membrane parameters, further studies are required to properly understand the effect of temperature on membrane parameters. Increases in water temperature were shown to enhance the power density in several studies [125,126]. However, the increase in power density with increasing water temperature was reported to be more profound at low draw solution concentrations [25]. Therefore, due to the noticeable effect of temperature on the PRO performance, changes in solution temperatures should be considered when designing PRO systems or when assessing their performance.

6. Conclusions and Recommendations

Despite being recognized as a promising technique to generate power, the commercialization of PRO is currently facing many obstacles. One major challenge is the suboptimal performance of the commercially available membranes when used in PRO systems, as some of these membranes were developed for other applications. Membrane selection in PRO is based on the concentration difference used, which determines the required operating pressure difference. With an increasing concentration difference, the optimal hydraulic pressure difference increases, requiring a membrane that can withstand high pressures. These membranes usually include a thick support layer, which can negatively affect the PRO performance by increasing the effect of the internal concentration polarization. Conversely, at a lower concentration difference, an optimal membrane would possess a thin support layer to reduce the effect of the internal concentration polarization. Therefore, a broad selection of commercially available membranes that are suitable for the wide range of concentration differences can substantially aid PRO commercialization.
Other important challenges to commercialization are the inadequacy of the available feed spacer and module designs for use in PRO systems. Due to the high pressure in the draw channel, feed spacers are compressed, decreasing the effective feed channel height, which significantly increases the pressure drop on the feed side. This phenomenon has not been thoroughly investigated, as it is specific to the PRO process while being absent in RO systems and negligible in FO due to the absence of the large hydraulic pressure difference. Although many 2D CFD simulations were conducted to model the performance of spacers, they could not accurately estimate the pressure drop for spacers with complex geometrical features. On the other hand, 3D models can be time-consuming and require high computational power. Thus, it is necessary to develop appropriate feed spacers that are able to withstand high draw pressures without significantly decreasing the effective feed channel gap. Similarly, the development of a suitable module design is required to facilitate PRO commercialization, as the commercially available spiral wound module is not adequate for PRO systems. This can be explained by the unfavorable flow pattern on the feed side, which results in a high pressure drop and the presence of areas of low velocity.
Although PRO commercialization is a remote prospect, notable advancements in membrane development are present. This can be seen in the abundance of lab-synthesized membranes developed for different concentration differences, which are yet to be commercialized. On the other hand, there is a lack of research efforts aimed at improving spacers and module designs for use in PRO. Therefore, future research should focus on developing appropriate spacers and modules to achieve successful PRO commercialization.

Funding

This research was funded by the Natural Sciences and Engineering Research Council of Canada (NSERC) grant #401366.

Data Availability Statement

The data presented in this study are available in (1) Abdelkader, B.A, 2022, Experimental and Numerical Investigations of Salinity Gradient Energy Harvesting by Pressure Retarded Osmosis, PhD thesis, University of Guelph, Guelph. https://hdl.handle.net/10214/27156 (accessed on 28 August 2022) and (2) Abdelkader, B.A.; Sharqawy, M.H. Temperature effects on salinity gradient energy harvesting and utilized membrane properties—Experimental and numerical investigation. Sustain. Energy Technol. Assess. 2021, 48, 101666. https://doi.org/10.1016/j.seta.2021.101666.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

AMembrane water permeabilitym/s·Pa
BMembrane salt permeabilitym/s
CConcentration g/kg
DDiffusion coefficient m2/s
JFlow flux m/s
KSolute resistivity s/m
kMass transfer coefficient m/s
PHydraulic pressurePa
RUniversal gas constant J/mol·K
SStructure parameterm
tMembrane thicknessm
WPower densityW/m2
Greek symbols
βvan’t Hoff coefficient-
εPorosity -
πOsmotic pressurePa
τTortuosity-
Subscripts
bBulk
DDraw solution
FFeed solution
icpActive–support layer interface
mMembrane surface
sSalt or solute
wWater

References

  1. Pattle, R.E. Production of Electric Power by mixing Fresh and Salt Water in the Hydroelectric Pile. Nature 1954, 174, 660. [Google Scholar] [CrossRef]
  2. Lee, C.; Chae, S.H.; Yang, E.; Kim, S.; Kim, J.H.; Kim, I.S. A comprehensive review of the feasibility of pressure retarded osmosis: Recent technological advances and industrial efforts towards commercialization. Desalination 2020, 491, 114501. [Google Scholar] [CrossRef]
  3. Jia, Z.; Wang, B.; Song, S.; Fan, Y. Blue energy: Current technologies for sustainable power generation from water salinity gradient. Renew. Sustain. Energy Rev. 2014, 31, 91–100. [Google Scholar] [CrossRef]
  4. Chung, T.-S. Membrane Technology for Osmotic Power Generation by Pressure Retarded Osmosis, 1st ed.; CRC Press: New York, NY, USA, 2020. [Google Scholar] [CrossRef]
  5. Hatzell, M.C.; Cusick, R.D.; Logan, B.E. Capacitive mixing power production from salinity gradient energy enhanced through exoelectrogen-generated ionic currents. Energy Environ. Sci. 2014, 7, 1159–1165. [Google Scholar] [CrossRef]
  6. Jones, A.; Finley, W. Recent development in salinity gradient power. In Proceedings of the Oceans 2003: Celebrating the Past—Teaming Toward the Future, San Diego, CA, USA, 22–26 September 2003; Volume 4, pp. 2284–2287. [Google Scholar] [CrossRef]
  7. Yip, N.Y.; Brogioli, D.; Hamelers, H.V.M.; Nijmeijer, K. Salinity Gradients for Sustainable Energy: Primer, Progress, and Prospects. Environ. Sci. Technol. 2016, 50, 12072–12094. [Google Scholar] [CrossRef] [PubMed]
  8. Sharma, M.; Das, P.P.; Chakraborty, A.; Purkait, M.K. Clean energy from salinity gradients using pressure retarded osmosis and reverse electrodialysis: A review. Sustain. Energy Technol. Assess. 2021, 49, 101687. [Google Scholar] [CrossRef]
  9. Yip, N.Y.; Elimelech, M. Comparison of Energy Efficiency and Power Density in Pressure Retarded Osmosis and Reverse Electrodialysis. Environ. Sci. Technol. 2014, 48, 11002–11012. [Google Scholar] [CrossRef]
  10. Helfer, F.; Lemckert, C.; Anissimov, Y.G. Osmotic power with Pressure Retarded Osmosis: Theory, performance and trends—A review. J. Membr. Sci. 2014, 453, 337–358. [Google Scholar] [CrossRef]
  11. Cath, T.Y.; Childress, A.E.; Elimelech, M. Forward osmosis: Principles, applications, and recent developments. J. Membr. Sci. 2006, 281, 70–87. [Google Scholar] [CrossRef]
  12. Sarp, S.; Li, Z.; Saththasivam, J. Pressure Retarded Osmosis (PRO): Past experiences, current developments, and future prospects. Desalination 2016, 389, 2–14. [Google Scholar] [CrossRef]
  13. Chung, T.-S.; Luo, L.; Wan, C.F.; Cui, Y.; Amy, G. What is next for forward osmosis (FO) and pressure retarded osmosis (PRO). Sep. Purif. Technol. 2015, 156, 856–860. [Google Scholar] [CrossRef] [Green Version]
  14. Bogler, A.; Lin, S.; Bar-Zeev, E. Biofouling of membrane distillation, forward osmosis and pressure retarded osmosis: Principles, impacts and future directions. J. Membr. Sci. 2017, 542, 378–398. [Google Scholar] [CrossRef]
  15. Altaee, A.; Sharif, A. Pressure retarded osmosis: Advancement in the process applications for power generation and desalination. Desalination 2015, 356, 31–46. [Google Scholar] [CrossRef]
  16. Abdelkader, B.A.; Sharqawy, M.H. Ecological Potential of Osmotic Power Generation by Pressure Retarded Osmosis in Ontario, Canada. In Proceedings of the 4th International Conference on Energy Harvesting, Storage, and Transfer (EHST’20), Ottawa, ON, Canada, 18–19 June 2019. [Google Scholar] [CrossRef]
  17. Zadeh, A.E.; Touati, K.; Mulligan, C.N.; McCutcheon, J.R.; Rahaman, S. Closed-loop pressure retarded osmosis draw solutions and their regeneration processes: A review. Renew. Sustain. Energy Rev. 2022, 159, 112191. [Google Scholar] [CrossRef]
  18. Yip, N.Y.; Elimelech, M. Thermodynamic and Energy Efficiency Analysis of Power Generation from Natural Salinity Gradients by Pressure Retarded Osmosis. Environ. Sci. Technol. 2012, 46, 5230–5239. [Google Scholar] [CrossRef]
  19. Loeb, S. Method and Apparatus for Generating Power Utilizing Pressure-Retarded-Osmosis. U.S. Patent US3906250A, 16 September 1975. [Google Scholar]
  20. Loeb, S. Large-scale power production by pressure-retarded osmosis, using river water and sea water passing through spiral modules. Desalination 2002, 143, 115–122. [Google Scholar] [CrossRef]
  21. Touati, K.; Tadeo, F.; Chae, S.H.; Kim, J.H.; Alvarez-Silva, O. Pressure Retarded Osmosis: Renewable Energy Generation and Recovery, 1st ed.; Academic Press: Cambridge, MA, USA, 2017. [Google Scholar]
  22. Kim, J.; Jeong, K.; Park, M.J.; Shon, H.K.; Kim, J.H. Recent Advances in Osmotic Energy Generation via Pressure-Retarded Osmosis (PRO): A Review. Energies 2015, 8, 11821–11845. [Google Scholar] [CrossRef]
  23. Sharqawy, M.H.; Zubair, S.M.; Lienhard, J.H. Energy utilization from disposed brine of desalination plants. In Proceedings of the International Mechanical Engineering Congress & Exposition, Houston, TX, USA, 9–15 November 2012; pp. 1–9. [Google Scholar]
  24. Bajraktari, N.; Hélix-Nielsen, C.; Madsen, H.T. Pressure retarded osmosis from hypersaline sources—A review. Desalination 2017, 413, 65–85. [Google Scholar] [CrossRef]
  25. Kim, Y.C.; Elimelech, M. Potential of osmotic power generation by pressure retarded osmosis using seawater as feed solution: Analysis and experiments. J. Membr. Sci. 2013, 429, 330–337. [Google Scholar] [CrossRef]
  26. Kim, Y.C.; Elimelech, M. Adverse Impact of Feed Channel Spacers on the Performance of Pressure Retarded Osmosis. Environ. Sci. Technol. 2012, 46, 4673–4681. [Google Scholar] [CrossRef]
  27. She, Q.; Hou, D.; Liu, J.; Tan, K.H.; Tang, C.Y. Effect of feed spacer induced membrane deformation on the performance of pressure retarded osmosis (PRO): Implications for PRO process operation. J. Membr. Sci. 2013, 445, 170–182. [Google Scholar] [CrossRef]
  28. She, Q.; Jin, X.; Tang, C.Y. Osmotic power production from salinity gradient resource by pressure retarded osmosis: Effects of operating conditions and reverse solute diffusion. J. Membr. Sci. 2012, 401–402, 262–273. [Google Scholar] [CrossRef]
  29. Kim, Y.C.; Kim, Y.; Oh, D.; Lee, K.H. Experimental Investigation of a Spiral-Wound Pressure-Retarded Osmosis Membrane Module for Osmotic Power Generation. Environ. Sci. Technol. 2013, 47, 2966–2973. [Google Scholar] [CrossRef] [PubMed]
  30. Nielsen, W.K. Progress in the Development of Osmotic Power. In Proceedings of the Quingdao International Conference on Desalination and Water Reuse, Qingdao, China, 20–23 June 2011. [Google Scholar]
  31. Kim, J.; Lee, J.; Kim, J.H. Overview of pressure-retarded osmosis (PRO) process and hybrid application to sea water reverse osmosis process. Desalin. Water Treat. 2012, 43, 193–200. [Google Scholar] [CrossRef]
  32. Fang, L.-F.; Cheng, L.; Jeon, S.; Wang, S.-Y.; Takahashi, T.; Matsuyama, H. Effect of the supporting layer structures on antifouling properties of forward osmosis membranes in AL-DS mode. J. Membr. Sci. 2018, 552, 265–273. [Google Scholar] [CrossRef]
  33. Yang, T.; Wan, C.F.; Xiong, J.Y.; Chung, T.-S. Pre-treatment of wastewater retentate to mitigate fouling on the pressure retarded osmosis (PRO) process. Sep. Purif. Technol. 2019, 215, 390–397. [Google Scholar] [CrossRef]
  34. Shi, Y.; Zhang, M.; Zhang, H.; Yang, F.; Tang, C.Y.; Dong, Y. Recent development of pressure retarded osmosis membranes for water and energy sustainability: A critical review. Water Res. 2020, 189, 116666. [Google Scholar] [CrossRef]
  35. Einarsson, S.J.; Wu, B. Thermal associated pressure-retarded osmosis processes for energy production: A review. Sci. Total Environ. 2020, 757, 143731. [Google Scholar] [CrossRef]
  36. NASA Earth Observatory. Sea Surface Temperatures. 2021. Available online: https://earthobservatory.nasa.gov/global-maps/MYD28M (accessed on 16 September 2022).
  37. Lee, K.; Baker, R.; Lonsdale, H. Membranes for power generation by pressure-retarded osmosis. J. Membr. Sci. 1981, 8, 141–171. [Google Scholar] [CrossRef]
  38. Achilli, A.; Cath, T.Y.; Childress, A.E. Power generation with pressure retarded osmosis: An experimental and theoretical investigation. J. Membr. Sci. 2009, 343, 42–52. [Google Scholar] [CrossRef]
  39. Elimelech, M.; Bhattacharjee, S. A novel approach for modeling concentration polarization in crossflow membrane filtration based on the equivalence of osmotic pressure model and filtration theory. J. Membr. Sci. 1998, 145, 223–241. [Google Scholar] [CrossRef]
  40. 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. Membr. Sci. 1987, 30, 153–170. [Google Scholar] [CrossRef]
  41. Yip, N.Y.; Elimelech, M. Performance Limiting Effects in Power Generation from Salinity Gradients by Pressure Retarded Osmosis. Environ. Sci. Technol. 2011, 45, 10273–10282. [Google Scholar] [CrossRef] [PubMed]
  42. Touati, K.; Hänel, C.; Tadeo, F.; Schiestel, T. Effect of the feed and draw solution temperatures on PRO performance: Theoretical and experimental study. Desalination 2015, 365, 182–195. [Google Scholar] [CrossRef]
  43. Abdelkader, B.A.; Sharqawy, M.H. Temperature effects on salinity gradient energy harvesting and utilized membrane properties—Experimental and numerical investigation. Sustain. Energy Technol. Assess. 2021, 48, 101666. [Google Scholar] [CrossRef]
  44. Zhang, S.; Chung, T.-S. Minimizing the Instant and Accumulative Effects of Salt Permeability to Sustain Ultrahigh Osmotic Power Density. Environ. Sci. Technol. 2013, 47, 10085–10092. [Google Scholar] [CrossRef]
  45. Han, G.; Wang, P.; Chung, T.-S. Highly Robust Thin-Film Composite Pressure Retarded Osmosis (PRO) Hollow Fiber Membranes with High Power Densities for Renewable Salinity-Gradient Energy Generation. Environ. Sci. Technol. 2013, 47, 8070–8077. [Google Scholar] [CrossRef]
  46. Sun, S.-P.; Chung, T.-S. Outer-Selective Pressure-Retarded Osmosis Hollow Fiber Membranes from Vacuum-Assisted Interfacial Polymerization for Osmotic Power Generation. Environ. Sci. Technol. 2013, 47, 13167–13174. [Google Scholar] [CrossRef]
  47. Gonzales, R.R.; Park, M.J.; Bae, T.-H.; Yang, Y.; Abdel-Wahab, A.; Phuntsho, S.; Shon, H.K. Melamine-based covalent organic framework-incorporated thin film nanocomposite membrane for enhanced osmotic power generation. Desalination 2019, 459, 10–19. [Google Scholar] [CrossRef]
  48. Cui, Z.; Jiang, Y.; Field, R. Fundamentals of Pressure-Driven Membrane Separation Processes. Membr. Technol. 2010, 1–18. [Google Scholar] [CrossRef]
  49. Naguib, M.F.; Maisonneuve, J.; Laflamme, C.B.; Pillay, P. Modeling pressure-retarded osmotic power in commercial length membranes. Renew. Energy 2015, 76, 619–627. [Google Scholar] [CrossRef]
  50. Lee, S. Performance Comparison of Spiral-Wound and Plate-and-Frame Forward Osmosis Membrane Module. Membranes 2020, 10, 318. [Google Scholar] [CrossRef] [PubMed]
  51. Chung, H.W.; Swaminathan, J.; Banchik, L.D.; Lienhard, J.H. Economic framework for net power density and levelized cost of electricity in pressure-retarded osmosis. Desalination 2018, 448, 13–20. [Google Scholar] [CrossRef] [Green Version]
  52. Chung, H.W.; Banchik, L.D.; Swaminathan, J. On the present and future economic viability of stand-alone pressure-retarded osmosis. Desalination 2017, 408, 133–144. [Google Scholar] [CrossRef] [Green Version]
  53. Straub, A.P.; Yip, N.Y.; Elimelech, M. Raising the Bar: Increased Hydraulic Pressure Allows Unprecedented High Power Densities in Pressure-Retarded Osmosis. Environ. Sci. Technol. Lett. 2013, 1, 55–59. [Google Scholar] [CrossRef]
  54. Hickenbottom, K.L.; Vanneste, J.; Elimelech, M.; Cath, T.Y. Assessing the current state of commercially available membranes and spacers for energy production with pressure retarded osmosis. Desalination 2016, 389, 108–118. [Google Scholar] [CrossRef] [Green Version]
  55. Plata, S.L.; Childress, A.E. Limiting power density in pressure-retarded osmosis: Observation and implications. Desalination 2019, 467, 51–56. [Google Scholar] [CrossRef]
  56. Madsen, H.T.; Nissen, S.S.; Muff, J.; Søgaard, E.G. Pressure retarded osmosis from hypersaline solutions: Investigating commercial FO membranes at high pressures. Desalination 2017, 420, 183–190. [Google Scholar] [CrossRef]
  57. Maisonneuve, J.; Laflamme, C.B.; Pillay, P. Experimental investigation of pressure retarded osmosis for renewable energy conversion: Towards increased net power. Appl. Energy 2015, 164, 425–435. [Google Scholar] [CrossRef]
  58. Thorsen, T.; Holt, T. The potential for power production from salinity gradients by pressure retarded osmosis. J. Membr. Sci. 2009, 335, 103–110. [Google Scholar] [CrossRef]
  59. Manzoor, H.; Selam, M.A.; Rahman, F.B.A.; Adham, S.; Castier, M.; Abdel-Wahab, A. A tool for assessing the scalability of pressure-retarded osmosis (PRO) membranes. Renew. Energy 2019, 149, 987–999. [Google Scholar] [CrossRef]
  60. Abdelkader, B.; Sharqawy, M.H. Temperature Effects and Entropy Generation of Pressure Retarded Osmosis Process. Entropy 2019, 21, 1158. [Google Scholar] [CrossRef] [Green Version]
  61. Foreman, G.E.; Worsley, P.K. Spiral Wound Membrane Module for Direct Osmosis Separations. U.S. Patent US4033878A, 5 July 1977. [Google Scholar]
  62. Attarde, D.; Jain, M.; Chaudhary, K.; Gupta, S.K. Osmotically driven membrane processes by using a spiral wound module—Modeling, experimentation and numerical parameter estimation. Desalination 2015, 361, 81–94. [Google Scholar] [CrossRef]
  63. Achilli, A.; Prante, J.L.; Hancock, N.T.; Maxwell, E.B.; Childress, A.E. Experimental Results from RO-PRO: A Next Generation System for Low-Energy Desalination. Environ. Sci. Technol. 2014, 48, 6437–6443. [Google Scholar] [CrossRef] [PubMed]
  64. Kakihana, Y.; Jullok, N.; Shibuya, M.; Ikebe, Y.; Higa, M. Comparison of Pressure-Retarded Osmosis Performance between Pilot-Scale Cellulose Triacetate Hollow-Fiber and Polyamide Spiral-Wound Membrane Modules. Membranes 2021, 11, 177. [Google Scholar] [CrossRef]
  65. Xu, Y.; Peng, X.; Tang, C.Y.; Fu, Q.S.; Nie, S. Effect of draw solution concentration and operating conditions on forward osmosis and pressure retarded osmosis performance in a spiral wound module. J. Membr. Sci. 2010, 348, 298–309. [Google Scholar] [CrossRef]
  66. Matta, S.M.; Selam, M.A.; Manzoor, H.; Adham, S.; Shon, H.K.; Castier, M.; Abdel-Wahab, A. Predicting the performance of spiral-wound membranes in pressure-retarded osmosis processes. Renew. Energy 2022, 189, 66–77. [Google Scholar] [CrossRef]
  67. Altaee, A.; Cipolina, A. Modelling and optimization of modular system for power generation from a salinity gradient. Renew. Energy 2019, 141, 139–147. [Google Scholar] [CrossRef]
  68. Binger, Z.M.; Achilli, A. Forward osmosis and pressure retarded osmosis process modeling for integration with seawater reverse osmosis desalination. Desalination 2020, 491, 114583. [Google Scholar] [CrossRef]
  69. Hong, S.-S.; Ryoo, W.; Chun, M.-S.; Lee, S.O.; Chung, G.-Y. Numerical studies on the pressure-retarded osmosis (PRO) system with the spiral wound module for power generation. Desalin. Water Treat. 2013, 52, 6333–6341. [Google Scholar] [CrossRef]
  70. Skilhagen, S.E. Osmotic power—A new, renewable energy source. Desalin. Water Treat. 2010, 15, 271–278. [Google Scholar] [CrossRef]
  71. Gerstandt, K.; Peinemann, K.-V.; Skilhagen, S.E.; Thorsen, T.; Holt, T. Membrane processes in energy supply for an osmotic power plant. Desalination 2008, 224, 64–70. [Google Scholar] [CrossRef]
  72. Alsvik, I.L.; Hägg, M.-B. Pressure Retarded Osmosis and Forward Osmosis Membranes: Materials and Methods. Polymers 2013, 5, 303–327. [Google Scholar] [CrossRef] [Green Version]
  73. Reddy, D.; Moon, T.Y.; Reineke, C.E. Counter Current Dual-Flow Spiral Wound Dual-Pipe Membrane Separation. U.S. Patent US5034126A, 23 July 1991. [Google Scholar]
  74. Mcginnis, R.L. Spiral Wound Membrane Module for Forward Osmotic Use. U.S. Patent US20110036774A1, 17 February 2011. [Google Scholar]
  75. Uda, Y.; Kobuke, M. Spiral-Wound Forward Osmosis Membrane Element and Forward Osmosis Membrane Module. U.S. Patent US9861938B2, 9 January 2018. [Google Scholar]
  76. Herron, J.R. Forward Osmosis and Pressure Retarded Osmosis Spacer. WO2014107464A1, 10 July 2014. [Google Scholar]
  77. Sun, G.; Yang, Z. Spiral Wound Membrane Rolls and Modules. WO2018153978A1, 30 August 2018. [Google Scholar]
  78. Lee, S.Y.; Sim, Y.J.; Lee, J.H. Porous Outflow Pipe for Forward Osmosis or Pressure-Retarded Osmosis, and Forward Osmosis or Pressure-Retarded Osmosis Module Comprising Same. U.S. Patent US 11,020,705 B2, 1 June 2021. [Google Scholar]
  79. Benton, C.; Bakajin, O. Separation Systems, Elements, and Methods for Separation Utilizing Stacked Membranes and Spacers. U.S. Patent US 2020/0094193 A1, 17 August 2020. [Google Scholar]
  80. Revanur, R.; Roh, I.; Klare, J.E.; Noy, A.; Bakajin, O. Thin Film Composite Membranes for Forward Osmosis, and Their Preparation Methods. U.S. Patent US20120080378A1, 5 April 2012. [Google Scholar]
  81. Tang, C.; She, Q.; Ma, N.; Wei, J.; Sim, S.T.V.; Fane, A.G. Reinforced Membranes for Producing Osmotic Power in Pressure Retarded Osmosis. U.S. Patent US20150217238A1, 6 August 2015. [Google Scholar]
  82. Tawalbeh, M.; Al-Othman, A.; Abdelwahab, N.; Alami, A.H.; Olabi, A.G. Recent developments in pressure retarded osmosis for desalination and power generation. Renew. Sustain. Energy Rev. 2020, 138, 110492. [Google Scholar] [CrossRef]
  83. AlZainati, N.; Saleem, H.; Altaee, A.; Zaidi, S.J.; Mohsen, M.; Hawari, A.; Millar, G.J. Pressure retarded osmosis: Advancement, challenges and potential. J. Water Process Eng. 2021, 40, 101950. [Google Scholar] [CrossRef]
  84. Cath, T.Y.; Elimelech, M.; Mccutcheon, J.R.; Mcginnis, R.L.; Achilli, A. Standard methodology for evaluating membrane performance in osmotically driven membrane processes. Desalination 2013, 312, 31–38. [Google Scholar] [CrossRef] [Green Version]
  85. Zimmerer, C.; Kottke, V. Effects of spacer geometry on pressure drop, mass transfer, mixing behavior, and residence time distribution. Desalination 1996, 104, 129–134. [Google Scholar] [CrossRef]
  86. Schock, G.; Miquel, A. Mass transfer and pressure loss in spiral wound modules. Desalination 1987, 64, 339–352. [Google Scholar] [CrossRef]
  87. 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, 33, 1845–1851. [Google Scholar] [CrossRef]
  88. Haidari, A.; Heijman, S.; van der Meer, W. Visualization of hydraulic conditions inside the feed channel of Reverse Osmosis: A practical comparison of velocity between empty and spacer-filled channel. Water Res. 2016, 106, 232–241. [Google Scholar] [CrossRef] [Green Version]
  89. Kavianipour, O.; Ingram, G.D.; Vuthaluru, H.B. Investigation into the effectiveness of feed spacer configurations for reverse osmosis membrane modules using Computational Fluid Dynamics. J. Membr. Sci. 2017, 526, 156–171. [Google Scholar] [CrossRef] [Green Version]
  90. Abdelkader, B.A.; Sharqawy, M.H. Pressure drop across membrane spacer-filled channels using porous media characteristics and computational fluid dynamics simulation. Desalin. Water Treat. 2022, 247, 28–39. [Google Scholar] [CrossRef]
  91. Ruiz-García, A.; De la Nuez, I. Feed Spacer Geometries and Permeability Coefficients. Effect on the Performance in BWRO Spriral-Wound Membrane Modules. Water 2019, 11, 152. [Google Scholar] [CrossRef] [Green Version]
  92. Haidari, A.H.; Heijman, S.G.J.; van der Meer, W.G.J. Optimal design of spacers in reverse osmosis. Sep. Purif. Technol. 2018, 192, 441–456. [Google Scholar] [CrossRef]
  93. Da Costa, A.; Fane, A.; Fell, C.; Franken, A. Optimal channel spacer design for ultrafiltration. J. Membr. Sci. 1991, 62, 275–291. [Google Scholar] [CrossRef]
  94. Thomas, D.G. Forced Convection Mass Transfer in Hyperfiltration at High Fluxes. Ind. Eng. Chem. Fundam. 1973, 12, 396–405. [Google Scholar] [CrossRef]
  95. Schwinge, J.; Wiley, D.E.; Fletcher, D.F. Simulation of the Flow around Spacer Filaments between Channel Walls. 2. Mass-Transfer Enhancement. Ind. Eng. Chem. Res. 2002, 41, 4879–4888. [Google Scholar] [CrossRef]
  96. Kerdi, S.; Qamar, A.; Vrouwenvelder, J.S.; Ghaffour, N. Fouling resilient perforated feed spacers for membrane filtration. Water Res. 2018, 140, 211–219. [Google Scholar] [CrossRef]
  97. Schwinge, J.; Neal, P.R.; Wiley, D.E.; Fletcher, D.F.; Fane, A.G. Spiral wound modules and spacers: Review and analysis. J. Membr. Sci. 2004, 242, 129–153. [Google Scholar] [CrossRef]
  98. Fárková, J. The pressure drop in membrane module with spacers. J. Membr. Sci. 1991, 64, 103–111. [Google Scholar] [CrossRef]
  99. Da Costa, A.R.; Fane, A.G.; Wiley, D.E. Spacer characterization and pressure drop modeling in spacer-filled channels.pdf. J. Membr. Sci. 1994, 87, 79–98. [Google Scholar] [CrossRef]
  100. Sagiv, A.; Xu, W.; Christofides, P.D.; Cohen, Y.; Semiat, R. Evaluation of osmotic energy extraction via FEM modeling and exploration of PRO operational parameter space. Desalination 2016, 401, 120–133. [Google Scholar] [CrossRef]
  101. Maisonneuve, J.; Pillay, P.; Laflamme, C.B. Pressure-retarded osmotic power system model considering non-ideal effects. Renew. Energy 2015, 75, 416–424. [Google Scholar] [CrossRef]
  102. Roy, Y.; Sharqawy, M.H.; Lienhard, J.H. Modeling of flat-sheet and spiral-wound nanofiltration configurations and its application in seawater nanofiltration. J. Membr. Sci. 2015, 493, 360–372. [Google Scholar] [CrossRef]
  103. Koutsou, C.; Yiantsios, S.; Karabelas, A. A numerical and experimental study of mass transfer in spacer-filled channels: Effects of spacer geometrical characteristics and Schmidt number. J. Membr. Sci. 2009, 326, 234–251. [Google Scholar] [CrossRef]
  104. Geraldes, V. Flow management in nanofiltration spiral wound modules with ladder-type spacers. J. Membr. Sci. 2002, 203, 87–102. [Google Scholar] [CrossRef]
  105. 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. Membr. Sci. 2017, 527, 78–91. [Google Scholar] [CrossRef]
  106. Santos, J.; Geraldes, V.; Velizarov, S.; Crespo, J. Investigation of flow patterns and mass transfer in membrane module channels filled with flow-aligned spacers using computational fluid dynamics (CFD). J. Membr. Sci. 2007, 305, 103–117. [Google Scholar] [CrossRef]
  107. Fimbres-Weihs, G.; Wiley, D. Review of 3D CFD modeling of flow and mass transfer in narrow spacer-filled channels in membrane modules. Chem. Eng. Process. Process Intensif. 2010, 49, 759–781. [Google Scholar] [CrossRef]
  108. Koutsou, C.P.; Karabelas, A.J. Shear stresses and mass transfer at the base of a stirred filtration cell and corresponding conditions in narrow channels with spacers. J. Membr. Sci. 2012, 399–400, 60–72. [Google Scholar] [CrossRef]
  109. Shakaib, M.; Hasani, S.; Mahmood, M. CFD modeling for flow and mass transfer in spacer-obstructed membrane feed channels. J. Membr. Sci. 2009, 326, 270–284. [Google Scholar] [CrossRef]
  110. Toh, K.; Liang, Y.; Lau, W.; Fletcher, D. CFD study of the effect of perforated spacer on pressure loss and mass transfer in spacer-filled membrane channels. Chem. Eng. Sci. 2020, 222, 115704. [Google Scholar] [CrossRef]
  111. Fimbres-Weihs, G.; Wiley, D. Numerical study of two-dimensional multi-layer spacer designs for minimum drag and maximum mass transfer. J. Membr. Sci. 2008, 325, 809–822. [Google Scholar] [CrossRef]
  112. Landslide damage to the Boar River water supply pipeline, Bromley Hill, Jamaica: Case study of a landslide caused by Hurricane Gilbert (1988): Ahmad, R.; Earle, A.H.; Hugues, P.; Maharaj, R.; Robinson, E. Int Assoc Engng Geol BullN47, April 1993, P59–70. Int. J. Rock Mech. Min. Sci. Géoméch. Abstr. 1994, 31, A46. [CrossRef]
  113. Wardeh, S.; Morvan, H. CFD simulations of flow and concentration polarization in spacer-filled channels for application to water desalination. Chem. Eng. Res. Des. 2008, 86, 1107–1116. [Google Scholar] [CrossRef]
  114. Schwinge, J.; Wiley, D.; Fletcher, D. A CFD study of unsteady flow in narrow spacer-filled channels for spiral-wound membrane modules. Desalination 2002, 146, 195–201. [Google Scholar] [CrossRef]
  115. Ahmad, A.; Lau, K.K.; Abu Bakar, M. Impact of different spacer filament geometries on concentration polarization control in narrow membrane channel. J. Membr. Sci. 2005, 262, 138–152. [Google Scholar] [CrossRef]
  116. Ranade, V.V.; Kumar, A. Fluid dynamics of spacer filled rectangular and curvilinear channels. J. Membr. Sci. 2006, 271, 1–15. [Google Scholar] [CrossRef] [Green Version]
  117. Saeed, A.; Vuthaluru, R.; Yang, Y.; Vuthaluru, H.B. Effect of feed spacer arrangement on flow dynamics through spacer filled membranes. Desalination 2012, 285, 163–169. [Google Scholar] [CrossRef]
  118. Picioreanu, C.; Vrouwenvelder, J.; van Loosdrecht, M. Three-dimensional modeling of biofouling and fluid dynamics in feed spacer channels of membrane devices. J. Membr. Sci. 2009, 345, 340–354. [Google Scholar] [CrossRef]
  119. Li, Y.-L.; Tung, K.-L. CFD simulation of fluid flow through spacer-filled membrane module: Selecting suitable cell types for periodic boundary conditions. Desalination 2008, 233, 351–358. [Google Scholar] [CrossRef]
  120. Li, Y.-L.; Tung, K.-L.; Lu, M.-Y.; Huang, S.-H. Mitigating the curvature effect of the spacer-filled channel in a spiral-wound membrane module. J. Membr. Sci. 2009, 329, 106–118. [Google Scholar] [CrossRef]
  121. Vrouwenvelder, J.S.; Picioreanu, C.; Kruithof, J.C.; Van Loosdrecht, M.C.M. Biofouling in spiral wound membrane systems: Three-dimensional CFD model based evaluation of experimental data. J. Membr. Sci. 2010, 346, 71–85. [Google Scholar] [CrossRef]
  122. Karabelas, A.J.; Koutsou, C.P.; Sioutopoulos, D.C. Comprehensive performance assessment of spacers in spiral-wound membrane modules accounting for compressibility effects. J. Membr. Sci. 2018, 549, 602–615. [Google Scholar] [CrossRef]
  123. Jeon, J.; Jung, J.; Lee, S.; Choi, J.Y.; Kim, S. A simple modeling approach for a forward osmosis system with a spiral wound module. Desalination 2018, 433, 120–131. [Google Scholar] [CrossRef]
  124. Touati, K.; Tadeo, F.; Hänel, C.; Schiestel, T. Effect of the operating temperature on hydrodynamics and membrane parameters in pressure retarded osmosis. Desalin. Water Treat. 2015, 57, 10477–10489. [Google Scholar] [CrossRef]
  125. Wang, Q.; Zhou, Z.; Li, J.; Tang, Q.; Hu, Y. Investigation of the reduced specific energy consumption of the RO-PRO hybrid system based on temperature-enhanced pressure retarded osmosis. J. Membr. Sci. 2019, 581, 439–452. [Google Scholar] [CrossRef]
  126. Sivertsen, E.; Holt, T.; Thelin, W.R. Concentration and Temperature Effects on Water and Salt Permeabilities in Osmosis and Implications in Pressure-Retarded Osmosis. Membranes 2018, 8, 39. [Google Scholar] [CrossRef] [Green Version]
  127. Touati, K.; Tadeo, F.; Schiestel, T. Impact of Temperature on Power Recovery in Osmotic Power Production by Pressure Retarded Osmosis. Energy Procedia 2014, 50, 960–969. [Google Scholar] [CrossRef]
Figure 1. Schematic of a basic PRO process.
Figure 1. Schematic of a basic PRO process.
Energies 15 07325 g001
Figure 2. Variations in water flux with hydraulic pressure difference for FO, RO, and PRO processes, as well as changes in power density for PRO.
Figure 2. Variations in water flux with hydraulic pressure difference for FO, RO, and PRO processes, as well as changes in power density for PRO.
Energies 15 07325 g002
Figure 3. PRO concentration profile for an asymmetric membrane with the active dense layer facing the draw solution.
Figure 3. PRO concentration profile for an asymmetric membrane with the active dense layer facing the draw solution.
Energies 15 07325 g003
Figure 4. Comparison between the water flux generated using different modifications of the solution diffusion model.
Figure 4. Comparison between the water flux generated using different modifications of the solution diffusion model.
Energies 15 07325 g004
Figure 5. A 3D schematic of the (a) plate-and-frame module, (b) hollow fiber module, (c) spiral wound module, and (d) tubular module.
Figure 5. A 3D schematic of the (a) plate-and-frame module, (b) hollow fiber module, (c) spiral wound module, and (d) tubular module.
Energies 15 07325 g005
Figure 6. PRO performance from experimental and numerical investigations of a single flat sheet membrane module [25,26,28,38,42,43,53,54,55,56,57,58,59,60]. The data used in this figure are summarized in Table 2.
Figure 6. PRO performance from experimental and numerical investigations of a single flat sheet membrane module [25,26,28,38,42,43,53,54,55,56,57,58,59,60]. The data used in this figure are summarized in Table 2.
Energies 15 07325 g006
Figure 7. An illustration of the flow path for the commercial SWM.
Figure 7. An illustration of the flow path for the commercial SWM.
Energies 15 07325 g007
Figure 8. PRO performance from experimental and numerical investigations of a spiral wound module [29,62,63,64,65,66,67,68,69]. The data used in this figure are summarized in Table 3.
Figure 8. PRO performance from experimental and numerical investigations of a spiral wound module [29,62,63,64,65,66,67,68,69]. The data used in this figure are summarized in Table 3.
Energies 15 07325 g008
Figure 9. Effect of the draw solution’s high pressure on the feed channel hydrodynamics.
Figure 9. Effect of the draw solution’s high pressure on the feed channel hydrodynamics.
Energies 15 07325 g009
Table 1. Solution diffusion model development.
Table 1. Solution diffusion model development.
ModelModel ExpressionAssumptions
Lee et al. [37] J w = A π D , b 1 C F , b C D , b exp J w K 1 + B J w exp J w K 1 Δ P C F , b = C F , m C D , b = C D , m
Achilli et al. [38] J w = A π D , b exp J w k π F , b exp J w K 1 + B J w exp J w K 1 Δ P C F , b = C F , m C D , b C D , m C F , b C D , m = π F , b π D , m
Yip et al. [41] J w = A π D , b exp J w k π F , b exp J w K 1 + B J w exp J w K exp J w k Δ P C F , b = C F , m C D , b C D , m Δ π m = Δ C m
Touati et al. [42] J w = A π D , b + B A 1 + A Δ P J w exp J w k D π F , b + B A 1 + A Δ P J w exp J w K exp J w k F Δ P C F , b C F , m C D , b C D , m π = β R T C k D k F
Table 2. Performance and key properties for commercially available membranes in a single flat sheet membrane module from both experimental and numerical investigations.
Table 2. Performance and key properties for commercially available membranes in a single flat sheet membrane module from both experimental and numerical investigations.
Membrane MaterialMembrane Area
(m2) × 10−4
Membrane PropertiesOperating ConditionsPower Density
(W/m2)
Ref.
A
(L/m2 hr bar)
B
(L/m2 hr)
S
(µm)
Concentration
(Feed/Draw)
Hydraulic Pressure Difference (bar)
Experimental investigationsCTA18.80.670.40682DI/0.6 M NaCl9.72.7[38]
CTA18.80.670.40682DI/1.026 M NaCl9.75.1[38]
CTA140.00.370.285900.01 M/1 M NaCl15.03.9[28]
CTA140.00.440.0713800.01 M/1 M NaCl14.03.0[28]
CTA140.00.750.014800.01 M/1 M NaCl12.04.5[28]
CTA20.01.232.624090.5 M/1 M NaCl12.50.9[25]
CTA20.01.232.625040.5 M/1.5 M NaCl12.52.8[25]
CTA20.01.232.625050.5 M/2 M NaCl12.54.7[25]
TFC20.02.490.39564DI/0.6 M NaCl13.87.5[53]
TFC20.02.490.39564DI/1 M NaCl20.714.1[53]
TFC20.02.490.39564DI/2 M NaCl41.439.4[53]
TFC20.02.490.39564DI/3 M NaCl48.359.7[53]
TFC139.01.503.74159DI/1 M NaCl10.57.0[54]
TFC139.01.941.99274DI/1 M NaCl10.56.8[54]
TFC139.01.631.42295DI/1 M NaCl21.08.3[54]
TFC139.01.631.42295DI/2 M NaCl35.011.5[54]
TFC139.01.631.42295DI/3 M NaCl41.015.5[54]
CTA139.00.512.19600DI/1 M NaCl19.03.8[54]
TFC120.02.781.36513DI/1.2 M NaCl21.015.3[55]
TFC120.02.290.30198DI/1.2 M NaCl21.012.5[55]
CTA120.00.770.29408DI/1.2 M NaCl21.09.8[55]
CTA120.00.770.29408DI/1.2 M NaCl12.06.9[55]
TFC120.03.351.18398DI/1.2 M NaCl12.07.9[55]
CTA33.70.420.291028DI/3 M NaCl40.010.5[56]
CTA33.70.760.44655DI/3 M NaCl40.030.0[56]
CTA33.70.690.34707DI/3 M NaCl60.032.0[56]
TFC33.71.250.19471DI/3 M NaCl35.027.0[56]
TFC33.7---DI/3 M NaCl10.05.0[56]
TFC87.54.211.43267DI/0.51 M NaCl6.05.0[57]
CTA10.00.130.028000.008 M/0.6 M NaCl13.02.3[42]
CTA10.00.380.0915500.008 M/0.6 M NaCl13.03.0[42]
CTA10.00.380.0915500.008 M/1.026 M NaCl24.05.0[42]
CA0.50.720.222500DI/0.4 M NaCl7.51.6[58]
TFC0.52.560.40670DI/0.48 M NaCl10.52.7[58]
CTA20.01.232.62689DI/0.5 M NaCl12.51.0[26]
CTA20.01.232.62689DI/1 M NaCl12.53.8[26]
CTA20.01.232.62689DI/2 M NaCl12.57.9[26]
CTA140.00.80.41652Tap water/0.17 M NaCl3.60.3[43]
CTA140.00.80.41652Tap water/0.34 M NaCl7.11.2[43]
CTA140.00.80.41652Tap water/0.51 M NaCl10.72.6[43]
Numerical investigationsTFC139.01.631.42295DI/1 M NaCl24.09.0[54]
TFC139.01.631.42295DI/2 M NaCl52.526.0[54]
TFC139.01.631.42295DI/3 M NaCl87.044.0[54]
CTA10.00.130.028000.008 M/0.6 M NaCl13.02.5[42]
CTA10.00.380.0915500.008 M/0.6 M NaCl13.03.5[42]
CTA10.00.380.0915500.008 M/1.026 M NaCl24.05.5[42]
CTA18.80.670.40682DI/0.6 M NaCl13.02.9[38]
CTA18.80.670.40646DI/1.026 M NaCl23.08.0[38]
CTA20.01.232.624090.5 M/1 M NaCl10.01.0[25]
CTA20.01.232.625040.5 M/1.5 M NaCl22.03.2[25]
CTA20.01.232.625050.5 M/2 M NaCl34.06.8[25]
TFC20.02.490.39564DI/0.6 M NaCl13.86.9[53]
TFC20.02.490.39564DI/1 M NaCl24.015.0[53]
TFC20.02.490.39564DI/2 M NaCl55.041.0[53]
TFC20.02.490.39564DI/3 M NaCl95.075.0[53]
TFC20.02.490.39564DI/1 M NaCl23.014.2[59]
TFC20.02.490.39564DI/3 M NaCl51.039.5[59]
TFC20.01.232.62689DI/1 M NaCl18.24.8[59]
TFC20.01.232.62689DI/0.5 M NaCl8.01.6[59]
CTA10.00.380.0915500.008 M/0.6 M NaCl12.02.5[60]
CTA10.00.380.0915500.008 M/1.2 M NaCl30.012.0[60]
Table 3. Performance and key properties for commercially available membranes in a spiral wound module from both experimental and numerical investigations.
Table 3. Performance and key properties for commercially available membranes in a spiral wound module from both experimental and numerical investigations.
Membrane MaterialMembrane Area (m2)Membrane PropertiesOperating ConditionsPower Density
(W/m2)
Ref.
A
(L/m2 hr bar)
B
(L/m2 hr)
S
(µm)
Concentration
(Feed/Draw)
Hydraulic Pressure Difference (bar)
Experimental investigationsTFC29.00.720.286450.001 M/0.6 M NaCl7.70.96[29]
TFC29.00.720.286450.001 M/1.2 M NaCl 15.62.10[29]
TFC29.00.720.286450.001 M/0.52 M NaCl9.80.82[29]
TFC29.00.720.286450.002 M/0.6 M NaCl9.81.00[29]
TFC29.00.720.286450.01 M/0.6 M NaCl9.80.91[29]
TFC29.00.720.286450.05 M/0.6 M NaCl9.80.49[29]
TFC29.00.720.286450.08 M/0.6 M NaCl9.80.08[29]
CTA0.50.630.424960.008 M/0.07 M NaCl1.00.02[62]
CTA0.50.630.424960.008 M/0.1 M NaCl2.30.07[62]
CTA0.50.630.424960.008 M/0.15 M NaCl3.40.15[62]
CTA0.50.630.424960.008 M/0.51 M NaCl4.00.56[62]
CTA0.50.630.424960.008 M/1.026 M NaCl4.01.10[62]
CTA0.9---DI/0.5 M NaCl4.50.40[65]
TFC15.3---Tap water/0.6 M NaCl12.01.40[64]
TFC15.3---Tap water/0.8 M NaCl12.01.64[64]
TFC15.3---Tap water/1.2 M NaCl16.02.33[64]
TFC4.25.100.09310DI/0.48 M NaCl4.51.12 *[63]
TFC4.25.100.09310DI/0.48 M NaCl9.53.51 *[63]
Numerical investigationsCTA0.50.630.424960.008 M/0.07 M NaCl1.50.03[62]
CTA0.50.630.424960.008 M/0.1 M NaCl2.30.07[62]
CTA0.50.630.424960.008 M/0.15 M NaCl3.40.15[62]
CTA0.50.630.424960.008 M/0.51 M NaCl10.20.95[62]
CTA0.50.630.424960.008 M/1.026 M NaCl21.03.25[62]
TFC4.25.100.09310DI/0.48 M NaCl8.05.10[68]
TFC4.25.100.09310DI/0.48 M NaCl10.59.50[68]
TFC4.25.100.09310DI/0.48 M NaCl9.59.00[68]
TFC4.25.100.09310DI/0.48 M NaCl11.011.00[68]
TFC4.25.100.09310DI/0.48 M NaCl10.010.80[68]
CTA30.00.400.307020.6 M/2.74 M NaCl50.09.90[66]
CTA30.00.400.30702DI/0.5 M NaCl9.01.75[66]
CTA8.43.420.31350DI/0.6 M NaCl13.016.00[69]
N/A24.01.232.601670.02 M/1.2 M NaCl27.512.10[67]
N/A24.01.232.601670.02 M/0.6 M NaCl13.83.70[67]
* The power density was calculated using the data given in the cited reference to account for the pressure drop.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Abdelkader, B.A.; Sharqawy, M.H. Challenges Facing Pressure Retarded Osmosis Commercialization: A Short Review. Energies 2022, 15, 7325. https://doi.org/10.3390/en15197325

AMA Style

Abdelkader BA, Sharqawy MH. Challenges Facing Pressure Retarded Osmosis Commercialization: A Short Review. Energies. 2022; 15(19):7325. https://doi.org/10.3390/en15197325

Chicago/Turabian Style

Abdelkader, Bassel A., and Mostafa H. Sharqawy. 2022. "Challenges Facing Pressure Retarded Osmosis Commercialization: A Short Review" Energies 15, no. 19: 7325. https://doi.org/10.3390/en15197325

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