1. Introduction
Solar energy is one of the most abundant and readily available renewable energy sources, with solar radiation being effectively harnessed to power various sustainable systems. Solar thermal energy can be absorbed and converted into electricity via photovoltaic (PV) panels [
1,
2], facilitating the operation of electrical devices. Alternatively, it can be directly utilized for thermal applications through different solar collector technologies. Flat plate collectors (FPCs) [
3] and evacuated tube collectors (ETCs) [
4] are commonly employed for low to medium temperature applications, whereas parabolic trough collectors (PTCs) [
5] are utilized for high temperature applications.
Among the various applications of solar energy, solar desalination is regarded as one of the most efficient and sustainable methods for freshwater production. Solar desalination systems are generally classified into two main categories: direct and indirect desalination [
6,
7,
8]. In direct solar desalination, both the evaporation and condensation processes occur within a single integrated unit, as seen in solar stills [
9,
10] and humidification–dehumidification (HDH) [
11] desalination systems. Conversely, the indirect solar desalination approach separates the solar energy collection unit from the desalination process. In this method, solar energy is either harvested using PV panels to power electrically or mechanically driven desalination technologies, such as electrodialysis (ED) [
12] and reverse osmosis (RO) [
13], or it is captured through various solar collectors to drive thermally powered desalination technologies, including multi-stage flash (MSF) [
14,
15] and multi-effect distillation (MED) [
16].
In addition to these solar desalination technologies, membrane distillation (MD) has emerged as a promising thermally driven, membrane-based desalination technique with significant potential for solar energy integration. MD operates by utilizing a hydrophobic membrane that permits the transport of water vapor while preventing the passage of liquid water. A key advantage of MD is its ability to operate with feed water temperatures below the boiling point, typically in the range of 40–90 °C, making it highly compatible with solar thermal energy sources [
17]. This characteristic enhances its feasibility for sustainable and energy-efficient freshwater production, particularly in regions with high solar irradiance. MD is available in several configurations, differentiated by the method of vapor collection and condensation within the permeate side. The four primary configurations include direct contact membrane distillation (DCMD) [
18], air gap membrane distillation (AGMD) [
19], where water vapor is condensed within the permeate channel, sweeping gas membrane distillation (SGMD) [
20], and vacuum membrane distillation (VMD) [
21], where the vapor is transported out of the module and condensed externally.
A recent advancement in MD technology is the water gap membrane distillation (WGMD) configuration [
22,
23,
24,
25,
26], which modifies the AGMD design by filling the air gap with stagnant distillate water. This approach enhances the overall productivity of the module compared to AGMD. Unlike DCMD, the stagnant distillate layer in WGMD eliminates convective heat transfer on the permeate side, thereby reducing conductive heat loss and improving thermal performance [
27]. Additionally, the separation of the permeate and cooling streams in WGMD allows for the use of saline water as a coolant, enabling partial recovery of thermal energy from the feed stream prior to main heating [
28]. Consequently, WGMD combines the high-water productivity characteristic of DCMD with the improved thermal efficiency of AGMD, offering a promising alternative for efficient solar-driven desalination.
Membrane distillation integrated with direct solar heating has recently emerged in various configurations as a compact and cost-effective desalination approach [
29]. Despite its potential, further enhancements are needed to significantly boost productivity and achieve sustainable operation. Elrakhawi et al. [
30] developed a one-dimensional finite difference model to analyze a novel sweeping air nanophotonic MD module. The system employed a nanophotonic photothermal coating on a hydrophobic membrane for efficient solar absorption and localized heating, with water vapor transported by a sweeping air stream and condensed on a cooling plate separating the permeate and cooling channels. Performance was tested outdoors in Houston, Texas (July 2019) with a 21,960 ppm saline feed, achieving a distillate flux of 0.65 ± 0.28 kg/(m
2h) under 784.4 ± 217 W/m
2 irradiance and 39 °C feed inlet temperature. Model predictions showed that concentrating solar irradiance by a factor of 7.5 increased productivity ninefold, while increasing feed channel thickness from 0.5 to 7.5 mm improved flux by 7%. Meo et al. [
31] conducted a systematic evaluation of a direct solar absorption DCMD module using an analytical model, where distributed solar irradiation was directly absorbed on the feed side to heat saline water of 35,000 ppm. The study investigated module performance under various flow configurations and operating conditions. For a counter-current feed and permeate configuration at a feed inlet temperature of 30 °C, Reynolds number of 56.8, and module length of 0.1 m, the permeate flux remained constant at approximately 0.58 kg/(m
2h) across all solar intensities. When the module length was increased to 1 m under the same conditions, the flux increased from 0.21 kg/(m
2h) (no solar input) to 0.44 kg/(m
2h) at 2000 W/m
2, achieving a 109.5% enhancement. The maximum flux of 0.9 kg/(m
2h) was observed at a higher feed Reynolds number of 567.7 and 0.1 m module length, but it declined to 0.55 kg/(m
2h) at 1 m length regardless of solar intensity for both lengths.
Conversely, several studies have investigated the transient performance of various membrane distillation configurations driving by indirect heating using external solar thermal collectors. In these systems, solar collectors are employed to absorb incident solar radiation and supply thermal energy to heat the MD feed water [
32,
33,
34]. Duong et al. [
35] developed a one-dimensional empirical model to simulate the performance of a solar-driven DCMD system integrated with an FPC of 22.6 m
2 for seawater desalination. The DCMD module featured an effective membrane area of 7.2 m
2. The system operated under transient solar irradiation in New South Wales, Australia, pending 1–5 January 1991. During that period, a maximum water flux of 2.7 kg/(m
2h) was recorded at a peak feed temperature of 52 °C and a feed flow rate of 900 kg/h. The average daily water production was 142 kg/day, equivalent to 6.3 kg/day per m
2 of the collector area and 19.7 kg/day per m
2 of the membrane area. Ma et al. [
36] theoretically analyzed the performance of a compact solar-powered VMD system for seawater desalination. The system was driven by solar irradiation, varying from 70 to 640 W/m
2, over a 12 h period on 1 August in Toulouse, France, using an FPC. A freshwater yield of 3.7 kg/day was achieved with an FPC area of 0.18 m
2, which increased to 96 kg/day with a collector area of 3 m
2.
Furthermore, extensive research has been dedicated to solar-powered AGMD systems due to their high thermal efficiency and potential for thermal energy recovery [
37,
38,
39,
40]. Sandid et al. [
41] conducted both experimental and theoretical studies on a spiral-wound AGMD module. The experiments were carried out with a membrane area of 14.4 m
2, powered by FPC field with an area of 12 m
2 side by side with ETC field with an area of 8 m
2 to desalinate seawater in Port Said, Egypt, on 28th and 29th of June 2020. Maximum water fluxes of 0.9 kg/(m
2h) and 1.25 kg/(m
2h) were achieved using FPC and ETC, respectively, at a maximum feed inlet temperature of 91 °C in June. In winter (December), at a reduced feed temperature of 70 °C, fluxes declined by 38.3% (FPC) and 44.4% (ETC). The AGMD module coupled with FPC produced 50 kg/day of freshwater in June, which increased by 57.7% when integrated with ETC. On the other hand, they used mathematical modeling to predict accumulated distillate throughout a complete year. Ruiz-Aguirre et al. [
42] conducted pilot-scale experiments on multi-channel spiral-wound AGMD modules (7.2 m
2 and 24 m
2 membrane areas) integrated with FPCs delivering up to 7 kW thermal power. They also developed an optimization model, validated experimentally, that could produce 2.44 kg/(m
2h) of freshwater with a specific energy consumption of 141.5 kWh/m
3 at 80 °C feed and 20 °C coolant inlet temperatures with a length of 3.72 m and 17.9 m
2 membrane area.
In contrast, there is a notable scarcity of research on WGMD systems integrated with solar thermal energy [
43,
44]. Alquraish et al. [
44] carried out experimental investigations using a pilot-scale spiral-wound WGMD module with a membrane surface area of 10 m
2, powered by a solar thermal system comprising a total collector area of 2 m
2. The experiments were conducted under the climatic conditions of Kairouan, Tunisia, over several days in August 2020. The results demonstrated a maximum freshwater production rate of 15.92 kg/day per m
2 of membrane surface on August 9th, corresponding to a peak solar irradiance of 978.3 W/m
2 and a maximum feed inlet temperature of 71.9 °C. Furthermore, the system’s specific thermal energy consumption (STEC) varied between 90 and 310 kWh/m
3.
The literature reveals a significant gap in studies addressing the performance of WGMD driven by solar energy, particularly for hollow fiber (HF) module configurations, which offer high compactness and scalability. To address this, the present work develops a computational fluid dynamics (CFD) model to simulate the transient performance of an HF-WGMD system integrated with a flat-plate solar collector system. A lumped-parameter transient model of the FPC is coupled with the CFD model to predict the temporal variation in feed water temperature based on the absorbed solar irradiation. The CFD model is validated against experimental data, showing strong agreement.
The system performance is evaluated over four representative days in March 2025 in Alexandria, Egypt, under varying solar intensities. A parametric analysis is conducted to assess the influence of key solar collector parameters, including total collectors’ area, overall width, number of solar tubes and mass flow rate of the working fluid, on the instantaneous and average water flux of the HF-WGMD module. Additionally, the impact of meteorological conditions (solar irradiation and ambient temperature) during the four days on feed inlet temperature and module productivity is examined. Finally, the study investigates multistage HF-WGMD configurations, analyzing their effects on collector area requirements, thermal energy recovery, energy consumption, overall solar desalination efficiency and freshwater productivity.
3. Theory
A two-dimensional axisymmetric transient model is developed to simulate the desalination process in an HF-WGMD module thermally integrated with a flat-plate solar collector, which serves as the primary heating source for the feed water. In the numerical setup, a single hollow fiber membrane enclosed within a cooling tube is modeled to estimate the permeate flux.
It can be observed that incorporating cooling tubes in the HF-WGMD configuration helps prevent potential interactions between the membrane hollow fibers. However, this aspect requires further experimental investigation to assess the influence of headers and other fittings on the overall module performance. In this study, the predicted performance is scaled according to the total number of hollow fibers in the module to estimate the total water production, assuming uniform performance for all individual hollow fibers enclosed within the cooling tubes.
The computational domain consists of five concentric, axisymmetric regions arranged radially from the center outward: the feed channel, the hollow fiber membrane, the water gap, the cooling tube wall and the coolant channel, as illustrated in
Figure 3. Additionally, the geometrical and operating conditions of HF-WGMD module domain are presented in
Table 1.
The model is based on the following assumptions:
Both feed and coolant flows are laminar.
The membrane is isotropic in porosity.
Fouling at the feed-membrane interface is neglected.
Heat losses to the surroundings are neglected.
Cooling water temperature is equal to the instantaneous ambient temperature.
The heat exchanger is ideal and has 100% effectiveness ().
The heat loss coefficient of FPC is constant at an average value.
3.1. Governing Equations
The two-dimensional cylindrical forms of the transient conservation equations for mass, momentum and energy are solved across the five computational domains of the HF-WGMD module. These equations are simultaneously coupled with the energy conservation equation of the flat-plate solar collector to accurately determine the transient feedwater temperature supplied to the HF-WGMD unit.
3.1.1. Mass Transport Equations
The concentration distributions in both the feed and membrane domains are determined by solving the transient mass transport equations in cylindrical coordinates, as summarized in
Table 2. In the feed domain, water vapor transport is governed by the combined effects of convection and diffusion. In contrast, transport across the membrane domain occurs solely via diffusion, in accordance with Fick’s law.
Table 2.
Transient mass conservation equations of feed and membrane domains in cylindrical coordinates.
Table 2.
Transient mass conservation equations of feed and membrane domains in cylindrical coordinates.
| Domain | Equation |
|---|
| Feed | | (1) |
| Membrane | | (2) |
The mutual diffusion coefficient of water and salt in the feed solution,
is estimated using the Wilke–Chang correlation [
45], expressed as:
where
is the feed solution molecular mass,
is the feed water dynamic viscosity in cP, and
is the water molecular volume in cm
3/mol.
In contrast, mass transport within the membrane domain is governed by a combination of diffusion mechanisms, namely ordinary molecular diffusion and Knudsen diffusion. The relative significance of these mechanisms is determined by evaluating the Knudsen number, which is calculated using the following expression [
46]:
where
is the average membrane pore diameter and
is the mean free path of vapor molecules, which can be estimated using the following equation:
where
is the Boltzmann constant,
is the average membrane temperature in K,
is the molecular collision diameter of air and
is the molecular collision diameter of water vapor [
47].
In the present study, the calculated Knudsen number lies within the transition regime
, confirming that mass transport through the membrane is simultaneously influenced by both molecular diffusion and Knudsen diffusion, with neither mechanism being negligible. Accordingly, the effective mass transport is represented as the combined contribution of these two mechanisms. The corresponding diffusion coefficients for molecular and Knudsen diffusion are evaluated as follows [
46]:
While, the effective diffusion coefficient, which accounts for the combined contribution of both mechanisms, is then calculated from:
To account for the structural characteristics of the membrane, namely porosity and pore tortuosity, the effective membrane diffusion coefficient is expressed as:
where ε is the membrane porosity and
is the membrane pore tortuosity which can be estimated by the following formula [
48].
3.1.2. Momentum Transport Equations
The continuity and Navier–Stokes equations are solved within the feed and cooling channel domains to obtain the transient pressure and velocity fields. These governing equations are formulated for two-dimensional, incompressible, unsteady, laminar flow in cylindrical coordinates. They are expressed as follows:
where
is density,
is dynamic viscosity and
is domain local pressure.
3.1.3. Energy Transport Equations
Temperature distribution plays a pivotal role in this study, as it directly governs the local saturation concentrations on both sides of the membrane and influences the temperature-dependent diffusion coefficients. To capture these effects, the energy conservation equation, whose mathematical formulations are summarized in
Table 3, is solved simultaneously with the mass and momentum transport equations across all five domains of the HF-WGMD module. In the feed and cooling channel domains, heat transfer occurs through a combination of convection and conduction. In contrast, heat transfer within the membrane, water gap and cooling tube wall domains is restricted to conduction, as also presented in
Table 3.
Table 3.
Transient energy conservation equations of HF-WGMD module domains in cylindrical coordinates.
Table 3.
Transient energy conservation equations of HF-WGMD module domains in cylindrical coordinates.
| Domain | Equation |
|---|
| Feed | | (14) |
| Coolant |
| Membrane | | (15) |
| Water gap |
| Cooling tube |
The effective thermal conductivity of the membrane,
, is estimated by accounting for both the solid polymer material and the vapor within the pores. It is expressed as a porosity-weighted combination of the membrane material conductivity and the vapor conductivity, as follows:
where
is the effective thermal conductivity of membrane domain,
is the water vapor thermal conductivity and
is the thermal conductivity of solid PVDF membrane material.
3.1.4. FPC Mathematical Model
The flat-plate collector system, comprising an array of parallel and series-connected collectors each with an aperture area of
, is modeled as a transient lumped-parameter system. The detailed specifications of the FPC system, operating with water as the working solar fluid, are summarized in
Table 4.
The width of the collectors’ array corresponds to the number of collectors connected in parallel. For example, an array width of 5 m represents five collectors arranged in parallel. Conversely, the length of the array represents the number of collectors connected in series.
The total area of the array is calculated as the product of its width and length. The number of tubes is counted along the width direction. For instance, in an array that is 5 m wide and 10 m long with a total of 40 tubes, each 1 m-wide collector contains 8 tubes. The total length of the connected tubes in series is therefore 10 m (representing 10 series-connected collectors).
The specific configuration of the collectors will be defined in each results section according to the requirements of the investigation.
The instantaneous useful heat gain by the solar fluid in the FPC, which represents the thermal energy transferred to the saline feed water of the desalination unit, is calculated using the following expression [
49]:
where
is the instantaneous useful heat gain by the solar fluid,
is the FPC area,
is the instantaneous solar fluid inlet temperature,
is the glass transmissivity,
is the plate absorptivity,
is the overall heat loss coefficient,
is the instantaneous solar irradiance on the FPC,
is the instantaneous ambient temperature and
is the heat removal factor which can be calculated using the equation [
49]:
where
is the solar fluid mass flow rate,
is the solar fluid specific heat at constant pressure and
is the collector efficiency factor and expressed as follows [
49]:
where
and
are the inside and outside solar tube diameters, respectively,
is the solar tube spacing in the FPC,
is the bond conductance,
is the solar fluid convection heat transfer coefficient and
is the standard fin efficiency that can be calculated using the following equation [
49]:
where
is the absorber plate thermal conductivity and
is the absorber plate thickness.
On the other hand, the ideal heat exchanger is assumed to transfer the thermal energy gained by the solar fluid directly to the feed water supplied to the HF-WGMD module. Accordingly, the instantaneous outlet temperature of the solar fluid from the FPC, and thus the inlet feed water temperature to the desalination system, is determined as follows:
where
is the instantaneous feed water inlet temperature to the desalination module.
3.2. Boundary Conditions
The governing equations described above are coupled with appropriate boundary conditions applied across all computational domains, thereby forming a complete mathematical model for numerical solution. These boundary conditions (BC) specify the physical constraints at the inlets, outlets, walls and the interfaces between adjacent domains. The subsequent sections present the detailed boundary conditions adopted for each of the conservation equations.
3.2.1. Mass Transport BC
Table 5 summarizes the boundary conditions imposed on the mass transport equations within the feed and membrane domains. Specifically, the concentration at the hot feed-membrane interface (
) and at the cold membrane-water gap interface (
) are prescribed according to the saturation concentration of water vapor corresponding to the local temperature at each interface.
The vapor concentrations at the hot and cold membrane interfaces are determined from the saturation vapor pressures evaluated at the local membrane-water gap interface temperature, using the standard Antoine equation. At the feed-membrane interface, however, the effect of feed water salinity is taken into account by employing a modified form of the Antoine equation, which reduces the saturation pressure relative to that of pure water [
50].
where
denotes the mole fraction of water and
represents the water activity coefficient. Both
and
are calculated using the following equations [
51]:
where
is the mole fraction of salt in the feed solution.
These saturation pressures are subsequently employed to calculate the corresponding vapor concentrations at both the feed–membrane and membrane–water gap interfaces, as expressed by the following equations:
where
is the saturation concentration of water vapor,
is the membrane local temperature,
is the specific volume of water vapor,
is the vapor mass content and
is the atmospheric pressure.
3.2.2. Momentum Transport BC
The boundary conditions applied to the momentum transport equations, governing the feed and coolant channel domains, are summarized in
Table 6.
3.2.3. Energy Transport BC
The heat energy conservation equation is solved simultaneously across all five domains of the HF-WGMD module. Accordingly, the external boundary conditions governing heat transport in the module are summarized in
Table 7.
Additionally, two internal boundary conditions are imposed within the feed channel and water gap domains. At the feed–membrane interface, a boundary heat sink is applied to account for the thermal energy consumed from the feed water during evaporation. Conversely, at the membrane-water gap interface, a boundary heat source is introduced to represent the release of latent heat to the distillate water during vapor condensation. In both cases, the corresponding amounts of latent heat flux are evaluated using the following expression:
where
denotes the distillate water flux and
represents the latent heat of vaporization (or condensation). Both quantities are calculated using the following relations:
where
is the membrane diffusion coefficient,
is the local vapor concentration in membrane,
is the water molecular mass and
is the membrane local temperature.
3.3. Performance Indicators of the Solar Desalination System
The performance of the proposed solar-driven HF-WGMD system is assessed through the evaluation of three key parameters: the specific thermal energy consumption, the overall system efficiency and the freshwater productivity. The instantaneous STEC of the HF-WGMD module is defined as the amount of thermal energy supplied to the feed stream per unit volume of distillate produced. It can be expressed mathematically as:
Here, is the distillate water density, is the feed water mass flow rate, is the feed water specific heat at constant pressure, is the instantaneous feed water inlet temperature, is the instantaneous cooling water outlet temperature, is the number of fibers inside the module, is the number of module stages, is the membrane outer radius, is the module effective length, is the feed channer radius, is the instantaneous distillate water flux and is the feed water density.
Moreover, the average STEC over the entire daytime operation period is employed to evaluate the overall thermal energy consumption of the HF-WGMD modules. It is determined as:
where
is the daytime operation period.
The solar desalination efficiency of the solar-driven HF-WGMD system is defined as the ratio of the thermal energy effectively utilized for water evaporation at the feed-membrane interface to the total solar energy incident on the flat-plate collector array. Accordingly, the instantaneous system efficiency is expressed as:
where
is the latent heat of vaporization,
is the instantaneous solar irradiance on the FPC and
is the solar collector area.
The average overall efficiency of the solar-driven HF-WGMD system over the daytime operation period is obtained by integrating the instantaneous efficiency with respect to time and normalizing by the total operation duration. It is expressed as:
On the other hand, different forms of productivity are considered in the present study to evaluate the performance of the proposed multistage HF-WGMD systems. The specific productivity (SP) per unit membrane area is defined as:
In contrast, the specific productivity normalized by the solar collector area quantifies the freshwater yield relative to the effective collector surface. It is mathematically expressed as:
The overall productivity represents the cumulative daily freshwater yield of the desalination system in kg/day and is mathematically expressed as:
3.4. Solution Technique
The numerical model couples the governing transport equations with supplementary algebraic relations, solved simultaneously in COMSOL Multiphysics 5.4 using the finite element method. Transient mass, momentum, and heat transport within the HF-WGMD module are resolved using the Transport of Diluted Species, Laminar Flow, and Heat Transfer models. The flat-plate solar collector is represented by a transient lumped-parameter model, which is applicable under adaptable Mathematics models, where the energy balance equations (Equations (18)–(23)) predict the time-dependent outlet temperature of the solar working fluid based on absorbed solar irradiation, ambient temperature, and collector characteristics. This predicted outlet temperature is directly applied as the inlet boundary condition for the feed stream entering the HF-WGMD module, thereby establishing a coupled simulation framework between the solar collector and CFD models.
Additional user-defined equations are implemented to account for vapor diffusion coefficients (Equations (3) and (6)–(9)), Antoine correlations for vapor pressure (Equations (24) and (25)), and temperature-dependent thermophysical properties. The complete set of equations is solved concurrently under specified initial and boundary conditions, with the distillate side assumed to be salt-free (zero solute concentration).
3.5. Grid Independence Study
A grid independence study is conducted to evaluate the influence of mesh density on the accuracy of the numerical solution while minimizing computational cost. Three grid resolutions are tested for the simulated geometry of a 300 mm HF-WGMD module integrated with a solar collector array of 50 m2 total area under transient conditions corresponding to solar irradiation and ambient temperature on Day 1. The water flux is evaluated at three different daytime hours (09:00, 12:00 and 15:00) for grids consisting of 112,715, 348,567 and 1,017,025 elements.
The results, presented in
Figure 4, indicate that increasing the number of grid elements beyond 348,567 yields no significant improvement in solution accuracy. Specifically, tripling the number of elements from 348,567 to 1,017,025 alters the predicted water flux by only 0.75%, 0.91% and 1.77% at 09:00, 12:00 and 15:00, respectively. Based on these findings, a grid size of 348,567 elements is adopted for the present transient CFD simulations to ensure accurate numerical predictions while minimizing computational time.