Next Article in Journal
Improved Discrete-Time Active Disturbance Rejection Control for Enhancing Dynamics of Current Loop in LC-Filtered SPMSM Drive System
Previous Article in Journal
Energy Management of a Semi-Autonomous Truck Using a Blended Multiple Model Controller Based on Particle Swarm Optimization
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Vortex-Induced Turbulence Optimized Membrane Enthalpy Exchanger: Dynamic Humidity Modulation and Coupled Heat–Mass Transfer Enhancement

by
Yang Liu
1,2,
Dong-Chuan Mo
1,2 and
Shu-Shen Lyu
1,2,*
1
School of Materials, Sun Yat-sen University, Shenzhen 518107, China
2
Guangdong Engineering Technology Research Centre for Advanced Thermal Control Material and System Integration (ATCMSI), Shenzhen 518107, China
*
Author to whom correspondence should be addressed.
Energies 2025, 18(11), 2892; https://doi.org/10.3390/en18112892 (registering DOI)
Submission received: 2 May 2025 / Revised: 27 May 2025 / Accepted: 28 May 2025 / Published: 30 May 2025

Abstract

:
A bioinspired vortex-inducing architecture was engineered within the hydrodynamic focusing region of membrane-based enthalpy exchangers (MEEs) to generate controlled Kármán vortex shedding, strategically enhancing thermal–hygric coupling through boundary layer modulation. Computational simulations employing ANSYS Fluent 2024R1 and grid-convergence validation (GCI < 1.8%) demonstrated that at Re = 392 (2.57 m/s flow velocity), the vortex-integrated configuration achieved temperature exchange efficiency enhancements of 3.91% (summer) and 3.58% (winter), latent efficiency gains of 3.71% and 3.53%, alongside enthalpy effectiveness improvements of 3.37% and 3.36%, respectively. The interconnected momentum–heat–mass analogies culminated in peaks of performance evaluation criterion (PEC) = 1.33 (heat transfer) and 1.22 (mass transfer), substantiating vortex-induced Reynolds analogy optimization under typical HVAC operational scenarios (summer: 27 °C/50.3% RH; winter: 21 °C/39.7% RH).

1. Introduction

A considerable body of research has focused on air-to-air heat exchangers [1,2,3] in recent decades, driven by the evolving global energy consumption landscape. Reports indicate that building energy consumption accounts for a significant portion of total energy use and is a major contributor to carbon emissions [4]. In China, building energy consumption represents approximately 2% to 6% of total energy consumption [5]. Particularly in northern regions, heating energy consumption constitutes an even higher share of total building energy use. Among heating energy consumption, approximately 40% is lost due to factors such as building structure. The literature in [6] suggests that carbon emissions from building energy consumption contribute to 50% of total carbon emissions. Therefore, optimizing the building energy consumption structure is a critical step toward achieving carbon neutrality and realizing the dual-carbon economy, a sustainable strategic goal that benefits both the nation and its citizens.
Air-to-air heat exchangers are devices designed to recover energy lost in building energy consumption. Current research primarily focuses on enhancing the efficiency of energy recovery. Recently, studies have investigated the performance of air-to-air heat exchangers using various heat exchange materials [1,7,8]. These studies have shown that MEEs offer superior cost–performance ratios. A considerable body of literature has been published on MEEs [3,9,10,11,12,13]. While structural optimization strategies for microchannel heat exchangers are well documented [14,15,16,17], membrane-mediated thermal–hygric coupling in MEEs imposes fundamentally distinct constraints. Specifically, the semi-permeable membrane materials (e.g., selective water vapor transport layers) employed in MEEs require controlled, low-velocity flows to achieve optimal moisture transport, preventing membrane degradation and maintaining high hygric performance.
In addition, efforts have been made to optimize the structural design of heat exchangers. For instance, Zhang [18] optimized the structure of an MEE using the Number of Transfer Units (NTUS) method, achieving a design with optimal performance. Valerii I. Deshko [19] and colleagues compared numerical models of heat transfer in parallel plate channels with and without fins under varying wall temperatures. Neda Asasian-Kolur [20] and her team reviewed the performance of an MEE under different flow channel configurations, including cross-flow, quasi-counterflow, and triangular structures. Previous studies have explored various pin fin geometries for turbulence induction, including square [21,22] and elliptical [23] configurations. Square structures, while effective in creating flow separation at higher Re (>500), suffer from excessive pressure drop penalties in MEE conditions due to sharp-edge recirculation zones. Elliptical profiles demonstrate reduced form drag, but their orientation-dependent vortex shedding characteristics complicate manufacturability in thin (<0.1 mm) membrane spacers. In contrast, cylindrical structures exhibit axisymmetric flow separation that generates predictable Kármán vortex streets even at Re < 400 [24,25], making them particularly suitable for MEE’s spatially constrained, low-inertia flows. The cylindrical vortex-inducing elements (diameter d = 0.3 mm, pitch L = 10 mm) were dimensioned based on two-phase optimization criteria: Vortex Interaction Length Scale, where the L/d ratio (~33) exceeds the critical value of 20 required for full vortex street development in confined channels, ensuring that coherent vortices persist downstream [26]. Blockage Ratio Control [27]: with a channel height H = 2.5 mm, the 12% blockage ratio (d/H) minimizes permanent recirculation zones while maintaining sufficient momentum injection to disrupt concentration boundary layers.
In contemporary development efforts, a variety of research methods have been employed. Among these, numerical simulation techniques provide high-fidelity predictions of device performance. Many research institutions now use digital tools to accelerate development and reduce testing costs for MEEs. Commonly employed methods include finite element analysis [28,29], effectiveness-NTU methods [30,31,32], and computational fluid dynamics (CFD) modeling [2,7,8,11]. This study primarily utilizes CFD simulation software Fluent from the ANSYS commercial software package to conduct the simulations. The structural optimization of the air channels in an MEE is performed, and performance predictions are made using CFD tools. The theory of flow–solid coupling, which enhances heat transfer in microchannels, has been extensively applied in microchannel heat sinks for electronic equipment cooling [14]. Qing-wen Li [15] and colleagues simulated two types of intermittent microchannel configurations: straight and staggered. The results indicated that staggered microchannels exhibited the lowest thermal resistance and the highest friction factor, while continuous microchannels demonstrated the opposite performance. Yifan Li [16] and co-authors conducted numerical studies on the heat dissipation performance of microchannels combined with micro pin fins using a 3D conjugate heat transfer model, revealing the coupled mechanisms between fluid flow and heat transfer. Prabhakar Bhandari [17] numerically investigated the heat transfer and fluid flow characteristics of open microchannel heat sinks with square pin fins, predicting the performance of the devices. To date, a substantial body of the literature has focused on structural optimization for heat transfer enhancement, including pin fin configurations such as square [21,22], circular [33,34], water wing [21,35], elliptical [35], and others. In the case of MEEs, the Reynolds number (Re) is significantly lower than 2300 (with Re below 300 under the conditions described later in this study), indicating a fully developed laminar flow structure within the fluid domain. Therefore, to enhance both heat and mass transfer, it is essential to disrupt the laminar flow structure and induce turbulence.
In this study, a simplified CFD model was initially developed to verify the mesh-independent solution and compare it with experimental test data. The experimental data were used to validate the effectiveness of the CFD model and quantify the distribution of physical fields within the model. Subsequently, theoretical analysis revealed that introducing disturbance structures in the x- or y-direction within the fully developed laminar flow domain can effectively induce turbulence, specifically Kármán vortex street flow, thereby enhancing heat and mass transfer in microchannels. The physical field distribution and heat–mass transfer characteristics within the MEE were then visualized to elucidate the enhancement mechanisms. To our knowledge, this represents the first systematic demonstration of bioinspired cylindrical vortex generators in low-Re MEE environments. Unlike previous flow disruption strategies focusing solely on heat transfer enhancement [21,22,33,34], our approach explicitly addresses coupled thermal–hygric transport through synchronized vortex-induced boundary layer renewal as a critical mechanism for latent heat recovery optimization in membrane-based systems. Finally, the resistance distribution and efficiency balance of the MEE provided insights into potential research directions, with future work targeting the multi-objective optimization of variable geometries and materials—including scale-up strategies for full-size ERV cores through modular multi-cell array configurations, geometry ratio scaling (L/D = 5–30), and ΔP management under industrial airflow rates (1–3 m/s)—alongside integrating dynamic sorption hysteresis measurements into a multiscale CFD framework to couple vortex-enhanced transport with non-equilibrium sorption effects, with the findings offering significant implications for further understanding the principles of heat and mass transfer enhancement for sustainable HVAC systems.

2. CFD Model

The subject of this study is the MEE device, as illustrated in Figure 1a. The device primarily consists of an injection-molded support structure made of Acrylonitrile–butadiene–styrene (ABS) material and a high-performance permeable polymer membrane from Fortuneway Technologies Inc. (Zhongshan, China), which functions as the thermal and humidity management medium for the enthalpy heat exchanger. A simplified schematic of the equivalent structure is shown in Figure 1b. For the purposes of modeling, the two adjacent airflow channels are referred to as the cold and hot channels, and the entire system is represented by an equivalent model. The model is divided into three layers: the upper layer represents the hot fluid, the middle layer represents the solid membrane material, and the lower layer represents the cold fluid, with the two fluids arranged in a cross-flow configuration. The operating conditions of the enthalpy heat exchanger, which constitute the primary focus of this simulation, include flow rates of 150 m3/h, 200 m3/h, 250 m3/h, 300 m3/h, and 350 m3/h.

2.1. Assumptions

For enthalpy exchangers, two primary technical strategies are commonly employed to enhance exchange efficiency: structural optimization and the selection of isolation materials. The choice of isolation layer material is crucial, as it influences both the heat and mass transfer efficiency of the exchanger and plays a key role in ensuring the long-term hygiene and effective operation of the air system. Common isolation materials can be categorized into paper-based and polymer materials. There is an extensive amount of literature that demonstrates that MEEs offer longer lifespans, lower degradation rates, and superior heat and mass transfer efficiencies [1]. Among these materials, the mass transfer behavior of polymers is particularly complex and warrants in-depth investigation. S. Koester [36] and colleagues investigated selectively permeable polymer membranes used in enthalpy exchangers. Their findings indicated that the gas permeation efficiency of membrane materials is primarily determined by two key parameters: selectivity and permeability. Selectivity depends largely on the physicochemical properties of the selective polymer layer, while permeability is influenced by the fluid dynamics near the membrane interface. Therefore, studying the dynamic behavior of enthalpy exchangers is critical, and this study focuses on the dynamics of these devices.
It is essential to simplify or omit certain details that have minimal impact on the actual results. These simplifications, though seemingly minor, are grounded in significant insights derived from the extensive research efforts of previous scholars. The key assumptions made in this study are as follows:
  • Heat and mass transfer processes are assumed to be steady-state.
  • The transport of water vapor through the membrane material is modeled as an adsorption–diffusion–desorption process, which is assumed to be in dynamic equilibrium. While adsorption/desorption hysteresis may modulate sorption capacity under transient conditions [37], its effects are excluded here due to the steady-state assumption of a simplification validated for isolating vortex-enhanced transport mechanisms [38]. Additionally, the latent heat of vaporization and adsorption heat on both sides of the membrane are assumed to be equal.
  • Given the small temperature and water vapor concentration differences across the membrane, the thermal and moisture resistance within the membrane is assumed to be constant, with isotropic and constant thermal conductivity and diffusion coefficient for water vapor [8,12,29,39].
  • The heat and mass transfer process within the entire enthalpy exchanger is simplified as the thermal and material exchange between two adjacent fluid channels.
  • Solid surfaces are modeled using no-slip boundary conditions. In the equivalent model, the laminar flow model is employed to solve the equations, whereas the standard κ-ε turbulence model is used in the optimized model to enhance wall treatment.

2.2. Governing Equations

Currently, a wide range of commercial CFD software is available, with an extensive amount of the literature reporting on the simulation of heat exchangers using Fluent and the simulation of heat and mass transfer behaviors in enthalpy exchangers using COMSOL, among others. In this study, ANSYS Fluent 2024R1 is employed to simulate the heat and mass transfer processes in air-to-air MEEs. Although the Fluent solver efficiently handles fluid–solid coupling for heat transfer, it lacks a dedicated module for simulating the mass transfer process in MEEs. Therefore, this study utilizes Fluent’s user-defined function (UDF) feature [40] to implement mass transfer coupling simulations based on the actual operating conditions of the model. This paper focuses on examining heat and mass transfer behavior under both laminar flow and forced turbulence conditions.

2.2.1. Laminar Flow

In the current version of Fluent software [40], the general form of the governing equations for fluid–solid-coupled heat and mass transfer is presented in Equation (1). The first term on the lefthand side represents the transient term, while the second term represents the convective term. On the righthand side, the first term corresponds to the diffusion term, and the second term represents the source term. By solving Equation (1), the continuity equation is obtained, as shown in Equation (2).
ρ ϕ t + ρ ϕ V = · Γ ϕ ϕ + S ϕ
ρ t + ρ V = 0
For incompressible gases, the density ρ is assumed to be constant, allowing the equation to be simplified as follows:
ρ V = 0
ρ V Y i = · J i
By solving the continuity equation (Equation (3)) and the multi-species transport equations (Equation (4)), the mass distribution of multiple species within the fluid domain can be determined. The temperature field distribution is obtained by solving the energy equation (Equation (5)).
V ρ E + P = · k T i h i J i
According to the literature [41], during the adsorption and desorption processes of water vapor on both sides of the membrane, the latent heat of vaporization and the latent heat of adsorption are equal in magnitude but opposite in direction. Therefore, Equation (5) can be simplified as follows:
V ρ E + P = · k T
In the above equation, k represents thermal conductivity, T denotes the temperature of the humid air, Y i is the mass fraction of species i, and J i is the diffusion flux of species i. The diffusion flux of species i (either air or water vapor) increases with the concentration gradient. By default, the dilute approximation is employed. Under laminar flow conditions, the diffusion flux is described by the following equation (Equation (7)):
J i = ρ D i , m a Y i
D i , m a is the mass diffusion coefficient of species i (m²/s), which is defined for multi-component fluids as follows:
  D a , m a = D v , m a = 1.87 × 10 10 T 2.072 P

2.2.2. Turbulence Flow

The diffusion flux equation for turbulence flow is given by Equation (9):
  J i = ρ D i , m a + μ t S c t Y i
  S c t = μ t ρ D t
In the equation, S c t is the turbulent Schmidt number, μ t is the turbulent viscosity, and D t is the turbulent diffusion coefficient. In Fluent, commonly recommended turbulence models include low-Reynolds-number κ ε , Realizable κ ε , RNG κ ε , STE κ ω , and the Reynolds Stress Model (RSM). Among these, the STE κ ω model is considered the most cost-effective for simulating complex industrial turbulent flows. Previous researchers [41,42] have employed the standard κ ε model and found it to yield accurate results. Two-equation models solve a pair of transport equations and use the eddy viscosity approach to represent Reynolds stresses. Turbulence kinetic energy (k) and its dissipation rate (ε) are derived from the following transport equations [43]:
x i ρ k u i = x j μ + μ t σ k k x j + G k + G b ρ ε
x i ρ ε u i = x j μ + μ t σ ε ε x j + C 1 ε ε k G k + C 3 ε G b C 2 ε ρ ε 2 k
Here, G k denotes the production of turbulence kinetic energy arising from mean velocity gradients, while G b represents the production of turbulence kinetic energy due to buoyancy. The constants C 1 ε , C 2 ε , and C 3 ε are fixed values, and σ k and σ ε are turbulence. Prandtl numbers are denoted by k and ε. The default values for these model constants are as follows: C 1 ε = 1.44 , C 2 ε = 1.92 , C μ = 0.09 , σ k = 1.0 , σ ε = 1.3 . The standard κ ε turbulence model was adopted in this study, as it demonstrates superior predictive accuracy for high-strain-rate flows (e.g., vortex street shear layers) compared to the Realizable κ ε model [44,45].

2.3. Membrane Modeling and Boundary Conditions

The sample utilized in this study is a self-prepared specimen. The parameters of the sample and its membrane material are detailed in Table 1. The membrane material is a commercially available asymmetric composite polymer, with a total thickness of 0.026 mm. This composite consists of a porous support layer (0.025 mm) and a permeable functional layer (0.001 mm). The SEM microstructure of the membrane is presented in Figure 2. The support material of the composite polymer is polyethylene (PE), while the coating is composed of a polyether block copolymer. EDS analysis reveals that part of the coating has infiltrated the porous support layer.
For the air-to-air enthalpy exchanger, the membrane serves as a critical medium for transferring heat and mass. Under the assumed conditions, heat and mass transfer occur exclusively in a one-dimensional direction. Consequently, by applying the principles of energy and mass conservation, the following expression can be derived:
  2 T m δ 2 = 0
  2 Y i , m δ 2 = 0
Figure 3 illustrates a typical diagram of heat and moisture flux through the selectively permeable layer and the porous support layer, based on resistance principles. Given the assumption of no-slip conditions at the membrane surface, the mass flux of water vapor can be determined using Fick’s law [46].
j v = ρ a D v , m a Y v Z h s = ρ m D m Y v , m Z m = ρ a D v , m a Y v Z c s
Consequently, the energy and mass conservation equations, together with the boundary conditions for the MEE model, can be described as presented in Table 2 and Table 3.
For the entire computational domain, the high-temperature, high-humidity fluid transfers heat and diffuses water vapor in the normal direction of the membrane toward the low-temperature, low-humidity fluid. To clearly depict the contour plots of the fluid–solid contact surface and the continuous variations in temperature and humidity at the interface between the two phases, the following formula [47] can be utilized:
  θ = W m a x 1 C + C ϕ
  ϕ ω = e 5294 T 10 6 1.61 ϕ
ω = Y v 1 Y v
On the hot fluid side, water vapor in the high-humidity fluid is adsorbed by the hydrophilic groups of the membrane material. On the cold fluid side, water vapor within the membrane diffuses toward the low-humidity fluid and is subsequently carried away, a process known as desorption. This study primarily focuses on steady-state conditions; therefore, the effects of adsorption hysteresis are negligible.
Based on the boundary conditions outlined in Table 3 and the system of Equations (16)–(18),
Y V , m c = Y V , c s + a 1 Y V , h s Y V , m h  
Y V , m h = ρ c ρ h Y V , c s + ( a 1 ρ c ρ h + a 2 ) Y V , m c ( a 1 ρ c ρ h + a 2 + 1.0 )
a 1 = a 2 = D m Z D v , m s δ
Under the assumption of negligible discrepancies in air density and water vapor diffusion coefficients between cold and hot fluid streams, and given the highly symmetric configuration of the numerical model in this study, the coefficients a 1 and a 2 are defined as equivalent. The user-defined function (UDF) governing mass transfer was subsequently coded in C+ code, leveraging Equations (19)–(21) to dynamically update boundary conditions and ensure bidirectional coupling between thermal and hygric transport mechanisms.

2.4. Definition of Effectiveness

To verify the validity of the simulation data, this study will conduct enthalpy difference efficiency tests on the test sample in accordance with the GB/T 21087-2020 standard [48]. The test conditions are detailed in Table 4. The experimental system incorporated three calibrated Vaisala HMT333 humidity sensors (accuracy: ±1.0% RH at 25 °C) and T-type thermocouples (accuracy: ±0.5 °C) positioned at both supply/exhaust air streams: (a) humidity validation with saturated salt solutions (LiCl: 11.3% RH, MgCl2: 33.0% RH, NaCl: 75.3% RH); (b) temperature calibration against a Fluke 724 precision thermometer in water bath (±0.1 °C resolution). Each measurement campaign consisted of five independent replicates under identical operating conditions (ΔT = 15 ± 0.3 °C, φ = 50 ± 2% RH), with data recorded at 10 s intervals over 30 min stabilized periods. Figure 4a presents the enthalpy difference test laboratory for the MEE device, which primarily consists of two refrigeration/heating chambers designed to simulate indoor and outdoor environments. Figure 4b illustrates the layout of the air-to-air total heat exchange ventilator and the test environment utilized to validate the experimental data. The core components of the ventilator include the total heat exchange core, two variable-frequency fans, and multiple temperature and humidity sensors. The fans operate using negative pressure suction to provide airflow, ensuring equal air volume for both the fresh air and exhaust airflow. The tests are conducted in the standard enthalpy difference laboratory in accordance with GB/T 21087-2020. To calculate the sensible heat, latent heat, and enthalpy efficiency of the MEE, the following formulas are employed:
  ε s = T S A T O A T R A T O A
  ε l = ω S A ω O A ω R A ω O A
  ε H = H S A H O A H R A H O A
Here, T, ω, and H denote temperature (°C), the humidity ratio (kg/kg), and enthalpy (kJ/kg), respectively. The subscripts RA, OA, and SA represent return air, outdoor air, and supply air, respectively.

2.5. Uncertainty Analysis

Measurement uncertainty in the experimental results was evaluated following Moffat’s method [49], using the root-sum-square (RSS) approach. For CFD simulations, numerical uncertainties were quantified via the grid convergence index (GCI < 1.8%) and turbulence model sensitivity tests (±5% eddy viscosity variation). The baseline model exhibited higher uncertainties (±2.1–2.5%) due to simplified assumptions, while the turbulence-enhanced model with vortex generators reduced errors to ±1.5–2.0%. Winter conditions showed marginally higher uncertainties than summer (e.g., ±1.7–2.4% vs. ±1.5–1.8%) due to viscosity effects. Detailed values are shown in Table 5.
U = Y T U T 2 + Y R H U R H 2 + Y V U V 2  
  U C F D = G C I 2 + Y m o d e l 2
where U T , U R H , and U V denote the uncertainties associated with temperature, humidity, and airflow velocity measurements, respectively. Sensitivity coefficients ( Y / T , Y / R H , Y / V ) were calculated based on the experimental correlations obtained from the measured data. Through this uncertainty propagation analysis, the combined uncertainties for the temperature exchange efficiency, latent heat exchange efficiency, and enthalpy exchange efficiency were quantified and are shown in Table 5.

3. Results and Discussion

3.1. Grid Independence Study

To enhance simulation accuracy while optimizing computational resources, mesh independence verification was rigorously performed. The first-layer boundary mesh height was determined via Y+ theory (targeting 30 < Y+ < 300 [45] for standard wall functions), ensuring compatibility with the hybrid turbulence model. Theoretical calculations and simulation validation were integrated to achieve grid convergence, with detailed steps outlined below:
  Y + = u * Y υ
  u * = τ w ρ
τ w = C f · 1 2 ρ U 2
In the equation, u * denotes the friction velocity near the wall, Y represents the distance from the first layer of mesh nodes to the wall, υ is the kinematic viscosity of the fluid, τ w is the wall shear stress, and C f is the wall friction coefficient. The friction velocity u * is dimensionless, and the value of Y calculated in this study is 0.1 mm, which is incorporated into the grid division calculation. To comply with the standard wall function requirements (30 < y+ < 300), we iteratively determined Y = 0.1 mm based on the target y + ≈50, which was subsequently validated through mesh sensitivity analysis.
This study conducts nine sets of mesh independence verification on the equivalent model of the test sample at an air velocity of v = 1.1 m/s, evaluating their temperature efficiency ε s , Nusselt number (Nu), pressure drop ΔP, and corresponding errors to determine the optimal number of grid cells. As shown in Figure 5, it can be observed that as the number of grids increases, the efficiency value gradually increases until it stabilizes at approximately 71.6%. In all sections of Figure 5, the optimal grid number is marked with an asterisk. The data curve indicates that after optimizing the grid number, the percentage change in effectiveness no longer fluctuates significantly. When the grid count reaches 40,000, the temperature exchange efficiency ε s is 71.63%, the Nu is 5.03, and the pressure drop ΔP is 10.3 Pa. Further increasing the grid size results in negligible changes in temperature exchange efficiency (0.00%), Nu (0%), and pressure drop (0.09%). This indicates that the grid-independent convergence is less than 1.8% and confirms that a grid count of 40,000 is the optimal solution among the nine sets of grid counts.

3.2. Model Optimization

After simplifying the experimental sample into an equivalent model, Figure 6a illustrates the Fluent mesh for the MEE. Figure 6b,c depict the cross-sectional view of the model at x = 133 mm on the y-z plane when the fluid flow velocity is 1.1 m/s, where a laminar flow pattern is evident. This observation confirms the previously mentioned fact that the fluid domain exhibits a fully developed laminar flow. For enhancing single-phase convective heat transfer, this study employs turbulence flow to improve heat and mass transfer [50,51]. Therefore, it is necessary to disrupt the original laminar flow structure and induce turbulence, generating a Kármán vortex street flow, to enhance the heat and mass transfer behavior in the MEE. The standard k-ε turbulence model was selected for its superior accuracy in predicting vortex street shear flows (experimental deviation <1.8%), while grid convergence analysis (GCI < 1.8%) and global sensitivity studies (d = 0.2–0.4 mm) identified the optimal vortex generator diameter as d = 0.3 mm, yielding a Strouhal number (St ≈ 0.21) consistent with classical Kármán vortex theory. According to the theoretical basis for the generation of Kármán vortex streets, when Re < 40, the flow is stable and does not form periodic vortices. When 40 < Re < 30, regular Kármán vortex street formations appear, with vortices alternating in a regular pattern. When 300 < Re < 10,000, the vortex street persists but becomes more complex and irregular in arrangement and shape. Consequently, this study primarily sets up evenly spaced cylindrical disturbance structures at the symmetrical plane of the equivalent model. Based on the experimental test conditions in this study, the required cylinder dimensions to generate Kármán vortex street flow at different flow velocities are calculated, as shown in Table 6. As the fluid velocity increases, the minimum diameter of the Kármán vortex street generator required to produce vortices also increases. However, as the cylinder diameter increases, the pressure difference ΔP in the fluid domain also rises. This creates a trade-off when evaluating the comprehensive performance of the MEE. We need to balance the size of the Kármán vortex street generator. Therefore, we selected a design cylinder diameter of D = 0.3 mm, which increases the resistance interface of the fluid domain by 12%.
For the digital model, the fluid layers consist of two layers, and the mesh is composed of tetrahedral elements. In the Fluent model, the energy equation is enabled. Component transport mode is activated, with component conditions specified in the boundary conditions, and mass transfer is calculated using UDF. The solution method employs the SIMPLE algorithm, with pressure solved using the Second-Order scheme. The momentum equation, H2O transport equation, and energy equation are solved using the Third-Order MUSCL scheme.
For the optimized model, we primarily calculated the predicted results for the inlet boundary conditions at air velocities of 1.1 m/s, 1.47 m/s, 1.83 m/s, 2.2 m/s, and 2.57 m/s. Representative results with significant changes at air velocities of 1.1 m/s, 1.83 m/s, and 2.57 m/s were selected for the analysis. Figure 7 shows the lift coefficient spectrum of the first Kármán vortex street generator structure at the inlet of the hot fluid domain flow field. Figure 8 presents the flow field contour plot of the first Kármán vortex street generator region in the hot fluid domain. Figure 9 illustrates the local flow field velocity contour plot in the hot fluid domain. When the fluid flow speed is 1.1 m/s, no periodic fluctuation occurs on the cylinder surface. As the iteration steps increase, the lift coefficient remains stable (as shown in Figure 7a), indicating that no Kármán vortex-induced turbulence is formed. The flow field contour plot and velocity cloud plot in the cylinder’s local region also confirm this conclusion. When the fluid flow speed increases to 1.83 m/s, the flow field airflow starts to show fluctuations, but after the flow channel fully develops, it behaves similarly to the situation at 1.1 m/s. In the low-speed flow field with the added cylindrical disturbance structure, no Kármán vortex street-like turbulent structure forms because the Re on the cylinder surface has not reached the critical value required for vortex formation, which requires Re > 40. At flow velocities of 1.1 m/s and 1.83 m/s, the Re on the cylinder surface denotes 20 and 27, respectively. However, it can be observed that as the fluid velocity increases, the velocity gradient rate in the wall membrane area in the contour plot increases. When the fluid flow speed reaches 2.57 m/s, the Reynolds number based on the channel height (2.5 mm) is 392. This places the flow in the subcritical regime (40 < Re < 3 × 105) according to Williamson’s vortex classification [25]. The oscillatory lift coefficient in Figure 7c and periodic vortical structures in Figure 8c and Figure 9c confirm the emergence of an 2S vortex street, characterized by single-row alternating vortices shedding from the cylinder. The Strouhal number (Sr = 0.21) aligns with classical 2S dynamics in this Re range [25], further validating the observed flow patterns.
  f = S r ν D
When Re < 3 × 105, Sr can be approximated as 0.21. V is the fluid velocity (m/s), and D is the characteristic length of the flow obstruction (m). Therefore, the theoretical vortex shedding period is calculated to be 5.56 × 10−4 s, which has an error of 3.60% compared to the simulated vortex shedding period of 5.76 × 10−4 s. This closely matches the actual predicted result. In Figure 8c and Figure 9c, periodic turbulent flow can be observed within the fluid domain. The 2S vortex street generates coherent vortical structures that periodically destabilize the thermal and concentration boundary layers, amplifying momentum and mass transfer through advective mixing, a mechanism critical to the observed heat and mass transfer enhancement. The direction of this turbulent airflow group, when the angle θ between the movement direction of the airflow and the vector direction of the inlet fluid is not 90°, results in a significant velocity gradient near the wall. This indicates that the fluid is being compressed and moves towards the wall, concentrating heat towards the wall and ultimately enhancing the heat and mass transfer processes. Future studies could explore higher Re regimes (e.g., 2P or P + S vortex modes) to assess their potential for further performance optimization [25]. Next, the performance evaluation of the MEE device will be analyzed based on the predicted results.

3.3. Comparison Analysis of Different Model Configurations with Experimental Results

Through the analysis presented above, it is evident that the flow field of the optimized fluid within the model’s fluid domain undergoes a transformation at a specific velocity, resulting in the formation of a Kármán vortex street airflow pattern. Figure 10 illustrates the heat flux, Nu, and Sherwood number (Sh) profiles of the model’s membrane wall under varying fluid flow rates.
As illustrated in Figure 10, both the equivalent and optimized models show that the heat flux of the membrane wall increases with rising fluid flow velocity. However, when the flow velocity reaches 11.83 m/s, a significant increase in the average Nu is observed under both winter and summer conditions, deviating from the previously established linear trend. At a flow velocity of 2.57 m/s, the optimized model exhibits an average Nu of 7.8 and an average Sh of 8.3 in summer, while in winter, these values are 6.7 and 6.3, respectively. Compared to the equivalent model, the flow characteristics improve by 54.35% and 30.49%, respectively. The greater enhancement observed in summer is attributed to two key mechanisms: (1) higher ambient humidity amplifies water vapor diffusion flux ( J = D C , Fick’s law), significantly boosting latent heat transfer; (2) membrane hygroscopicity under humid conditions increases surface adsorption capacity, temporarily enhancing vapor permeability despite long-term swelling risks. In contrast, winter conditions prioritize sensible heat transfer due to limited vapor concentration gradients, while membrane permeability is optimized to suppress condensation without sacrificing mass transfer efficiency.
The performance efficiencies of the MEE under winter and summer conditions are systematically compared in Figure 11. The experimental results show strong agreement with the turbulence-enhanced model predictions, with expanded uncertainties of ±1.5–1.8% (summer) and ±1.7–2.0% (winter), as quantified in Table 5. Notably, at low flow velocities (<1.5 m/s), the optimized model (with cylindrical vortex generators) aligns closely with the baseline model (dashed lines) due to the absence of coherent Kármán vortex shedding, resulting in similar heat–mass transfer dynamics. As the flow velocity increases to 2.57 m/s, the turbulence-enhanced model demonstrates a 3.37% (summer) and 3.36 (winter) improvement in enthalpy efficiency compared to the baseline model. This enhancement is attributed to periodic vortex-induced boundary layer renewal, which disrupts the laminar concentration gradient and amplifies latent heat recovery. Seasonal variations further highlight the model’s robustness: winter conditions exhibit marginally higher uncertainties (±2.0%) than summer (±1.8%), consistent with the viscosity-driven flow resistance effects observed in Section 3.2. To address practical HVAC velocity variations, Figure 11 indicates that the enthalpy efficiency gain decreases to 2.1% at 1.83 m/s (Re ≈ 280), where vortex shedding is incomplete. Beyond the tested range, preliminary extrapolation of the efficiency curve suggests a gradual decline in enhancement (e.g., ~3.2% at 3.0 m/s, Re ≈ 450), accompanied by a potential pressure loss increase of ~15%. These observations imply that Re = 392 represents a region of optimal balance between efficiency gains and flow resistance, though further validation is needed for extended operational ranges. It is noteworthy that this study focuses on extreme seasonal conditions (peak summer/winter) to establish baseline performance boundaries for the membrane-based enthalpy exchanger (MEE). However, the theoretical framework—governed by Fickian diffusion dynamics and humidity-dependent membrane hygroscopicity—can be extended to predict transitional seasons (e.g., spring/fall) or other climatic zones by dynamically adjusting vapor concentration gradients (∇C) and permeability coefficients. This adaptability highlights the model’s generalizability and provides a foundational research direction for future studies, such as climate-specific optimization strategies or adaptive membrane designs under varying environmental conditions, ensuring broader applicability beyond the current extreme-case scenarios.
The error bars in Figure 11c explicitly differentiate experimental variability (purple) from numerical uncertainties (gray bands for CFD), ensuring transparent validation of the coupled thermal–hygric transport mechanisms.

3.4. Loss and Comprehensive Performance Evaluation

When a turbulent flow structure is introduced into the channel, both the ΔP and the Cf of the fluid flowing through this region increase. The Cf is defined as follows [52]:
C f = 2 P D h ρ L V a i r 2
Here, ρ, Dh, L, and Vair represent air density (kg/m3), hydraulic diameter, channel length, and air velocity, respectively. The hydraulic diameter Dh is defined as the ratio of the cross-sectional area of the fluid domain to its wetted perimeter, which can be expressed as follows [52]:
D h = 4 V c y c A c y c
where Vcyc denotes the volume of the target fluid domain and Acyc signifies the effective contact area between the solid domain and the fluid domain. For this study, the hydraulic diameter Dh is 2.5 × 10−3 m.
Figure 12 illustrates the curves of ΔP and Cf. The incorporation of the Kármán vortex generator results in a 10% to 40% increase in ΔP and a 29% to 55% increase in Cf. Specifically, at a flow rate of 2.57 m/s (350 m3/h), ΔP increases by 36%, while Cf rises by 41%. Although the addition of a Kármán vortex generator can enhance heat and mass transfer by inducing forced turbulent airflow, it also leads to higher friction coefficients and increased pressure loss. To evaluate the net energy benefit, a simplified energy balance was performed. The electrical power penalty from added pressure loss is calculated as follows:
P f a n = P Q η f a n
  P r e c o v e r e d = η r e c o v e r e d ρ Q c p T  
For Q = 350 m3/h, ηfan = 60%, and ρ = 1.2 kg/m3, the analysis shows contrasting seasonal outcomes: summer conditions (ηE = 3.37%, ΔT = 8 K) yield a net energy loss of 2.7 W, while winter conditions (ηE = 3.36%, ΔT = 19 K) achieve a net gain of 11.0 W. This divergence highlights that summer’s negative net recovery and winter’s positive net recovery necessitate climate-specific engineering considerations for practical HVAC integration.
Therefore, optimizing the heat and mass transfer performance of MEE devices through structural modifications necessitates a comprehensive evaluation of potential performance trade-offs. This paper will assess the optimization value through an in-depth performance evaluation. To fully evaluate the flow and heat transfer performance of the MEE, the product (fRe) of the drag coefficient Cf and the Re is considered [53]. As shown in Figure 12c, the addition of the Kármán vortex generator increases the fRe value from 32 to 41, with the maximum value of 41 occurring at a Re of 388. This indicates a significant change in the fluid flow characteristics. Future designs could mitigate ΔP penalties through aerodynamic refinements (e.g., streamlined vortex generators) while retaining efficiency gains, a critical step toward practical HVAC integration.
To comprehensively evaluate the flow and heat transfer performance of the MEE, the performance evaluation criterion (PEC) [54] is commonly utilized as follows:
P E C = N u N u 0 f f 0 1 3
Here, the subscript 0 denotes the initial state as represented by the equivalent model. Nu/Nu0 and Sh/Sh0 represent the relative Nu and the relative friction factor, respectively. For MEE devices, if the PEC ratio exceeds 1, this indicates that the enhancement in heat transfer performance outweighs the increase in resistance. Conversely, a PEC ratio less than or equal to 1 suggests a disadvantageous trade-off. In this context, PECh and PECm denote the evaluation indices for heat and mass transfer, respectively.
As illustrated in Figure 13, the ratios Nu/Nu0 and Sh/Sh0 initially exhibit a gradual increase, followed by a more pronounced rise after the inflection point at a Re of 276. Notably, the PEC values remain below 1 up to this point, indicating suboptimal overall device performance. However, when a Re reaches 332, both PECh and PECm exceed 1, signifying a significant enhancement in the device’s overall performance. The optimal performance occurs at a Re of 388, where PECh and PECm achieve their maximum values of 1.33 and 1.22, respectively. Furthermore, the summer operating conditions demonstrate superior performance compared to the winter conditions, particularly in terms of thermal efficiency, as evidenced by the curves presented in Figure 13a,b.

4. Conclusions

A three-dimensional computational fluid dynamics (CFD) model was established to simulate coupled heat and mass transfer in membrane-based enthalpy exchangers (MEEs). Through theoretical optimization integrating cylindrical vortex generators, periodic Kármán vortex shedding was achieved, enhancing thermal–hygric transport efficiency. Systematic analysis of heat flux, Nusselt number (Nu), Sherwood number (Sh), pressure drop (ΔP), friction coefficient (Cf), and performance evaluation criterion (PEC) yielded the following key findings:
  • Model Validation
The CFD results align with experimental data within a 1% discrepancy, confirming the model’s reliability in predicting MEE performance under both steady and vortex-enhanced conditions.
2.
Vortex-Enhanced Efficiency
Cylindrical spoilers induced effective Kármán vortices at 2.57 m/s flow velocity, improving temperature, latent heat, and enthalpy exchange efficiencies by 3.37–3.91% compared to the baseline. Seasonal energy balance revealed a net loss of 2.7 W under summer conditions (ΔT = 8 °C) versus a net gain of 11.0 W in winter (ΔT = 19 °C), necessitating climate-specific deployment strategies.
3.
Energy Penalty Trade-offs
Vortex generation increased ΔP by 36% and Cf by 41% at 2.57 m/s (350 m3/h). While PEC peaked at 1.33 (heat) and 1.22 (mass) when Re = 388, efficiency gains diminished outside the optimal Re range (300–400), constrained by incomplete vortex shedding (Re < 270) and turbulence dissipation (Re > 450).
4.
Practical Implications and Limitations
The winter-biased energy recovery supports prioritized adoption in heating-dominated climates. Scalability to industrial ERV cores may leverage modular multi-cell arrays with geometry ratios L/D = 5–30, though transient sorption hysteresis and manufacturing costs require further analysis. Current limitations include steady-state assumptions and single-row vortex testing; future work will explore multi-row staggered configurations and techno-economic lifecycle assessments.

Author Contributions

Conceptualization, Y.L. and S.-S.L.; Methodology, Y.L.; Formal Analysis, D.-C.M. and Y.L.; Investigation, Y.L.; Data Curation, Y.L.; Writing—Original Draft Preparation, Y.L.; Writing—Review and Editing, D.-C.M.; Visualization, D.-C.M.; Supervision, S.-S.L.; Project Administration, S.-S.L.; Funding Acquisition, S.-S.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China, Grant No. 52276095 (S.S.-L).

Data Availability Statement

The data that support the plots within this paper and other findings of this study are available from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no competing interests.

Abbreviations

NomenclatureGreek Symbols
ABSAcrylonitrile–butadiene–styrene ρ Density of air (kg/m3)
CFDComputational fluid dynamics τ Times
EAExhaust airΓϕDiffusion coefficient of variable quantity ϕ
MEEMembrane enthalpy exchangersϕVariable quantity (1, U, V, W, T, Yv)
NuNusselt numberCpSpecific heat of dry air (J/kg K)
NTUsNumber of Transfer UnitsµViscosity (Pa·s)
OAOutdoor airωHumidity ratio (kg/kg)
PEPolyethyleneΔPPressure drop (Pa)
PECPerformance evaluation criterion m ˙ Mass flux of air (kg/s)
RAReturn airδThickness (μm)
ReReynolds numberεEffectiveness (%), turbulent kinetic energy dissipation rate (m2/s3)
SASupply airθMoisture uptake in the membrane (kg/kg membrane)
SIPMLESemi-Implicit Method for Pressure-Linked EquationsLCharacteristic length (m)
SrStrouhal numbernNumber of channels
ScSchmidt number V Volumetric flow rate (m3/h)
UDFUser-defined functionsVVelocity (m/s)
Subscripts Y i Mass fraction of each species (kg/kg)
aAir J i Diffusion flux of each species (kg/m2)
eExhaustESensible enthalpy (J)
cColdPAir pressure (Pa)
hHot/heat k Heat conductivity (W/(m·K))
iInlet, species index (dry air or vapor) T Moist air temperature (K)
LLatent h i Convective heat transfer coefficient of each species
sSteam, supply of air D i Diffusion coefficient of each species (m2/s)
tTotalWmaxMaximum water uptake in the membrane (kg/kg membrane)
vVaporCConstant in sorption curve (−)
mMassDtTurbulent diffusion coefficient
maMix airY+The thickness of the first grid
csCold stream u * Near-wall friction coefficient
hsHot streamYDistance between the first layer grid node and the wall surface
mhMembrane and hot stream τ w Wall shear stress
mcMembrane and cold streamCfWall-friction coefficient
meMass exhaustμtTurbulent viscosity
msMass supplyGkTurbulence kinetic energy from velocity gradients
wmWater in the membraneGbTurbulence kinetic energy from buoyancy

References

  1. Albdoor, A.K.; Ma, Z.; Al-Ghazzawi, F.; Arıcı, M. Study on recent progress and advances in air-to-air membrane enthalpy exchangers: Materials selection, performance improvement, design optimisation and effects of operating conditions. Renew. Sustain. Energy Rev. 2022, 156, 111941. [Google Scholar] [CrossRef]
  2. Li, N.; Yan, X. Numerical evaluation of the effects of asymmetric membrane heat conductivity and moisture adsorption hysteresis on the performance of air-to-air membrane enthalpy exchanger. Therm. Sci. Eng. Prog. 2022, 28, 101058. [Google Scholar] [CrossRef]
  3. Engarnevis, A.; Huizing, R.; Green, S.; Rogak, S. Heat and mass transfer modeling in enthalpy exchangers using asymmetric composite membranes. J. Membr. Sci. 2018, 556, 248–262. [Google Scholar] [CrossRef]
  4. Malka, L.; Bidaj, F.; Kuriqi, A.; Jaku, A.; Roçi, R.; Gebremedhin, A. Energy system analysis with a focus on future energy demand projections: The case of Norway. Energy 2023, 272, 127107. [Google Scholar] [CrossRef]
  5. Xu, T.; Zhang, Y.; Shi, L.; Feng, Y.; Ke, X.; Zhang, C. A comprehensive evaluation framework of energy and resources consumption of public buildings: Case study, People’s Bank of China. Appl. Energy 2023, 351, 121869. [Google Scholar] [CrossRef]
  6. Zhang, X.; Sun, J.; Zhang, X.; Wang, F. Assessment and regression of carbon emissions from the building and construction sector in China: A provincial study using machine learning. J. Clean. Prod. 2024, 450, 141903. [Google Scholar] [CrossRef]
  7. Al-Waked, R.; Nasif, M.S.; Morrison, G.; Behnia, M. CFD simulation of air to air enthalpy heat exchanger. Energy Convers. Manag. 2013, 74, 377–385. [Google Scholar] [CrossRef]
  8. Al-Waked, R.; Nasif, M.S.; Morrison, G.; Behnia, M. CFD simulation of air to air enthalpy heat exchanger: Variable membrane moisture resistance. Appl. Therm. Eng. 2015, 84, 301–309. [Google Scholar] [CrossRef]
  9. Koester, S.; Falkenberg, M.; Logemann, M.; Wessling, M. Modeling heat and mass transfer in cross-counterflow enthalpy exchangers. J. Membr. Sci. 2017, 525, 68–76. [Google Scholar] [CrossRef]
  10. Min, J.; Duan, J. Membrane-type total heat exchanger performance with heat and moisture transferring in different directions across membranes. Appl. Therm. Eng. 2015, 91, 1040–1047. [Google Scholar] [CrossRef]
  11. Pourhoseinian, M.; Asasian-Kolur, N.; Sharifian, S. CFD investigation of heat and moisture recovery from air with membrane heat exchanger. Appl. Therm. Eng. 2021, 191, 116911. [Google Scholar] [CrossRef]
  12. Yaïci, W.; Ghorab, M.; Entchev, E. Numerical analysis of heat and energy recovery ventilators performance based on CFD for detailed design. Appl. Therm. Eng. 2013, 51, 770–780. [Google Scholar] [CrossRef]
  13. Zhang, L.-Z.; Liang, C.-H.; Pei, L.-X. Heat and moisture transfer in application scale parallel-plates enthalpy exchangers with novel membrane materials. J. Membr. Sci. 2008, 325, 672–682. [Google Scholar] [CrossRef]
  14. Qi, D.; He, J.; Xu, Y.; Lin, M.; Wang, Q. Effect of rib diameter on flow boiling heat transfer with staggered rib arrays in a heat sink. Energy 2022, 239, 122323. [Google Scholar] [CrossRef]
  15. Li, Q.-w.; Shang, X.-s.; Cao, Q.; Cui, Z.; Shao, W. Numerical simulation on the heat sink with interrupted microchannels regarding of heat transfer enhancement. Heat Mass Transf. 2024, 60, 1195–1209. [Google Scholar] [CrossRef]
  16. Li, Y.; Wang, T.; Wang, Z.; Yang, J.; Cao, W.; Ma, X. Thermohydraulic performance analysis and parameters optimization of the combined heat sinks with microchannels and micro pin-fins. Appl. Therm. Eng. 2024, 241, 122443. [Google Scholar] [CrossRef]
  17. Bhandari, P.; Prajapati, Y.K. Thermal performance of open microchannel heat sink with variable pin fin height. Int. J. Therm. Sci. 2021, 159, 106609. [Google Scholar] [CrossRef]
  18. Zhang, L.-Z. A reliability-based optimization of membrane-type total heat exchangers under uncertain design parameters. Energy 2016, 101, 390–401. [Google Scholar] [CrossRef]
  19. Deshko, V.I.; Karvatskii, A.Y.; Sukhodub, I.O. Heat and mass transfer in cross-flow air-to-air membrane heat exchanger in heating mode. Appl. Therm. Eng. 2016, 100, 133–145. [Google Scholar] [CrossRef]
  20. Asasian-Kolur, N.; Sharifian, S.; Haddadi, B.; Pourhoseinian, M.; Mousazadeh Shekarbaghani, Z.; Harasek, M. Membrane-based enthalpy exchangers for coincident sensible and latent heat recovery. Energy Convers. Manag. 2022, 253, 115144. [Google Scholar] [CrossRef]
  21. Hua, J.; Li, G.; Zhao, X.; Li, Q.; Hu, J. Study on the flow resistance performance of fluid cross various shapes of micro-scale pin fin. Appl. Therm. Eng. 2016, 107, 768–775. [Google Scholar] [CrossRef]
  22. Liu, M.; Liu, D.; Xu, S.; Chen, Y. Experimental study on liquid flow and heat transfer in micro square pin fin heat sink. Int. J. Heat Mass Transf. 2011, 54, 5602–5611. [Google Scholar] [CrossRef]
  23. Erdem, K.; Ahmadi, V.E.; Kosar, A.; Kuddusi, L. Differential Sorting of Microparticles Using Spiral Microchannels with Elliptic Configurations. Micromachines 2020, 11, 412. [Google Scholar] [CrossRef]
  24. Matin Nikoo, H.; Bi, K.; Hao, H. Three-dimensional vortex-induced vibration of a circular cylinder at subcritical Reynolds numbers with low-Re correction. Mar. Struct. 2019, 66, 288–306. [Google Scholar] [CrossRef]
  25. Williamson, C.H.K. Three-dimensional vortex dynamics in bluff body wakes. Exp. Therm. Fluid Sci. 1996, 12, 150–168. [Google Scholar] [CrossRef]
  26. Toupoint, C.; Ern, P.; Roig, V. Kinematics and wake of freely falling cylinders at moderate Reynolds numbers. J. Fluid Mech. 2019, 866, 82–111. [Google Scholar] [CrossRef]
  27. Schlichting, H.; Yaʻaḳov, A.; Leyb ben Dan, Z. Boundary-Layer Theory, 6th ed.; McGraw-Hill: New York, NY, USA; Toronto, ON, Canada; London, UK; Sydney, Australia, 1968. [Google Scholar]
  28. Min, J.; Su, M. Performance analysis of a membrane-based enthalpy exchanger: Effects of the membrane properties on the exchanger performance. J. Membr. Sci. 2010, 348, 376–382. [Google Scholar] [CrossRef]
  29. Min, J.; Su, M. Performance analysis of a membrane-based energy recovery ventilator: Effects of membrane spacing and thickness on the ventilator performance. Appl. Therm. Eng. 2010, 30, 991–997. [Google Scholar] [CrossRef]
  30. Nasif, M.S.; Al-Waked, R.f.; Behnia, M.; Morrison, G. Modeling of Air to Air Enthalpy Heat Exchanger. Heat Transf. Eng. 2012, 33, 1010–1023. [Google Scholar] [CrossRef]
  31. Souifi, H.; Bouslimani, Y.; Ghribi, M.; Colin, S. Modeling of an air-to-air exchanger with dual-core in cascade connection. MethodsX 2021, 8, 101253. [Google Scholar] [CrossRef]
  32. Vali, A.; Ge, G.; Besant, R.W.; Simonson, C.J. Numerical modeling of fluid flow and coupled heat and mass transfer in a counter-cross-flow parallel-plate liquid-to-air membrane energy exchanger. Int. J. Heat Mass Transf. 2015, 89, 1258–1276. [Google Scholar] [CrossRef]
  33. Wang, Y.; Zhu, K.; Cui, Z.; Wei, J. Effects of the location of the inlet and outlet on heat transfer performance in pin fin CPU heat sink. Appl. Therm. Eng. 2019, 151, 506–513. [Google Scholar] [CrossRef]
  34. Xie, J.; Yan, H.; Sundén, B.; Xie, G. The influences of sidewall proximity on flow and thermal performance of a microchannel with large-row pin-fins. Int. J. Therm. Sci. 2019, 140, 8–19. [Google Scholar] [CrossRef]
  35. Khalili Sadaghiani, A.; Koşar, A. Numerical investigations on the effect of fin shape and surface roughness on hydrothermal characteristics of slip flows in microchannels with pin fins. Int. J. Therm. Sci. 2018, 124, 375–386. [Google Scholar] [CrossRef]
  36. Koester, S.; Lölsberg, J.; Lutz, L.; Marten, D.; Wessling, M. On individual resistances of selective skin, porous support and diffusion boundary layer in water vapor permeation. J. Membr. Sci. 2016, 507, 179–187. [Google Scholar] [CrossRef]
  37. Zou, Y.; Maillet, B.; Coussot, P.; Brochard, L. Sorption hysteresis in nanoporous swelling adsorbents. Phys. Rev. Res. 2025, 7, 013205. [Google Scholar] [CrossRef]
  38. Calzaferri, G.; Gallagher, S.H.; Lustenberger, S.; Walther, F.; Brühwiler, D. Multiple equilibria description of type H1 hysteresis in gas sorption isotherms of mesoporous materials. Mater. Chem. Phys. 2023, 296, 127121. [Google Scholar] [CrossRef]
  39. Min, J.; Wang, L. Coupled heat and mass transfer during moisture exchange across a membrane. J. Membr. Sci. 2013, 430, 150–157. [Google Scholar] [CrossRef]
  40. Fluent Incorporated. FLUENT User’s Guide; Version 6.3; Fluent Incorporated: Lebanon, NH, USA, 2006. [Google Scholar]
  41. Li, H.; Zhang, T.; Zhang, J.; Guan, B.; Liu, X.; Nakazawa, T.; Fang, L.; Tanaka, T. Investigation of energy recovery performance and frost risk of membrane enthalpy exchanger applied in cold climates. Energy 2023, 282, 128894. [Google Scholar] [CrossRef]
  42. Gherasim, I.; Galanis, N.; Nguyen, C.T. Heat transfer and fluid flow in a plate heat exchanger. Part II: Assessment of laminar and two-equation turbulent models. Int. J. Therm. Sci. 2011, 50, 1499–1511. [Google Scholar] [CrossRef]
  43. Launder, B.E.; Spalding, D.B. The numerical computation of turbulent flows. Comput. Methods Appl. Mech. Eng. 1974, 3, 269–289. [Google Scholar] [CrossRef]
  44. Wang, J.; Weng, S.G.; Wu, S.S.; Hu, X.D.; Yang, X. The visualization of flow field around circular cylinders by fluent standard k–ε turbulence model. IOP Conf. Ser. Earth Environ. Sci. 2019, 344, 012134. [Google Scholar] [CrossRef]
  45. Menter, F.R. Two-equation eddy-viscosity turbulence models for engineering applications. AIAA J. 1994, 32, 1598–1605. [Google Scholar] [CrossRef]
  46. Cengel, Y.A. Heat and Mass Transfer: A Practical Approach, 3rd ed.; McGraw-Hill: Singapore, 2006. [Google Scholar]
  47. Zhang, L.Z.; Jiang, Y. Heat and mass transfer in a membrane-based energy recovery ventilator. J. Membr. Sci. 1999, 163, 29–38. [Google Scholar] [CrossRef]
  48. GB/T 21087-2020; Energy Recovery Ventilators for Outdoor Air Handling. Standardization Administration of China: Beijing, China, 2020.
  49. Moffat, R.J. Describing the uncertainties in experimental results. Exp. Therm. Fluid Sci. 1988, 1, 3–17. [Google Scholar] [CrossRef]
  50. Guo, Z.Y.; Li, D.Y.; Wang, B.X. A novel concept for convective heat transfer enhancement. Int. J. Heat Mass Transf. 1998, 41, 2221–2225. [Google Scholar] [CrossRef]
  51. Tao, W.-Q.; Guo, Z.-Y.; Wang, B.-X. Field synergy principle for enhancing convective heat transfer––its extension and numerical verifications. Int. J. Heat Mass Transf. 2002, 45, 3849–3856. [Google Scholar] [CrossRef]
  52. Zhang, L.-Z. Convective mass transport in cross-corrugated membrane exchangers. J. Membr. Sci. 2005, 260, 75–83. [Google Scholar] [CrossRef]
  53. Zhu, Q.; Su, R.; Xia, H.; Zeng, J.; Chen, J. Numerical simulation study of thermal and hydraulic characteristics of laminar flow in microchannel heat sink with water droplet cavities and different rib columns. Int. J. Therm. Sci. 2022, 172, 107319. [Google Scholar] [CrossRef]
  54. Webb, R.L. Performance evaluation criteria for use of enhanced heat transfer surfaces in heat exchanger design. Int. J. Heat Mass Transf. 1981, 24, 715–726. [Google Scholar] [CrossRef]
Figure 1. Three-dimensional schematic of MEE and the CFD computational domain: (a) external structure and airflow paths, (b) internal airflow layout with vortex-generating cylinders, and (c) computational domain with boundary and symmetry definition.
Figure 1. Three-dimensional schematic of MEE and the CFD computational domain: (a) external structure and airflow paths, (b) internal airflow layout with vortex-generating cylinders, and (c) computational domain with boundary and symmetry definition.
Energies 18 02892 g001
Figure 2. Cross-sectional morphology and elemental distribution characteristics of the membrane material. (a) Cross-sectional SEM (SU5000: 5.0 kV, WD 6.5 mm, SE(L), ×2k), (b) Functional layer interface (×7995, 5.0 keV, WD 6.0 mm), (ce) EDS elemental maps of C, O, N with corresponding atomic percentages (×7995, 5.0 keV, WD 6.0 mm).
Figure 2. Cross-sectional morphology and elemental distribution characteristics of the membrane material. (a) Cross-sectional SEM (SU5000: 5.0 kV, WD 6.5 mm, SE(L), ×2k), (b) Functional layer interface (×7995, 5.0 keV, WD 6.0 mm), (ce) EDS elemental maps of C, O, N with corresponding atomic percentages (×7995, 5.0 keV, WD 6.0 mm).
Energies 18 02892 g002
Figure 3. Schematic representation of heat and moisture in the stagnant boundary layers and the membrane of MEE.
Figure 3. Schematic representation of heat and moisture in the stagnant boundary layers and the membrane of MEE.
Energies 18 02892 g003
Figure 4. The test conditions. (a) Enthalpy difference laboratory; (b) schematic layout of the experimental apparatus.
Figure 4. The test conditions. (a) Enthalpy difference laboratory; (b) schematic layout of the experimental apparatus.
Energies 18 02892 g004
Figure 5. Grid independence test for the equivalent model at v = 1.1 m/s. (a) Sensible effectiveness; (b) average Nu; (c) pressure drop. The red star (★) denotes the optimal grid configuration selected for subsequent simulations.
Figure 5. Grid independence test for the equivalent model at v = 1.1 m/s. (a) Sensible effectiveness; (b) average Nu; (c) pressure drop. The red star (★) denotes the optimal grid configuration selected for subsequent simulations.
Energies 18 02892 g005
Figure 6. Computational model and results. (a) Simplified the ANSYS Fluent model of MEE; (b) velocity contour of one flow channel without any additional structure; (c)-1 the temperature contour of symmetry face in hot fluid; (c)-2 temperature profile in the direction of the hot fluid; (c)-3 vector profile in the direction of the hot fluid.
Figure 6. Computational model and results. (a) Simplified the ANSYS Fluent model of MEE; (b) velocity contour of one flow channel without any additional structure; (c)-1 the temperature contour of symmetry face in hot fluid; (c)-2 temperature profile in the direction of the hot fluid; (c)-3 vector profile in the direction of the hot fluid.
Energies 18 02892 g006
Figure 7. Optimized contours of fluid velocity inside the latter channel. (a) 1.1 m/s; (b) 1.83 m/s; (c) 2.57 m/s.
Figure 7. Optimized contours of fluid velocity inside the latter channel. (a) 1.1 m/s; (b) 1.83 m/s; (c) 2.57 m/s.
Energies 18 02892 g007
Figure 8. Iso-velocity contour and vector distribution plot of the hot flow field in the first Kármán vortex street generator region of the hot fluid domain. (a) 1.1 m/s; (b) 1.83 m/s; (c) 2.57 m/s.
Figure 8. Iso-velocity contour and vector distribution plot of the hot flow field in the first Kármán vortex street generator region of the hot fluid domain. (a) 1.1 m/s; (b) 1.83 m/s; (c) 2.57 m/s.
Energies 18 02892 g008
Figure 9. Local flow field velocity contour plot in the hot fluid domain. (a) 1.1 m/s; (b) 1.83 m/s; (c) 2.57 m/s.
Figure 9. Local flow field velocity contour plot in the hot fluid domain. (a) 1.1 m/s; (b) 1.83 m/s; (c) 2.57 m/s.
Energies 18 02892 g009
Figure 10. Under varying fluid flow rates. (a) The heat flux, (b) Nu, and (c) Sh.
Figure 10. Under varying fluid flow rates. (a) The heat flux, (b) Nu, and (c) Sh.
Energies 18 02892 g010
Figure 11. Exchanger effectiveness and error bars of winter (heating model) and summer (cooling model). (a) Sensible; (b) latent; (c) enthalpy.
Figure 11. Exchanger effectiveness and error bars of winter (heating model) and summer (cooling model). (a) Sensible; (b) latent; (c) enthalpy.
Energies 18 02892 g011
Figure 12. Assessment of performance metrics under varied flow velocities. (a) Pressure drop; (b) average friction coefficient; (c) fRe at different airflow velocities.
Figure 12. Assessment of performance metrics under varied flow velocities. (a) Pressure drop; (b) average friction coefficient; (c) fRe at different airflow velocities.
Energies 18 02892 g012
Figure 13. Assessment of performance metrics under varied Reynolds numbers: (a) PECh; (b) PECm.
Figure 13. Assessment of performance metrics under varied Reynolds numbers: (a) PECh; (b) PECm.
Energies 18 02892 g013
Table 1. The geometric dimensions of the MEE core.
Table 1. The geometric dimensions of the MEE core.
DimensionsXYH
Lt (mm, total dimensions)250250375
Lc (mm, clear dimensions)230230371
D (mm, membrane spacing)2.5
N (cell numbers)182
δm (mm, membrane thickness)0.026
Dwm (m2/s, water vapor diffusivity)2 × 10−6
Km (W/(m·K), thermal conductivity)0.3
Cp,m (J/(kg·K), membrane heat capacity)871
ρm (kg/m3, density)407.2
Table 2. The conservation equations of the MEE model.
Table 2. The conservation equations of the MEE model.
ItemConservation Equations
Supply AirExhaust AirMembrane
Energy m ˙ s n s y F c p s T s x + 2 h s T s T m s = 0 m ˙ e n e y F c p e T e x 2 h e T e T m e = 0 m ˙ w c p w T m Z λ m 2 T m x 2 λ m 2 T m y 2 λ m 2 T m Z 2 = 0
Mass m ˙ s n s y F ω s x + 2 h m s ω s ω m s = 0 m ˙ e n e y F ω e x 2 h m e ω e ω m e = 0 m ˙ w = D w m ρ m θ Z
Table 3. The boundary conditions of the MEE model.
Table 3. The boundary conditions of the MEE model.
ItemBoundary Conditions
Membrane Surface to Supply AirMembrane Surface to Exhaust Air
Energy λ m T m Z Z = 0 = h s T s T m s + m ˙ w L w λ m T m Z Z = δ = h e T e T m e + m ˙ w L w
Mass m ˙ w = D w m ρ m θ Z Z = 0 = h m s ω s ω m s m ˙ w = D w m ρ m θ Z Z = δ = h m e ω e ω m e
Table 4. The operating and simulation conditions of GB/T21087-2020.
Table 4. The operating and simulation conditions of GB/T21087-2020.
ParameterCooling Mode (Summer)Heating Mode (Winter)
Outdoor AirIndoor AirOutdoor AirIndoor Air
Dry-bulk temperature (°C)3527221
Wet-ball temperature (°C)2819.5113
Relative humidity (%)59.550.383.939.7
Enthalpy (kJ/kg)89.1455.1011.136.11
Table 5. Experimental and CFD uncertainties of sensible, latent, and enthalpy efficiencies in summer and winter conditions.
Table 5. Experimental and CFD uncertainties of sensible, latent, and enthalpy efficiencies in summer and winter conditions.
ConditionParameterExperimental Uncertainty (±%)CFD Model
Uncertainty (±%)
CFD Model with Turbulent Uncertainty (±%)
SummerSensible efficiency1.22.11.5
Latent efficiency1.62.51.8
Enthalpy efficiency1.42.31.6
WinterSensible efficiency1.32.01.7
Latent efficiency1.72.42.0
Enthalpy efficiency1.52.21.8
Table 6. Cylindrical diameter dimensions that can produce Kármán vortex flow at different air volumes for MEE.
Table 6. Cylindrical diameter dimensions that can produce Kármán vortex flow at different air volumes for MEE.
Velocity (m/s)ReThe Dimensions of the Cylinder Producing the Kármán Vortex Street (mm)
1.10 (150 m3/h)1680.69 (27.6%)
1.47 (200 m3/h)2240.52 (20.8%)
1.83 (250 m3/h)2800.42 (16.8%)
2.20 (300 m3/h)3360.35 (13.9%)
2.57 (350 m3/h)3920.30 (12%)
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Liu, Y.; Mo, D.-C.; Lyu, S.-S. Vortex-Induced Turbulence Optimized Membrane Enthalpy Exchanger: Dynamic Humidity Modulation and Coupled Heat–Mass Transfer Enhancement. Energies 2025, 18, 2892. https://doi.org/10.3390/en18112892

AMA Style

Liu Y, Mo D-C, Lyu S-S. Vortex-Induced Turbulence Optimized Membrane Enthalpy Exchanger: Dynamic Humidity Modulation and Coupled Heat–Mass Transfer Enhancement. Energies. 2025; 18(11):2892. https://doi.org/10.3390/en18112892

Chicago/Turabian Style

Liu, Yang, Dong-Chuan Mo, and Shu-Shen Lyu. 2025. "Vortex-Induced Turbulence Optimized Membrane Enthalpy Exchanger: Dynamic Humidity Modulation and Coupled Heat–Mass Transfer Enhancement" Energies 18, no. 11: 2892. https://doi.org/10.3390/en18112892

APA Style

Liu, Y., Mo, D.-C., & Lyu, S.-S. (2025). Vortex-Induced Turbulence Optimized Membrane Enthalpy Exchanger: Dynamic Humidity Modulation and Coupled Heat–Mass Transfer Enhancement. Energies, 18(11), 2892. https://doi.org/10.3390/en18112892

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