1. Introduction
Hydrogen energy (123 MJ/kg) enables carbon neutrality via renewable integration [
1,
2].Green hydrogen from alkaline electrolysis dominates industries due to cost and durability [
3,
4]. However, scaling AWE (Alkaline Water Electrolysis) systems to multi-megawatt capacities introduces critical hydrodynamic bottlenecks. Maldistributed electrolyte flow, exacerbated by geometric imperfections in bipolar plates, induces stagnation zones where gas bubbles accumulate, elevating ionic resistance and localizing current densities beyond optimal ranges [
5,
6]. These phenomena collectively degrade Faradaic efficiency and accelerate electrode corrosion, posing formidable barriers to gigawatt-scale commercialization [
7,
8]. To address these challenges, resolving flow inhomogeneity is imperative to align AWE’s operational performance with its theoretical potential.
To mitigate these hydrodynamic limitations, experimental research has progressively unraveled multiphase flow complexities in AWE systems. Early studies by Eigeldinger and Vogt [
9] established quantitative relationships between electrolyte velocity (0.1–1.0 m/s) and bubble detachment frequencies, while Riegel et al. [
10] correlated gas holdup (5–20%) with current density fluctuations (±15%) in vertical electrodes. Building on these foundations, modern diagnostic tools, such as laser-induced fluorescence (LIF) and synchronized high-speed imaging, now enable micron-scale resolution of transient gas–liquid interfaces [
11,
12]. For example, Hine and Murakami [
13] demonstrated that forced convection (Re > 2000) reduces bubble-induced voltage overpotentials by 18% compared to natural convection regimes. Despite these advancements, empirical design practices—often reliant on trial-and-error iterations—remain prevalent, resulting in suboptimal flow uniformity at industrial scales. Notably, current studies predominantly focus on laboratory-scale systems (electrode diameters < 0.5 m), leaving a critical gap in understanding flow optimization for electrode plates exceeding 2 m in diameter, where inertial forces and multiphase interactions dominate flow inhomogeneity. A critical gap lies in the inability of conventional experimental setups to replicate dynamic operating conditions (e.g., fluctuating renewable energy inputs), which amplify remixing and electrolyte stratification [
14,
15]. This limitation underscores the necessity of integrating computational approaches to bridge laboratory-to-industry knowledge gaps.
Building on the experimental challenges in scaling to industrial dimensions, computational fluid dynamics (CFD) has emerged as an indispensable tool for deconvoluting multiscale flow phenomena in large-scale AWE systems. The Euler–Euler framework, treating gas–liquid phases as interpenetrating continua, accurately predicts macroscopic velocity profiles (error < 8% against PIV data) and residence time distributions [
16,
17]. Conversely, the Euler–Lagrange approach resolves discrete bubble dynamics, enabling precise modeling of coalescence and mass transfer at gas-evolving electrodes [
18]. Hybrid methodologies such as Delayed Detached Eddy Simulation (DDES) integrate Reynolds-Averaged Navier–Stokes (RANS) and Large Eddy Simulation (LES), optimizing the trade-off between computational accuracy and cost. For instance, Ziegler and Evans [
19] demonstrated that DDES achieves 92% agreement with experimental stagnation zone geometries, validating its efficacy in industrial-scale simulations. Notably, Mat [
20] utilized DDES to optimize baffle designs, reducing dead zones by 22% while maintaining pressure drops below 5 kPa. However, persistent challenges include the prohibitive computational cost of resolving turbulence spectra in multiphase flows and the lack of universal correlations for diverse electrolyzer geometries [
12]. These unresolved issues highlight the need for experimentally validated, reduced-order models tailored to industrial AWE scaling.
To bridge these gaps, this study addresses the aforementioned challenges through a dual experimental–computational strategy targeting flow uniformity in industrial AWE systems. First, fluorescent tracer experiments quantify electrolyte retention dynamics across a broad operational range (240–600 L/h; 0.5–1.25 m/s inlet velocities), employing sodium fluorescein and UV-PIV to map stagnation zones and remixing intensities. Concurrently, high-fidelity CFD simulations deploy a hybrid DDES model—validated against experimental velocity fields to evaluate novel inlet restrictors and hierarchical baffle geometries (e.g., radial vs. helical deflectors). Key metrics include velocity inhomogeneity indices, defined as the standard deviation of Z-direction velocities normalized by the mean, and residence time distribution (RTD) variances, calculated via discretized Lagrangian particle tracking. The resultant framework not only advances fundamental understanding of AWE hydrodynamics but also delivers actionable guidelines for minimizing energy losses in next-generation hydrogen infrastructures.
2. Equipment and Methods
2.1. Experimental Equipment
To enhance the analysis and visualization of fluid flow patterns within a large bipolar plate, we have established a testing system to investigate flow behavior in a Large Alkaline Water Electrolyzer. This system utilizes plate data provided through collaboration with an external research institute (see
Figure 1).
Figure 1 shows the 3D modeling structure and
Figure 1b shows the real experimental setup. The key component of this system is a full-scale, 1:1 replica of an existing Alkaline Water Electrolyzer plate. The experimental setup consists of several assembled components, including a transparent acrylic plate, external frame, rubber gasket, structured liner, and back frame plate. As shown in
Figure 2, the transparent acrylic plate serves as a visual observation window, with an overall diameter of 2000 mm, enabling internal flow visualization. The external frame shares the same outer diameter (2000 mm) as the acrylic plate, with an inner diameter of 1880 mm. Circular connection holes are uniformly distributed along the circumference at 15° intervals to secure the acrylic plate to the back frame.
The rubber gasket is used to seal the assembly, matching the outer and inner diameters of the external frame. Its uncompressed thickness is 5 mm, and it compresses to approximately 4 mm after sealing, effectively ensuring both gas and liquid tightness.
The geometric parameters of the structured liner are shown in
Figure 2. The liner has an overall diameter of 1880 mm and a thickness of 4 mm. The surface of the liner is uniformly covered with hemispherical protrusions, each with a radius of 5 mm, forming a disturbance structure to enhance flow behavior. There are eight inlet ports located at the lower part, with a central angle of 15° between adjacent inlets, and four outlet ports at the upper part, also spaced at 15° intervals. Fluid enters through the inlet ports, fills the entire flow structure from bottom to top, and exits through the outlet ports.
The components of the Large Alkaline Water Electrolyzer Internal Flow Visualization and Performance Test System are bolted together to ensure efficient sealing.
The replica preserves the active area (400 cm2), channel width, depth, and flow field pattern, which are critical for replicating realistic electrolyte flow behavior. The serpentine parallel-channel structure and manifold configurations align with standard industrial practices commonly observed in commercial AWE systems.
The flow configuration (vertical upward flow) and channel network are representative of commercial designs aimed at promoting effective gas–liquid separation. While the replica focuses primarily on hydrodynamic characteristics rather than electrochemical performance, ensuring geometric and hydraulic similarity enables meaningful insights into flow uniformity, potential dead zones, and preferential pathways, which are closely linked to gas purity and energy efficiency in industrial operations.
It is important to note that active electrochemical reactions, gas evolution, and long-term durability tests were not conducted in this initial study. Also, it is acknowledged that some industrial systems adopt horizontal flow setups to meet specific design and operational requirements.
Also, one limitation of the current study is that all experiments and simulations were conducted under ambient pressure conditions. While this approach ensures experimental safety and reproducibility, it does not fully replicate the pressurized environments (typically 10–30 bar) commonly used in industrial AWE operations to enhance gas purity and system efficiency. Future work will integrate electrochemical operation and performance evaluation to further validate the representativeness of the design under realistic operating conditions and will implement high-pressure experimental setups and corresponding numerical simulations, allowing for a more comprehensive evaluation of flow behavior and performance under realistic operating pressures.
2.2. Introduction to the Fluorescent Tracer Method and Experimental Procedure
2.2.1. Fluorescent Tracer Method
Given the scale of the internal flow field and the complexities in visualizing extracted data, a fluorescent tracer method was applied to facilitate detailed flow visualization, crucial for subsequent data analysis.
In this study, fluorescent dye visualization was employed primarily to qualitatively identify the dominant flow paths, recirculation zones, and large-scale distribution patterns within the electrolyzer channels. Sodium fluorescein was dissolved in deionized water at a mass ratio of 1:200 and stirred at 600 rpm for 30 min using a magnetic stirrer to ensure complete dissolution, producing a homogeneous fluorescent tracer solution. This tracer emits green fluorescence when excited by a 365 nm ultraviolet light source. Once the fluid flow within the bipolar plate reached a stable state, the fluorescent tracer was introduced at a steady rate, matching the bulk flow velocity, enabling clear visualization of internal flow dynamics, and supporting precise tracer analysis. Observations focused on the early stages of dye propagation, where convective transport dominates over molecular diffusion. It is acknowledged that while dye visualization provides valuable qualitative insights, it does not offer fully quantitative flow measurements. To achieve more detailed and quantitative flow field characterization, future work will incorporate particle image velocimetry (PIV) techniques. It should be noted that in this study, the experimental setup was operated under steady-state conditions to establish a baseline understanding of the flow distribution and to enable direct validation with steady-state numerical simulations. While dynamic operating conditions—such as fluctuating inlet flow rates representing renewable energy variability—are recognized as important factors affecting electrolyte remixing and stratification, they were not investigated in the current work. Future studies are planned to incorporate dynamic boundary conditions to evaluate system robustness under transient operating scenarios. In addition to steady-state tests, future work will explore dynamic operation scenarios to further validate the system’s performance.
2.2.2. Experimental Procedure
The experimental setup of this study is shown in
Figure 2, which includes a pure water storage tank, a fluorescent dye reservoir, a pump, a controller, a camera, a flow control mechanism, and an ultraviolet lamp. The controller is designed to independently regulate the pump, enabling precise control over the injection of pure water from the storage tank and fluorescent dye from the dye reservoir at specific flow rates and velocities into the infusion-style flow field assembly. In this configuration, the liquid enters the flow field through eight inlet ports located at the bottom of the bipolar plate, traverses the flow field in a vertical, bottom-up direction, and exits through four outlet ports positioned at the top. This bottom-to-top flow arrangement allows for a comprehensive assessment of flow dynamics within the flow field.
During experimental validation, pure water was initially pumped into the flow field to fully saturate the internal channels. Once a stable flow was established, the fluorescent dye was introduced using the flow control device, and the UV lamp was activated under dark conditions. Photographs were then captured through the camera to observe and document the distribution of the fluorescent dye, providing a simulated visualization of electrolyte flow behavior within the conductive-type flow field component.
2.3. Evaluation Methods for Analyzing the Results of Experiments
Accurately evaluating non-ideal flow phenomena, including stagnation zones and remixing, is essential for enhancing flow uniformity in electrolytic cells. In this study, the analysis of experimental imaging results plays a pivotal role. We evaluated the flow characteristics inside the electrode plate by analyzing the occupied area and distribution characteristics of green fluorescent dye in the experimental photos frame by frame and then presented the residence time distribution. This method is called the area analysis method, and the detailed steps are as follows:
As shown in
Figure 3, the first image in the set to be processed is used for scale calibration to determine the true size of the area being analyzed, and this scale is applied to the subsequent processing methods.
- 2.
Determination of the maximum area and calibration of the fluorescence generation area;
In order to qualitatively analyze the internal flow state, the theoretical maximum flow area and fluorescence appearance area were ticked off and recorded using the ticking tool, as shown in
Figure 4, taken from the first image where the green fluorescent dye appeared.
Subsequently, the time interval, , was determined, and the fluorescence emergence area was determined at the same time interval until a set of experimental photo data were analyzed.
- 3.
Data processing methods;
To quantitatively analyze the area data, the fluorescence area recorded in each measurement image was compared to the theoretical maximum flow area to obtain the ratio of fluorescence area to total flow area. This ratio was then plotted against time, with the ratio as the vertical axis and time as the horizontal axis, to characterize the temporal evolution of the dyed region within the system. The resulting analysis was used to assess the impact of different structural designs on the optimization of internal flow behavior.
3. Simulation Details
The model described in this study is based on the following assumptions: (1) This study focuses solely on the flow conditions of the main liquid within the polar plate, considering only the liquid phase in the system. (2) According to the experimental methodology employed, the physical properties of the dye used are essentially identical to those of pure water; therefore, the two liquids considered in the calculations share the same physical properties. (3) The flow is assumed to be adiabatic, with negligible heat loss.
To avoid the extremely high grid resolutions typically required in Large Eddy Simulations, this study employs a hybrid Delayed Detached Eddy Simulation (DDES) model. This model integrates elements from both Reynolds-Averaged Navier–Stokes (RANS) and Large Eddy Simulation (LES) approaches. Within this framework, the wall boundary layer is modeled using RANS, while the larger separation regions are addressed using LES, enabling the resolution of parts of the turbulence spectrum in both time and space. The DDES model facilitates the transition between RANS and LES by comparing turbulence length scales with the grid spacing. For the k-ε, BSL, and SST models, the DDES function has been recalibrated to enhance the protection of the boundary layer. In this study, the hybrid Delayed Detached Eddy Simulation (DDES) framework was selected to balance computational cost and flow structure resolution. While DDES does not achieve the full fidelity of DNS or LES methods, it captures key turbulent features with significantly lower computational expense, making it suitable for simulating large-scale industrial flow fields such as those in AWE cells.
A detailed benchmarking of computational cost against other modeling strategies (e.g., RANS, LES, DNS) was not conducted, as this would require an extensive separate study. Future work will explore a systematic cost–performance comparison across different turbulence modeling approaches to further validate the computational advantages of DDES.
The specific equations are provided below.
Momentum conservation equation:
Species transport equation:
where
is the net rate of production of species
i by the chemical reaction (described later in this section) and
is the rate of creation by addition from the dispersed phase plus any user-defined sources. An equation of this form will be solved for the
N-1 species where
N is the total number of fluid phase. In this simulation, both
and Y are
.
In turbulent flows, ANSYS 2023 R1 Fluent computes the mass diffusion in the following form:
where
, is the turbulent Schmidt number (
, where
is the turbulent viscosity and
is the turbulent diffusivity). The default
is 0.7 and
is 2 × 10
−9 m
2/s. Note that turbulent diffusion generally overwhelms laminar diffusion, and the specification of detailed laminar diffusion properties in turbulent flows is generally not necessary.
We have simulated the flow within the polar plate using a low Reynolds number corrected SST
k-ω turbulence model. The turbulence kinetic energy,
k, and the specific dissipation rate,
ω, are obtained from the following transport equations:
and
In these equations, represents the generation of turbulence kinetic energy due to mean velocity gradients. represents the generation of ω. and represent the effective diffusivity of k and ω, respectively. and represent the dissipation of k and ω due to turbulence.
The effective diffusivities for the
k-
ω model are given by the following:
where
and
are the turbulent Prandtl numbers for
k and
ω, respectively, and
is the turbulent velocity in the SST model. The computational equations are as follows:
The coefficient
damps the turbulent viscosity causing a low-Reynolds number correction. It is given by the following:
where
where
is the strain rate magnitude and
is defined in Equation a.
is given by
where
y is the distance to the next surface.
In this study, a single-phase liquid flow model was employed to characterize the baseline hydrodynamic behavior within the electrolyzer channels. The modeling approach intentionally excluded gas generation and electrochemical reactions in order to isolate the effects of geometry and inlet conditions on flow distribution under controlled conditions.
While this simplification enables a clearer analysis of the fundamental flow structures, it is recognized that gas bubbles generated during electrolysis can influence flow uniformity, pressure drop, and mass transport by altering the effective flow area and introducing buoyancy-driven effects.
Future work will extend the current model to incorporate multiphase flow dynamics, including gas evolution, coalescence, and detachment mechanisms, to more accurately replicate real-world operational scenarios in industrial-scale AWE cells.
5. Conclusions
This study establishes a systematic framework for optimizing hydrodynamic uniformity in industrial-scale Alkaline Water Electrolyzers (AWE), combining fluorescent tracer experiments with high-fidelity DDESs. Three critical design aspects were investigated to enhance electrolyte flow dynamics and mitigate stagnation/gas retention:
Increasing total electrolyte flow rates (240–360 L/h) substantially reduces velocity inhomogeneity (35–40%) and shrinks low-speed stagnation zones by enhancing forced convection. Beyond 420 L/h, marginal gains in flow uniformity occur due to turbulent energy saturation. Symmetric inlet scheduling (e.g., alternating high/low velocities in Plan 6) further refined flow allocation, lowering velocity disparities by 28% and balancing residence time distributions. These findings highlight the dual role of total flow magnitude and dynamic inlet adjustments in suppressing remixing and gas holdup.
Radially deflecting inlet homogenizers (IS5) and helical flow distributors (IS6) demonstrated superior performance, eliminating 95% of stagnation regions by optimizing momentum transfer. IS5 achieved a 22% reduction in residence time variance by redirecting inertial forces toward lateral zones, while IS6’s swirling flow enhanced gas detachment. Both designs reduced current density localization, offering a scalable strategy for industrial electrodes exceeding 2 m diameter.
Staggered baffle arrays in the VBS3 (Vertical Baffle Systems) configuration suppressed vertical recirculation, reducing velocity inhomogeneity by 45% through compartmentalized flow channels. The HBS2 (Horizontal Baffle Systems) design minimized lateral flow deviations, achieving near-uniform residence times (σ < 0.8 s) via segmented flow paths. Its symmetric blocking structures redirected stagnant fluid to active zones, improving gas–liquid separation efficiency.