3.2. OpenFOAM Post-Processing
Following the simulation process using OpenFOAM and the ParaView visualization tool, velocity contour plots were obtained based on the velocity field’s magnitude and directional components (x, y, and z). Concurrently, additional hydrodynamic parameters, such as water volume fraction, dynamic pressure, and static pressure, were evaluated. The simulation replicated the fluid–structure interaction along a defined channel, modeling the wave flow and its impact with the OWC structure over a total simulation duration of 120 s.
The pre-configuration parameters used in OpenFOAM are presented in
Table 2. The system commands listed were selected to achieve accurate simulation results. These configurations include definitions for constants such as gravitational acceleration, conserved mass transport schemes, turbulence modeling parameters, and the thermophysical properties of water as the working fluid. The simulation outputs primarily evaluate the system’s mass flow rate, pressure distribution, and velocity profiles. The OWC structure was defined as a wall-type obstacle for the boundary conditions, allowing it to be interpreted as a solid geometry within the computational domain. The pneumatic outlet interface between the OWC structure and the environment was also designated a wall-type boundary.
The multiphase solver in OpenFOAM was employed to simulate multiphase flow bevahior. The specific path used was multiphase/interFoam/laminar/waves/stokesV, where interFoam refers to the solver used for simulating multiphase flows based on the Volume of Fluid method. The laminar flow type was selected, given that the scenario involves regular wave patterns. The waves directory indicates the simulation of wave phenomena, while stokesV corresponds to a specific case representing Stokes wave conditions.
The blockMeshDict configuration, located in the system directory, was used to generate the computational mesh using hexahedral blocks. Vertices were defined to delimit the desired computational domain. Within the boundary conditions, the Oscillating Water Column structure was modeled as a wall-type obstacle, ensuring its recognition as part of the computational geometry. Similarly, the outlet boundary was also defined as a wall-type obstacle for the same purpose. As previously stated, the domain depth was set to a nearly negligible value of 0.04 m, supporting the assumption of a two-dimensional simulation.
To enhance accuracy in critical regions, a refined mesh was generated in areas of interest, such as the free surface, the wave–air interface, and the interior of the OWC chamber, to better quantify wave impacts and internal effects. This detailed meshing was performed using the snappyHexMeshDict configuration, which further subdivides the base mesh to improve resolution. Additional refinement was conducted using the meshQualityDict and refineMeshDict utilities. The resulting mesh characteristics are presented in
Table 3.
The controlDict file defines the main simulation control parameters, specifying a total simulation of 120 s, a time step of 0.01 s, and a data acquisition interval of 1 s. The decomposeParDict file is used to decompose the computational domain into multiple subdomains, enabling parallel processing across several CPU cores. This configuration is particularly useful for optimizing computational efficiency and reducing simulation time. Additionally, the fvSchemes and fvSolution files contain the numerical discretization schemes and solver settings, respectively. These defines the mathematical procedures and solution algorithms applied during the simulation, depending on the selected physical model. Finally, the setFieldsDict file initializes the fluid region within the computational domain. In this case, it sets the water level inside the wave channel, covering an x-directional range from 0 to 104 m and a water depth of 9 m.
The boundary conditions are presented in
Table 4. Wave generation and absorption are handled with the interFoam solver, applying different specifications at the inlet and outlet of the computational domain. At the inlet (left boundary), the volume fraction and velocity are prescribed according to OWC design. At the outlet (right boundary), a shallow water absorption model is imposed. The seabed, the structural surfaces of the OWC, and the outlet structure are treated as no-slip walls, labeled grounds, and obstacles, respectively. The top boundary is assigned atmospheric conditions, and the lateral boundaries are defined as empty to emulate two dimensional flow.
The physical properties defined for the simulation are as follows: Gravity was applied in the z di-rection with a magnitude of 9.81 m/s2. The kinematic viscosity was set to 1.0 × 10−6 m2/s for water and 1.48 × 10−5 m2/s for air. The density values used were 1025 kg/m3 for sea-water and 1.2 kg/m3 for air. Additionally, a surface tension coefficient of 0.07 N/m was defined between the water and air phases. Wave generation was configured with a wave height of 2 m, based on the average wave conditions at the selected site, and a wave period of 12 s, with a phase angle of 0°. A ramp time of 4 s was established to allow the wave amplitude to increase gradually, reaching the full height of 1 m. This approach minimizes transients and abrupt disturbances at the start of wave generation. Active wave absorption was also implemented at the inlet to reduce unwanted reflections. At the channel outlet, a shallow water absorption model was employed to dissipate wave energy and prevent wave reflections from re-entering the domain, thereby ensuring more realistic flow behavior and avoiding interference with the internal wave dynamics.
The computational domain can be seen in
Figure 8 and was constructed using hexahedral mesh blocks, with a bottom element size of 0.04 m to ensure adequate resolution near the seabed. The complete domain size is 104.5 m in length and 20 m in height. The water phase within the domain ranges from a minimum wave through height of 5.5 m in width and 10 m in height, with a wall thickness of 0.5 m. The outlet duct, which channels compressed air toward the turbine, has a width of 0.8 m. It is important to note that one of the key assumptions is the use of two-dimensional (2D) geometry in the OpenFOAM simulations. The three-dimensional (3D) modeling will be developed in the second stage, based on the assumption that the results are proportional across each unit of transverse area.
Figure 9 presents the velocity magnitude contours at 60 s and 120 s at two specific time instants. As this is a dynamic analysis, these time points were selected to assess the development of the flow field and observe system behavior approaching the end of the simulation period. At 60 s, a concentrated region of elevated water velocity is evident within the chamber, corresponding to the formation of the oscillating water column. Additionally, the incident waves can identify high-velocity profiles before reaching the structure’s front wall. By 120 s, the influence of the front wall on flow acceleration becomes more pronounced. The seabed also contributes to flow acceleration through boundary effects. The parameter under analysis shows a marked increase in velocity near the structure, primarily attributed to the constructive interference of incoming waves. This interaction facilitates a partial energy transfer to the structure, with the most significant effect being the localized rise in velocity both around and within the internal chamber of the system.
Figure 10 illustrates the distribution of velocity components within the channel, where the flow is predominantly confined to the XZ plane. This configuration allows for a two-dimensional analysis of the flow behavior. The velocity component along the y-axis can be considered negligible in the overall velocity estimation, as it exerts minimal influence on the flow dynamics. In contrast, the X and Z axes reveal the regions with the highest velocity magnitudes. On the other hand, turbulence phenomena are observed to reach velocities exceeding 4 m/s near the rear boundary of the computational domain and adjacent to the structural elements. The visual output corresponds to a single wave cycle analysis, wherein one wave crest is visible outside the structure. In contrast, the subsequent crest is captured within the pneumatic chamber.
Figure 11 illustrates the water volume fraction obtained from the two-phase simulation. Regions in red, corresponding to a numerical value of 1, represent the water phase, while regions in blue, with a value of 0, denote the air phase. The interface between these two fluids can be approximated as a pneumatic piston, effectively simulating a horizontal boundary within the chamber. The simulation and analysis were conducted over a single wave period. Results indicate that the mass distribution of water inside and outside the structure remains acceptable, preventing overtopping and flooding of the pneumatic chamber. The simulations are parameterized by wave period or frequency, which facilitates accurate modeling of wave height and the dynamic motion of the pneumatic piston within the chamber. The observed height differential between internal and external water levels also supports forming a linear profile, enabling a stable and proportionally consistent pneumatic airflow.
Figure 12 and
Figure 13 in the final analysis present static and dynamic pressure distributions, respectively.
Figure 12 shows that the maximum pressure values are concentrated in the lower region, indicating that the flow exerts a significant hydrostatic load on the analyzed volume. Static pressure increases moderately with the passage of waves and provides a critical support element for the structure by maintaining a relatively constant pressure gradient over time at the lower inlet. This stability facilitates the formation of a piston-like water level within the chamber. Conversely, reducing pressure on the front wall leads to more turbulent flow conditions. The regions of high dynamic pressure are located near the structure at the moment of wave impact, indicating that an increase in wave velocity directly corresponds to an increase in dynamic pressure within the zone of influence.
Conversely, static pressure remains relatively stable over time, as the water and air levels within the chamber tend to stabilize due to the periodic nature of the bidirectional flow. It is important to emphasize that the oscillation frequency of the water column within the chamber must resonate with the frequency of the incident waves. This resonance amplifies the oscillatory motion, enhancing the pneumatic compression and decompression, thereby maximizing overall energy conversion efficiency.
3.3. Computational Fluid Dynamics Post-Processing
In the second stage of the analysis, emphasis is placed on the use of Computational Fluid Dynamics (CFD) software. The data obtained from the first stage, using OpenFOAM, specifically velocity, pressure, and mass flow rate, serve as time-dependent input conditions for this phase. These parameters vary according to wave period and are imported into the CFD tool to assess airflow dynamics. With the theoretical framework established, the design of the Wells turbine and the corresponding flow redirectors was developed to evaluate the interaction of air with the turbine blade profiles.
The structural design of the wave power plant includes a 10-m-high front wall, of which 3.5 m are submerged and impacted by incoming waves. The wall thickness is 0.5 m. The internal air chamber measures 5 m in width, 3 m in length, and an average height of 6 m. These dimensions are based on optimization ratios and the reference design of the Mutriku wave power plant in Spain. At the top of the chamber, the air duct has a diameter of 0.8 m. It is concentrically connected to the turbine compartment, which has a rectangular prism shape with a square base, measuring 1.2 m in height and 1.1 m in width. The turbine duct was designed within a computational domain of an internal cylindrical geometry enclosed by an external cubic volume. The cylinder’s diameter corresponds to the outlet diameter derived from the OpenFOAM simulation. A flow director was incorporated to guide the airflow trajectory, ensuring it effectively impacts the turbine blades. This component is essential to prevent adverse effects such as flow recirculation, backpressure, and excessive compressive stress on the turbine shaft. Lastly, structural supports were included to secure the energy conversion assembly, ensuring operational stability.
The configuration settings within the CFD tool were defined to simulate internal flow, incorporating gravity effects, rotational dynamics, thermodynamic properties, and the boundaries of the computational domain. Boundary conditions were established based on the inlet parameters derived from the OpenFOAM simulations, while the outlet was set to atmospheric pressure.
Figure 14 illustrates the behavior and pressure phenomena on a shear plane across the Wells turbine blades. The atmospheric pressure, defined as the upper boundary condition, creates a region of vector field stability. The lower section of the turbine shows high pressure because the turbine is in the impulse phase. As a result, the opposite side of the turbine experiences a low-pressure (vacuum) region, and this pressure differential generates the rotational inertia required to drive the turbine. This same effect can be observed when the pressure zones are reversed, due to the return motion of the ocean wave, which induces a vacuum state within the air chamber. However, due to the symmetrical airfoil profile of the Wells turbine blades, the rotation is maintained in a single direction, regardless of the airflow direction. Due to the idealized simulation conditions, a relatively low pressure distribution is seen throughout the domain except in the vicinity of the turbine blades, where localized variations are more pronounced.
Figure 15 shows the velocity vector field and its interaction with the duct and the Wells turbine. A change in velocity can be observed after the airflow comes into contact with the turbine blades, indicating a reduction in air energy. The shape and arrangement of the NACA airfoil enable the observation of an inertial rotational motion. Additionally,
Figure 15 presents a velocity shear profile along the blade section. It is evident that the maximum air velocity occurs at the leading and trailing edges of the profile, which enhances the rotational tendency of the shaft.