This section is divided by subheadings. It provides a concise and precise description of the experimental results, their interpretation, and the experimental conclusions that can be drawn.
3.1. Application of Res-PINNs to the Flow Around a Cylinder Problem
The flow around a cylinder problem is a classic case in Computational Fluid Dynamics (CFD), where different physical phenomena emerge in the flow field at various Reynolds numbers. In this case study, the objective is to reconstruct the flow field state of a two-dimensional flow around a cylinder using limited, discrete flow field data and a neural network model embedded with physical knowledge. A ROM for the flow around the cylinder is established, and the reconstruction results of the proposed Res-PINNs model are compared with those of the standard PINN model to validate the effectiveness of the network structure. In the flow around a cylinder problem, the Navier–Stokes (N-S) equations and the continuity equation typically serve as constraints. For two-dimensional, unsteady, incompressible fluid flow, the governing equations are generally expressed as follows:
Here, (x,y) are the spatial coordinates of points in the flow field, and t represents time. The variables u and v denote the velocity components of the fluid: u(x,y,t) is the velocity in the direction of the fluid inflow (denoted as IL), and v(x,y,t) is the velocity perpendicular to the fluid inflow direction (denoted as CF). The variable p represents the absolute pressure at each spatial point in the flow field, p(x,y,t). The term Re stands for the Reynolds number, which is a dimensionless parameter characterizing the flow regime, indicating the relative influence of inertial forces to viscous forces in the fluid.
In this case study, the numerical simulation method from CFD is used to obtain training data samples. The computational domain is illustrated in
Figure 4, where the overall flow field region is a rectangular area. The left side serves as the velocity inlet, while the right side is the pressure outlet. The upper and lower boundaries are set with no-slip boundary conditions. The inlet velocity on the left is set to
u0 = 0.09 m/s, and the absolute pressure at the right outlet is 0 Pa. The dynamic viscosity of the fluid is
μ = 1.5 × 10
−5 kg/m·s, the density is set to 1 kg/m
3, and the cylinder diameter is
D = 0.05, resulting in a Reynolds number of
Re = 300; we chose the laminar flow model in the Ansys simulation model for the computational analysis.
An O-type grid method was employed for the grid partitioning of the flow field region, as show in
Figure 5. The grid cell size was specified as 7.5 mm, and the height of the first layer in the boundary layer grid was set to 2 mm, with a total of seven layers and a growth rate of 1.15. The computational domain was divided into 40,064 cells in total. According to the grid quality evaluation results, 98.08% of the grids were classified as high quality. This grid partitioning approach ensures both the accuracy and stability of the calculations, providing a solid foundation for subsequent numerical simulations. The simulation time step is set to 0.005 s, and the PISO algorithm is used for transient solution calculations.
After 50 s from the start of the simulation, a stable vortex shedding phenomenon is observed. The flow field state information between 60 to 65 s is selected as the data collection period for training purposes. A training dataset is constructed using time slices of 0.1 s each. The spatial coordinates at each time instant are discretized, and the Latin hypercube sampling (LHS) algorithm is used to sample the spatial coordinates of the flow field region. During random sampling of the data, each time slice is processed independently. For each time slice, 1000 coordinate points are randomly sampled, and the corresponding flow field state information is used as training data for input into the training model. The data are then reset to a time frame of 0 to 5 s. The sampling results are shown in
Figure 6.
For both the PINN model and the Res-PINNs model, the input consists of the spatiotemporal coordinates
of the sparse flow field, and the output includes the velocity and pressure information
at each coordinate point within the flow field. After transformation, the equation for the loss function is expressed as follows:
The total loss function is composed of two parts: the loss for the flow field state information and the loss for the governing control equations. The flow field state information includes the velocity components in the
IL direction and
CF direction, as well as the pressure information. The total loss function can be expressed as follows:
In the above expressions,
denotes the grid points within the flow field, at the initial time, and along the boundary.
represents the data obtained from neural network training and the reference data.
denotes the residuals of the governing equations.
represents the flow field information at interior points, initial time flow field information, boundary point flow field information, and the weights for the loss function. Based on the analysis results from reference [
22], the loss function weights for the Res-PINNs neural network are set to
, while the loss function weights for the PINN neural network are set to
. These loss functions are then incorporated into the neural networks for model training. To avoid overfitting during the network training process, a dropout layer was added before the network output to reduce the interdependence between neurons, with the dropout rate set to 0.3. The DNN are trained using the machine learning framework TensorFlow on an NVIDIA GeForce RTX 3080 Ti GPU. After approximately 11 h of training, the model converged.
Figure 7 shows the training loss as a function of the number of training steps. After 14,000 iterations, the training error converged to less than 0.01.
In this case study, the model’s effectiveness is evaluated by reconstructing the flow field state at 5.5 s using sparse flow field information learned from 0 to 5 s. The reconstruction results are shown in
Figure 8, where
Figure 8a presents the reconstruction results using the PINN model;
Figure 8b shows the reconstruction results using the Res-PINNs model. Within both
Figure 8a,b, the first column displays the simulation results obtained from Ansys Fluent, which serve as a reference for evaluating the model’s accuracy. The second column shows the reconstructed flow field results from the ROM. The third column illustrates the error between the reference simulation results and the ROM reconstructions. From top to bottom, each row represents the pressure field contour plot, the velocity field contour plot in the
IL direction, and the velocity field contour plot in the
CF direction.
Based on the computational analysis of the results, for the PINN model, the relative error in the pressure field reconstruction is 10.9%; the relative errors in the velocity fields along the IL and CF directions are 38.8% and 39.2%, respectively; and the larger errors in the velocity fields are mainly concentrated around the cylinder wall region and the wake region. This discrepancy may be due to gradient vanishing or exploding during training, causing minimal changes in the loss function value as the number of iterations increases. Consequently, the network becomes less sensitive to learning the flow field states. In contrast, for the Res-PINNs model proposed in this study: The relative error in the pressure field is reduced to 7.8%; the relative errors in the velocity fields along the IL and CF directions are significantly reduced to 4.1% and 6.3%, respectively; and the Res-PINNs model demonstrates a more accurate reconstruction of the flow field state, particularly around the cylinder and in capturing vortex shedding features. This indicates that the model achieves more comprehensive learning of the flow characteristics, offering more reliable results for flow field modeling and reconstruction.
Overall, the results show that the Res-PINNs model significantly outperforms the standard PINN model in terms of accuracy and error reduction, particularly in regions with complex flow features. This highlights the effectiveness of the Res-PINNs framework in providing more accurate and stable solutions for fluid dynamics problems.
The analysis above indicates that the Res-PINNs model performs significantly better than the standard PINN model when dealing with complex flow fields. The Res-PINNs model demonstrates a more sensitive and comprehensive ability to capture state changes within the flow field. Although there are still some errors in certain regions of the flow field, overall, the Res-PINNs model exhibits a superior performance in flow field reconstruction compared to the standard PINN model.
Based on the reconstructed flow field data, the pressure and velocity gradient distributions on the cylinder surface can be integrated to compute the lift coefficient and drag coefficient of the cylinder. These coefficients are key parameters in characterizing the aerodynamic forces acting on the cylinder. The formulas for calculating these coefficients are as follows:
In the equations,
represents the boundary of the cylinder’s surface.
and
are the components of the unit normal vector on the cylinder surface. The dimensionless formulas for calculating the lift coefficient and drag coefficient are as follows:
Figure 9 illustrates the reconstruction results of the pressure and lift coefficient based on the flow field data. In the figure, the black line represents the results obtained from CFD simulation. The blue line indicates the predicted values from the Res-PINNs model. The red line shows the predicted results from the PINN model. From the results, it can be observed that for the prediction of
CD, the ROM results have a smaller error compared to the CFD simulation results. In contrast, the PINN model shows a larger amplitude difference compared to the CFD simulation results. This may be due to the gradient vanishing problem in the neural network model during the flow field reconstruction process, making it less sensitive to changes in the flow field state around the cylinder wall and failing to learn relevant flow characteristics, resulting in biased predictions in subsequent processes. The Res-PINNs prediction results are closer to the CFD simulation results; for the prediction of
CL, both the PINN and Res-PINNs models show some phase differences compared to the CFD simulation results. However, the Res-PINNs model, being more sensitive to changes in the flow field gradients, yields predictions that are closer to the CFD simulation results compared to the PINN model. The above analysis leads to the conclusion that, compared to the PINN model, the results of the flow field state reconstructed by the Res-PINNs model are in better agreement with the CFD simulation results.
3.2. Application of Res-PINNs to the Vortex-Induced Vibration Problem
VIV is a coupled fluid–structure interaction problem that occurs when a structure is placed in a flow field at a certain velocity. Periodic pulsating fluid forces along the flow direction and perpendicular to it cause the elastic structure to undergo periodic oscillations. This interaction alters the vortex shedding pattern of the fluid, resulting in coupled variations in fluid flow, structural displacement, and forces acting on the structure.
In this case study, the Reynolds averaged Navier–Stokes (RANS) equations are embedded into the Res-PINNs framework to solve high Reynolds number VIV problems, validating the model’s effectiveness. An incompressible fluid is governed by the N-S equations and the mass conservation equation, as detailed in
Section 3.1. However, as the Reynolds number increases, turbulence becomes significant, which not only raises the computational cost but can also lead to ill-conditioned matrices during numerical solutions. To address the challenges posed by turbulence, the RANS equations are introduced. The RANS equations provide a time-averaged approach to deal with turbulent flow, reducing the complexity of modeling turbulence while retaining the essential flow characteristics. The equations are expressed as follows:
Here,
represents the turbulent eddy viscosity coefficient. The RANS equations simulate turbulence by establishing a relationship between the eddy viscosity coefficient and the time-averaged parameters of the turbulence. The structural vibration equation can be modeled using a typical mass-spring-damper system. The equations of motion for the cylinder in the
IL direction and the
CF direction can be expressed as follows:
In these equations,
m represents the mass of the cylinder.
c and
k denote the damping coefficient and stiffness coefficient of the cylinder, respectively.
FL(
t) and
FD(
t) represent the lift and drag forces.
indicate the displacements along the
IL and
CF directions. At the initial state, the boundary conditions for the cylinder are as follows:
By discretizing the structural equations using the fourth-order Runge–Kutta method, the velocity and displacement responses of the cylinder can be solved. For solving the velocity and displacement in the
IL direction, the equations can be expressed as follows:
In the formula, k1, k2, k3, and k4 are the coefficients used in the Runge–Kutta method, and represents the simulation time step. The displacement and velocity responses along the CF direction are solved in a manner similar to the equations for the IL direction.
In the VIV case, CFD simulation technology is used to obtain the training data. The overall flow field is modeled and simulated using the Ansys Fluent software, and a schematic diagram of the entire flow field region is shown in
Figure 10. The structure consists of a 2-DOF two-dimensional cylindrical body with a spring. The left side is a velocity inlet with a magnitude of
u0 = 1 m/s, the right side is a 0 MPa absolute pressure outlet, the upper and lower boundaries are no-slip boundaries, and the cylinder surface is a no-slip wall. The overset mesh region is set to a size of 3D, and an unstructured
O-type mesh is used to divide the refined region. When the mesh is updated, the internal region of the
O-type mesh remains unchanged during the movement of the cylinder, while the area outside the
O-type mesh changes shape. The mesh division results are shown in
Figure 11. The size of the entire computational domain is 18
D × 12
D. The grid cell size was set to 3 mm, with the height of the first layer set to 1 mm and a growth rate of 1.3. The grid consisted of 60,040 cells in total, with 97.32% of the grids being categorized as high quality, according to the grid quality evaluation results. The Reynolds number is set to
Re = 1000, with a fluid density of 1 kg/m
3 and a dynamic viscosity of
μ = 0.5 × 10
−4 kg/(m·s). The cylinder’s mass is set to
m = 2 kg, the damping coefficient
c = 0.084, and the stiffness
k = 2.2020. During the simulation, the velocity, pressure, and other information at each grid point in the flow field are first computed, and the forces acting on the cylinder surface are calculated. This force information is then transferred to a compiled UDF, where the fourth-order Runge–Kutta method is used to solve the equations of motion for the structure. This process determines the displacement and velocity of the cylinder at the next time step, and the results are fed back into Fluent to update the flow field state for the subsequent time-step calculations, and the simulation time step is set to 0.005 s. The simulation model uses the
SST k−ω model, primarily because this model provides relatively accurate turbulence predictions in complex flow problems, especially in the near-wall region and in capturing flow separation. Compared to DNS and experimental flow types, the
SST k−ω model offers a reasonable trade-off between computational efficiency and accuracy.
Once the VIV motion of the cylinder reaches a steady state, the velocity field, pressure field, and cylinder motion information are extracted from the CFD simulation results as training data. The period from 20 to 25 s, after the motion has stabilized, is selected as the training dataset, with a time sampling interval of 0.1 s. To ensure that the data used are sufficient to support model training, we focused on its diversity and coverage during the data collection process, a Latin hypercube sampling method is employed to sample the flow field region, and 7000 data points are chosen as training data for each time step [
1].
In this case, the ROM based on Res-PINNs is used to reconstruct the entire flow field state using sparse flow field information. The neural network structure adopts the same architecture as the previous case, with the network input being the spatiotemporal data
of the sparse flow field. When solving for VIV, due to the coupling between the structure and the fluid, the structural vibration is also considered as a feature for both training and prediction. Additionally, the turbulent eddy viscosity coefficient is treated as an unknown parameter to be solved, making the network output
. For simplicity, the loss function of this network omits the overline notation from Equation (12). By embedding the horizontal and vertical displacements of the cylinder into Equation (12), the equation’s loss function is formulated as shown in Equation (17).
The overall loss function of the neural network model is the sum of the training data loss and the equation loss, as shown in Equation (18).
where
,
,
,
, and
,
represent the output results after neural network training. The loss is calculated, and parameter updates are performed through backpropagation using the Adam optimizer and the L-BFGS optimization method. This completes one training iteration of the neural network. Similarly, to prevent overfitting, a dropout layer was also added before the network output, with the dropout rate set to 0.5. The training is conducted using the same simulation environment and configuration as in Case 3.1. After 150,000 iterations, the network converges, and the loss calculation results are shown in
Figure 12.
The effectiveness of the model is evaluated by comparing the flow field reconstruction results at the 30 s mark using the ROM of Res-PINNs and PINN. The results are presented in
Figure 13, where
Figure 13a shows the reconstruction from the PINN model, and
Figure 13b displays the reconstruction from the Res-PINNs model. In both figures, the first column provides the Ansys Fluent simulation results as a reference, the second column shows the reconstructed results from the ROM, and the third column illustrates the error between the reference simulation and the reconstructed results. From top to bottom, the rows represent the pressure field contour plot, the velocity field contour plot along the
IL direction, and the velocity field contour plot along the
CF direction. Calculations of the prediction results for the two models reveal that the velocity field reconstructed by the Res-PINNs model shows errors of 17.7% and 17.2% relative to the CFD simulation results, while the errors for the PINN model are 40.3% and 30.8%, respectively. For the pressure field, the reconstruction errors of the Res-PINNs and PINN models are 14% and 35.1%, respectively. The areas with significant errors are primarily located in the
O-type mesh and its vicinity. This discrepancy is likely due to the limitations of the PINN model’s network architecture, which cannot fully analyze high-dimensional and nonlinear problems, resulting in a reduced ability to learn the flow field state features. In contrast, the Res-PINNs model, with its more complex network structure, can more accurately capture the velocity distribution of the flow field. The reconstruction results indicate that the Res-PINNs model outperforms the PINN model in the wake region, mitigating the shortcomings found in the PINN model.
Figure 14 shows the reconstruction results for the lift coefficient, drag coefficient, and structural displacement of the flow field from 25 to 28 s. In the figure, the black line represents the exact values obtained from the CFD simulation, the blue line represents the reconstruction results from the Res-PINNs model, and the red line represents the reconstruction results from the PINN model. Analysis of the reconstruction results reveals significant differences between the reconstructed structural vibration responses and the CFD simulation results. These discrepancies are more noticeable at the peaks and troughs of the response states, and there is a clear phase difference when compared to the CFD results. However, the overall trends remain consistent with the CFD results. A comparison with other studies [
23,
24] suggests that this discrepancy may result from the relatively low resolution of the data slicing, which impedes the model’s ability to capture the dynamic coupling between the flow field and the structure. Consequently, the model fails to accurately capture the transient behavior of the flow field and structural dynamics, particularly during more intense dynamic responses. The insufficient temporal information further causes phase deviations in the prediction results along the time axis, impairing the model’s ability to accurately track changes in the flow field. This discrepancy suggests that a greater amount of time-sliced data might be required for the model to learn and compute more accurately when constructing the training dataset. For the reconstruction of the lift and drag coefficients, the Res-PINNs model provides results that are more accurate and closer to the CFD simulation outcomes compared to the PINN model.
Based on the above analysis, it can be concluded that both the Res-PINNs and PINN models exhibit significant differences from the CFD simulation results when reconstructing the structural response in VIV problems. However, for the reconstruction of the lift and drag coefficients, the Res-PINNs model demonstrates superior performance.