Next Article in Journal
Dynamic Characteristics of an Underwater Ventilated Vehicle Exiting Water in an Environment with Scattered Ice Floes
Previous Article in Journal
Vertical Monotonic and Cyclic Responses of a Bucket in Over-Consolidated Clay
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Investigation of Physics-Informed Neural Networks to Reconstruct a Flow Field with High Resolution

1
State Key Laboratory of Ocean Engineering, Shanghai Jiao Tong University, Shanghai 200240, China
2
School of Naval Architecture, Ocean & Civil Engineering, Shanghai Jiao Tong University, Shanghai 200240, China
3
Shanghai Key Laboratory of Mechanics in Energy Engineering, Shanghai Institute of Applied Mathematics and Mechanics, School of Mechanics and Engineering Science, Shanghai University, Shanghai 200240, China
*
Author to whom correspondence should be addressed.
J. Mar. Sci. Eng. 2023, 11(11), 2045; https://doi.org/10.3390/jmse11112045
Submission received: 23 September 2023 / Revised: 19 October 2023 / Accepted: 22 October 2023 / Published: 25 October 2023
(This article belongs to the Section Ocean Engineering)

Abstract

:
Particle image velocimetry (PIV) is a widely used experimental technique in ocean engineering, for instance, to study the vortex fields near marine risers and the wake fields behind wind turbines or ship propellers. However, the flow fields measured using PIV in water tanks or wind tunnels always have low resolution; hence, it is difficult to accurately reveal the mechanics behind the complex phenomena sometimes observed. In this paper, physics-informed neural networks (PINNs), which introduce the Navier–Stokes equations or the continuity equation into the loss function during training to reconstruct a flow field with high resolution, are investigated. The accuracy is compared with the cubic spline interpolation method and a classic neural network in a case study of reconstructing a two-dimensional flow field around a cylinder, which is obtained through direct numerical simulation. Finally, the validated PINN method is applied to reconstruct a flow field measured using PIV and shows good performance.

1. Introduction

Particle image velocimetry (PIV) is a transient and noncontact fluid mechanics velocity measurement method that uses tracers and a high-speed camera. It has a wide range of applications within the field of offshore engineering. For instance, it is used to observe the wakes around a cylinder in a uniform flow to reveal the mechanics of vortex-induced vibration and many new vortex-shedding modes that are very important in the flow-induced load prediction of marine rises [1,2,3]. PIV is also applied to the study of the wakes behind wind turbines or ship propellers and can be used for the design of the airfoil profiles of such structures [4,5,6]. However, due to the restrictions of the experimental apparatus, the flow field measured using PIV does not always have very high resolution.
Over the years, to further study the mechanism behind the complex phenomena of the flow field, many scholars have devoted considerable effort to finding methods to enhance the resolution of the data obtained through PIV. The most convenient approach would be to interpolate the low-resolution flow field with methods such as Lagrange interpolation, Newton interpolation, and Hermite interpolation. However, the performance is not always satisfactory for cases in which the turbulent flow field has a complex vortex morphology.
In recent years, machine learning has developed rapidly and has been widely used in a variety of fields. Some scholars have applied it to fluid dynamics. Rao et al. [7] applied deep learning to low-Reynolds-number steady-state and transient laminar flow simulations. Fukami et al. [8] used a convolutional neural network (CNN) model and a hybrid downsampled skip-connection multiscale (DSC/MS) machine learning model to reconstruct laminar flow from low-resolution flow field data. Meng et al. [9] predicted the drag and lift coefficients of blunt body flow and cylindrical flow based on a CNN model, which significantly improved the prediction accuracy. However, the traditional neural network is normally considered a “black box” that cannot reflect the physical properties of the flow field. To address this problem, Karpatne et al. [10] developed physics-guided neural networks that can consider the physical relationships between temperature, density, and water depth. Raissi et al. [11,12,13,14,15] proposed the concept of physics-informed neural networks (PINNs) for nonlinear partial differential equations. They also conducted research on deep learning of vortex-induced vibration (VIV) and hidden fluid mechanics. The framework of PINNs imposes physical information constraints on the loss function of the neural network so that a more generalized model can be trained with fewer data samples. The primary advantage of using PINNs lies in their superior extrapolation capabilities compared to other interpolation methods.
Many researchers have applied PINNs to various fields where the physical information involved is different depending on the specific research objectives. Cai et al. [16,17] trained a neural network by minimizing a loss function composed of a data mismatch term and residual terms associated with the coupled Navier–Stokes and heat transfer equations. They inferred a complete continuous three-dimensional velocity field and pressure field from the three-dimensional temperature field snapshot obtained from Tomo-BOS imaging. Mao et al. [18] used a PINN to approximate Euler’s equations for simulating high-speed aerodynamic flows. Pang et al. [19] extended PINNs to fractional PINNs to solve space–time fractional advection–diffusion equations and analyzed the convergence systematically. Cuomo et al. [20] made a comprehensive review of scientific machine learning through PINNs and compared its advantages and disadvantages with traditional neural networks.
In this study, the performance of a PINN to reconstruct a flow field with high resolution was investigated. The ability of the PINN to deal with the difficulties induced by the sparseness and noise of the data was mainly evaluated. The flow around a cylinder in a uniform from direct numerical simulation (DNS) and experimental data measured by PIV were used in this case study. The predicted accuracy was compared with that of the cubic spline interpolation method and the classic neural network. It was found that the PINN yielded a better reconstructed flow field in all the cases studied in this paper.

2. Methodology

The classic case of a uniform flow past a cylinder in a two-dimensional space was studied. In this case study, the flow field is represented by six physical quantities: the time t, the coordinates x and y, the velocity of the flow particle u and v, and the pressure p.

2.1. Classic Neural Networks

The topology of the classic multilayer perceptron (MLP) neural network [21] applied in this case study is shown in Figure 1, consisting of input, hidden, and output layers. The time t and the coordinates x and y are selected as the three features for the input layer, while the velocity of the flow particle u and v and the pressure p are the features for the output layer. Therefore, both the input and output layers are composed of three nodes. Different numbers of hidden layers and different numbers of nodes in the layers are tested to determine the best topology. Each node is connected to all the nodes in front and behind it, thus forming a neural network. The connection of the node can be expressed as:
a i , j = f n = 1 j w i , j , n a i 1 , n + b i , j
where ai,j refers to the jth node in the ith layer in the neural network. Wi,j,n and bi,j are the weight and bias at the corresponding node, respectively. F(x) refers to the activation function.
The mean square error (MSE) is defined as the loss function, which can be expressed as:
L 1 = M S E u , v , p = 1 3 n i = 1 n j = 1 3 Y i , j o u t Y i , j t r u e 2
Here, n refers to the number of samples. Yi,1, Yi,2, and Yi,3 represent the ith sampling data of u, v, and p, respectively. The superscripts true and out refer to the true and predicted values, respectively.

2.2. Physics-Informed Neural Networks

The neural network introduced above is considered a “black box” since it only considers the features in the input and output layers to be independent numbers. However, there sometimes exist some physical relations among them. In this case study, the continuity equation and Navier–Stokes equations should be satisfied for viscous incompressible Newtonian fluids and can be written as [22]:
u x + v y = 0 u t + u u x + v v y = 1 ρ p x + ν u x x + u y y v t + u v x + v v y = 1 ρ p y + ν v x x + v y y
Here, the subscripts t, x, and y refer to the partial derivative with respect to the time and the coordinate, while the subscripts xx and yy represent the second-order partial derivative to the coordinate. These physical relations can be considered in the neural network by defining a new loss function. In the MLP, the loss function is defined as the difference between the true value and the predicted value of the velocity and pressure. Here, in the PINN, another three components are introduced into the loss function that can be expressed as:
L 2 = e 0 + β 1 e 1 + β 2 e 2 + β 3 e 3 e 0 = M S E u , v , p e 1 = u x + v y e 2 = u t + u u x + v v y + 1 ρ p x ν u x x + u y y e 3 = v t + u v x + v v y + 1 ρ p y ν v x x + v y y
where e0 represents the mean square error between the output values and true values of u, v, and p, as also described in Equation (2). E1, e2, and e3 refer to the difference between the left and right sides of Equation (3). Βi corresponds to a weight for each term. The topology of the neural network is shown in Figure 2. It is termed PINN #1 in this paper, where β1, β2, and β3 are set to be equal to one. This means that the weight of satisfaction of the continuity equation and Navier–Stokes equations are considered to be as equally important as the mean square error of the velocity and pressure field.
For the DNS training data, the pressure can be obtained directly from the numerically simulated results. However, for the PIV data, only the velocity field can be measured, and the reconstruction of the pressure field based on the velocity field with noise can introduce uncertain error [23]. Therefore, a PINN that only considers the continuity equation, as shown in Figure 3, denoted by PINN #2 in this paper, was also studied. In PINN #2, β2 and β3 are set to zero. The loss function L3 is written as:
L 3 = e 0 + e 1 e 0 = M S E u , v e 1 = u x + v y
Here, e 0 represents the mean square error between the output values and true values of u and v.

3. Training Data

Two sets of data of the flow field around a two-dimensional cylinder [24] were studied, which were numerically simulated based on the DNS method and experimentally measured using PIV. Since the numerical results already had a very high resolution and no noise, the data were first diluted and then added with some white Gaussian noise. The reconstruction of the processed field data could be used to evaluate the performance of different neural networks.

3.1. Original and Diluted DNS Data

The computational efforts for the DNS increased rapidly with an increasing Reynolds number. In this case, the Reynolds number was 200, and the data were obtained based on the finite volume method described in [25,26]. The center of the cylinder was at the origin of the coordinates, as shown in Figure 4. The sampling area was −3D to 8D in the x direction and −3D to 3D in the Y direction, where D refers to the diameter of the cylinder. The length of the selected data was 50 s, with a sampling frequency of 4 Hz. Meanwhile, at each time step, the distribution of the sampling spatial points was inhomogeneous, and there were more sampling points near the cylinder and in the wake field, as shown in Figure 4, with a total number of 9957. The flow field of the vortex shedding was periodic. The contour and time series of the velocity and pressure at positions A, B, and C are depicted in Figure 5.
As mentioned above, the DNS data already had a high resolution. Therefore, the data were first diluted, which means that a large amount of the data was removed. The remaining data were used to train the neural network for reconstructing the missing flow field. The reconstructed result was compared with the original removed data to validate the accuracy of the proposed methods. As 50%, 20%, 10%, and 5% of the sampling points were left, their spatial distribution is shown in Figure 6.

3.2. PIV Data

The quality of the PIV data depends heavily on the density of the tracer particles and the performance of the high-speed camera. Normally, the data have a lower resolution, which may not meet research requirements. In this case, PIV data were gathered from a recirculation water channel with a test section measuring 300 mm × 250 mm × 2000 mm. The flow was sufficiently seeded by hollow glass spheres, with an average diameter of 20 μm and a density of 1.03 g/mm2, which were well mixed by recirculating in the water channel. Then, during the experiment, the images were recorded using a charge-coupled device (CCD) camera. Details of the experimental apparatus can be seen in [27]. The cylinder diameter was 15 mm, which was far less than the cylinder length. The incoming flow velocity was 0.111 m/s. With the kinematic viscosity of water ν = 1.111 × 10−6 m2/s, the Reynolds number was 1500. The range of the measurement area by PIV was 96 mm × 80 mm. The sampling points tended to be evenly distributed, with a total of 152 × 127 velocity vectors. The pressure could not be measured using PIV directly. In addition, only a rectangular area of 60 mm × 40 mm near the cylinder was selected for the training. The center of the cylinder was at the origin of the coordinates, which was out of the view of the PIV camera. The sampling area was 2.7D to 6.6D in the x direction and −1.3D to 1.3D in the y direction, where D refers to the diameter of the cylinder, as shown in Figure 7. The length of the measured data was 600 s, with a sampling frequency of 5 Hz. The time slot when vortex shedding occurred was selected.

4. Results and Discussion

The performance of the neural network was strongly related to the hyperparameters selected, including, for instance, the number of hidden layers, the number of nodes in each layer, and the activation function. Therefore, a hyperparameter analysis was first performed. Then, the determined topology of the neural network was used for the reconstruction of the diluted DNS data and measured PIV data.

4.1. Hyperparameter Analysis

Different combinations of the hyperparameters listed in Table 1 were tested for the DNS data with a dilution rate of 50%. The performance of the trained network was evaluated based on the mean square error between the initial and predicted values, as well as the training efficiency.
During the training process, the Adam optimizer was adopted, which divided the overall epochs into three parts, and each part reduced the learning rate to 10% of the previous part. The number of epochs remained the same for different combinations of hyperparameters. It was trained until the loss function converged. An example of the loss function during the training is given in Figure 8.
Considering the accuracy and training time needed, the most appropriate hyperparameters were selected as follows: the number of neural network layers was 10, the number of nodes in each layer was 96, and the activation function was the Sin function. The predicted results of the 9957 sampling spatial points described in Figure 4 are all plotted in Figure 9. It showed very good accuracy. The hyperparameter combination was also used later for the PINN. The time series of the DNS data and the predicted values of u, v, and p at three selected positions are also plotted in Figure 10. They coincided with each other very well in this case.

4.2. Reconstruction of the DNS Data with Different Dilution Rates

The flow field was diluted with only 50%, 30%, 20%, 10%, 5%, and 2% of the data left, which were used to train the neural network. The training of the neural network was conducted on a laptop computer. It had a twelve-core 2.6GHz CPU, an RTX1660Ti display card, and a 16G RAM. Taking the case of the reconstruction of 5% diluted DNS data as an example, to guarantee the convergence of the training, the MLP, PINN #1, and PINN #2 took 18 min, 5.1 h, and 54 min, respectively. The process of the dilution of the flow field simulated using DNS to low resolution and the reconstruction of the diluted flow field to high resolution is shown in Figure 11.
The reconstructed velocity and pressure fields using different neural networks in different cases are compared in Figure 12. In this figure, the initial and predicted flow fields are smoothed. When the flow field was diluted to approximately 10%, the classic MLP neural network was still able to predict the missing data with quite a high accuracy. It was difficult to see the difference from the contour plot. With a further decrease in the dilution rate of 5% and 2%, the MLP neural network showed poor performance, where the predicted result was quite different from the initial values, especially for the velocity along the flow direction. Moreover, when the velocity gradient changed greatly, the prediction accuracy became worse, and the reconstructed field was prone to be inconsistent to some extent. However, both the PINN #1 and PINN #2 performed very well, even for the case where only 5% and 2% of the data were left for training. This can also be seen from the comparison of the MSE values based on the MLP and the PINNs, as shown in Figure 13. For the case of a dilution rate of 2%, the MSE for the classic neural network was more than ten times that of the MSEs for the PINNs. In addition, PINN #2, which only used the continuity equation as the loss function, showed a better performance than PINN #1, which was unexpected. It could have been attributed to the fact that the pressure could actually be determined through the pressure Poisson equation for a two-dimensional flow.

4.3. Reconstruction of the DNS Data with White Noise

The flow field measured using PIV in water channels always has noise due to the experimental environment or the test apparatus. To better simulate reality and evaluate the anti-interference ability of the neural network, a certain amount of noise was added to the flow velocity and pressure from the DNS. The noise selected in this case study followed a Gaussian distribution, with a mean value of zero and a standard deviation of 0.3, which can be written as:
Y j = Y j 1 + ε , j = 1 , 2 , 3 ε N μ , σ 2 ,   where   μ = 0 ; σ = 0.3
Here, Y1, Y2, and Y3 represent the original data of u, v, and p, respectively. Y j refers to the data with added white noise ε. The contour plots of u, v, and p with noise are shown in Figure 14.
The performance of the MLP, PINN #1, and PINN #2 on the reconstruction of the flow field with white noise was compared. It can be seen very clearly from the contour plot of the velocity along the flow direction that the MLP yielded a bad prediction. In contrast, it is difficult to see the difference between the predicted data and the DNS data of the velocity and pressure from the contour plot. Therefore, the distribution of the error was also plotted. The MSE of the velocity and pressure based on the PINNs was much smaller than the MSE based on the MLP. In addition, the spatial distribution of the error based on the MLP was more stochastic in the whole flow field, while the error was normally very large in the areas where the velocity gradient of the wake varied greatly. This could be attributed to the continuity equation or the Navier–Stokes equations embedded in the loss function.
To see the accuracy of the reconstruction of the flow field more clearly, the reconstructed velocity and pressure along the x-axis at y = 0.6D and y = 0 are plotted in Figure 15. The noise could be filtered by both the MLP and PINN methods. A better curve fitting was achieved by using the PINNs than the MLP from the comparison of the correlation coefficients.

4.4. Reconstruction Based on PIV Data

The flow field around a two-dimensional cylinder measured with PIV is shown in the first row in Figure 16. The resolution was not as high as that of the DNS numerical simulation data, and noise could be seen very clearly in the original data. The reconstruction of the flow field by using the MLP and PINN #2 is also depicted in this figure. The localized structural similarity index (SSIM) [28] between the original and reconstructed flow fields measured via PIV as well as the MSE value were computed. PINN #2 yielded a larger SSIM and a smaller MSE value, which means it yielded a better reconstructed flow field. In addition, the reconstructed velocities u and v along the x-axis at y = 0 and y = 0.6D were plotted and compared with the measured PIV data, as shown in Figure 17. It was very clear that PINN #2 gave a better curve fit of the PIV data than the classic neural network. Hence, PINN #2 was applied to reconstruct the velocity field measured by PIV to a high-resolution field at different time steps, as shown in Figure 18.

5. Conclusions

In this paper, the performances of a classic neural network and two physically informed neural networks in reconstructing a high-resolution flow field were evaluated and compared. The PINNs incorporated the continuity equation or Navier–Stokes equations in the loss function. The flow around a circular cylinder was used in this case study. The methodology considered can also be used in other cases, such as the wake fields behind wind turbines or ship propellers.
A hyperparameter analysis was first carried out to determine an appropriate topology of the neural network, i.e., the number of hidden layers, the number of nodes in each layer, and the activation function.
The effects of the sparseness and noise of the training data on the reconstruction quality were mainly studied. Both the numerical simulation data based on DNS and the experimental data measured by PIV were analyzed.
The effect of sparseness was discussed by processing the DNS data with different dilution rates. It was found that all three neural networks worked very well when the dilution rate was larger than 10%. As the dilution rate decreased to 5% and 2%, the flow field reconstructed by the MLP neural network was very different from the true value, while PINN #1 and PINN #2 still worked very well. In the case with a dilution rate of 2%, the MSE for the classic neural network was more than ten times that of the PINNs.
The influence of noise was studied for the DNS data with added white noise and the PIV data. Both the classic neural network and the PINNs showed their ability to filter out noise, and overfitting did not occur in this case study. In the case of the DNS data, the MSE of the flow velocity u based on the classic neural network was more than five times the MSE using the PINN. For the reconstruction of the PIV data, the curve fit of the velocity in the whole field using the PINN was also much closer to the true value. Finally, the PINN was applied to reconstruct the flow field measured using PIV to a high-resolution field without noise.
It should also be noted that the cases in this paper were characterized by a low Reynolds number. The reconstruction of fine vortex structures was also not within the scope of this paper. However, some published research has proved the great potential of PINNs in analyzing cases with a series of high Reynolds numbers up to 104 [29,30]. In these studies, the numerical data simulated by the Reynolds average Navier–Stokes (RANS) method were used for training a neural network. In future studies, efforts are still required to be devoted to dealing with the reconstruction of much more turbulent flow fields and much smaller-scale vortex structures.

Author Contributions

Methodology, Z.Y. and Y.X.; Investigation, Z.Y., Y.X., J.J., X.F. and M.Z.; Writing—Original Draft Preparation, Z.Y.; Writing—Review and Editing, H.R. and T.S.; Resources, B.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (grant No. 52271283, No. 52111530135, No. 52088102), the Science and Technology Commission of Shanghai Municipality (22ZR1434100), the Key R & D program of Shandong Province (grant No. 2021SFGC0701), and the Shenlan Project (grant No. SL2021MS018).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

This article does not contain any studies with human participants or animals performed by any of the authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sun, C.; Zhou, T.; An, H.; Zhu, H.; Cheng, L. On the study of vortex-induced vibration of circular cylinders covered with different roughness. Appl. Ocean. Res. 2022, 124, 103215. [Google Scholar] [CrossRef]
  2. Hsieh, S.C.; Low, Y.M.; Chiew, Y.M. Flow Characteristics around a Circular Cylinder Subjected to Vortex-Induced Vibration near a Plane Boundary. J. Fluids Struct. 2016, 65, 257–277. [Google Scholar] [CrossRef]
  3. Korkischko, I.; Meneghini, J.R. Suppression of Vortex-Induced Vibration Using Moving Surface Boundary-Layer Control. J. Fluids Struct. 2012, 34, 259–270. [Google Scholar] [CrossRef]
  4. Paik, B.G.; Kim, J.; Park, Y.H.; Kim, K.S.; Yu, K.K. Analysis of wake behind a rotating propeller using PIV technique in a cavitation tunnel. Ocean. Eng. 2007, 34, 594–604. [Google Scholar] [CrossRef]
  5. Jackson, R.S.; Amano, R. Experimental Study and Simulation of a Small-Scale Horizontal-Axis Wind Turbine. ASME J. Energy Resour. Technol. 2017, 139, 051207. [Google Scholar] [CrossRef]
  6. Dobrev, I.; Massouh, F. CFD and PIV investigation of unsteady flow through Savonius wind turbine. Energy Procedia 2011, 6, 711–720. [Google Scholar] [CrossRef]
  7. Rao, C.P.; Hao, S.; Yang, L. Physics-informed deep learning for incompressible laminar flows. Theor. Appl. Mech. Lett. 2020, 10, 207–212. [Google Scholar] [CrossRef]
  8. Fukami, K.; Fukagata, K.; Taira, K. Super-resolution reconstruction of turbulent flows with machine learning. J. Fluid Mech. 2019, 870, 106–120. [Google Scholar] [CrossRef]
  9. Meng, Y. Research on the Combined Method of Physical Model and Machine Learning of Flow around Circular Cylinder and Vortex-Induced Vibration. Master’s Thesis, Dalian Maritime University, Dalian, China, 2020. [Google Scholar]
  10. Karpatne, A.; Watkins, W.; Read, J.; Kumar, V. How can physics inform deep learning methods in scientific problems?: Recent progress and future prospects. In Deep Learning for Physical Sciences Workshop; NIPS 2017; Long Beach Convention Center: Long Beach, CA, USA, 2017. [Google Scholar]
  11. Raissi, M.; Perdikaris, P.; Karniadakis, G.E. Physics Informed Deep Learning (Part I): Data-driven Solutions of Nonlinear Partial Differential Equations. arXiv 2017, arXiv:1711.10561. [Google Scholar]
  12. Raissi, M.; Wang, Z.; Triantafyllou, M.; Karniadakis, G. Deep learning of vortex-induced vibrations. J. Fluid Mech. 2019, 861, 119–137. [Google Scholar] [CrossRef]
  13. Raissi, M.; Yazdani, A.; Karniadakis, G.E. Hidden Fluid Mechanics: A Navier-Stokes Informed Deep Learning Framework for Assimilating Flow Visualization Data. arXiv 2018, arXiv:1808.04327. [Google Scholar]
  14. Raissi, M.; Perdikaris, P.; Karniadakis, G.E. Physics-informed neural networks: A deep learning framework for solving forward and inverse problems involving nonlinear partial differential equations. J. Comput. Phys. 2019, 378, 686–707. [Google Scholar] [CrossRef]
  15. Raissi, M.; Yazdani, A.; Karniadakis, G.E. Hidden fluid mechanics: Learning velocity and pressure fields from flow visualizations. Science 2020, 367, 1026–1030. [Google Scholar] [CrossRef] [PubMed]
  16. Cai, S.; Wang, Z.; Fuest, F.; Jeon, Y.J.; Gray, C.; Karniadakis, G.E. Flow over an espresso cup: Inferring 3-D velocity and pressure fields from tomographic background oriented Schlieren via physics-informed neural networks. J. Fluid Mech. 2021, 915, A102. [Google Scholar] [CrossRef]
  17. Cai, S.; Wang, Z.; Wang, S.; Perdikaris, P.; Karniadakis, G.E. Physics-informed neural networks for heat transfer problems. J. Heat Transf. 2021, 143, 060801. [Google Scholar] [CrossRef]
  18. Mao, Z.; Jagtap, A.D.; Karniadakis, G.E. Physics-informed neural networks for high-speed flows. Comput. Methods Appl. Mech. Eng. 2020, 360, 112789. [Google Scholar] [CrossRef]
  19. Pang, G.; Lu, L.; Karniadakis, G.E. fPINNs: Fractional physics-informed neural networks. SIAM J. Sci. Comput. 2019, 41, A2603–A2626. [Google Scholar] [CrossRef]
  20. Cuomo, S.; Di Cola, V.S.; Giampaolo, F.; Rozza, G.; Raissi, M.; Piccialli, F. Scientific machine learning through physics–informed neural networks: Where we are and what’s next. J. Sci. Comput. 2022, 92, 88. [Google Scholar] [CrossRef]
  21. Rumelhart, D.; Hinton, G.; Williams, R. Learning representations by back-propagating errors. Nature 1986, 323, 533–536. [Google Scholar] [CrossRef]
  22. Kundu, P.; Cohen, I.M. Fluid Mechanics; Academic Press: Cambridge, MA, USA, 2008; Volume 3. [Google Scholar]
  23. Liu, S.; Xu, J.L.; Yu, K.K. Implementation and research on the reconstruction algorithms of pressure fields based on PIV. J. Exp. Fluid Mech. 2016, 30, 56–65. [Google Scholar]
  24. Rajani, B.; Kandasamy, A.; Majumdar, S. Numerical simulation of laminar flow past a circular cylinder. Appl. Math. Model. 2009, 33, 1228–1247. [Google Scholar] [CrossRef]
  25. Fu, X.; Fu, S.; Zhang, M.; Han, Z.; Ren, H.; Xu, Y.; Zhao, B. Frequency capture phenomenon in tandem cylinders with different diameters undergoing flow-induced vibration. Phys. Fluids 2022, 34, 085120. [Google Scholar] [CrossRef]
  26. Fu, X.; Fu, S.; Han, Z.; Niu, Z.; Zhang, M.; Zhao, B. Numerical simulations of 2-DOF vortex-induced vibration of a circular cylinder in two and three dimensions: A comparison study. J. Ocean. Eng. Sci. 2023. [Google Scholar] [CrossRef]
  27. Zhou, J.K.; Qiu, X.; Li, J.H.; Liu, Y.L. The gap ratio effects on vortex evolution behind a circular cylinder placed near a wall. Phys. Fluids 2021, 33, 037112. [Google Scholar] [CrossRef]
  28. Anatharaman, V.; Feldkamp, J.; Fukami, K.; Taira, K. Image and video compression of fluid flow data. Theor. Comput. Fluid Dyn. 2023, 37, 61–82. [Google Scholar] [CrossRef]
  29. Bai, X.D.; Zhang, W. Machine learning for vortex induced vibration in turbulent flow. Comput. Fluids 2022, 235, 105266. [Google Scholar] [CrossRef]
  30. Cheng, C.; Meng, H.; Li, Y.-Z.; Zhang, G.-T. Deep learning based on PINN for solving 2 DOF vortex induced vibration of cylinder. Ocean. Eng. 2021, 240, 109932. [Google Scholar] [CrossRef]
Figure 1. The topology of the MLP neural network.
Figure 1. The topology of the MLP neural network.
Jmse 11 02045 g001
Figure 2. Topology structure and loss function of PINN #1.
Figure 2. Topology structure and loss function of PINN #1.
Jmse 11 02045 g002
Figure 3. Topology structure and loss function of PINN #2.
Figure 3. Topology structure and loss function of PINN #2.
Jmse 11 02045 g003
Figure 4. Distribution of the sampling spatial points. A, B, and C are sampling points in the flow field with coordinates (−1.5, 0), (0, 1.5), and (4, 0), respectively.
Figure 4. Distribution of the sampling spatial points. A, B, and C are sampling points in the flow field with coordinates (−1.5, 0), (0, 1.5), and (4, 0), respectively.
Jmse 11 02045 g004
Figure 5. Contours and time series of the velocity and pressure.
Figure 5. Contours and time series of the velocity and pressure.
Jmse 11 02045 g005
Figure 6. Distribution of sampling of spatial points at different dilution rates.
Figure 6. Distribution of sampling of spatial points at different dilution rates.
Jmse 11 02045 g006
Figure 7. Velocity field measured using PIV in a period.
Figure 7. Velocity field measured using PIV in a period.
Jmse 11 02045 g007
Figure 8. The decline in the loss function during training.
Figure 8. The decline in the loss function during training.
Jmse 11 02045 g008
Figure 9. The relationship between the DNS data and predicted values of u, v, and p for the selected neural network.
Figure 9. The relationship between the DNS data and predicted values of u, v, and p for the selected neural network.
Jmse 11 02045 g009
Figure 10. Time series of the DNS data and the predicted values of u, v, and p. A, B, and C are sampling points in the flow field with coordinates (−1.5, 0), (0, 1.5), and (4, 0), respectively.
Figure 10. Time series of the DNS data and the predicted values of u, v, and p. A, B, and C are sampling points in the flow field with coordinates (−1.5, 0), (0, 1.5), and (4, 0), respectively.
Jmse 11 02045 g010
Figure 11. Reconstruction of the flow field from low to high resolution.
Figure 11. Reconstruction of the flow field from low to high resolution.
Jmse 11 02045 g011
Figure 12. Reconstruction of the DNS data with different dilution rates using MLP, PINN #1, PINN #2, and cubic interpolation method.
Figure 12. Reconstruction of the DNS data with different dilution rates using MLP, PINN #1, PINN #2, and cubic interpolation method.
Jmse 11 02045 g012
Figure 13. Comparison of the MSE values based on MLP and PINN.
Figure 13. Comparison of the MSE values based on MLP and PINN.
Jmse 11 02045 g013
Figure 14. DNS data, predicted data, and error of various reconfiguration methods for 30% noise.
Figure 14. DNS data, predicted data, and error of various reconfiguration methods for 30% noise.
Jmse 11 02045 g014
Figure 15. The reconstructed velocity and pressure along the x direction at y = 0.6D and y = 0.
Figure 15. The reconstructed velocity and pressure along the x direction at y = 0.6D and y = 0.
Jmse 11 02045 g015
Figure 16. Prediction results for PIV data.
Figure 16. Prediction results for PIV data.
Jmse 11 02045 g016
Figure 17. Comparison of various methods for PIV data prediction.
Figure 17. Comparison of various methods for PIV data prediction.
Jmse 11 02045 g017aJmse 11 02045 g017b
Figure 18. Reconstruction of the velocity field measured by PIV.
Figure 18. Reconstruction of the velocity field measured by PIV.
Jmse 11 02045 g018
Table 1. List of the hyperparameters analyzed.
Table 1. List of the hyperparameters analyzed.
No.HyperparameterValue
1Number
of hidden layers
5
210
315
4Number of nodes in each hidden layer72
596
6120
7Activation functionSin
8Tanh
9ReLU
10Leaky ReLU
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

Yang, Z.; Xu, Y.; Jing, J.; Fu, X.; Wang, B.; Ren, H.; Zhang, M.; Sun, T. Investigation of Physics-Informed Neural Networks to Reconstruct a Flow Field with High Resolution. J. Mar. Sci. Eng. 2023, 11, 2045. https://doi.org/10.3390/jmse11112045

AMA Style

Yang Z, Xu Y, Jing J, Fu X, Wang B, Ren H, Zhang M, Sun T. Investigation of Physics-Informed Neural Networks to Reconstruct a Flow Field with High Resolution. Journal of Marine Science and Engineering. 2023; 11(11):2045. https://doi.org/10.3390/jmse11112045

Chicago/Turabian Style

Yang, Zhou, Yuwang Xu, Jionglin Jing, Xuepeng Fu, Bofu Wang, Haojie Ren, Mengmeng Zhang, and Tongxiao Sun. 2023. "Investigation of Physics-Informed Neural Networks to Reconstruct a Flow Field with High Resolution" Journal of Marine Science and Engineering 11, no. 11: 2045. https://doi.org/10.3390/jmse11112045

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