A Clustering-Based Dimensionality Reduction Method Guided by POD Structures and Its Application to Convective Flow Problems
Abstract
1. Introduction
- (1)
- Modal decomposition projection methods (e.g., POD-Galerkin [5]) construct low-dimensional subspaces by extracting dominant modes and projecting governing equations onto these subspaces.
- (2)
- Balanced truncation methods [6] utilize controllability and observability analysis to retain state variables most significant to input–output behavior.
- (3)
- Harmonic balance methods [7] address periodic flows by approximating steady-state solutions in the frequency domain using a finite set of Fourier basis functions.
- (1)
- (2)
- (3)
- Colanera [15] proposed a robust spectral POD (SPOD) integrating robust principal component analysis for improved noise resilience.
- Gu [16] introduced frequency-domain POD (FD-POD) to reconstruct unsteady flows using embedded frequency information.
- Bui [17] developed POD-ISAT, combining POD with in situ adaptive tabulation for efficient steady-state PDE approximation.
- A clustering-based dimensionality reduction framework guided by POD, improving stability in nonlinear mode extraction.
- A novel ECEPM-based probabilistic mode-ranking method.
- Empirical validation demonstrating improved interpretability and reconstruction accuracy over traditional POD methods.
2. C-POD Method and Modal Sorting Method
2.1. C-POD Method
- (1)
- Arrange the responses of m data points under n different operating conditions as columns to form the database matrix D:
- (2)
- Determining initial cluster centers based on the snapshot POD method [23]. First, the correlation matrix of the snapshot matrix D is constructed:
- (3)
- Sample assignment. Each column vector dⱼ (j = 1, 2, …, n) from the data matrix D is assigned to the nearest cluster center, forming K clusters:
- (4)
- Updating the cluster centers. The cluster centers are updated by calculating the mean of all vectors in each cluster Si, and this mean becomes the new cluster centroid:
- (5)
- Repeat steps (6) and (7) until the following convergence criterion is met:
- (6)
- To obtain the modal coefficients, project the original data matrix D onto the cluster bases μ by solving the following least squares problem [24]:
| Algorithm 1: C-POD Method |
| Input: //Snapshot matrix K //Number of clusters Output: //Cluster basis A = [a1 a2 … an] //Modal coefficients Steps: 1. POD Initialization - Solving the eigenvalue problem: - Select first K eigenvectors as the initial cluster centers: 2. Clustering - Assign each snapshot vector to its closest cluster center - Compute cluster centroids: 3. Compute modal coefficients - Modal coefficient vector ar ∈ Rk: 4. Return , A |
2.2. Entropy-Controlled Euclidean-to-Probability Mapping Modal Sorting Method
- (1)
- Euclidean distance matrix construction: Assume the database matrix , clustering bases , and the column vectors are and , where .
- (2)
- Compute the column-wise sum of squares for and , as shown in Equations (11) and (12):
- (3)
- Construct the inner product matrix G:
- (4)
- Construct the squared Euclidean distance matrix:where 1 is a matrix with all elements equal to 1, so the Euclidean distance matrix is:
- (1)
- Take the negative of the distance matrix C and shift each column, to avoid overflow in exponential calculations, resulting in the matrix Z:where , , and represents the outer product.
- (2)
- Perform the exponential operation on the shifted matrix and introduce the temperature parameter , resulting in:
- (3)
- Normalize the exponential matrix column-wise to obtain the final probability weight matrix W:where represents element-wise division, is the sum of each column, and extends the column sum to a matrix where all columns are identical. In the construction of the probability mapping function, the temperature parameter is used to adjust the entropy of the probability distribution. The temperature parameter plays a critical role in shaping the entropy level of the probability mapping. A smaller (e.g., ≈ 0.1) leads to a more peaked distribution, favoring one dominant mode per snapshot, which increases mode separability but may result in instability. In contrast, a larger (e.g., > 5) produces a flatter distribution, balancing contributions across multiple modes and improving robustness but potentially reducing interpretability.
| Algorithm 2: ECEPM |
| Input: //Snapshot data matrix //Clustering bases (b modes) //Temperature parameter for entropy control Output: //Probability weight matrix Steps: 1. Compute the squared norms of D and μ: - - 2. Compute the inner product matrix 3. Compute the squared Euclidean distance matrix: - - 4. Shift the distance matrix column-wise to prevent numerical overflow: //Broadcasting to each row 5. Compute exponentials with entropy scaling: 6. Normalize E column-wise to obtain the probability matrix: //Element-wise division, columns sum to 1 Return: W //Each column W_j contains probabilities of modes for sample j |
3. Burgers’ Equation and C-POD Method
3.1. Introduction to the One-Dimensional Burgers’ Equation
3.2. Database Construction
- Initial Conditions:
- Spatial and Temporal Discretization:
- Independent Variable Settings:
3.3. C-POD Method Accuracy Testing and Modal Decomposition
4. Cylinder Wake Flow Case Introduction
4.1. Model Introduction
- (1)
- Using the alternating direction implicit (ADI) method, the equation was split into two one-dimensional equations. The implicit solution was first performed in the x direction, followed by the implicit solution in the y direction:
- (2)
- The pressure field was obtained by solving the Poisson equation using the fast Fourier transform (FFT):
- (3)
- We set a fixed flow velocity at the left boundary x = 0 and applied a no-slip boundary condition on the surface of the cylinder.
- (4)
- To quantify the rotational behavior in the flow field around the cylinder, the calculation of vorticity was introduced in the two-dimensional incompressible flow field. The vorticity is defined as:
4.2. Grid Independence and Numerical Method Validation
4.3. Presentation of Computational Results
- (1)
- From t = 1 to 9 s, a pair of side vortices formed on either side of the cylinder, gradually creating a pair of vortices with opposite rotational directions behind the cylinder.
- (2)
- From t = 9 to 33 s, the counter-rotating vortices behind the cylinder gradually moved and diffused in the direction of the flow.
- (3)
- From t = 33 to 110 s, the counter-rotating vortices behind the cylinder disappeared, and the flow became more stable.
- (4)
- For times exceeding 110 s, the flow pattern underwent notable changes, with vortices detaching periodically to form a stable vortex street. Multiple regions of alternating positive and negative vorticity emerged, indicating pronounced shear and recirculation phenomena within the flow field—hallmarks of the classic Kármán vortex street.
4.4. Cylinder Wake Flow Field Dimensionality Reduction and Reconstruction Accuracy, and Modal Decomposition
4.4.1. Dataset and Test Set Construction
4.4.2. Dimensionality Reduction and Reconstruction Accuracy of Cylinder Wake Flow Field
4.4.3. Comparison of Modal Decomposition Results
- (1)
- Type 1 vortex: Symmetrical side vortices extending downstream from both sides of the cylinder into the flow region.
- (2)
- Type 2 vortex: Counter-rotating pair of vortices formed at the rear of the cylinder due to recirculation, resembling a “horseshoe” shape.
- (3)
- Type 3 vortex: Developing horseshoe-shaped vortices.
- (4)
- Type 4 vortex: Asymmetric side vortices on both sides of the cylinder.
4.5. Evolutionary Characteristics of Modes Extracted by C-POD over Time
- For t < 9 s:
- First mode: Dominated by third-type vortex, which weakened over time.
- Second mode: Transitioned from irregular fourth-type vortex to first-type vortex, with rotation changes observed at t = 1 s, 5 s, and 9 s.
- For 9 s < t < 14.5 s:
- First mode: Evolved from third-type to first-type vortex, extending its influence across the field.
- Second mode: Underwent gradual rotational transition.
- For 13.5 s < t < 33 s:
- The vorticity distribution became more stable.
- Second-mode rotation direction changed.
- For 33 s < t < 110 s:
- System entered a quasi-stable regime, with negligible changes in vortex structures and modal vorticity values.
- For 110 s < t < 119.5 s:
- Flow transitioned toward vortex street formation.
- Side vortices began shedding at the tail, and second-mode rotation altered again.
- For t > 119.5 s:
- The system reached a state of periodic vortex street generation.
- Mode coefficients showed periodic oscillation (Figure 15).
5. Application Value of C-POD ROM in Inverse Problems
Limitations and Future Improvements
- Computational Cost
- Sensitivity to Sensor Layout in Inverse Reconstruction
- Hyperparameter Dependency
- (1)
- Developing lightweight clustering algorithms or incremental learning schemes to reduce computational overhead.
- (2)
- Exploring adaptive sensor deployment strategies.
- (3)
- Investigating robust and data-driven hyperparameter tuning frameworks.
6. Conclusions
Author Contributions
Funding
Data Availability Statement
Conflicts of Interest
References
- Guo, T.; Wu, W.-L.; Luo, Z.-M.; Zhang, Y.-N. Mode Transition and Drag Characteristics of Non-Circular Cylinders in a Uniform Flow. Ocean Eng. 2024, 297, 117025. [Google Scholar] [CrossRef]
- Santana, G.M.; Fabro, A.T.; Miserda, R.F.B. Analysis of the Dynamic Modes of the Transonic Flow around a Cylinder. J. Braz. Soc. Mech. Sci. Eng. 2024, 46, 414. [Google Scholar] [CrossRef]
- Copeland, D.M.; Cheung, S.W.; Huynh, K.; Choi, Y. Reduced Order Models for Lagrangian Hydrodynamics. Comput. Methods Appl. Mech. Eng. 2022, 388, 114259. [Google Scholar] [CrossRef]
- Knockaert, L.; De Zutter, D. Stable Laguerre-SVD Reduced-Order Modeling. IEEE Trans. Circuits Syst. I Fundam. Theory Appl. 2003, 50, 576–579. [Google Scholar] [CrossRef]
- Couplet, M.; Basdevant, C.; Sagaut, P. Calibrated Reduced-Order POD-Galerkin System for Fluid Flow Modelling. J. Comput. Phys. 2005, 207, 192–220. [Google Scholar] [CrossRef]
- Besselink, B.; van de Wouw, N.; Scherpen, J.M.A.; Nijmeijer, H. Model Reduction for Nonlinear Systems by Incremental Balanced Truncation. IEEE Trans. Autom. Control 2014, 59, 2739–2753. [Google Scholar] [CrossRef]
- Grolet, A.; Thouverez, F. On a New Harmonic Selection Technique for Harmonic Balance Method. Mech. Syst. Signal Process. 2012, 30, 43–60. [Google Scholar] [CrossRef]
- Bagheri, M.H.; Esmailpour, K.; Hoseinalipour, S.M.; Mujumdar, A.S. Numerical Study and POD Snapshot Analysis of Flow Characteristics for Pulsating Turbulent Opposing Jets. Int. J. Numer. Methods Heat Fluid Flow 2019, 29, 2009–2031. [Google Scholar] [CrossRef]
- Bistrian, D.A.; Navon, I.M. Randomized Dynamic Mode Decomposition for Nonintrusive Reduced Order Modelling. Int. J. Numer. Methods Eng. 2017, 112, 3–25. [Google Scholar] [CrossRef]
- Phillips, T.R.F.; Heaney, C.E.; Smith, P.N.; Pain, C.C. An Autoencoder-Based Reduced-Order Model for Eigenvalue Problems with Application to Neutron Diffusion. Int. J. Numer. Methods Eng. 2021, 122, 3780–3811. [Google Scholar] [CrossRef]
- Zhang, X.; Ji, T.; Xie, F.; Zheng, C.; Zheng, Y. Data-Driven Nonlinear Reduced-Order Modeling of Unsteady Fluid–Structure Interactions. Phys. Fluids 2022, 34, 053608. [Google Scholar] [CrossRef]
- Shen, X.; Du, C.; Jiang, S.; Sun, L.; Chen, L. Enhancing Deep Neural Networks for Multivariate Uncertainty Analysis of Cracked Structures by POD-RBF. Theor. Appl. Fract. Mech. 2023, 125, 103925. [Google Scholar] [CrossRef]
- Yuan, L.; Ni, Y.-Q.; Deng, X.-Y.; Hao, S. A-PINN: Auxiliary Physics Informed Neural Networks for Forward and Inverse Problems of Nonlinear Integro-Differential Equations. J. Comput. Phys. 2022, 462, 111260. [Google Scholar] [CrossRef]
- Taassob, A.; Kumar, A.; Gitushi, K.M.; Ranade, R.; Echekki, T. A PINN-DeepONet Framework for Extracting Turbulent Combustion Closure from Multiscalar Measurements. Comput. Methods Appl. Mech. Eng. 2024, 429, 117163. [Google Scholar] [CrossRef]
- Colanera, A.; Schmidt, O.T.; Chiatto, M. Robust Spectral Proper Orthogonal Decomposition. Comput. Phys. Commun. 2025, 307, 110056. [Google Scholar] [CrossRef]
- Gu, X.; Xu, C.; Liu, M.; Mao, Y. Frequency-Domain Proper Orthogonal Decomposition for Efficient Reconstruction of Unsteady Flows. Phys. Fluids 2025, 37, 025109. [Google Scholar] [CrossRef]
- Bui, D.; Hamdaoui, M.; De Vuyst, F. POD-ISAT: An Efficient POD-Based Surrogate Approach with Adaptive Tabulation and Fidelity Regions for Parametrized Steady-State PDE Discrete Solutions. Int. J. Numer. Methods Eng. 2013, 94, 648–671. [Google Scholar] [CrossRef]
- Burkardt, J.; Gunzburger, M.; Lee, H.-C. Centroidal Voronoi Tessellation-Based Reduced-Order Modeling of Complex Systems. SIAM J. Sci. Comput. 2006, 28, 459–484. [Google Scholar] [CrossRef]
- Iqbal, M.M.; Xavier, R.J. Development of Optimal Reduced-Order Model for Gas Turbine Power Plants Using Particle Swarm Optimization Technique. Int. Trans. Electr. Energy Syst. 2020, 30, e12278. [Google Scholar] [CrossRef]
- Huera-Huarte, F.J.; Vernet, A. Vortex Modes in the Wake of an Oscillating Long Flexible Cylinder Combining POD and Fuzzy Clustering. Exp. Fluids 2010, 48, 999–1013. [Google Scholar] [CrossRef]
- D’aDamo, J.; Collaud, M.; Sosa, R.; Godoy-Diana, R. Wake and Aeroelasticity of a Flexible Pitching Foil. Bioinspiration Biomim. 2022, 17, 045002. [Google Scholar] [CrossRef]
- Wei, Z.; Yang, Z.; Xia, C.; Li, Q. Cluster-Based Reduced-Order Modelling of the Wake Stabilization Mechanism behind a Twisted Cylinder. J. Wind. Eng. Ind. Aerodyn. 2017, 171, 288–303. [Google Scholar] [CrossRef]
- Sirovich, L.; Kirby, M. Turbulence and the Dynamics of Coherent Structures. Part1: Coherent Structures. Q. Appl. Math. 1987, 45, 561–571. [Google Scholar] [CrossRef]
- Guo, H.B.; Renaut, R.A. A Regularized Total Least Squares Algorithm. In Total Least Squares and Errors-in-Variables Modeling: Analysis, Algorithms and Applications; Van Huffel, S., Lemmerling, P., Eds.; Springer: Berlin/Heidelberg, Germany, 2002; pp. 57–66. [Google Scholar]
- Yuan, S.; Blömker, D.; Duan, J. Stochastic Turbulence for Burgers Equation Driven by Cylindrical Lévy Process. Stoch. Dyn. 2022, 22, 2240004. [Google Scholar] [CrossRef]
- Li, L.; Ong, K.W. Dynamic Transitions of Generalized Burgers Equation. J. Math. Fluid Mech. 2016, 18, 89–102. [Google Scholar] [CrossRef]
- Senechal, M. Spatial Tessellations: Concepts and Applications of Voronoi Diagrams. Science 1993, 260, 1170. [Google Scholar] [CrossRef]
- Williamson, C.H.K. Oblique and Parallel Modes of Vortex Shedding in the Wake of a Circular Cylinder at Low Reynolds Numbers. J. Fluid Mech. 1989, 206, 579–627. [Google Scholar] [CrossRef]
- Singha, S.; Sinhamahapatra, K.P. High-Resolution Numerical Simulation of Low Reynolds Number Incompressible Flow About Two Cylinders in Tandem. J. Fluids Eng. 2010, 132, 011101. [Google Scholar] [CrossRef]
- Qu, L.; Norberg, C.; Davidson, L.; Peng, S.-H.; Wang, F. Quantitative Numerical Analysis of Flow Past a Circular Cylinder at Reynolds Number Between 50 and 200. J. Fluids Struct. 2013, 39, 347–370. [Google Scholar] [CrossRef]
- Everson, R.; Sirovich, L. Karhunen–Loève Procedure for Gappy Data. J. Opt. Soc. Am. A 1995, 12, 1657–1664. [Google Scholar] [CrossRef]
- Deus, J.; Martin, E. Efficient Cardiovascular Parameters Estimation for Fluid-Structure Simulations Using Gappy Proper Orthogonal Decomposition. Ann. Biomed. Eng. 2024, 52, 3037–3052. [Google Scholar] [CrossRef] [PubMed]
- Yuan, Q.; Guo, X.; Han, J.; Han, B.; Zhang, B.; Lan, T.; Ariyur, K. Correlation Coefficient Filtering Method for Optimal Sensor Layout Strategies in Inverse Problems. J. Eng. 2025, 2025, 5142118. [Google Scholar] [CrossRef]























| Operating Condition Name | Total Number of Grids | CD | St | |
|---|---|---|---|---|
| A1 | 100 × 40 | 0.01 | 0.215 | 0.16 |
| A2 | 200 × 80 | 0.01 | 0.225 | 0.164 |
| A3 | 400 × 160 | 0.01 | 0.229 | 0.165 |
| A4 | 600 × 240 | 0.01 | 0.229 | 0.165 |
| A5 | 800 × 320 | 0.01 | 0.229 | 0.165 |
| A6 | 400 × 160 | 0.005 | 0.229 | 0.165 |
| A7 | 400 × 160 | 0.02 | 0.231 | 0.167 |
| Experiment from [28] | 0.165 | |||
| CFD from [29] | 0.226 | 0.165 | ||
| CFD from [30] | 0.233 | 0.166 |
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. |
© 2025 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/).
Share and Cite
Yuan, Q.; Zhang, B. A Clustering-Based Dimensionality Reduction Method Guided by POD Structures and Its Application to Convective Flow Problems. Algorithms 2025, 18, 366. https://doi.org/10.3390/a18060366
Yuan Q, Zhang B. A Clustering-Based Dimensionality Reduction Method Guided by POD Structures and Its Application to Convective Flow Problems. Algorithms. 2025; 18(6):366. https://doi.org/10.3390/a18060366
Chicago/Turabian StyleYuan, Qingyang, and Bo Zhang. 2025. "A Clustering-Based Dimensionality Reduction Method Guided by POD Structures and Its Application to Convective Flow Problems" Algorithms 18, no. 6: 366. https://doi.org/10.3390/a18060366
APA StyleYuan, Q., & Zhang, B. (2025). A Clustering-Based Dimensionality Reduction Method Guided by POD Structures and Its Application to Convective Flow Problems. Algorithms, 18(6), 366. https://doi.org/10.3390/a18060366

