Error Quantification of Gaussian Process Regression for Extracting Eulerian Velocity Fields from Ocean Drifters
Abstract
:1. Introduction
2. Materials and Methods
2.1. Theory and Calculation
2.1.1. Gaussian Process Regression
2.1.2. Optimization of Hyperparameters
2.1.3. Error Estimation and Density Calculation
2.2. Datasets and Test Configuration
2.2.1. Time-Periodic Double-Gyre Simulation
2.2.2. Navy Coastal Ocean Model (NCOM)—Convergence Region
3. Results
3.1. Double-Gyre Model
3.1.1. Performance Metrics and Temporal Trends
- High predictive accuracy:
- −
- Metrics such as , D, and consistently approach 1 across all time steps, indicating excellent agreement between predicted and reference velocities.
- −
- Error metrics (RMSE, MAE, MBE) remain close to zero throughout the time series, underscoring the robustness of the GPR model in reconstructing velocity fields.
- Temporal variations:
- −
- Model performance peaks during the middle observational period when (e.g., steps 9–14) the GPR reconstruction benefits from past and future observations. This results in lower errors and higher accuracy during this period.
- −
- The GPR reconstruction exhibits slightly higher errors when one-sided (either future or past) observations are available.) Predictions at these steps primarily rely on data from one temporal direction, reducing the overall information available for inference.
- Component-wise performance:
- −
- The meridional velocity (v) reconstruction has consistently smaller errors than the zonal velocity (u), as reflected in both accuracy and error metrics.
- −
- This disparity arises from the double-gyre dynamics, where perturbations are concentrated along the zonal direction (x-axis), leading to greater variability and complexity for u.
- Model robustness:
- −
- Despite minor temporal variations, the GPR model demonstrates consistent performance across all time steps, showcasing its ability to accurately reconstruct temporal velocity field dynamics.
3.1.2. Velocity Field Comparisons
3.1.3. Error Analysis
- Zonal velocity ():
- −
- Predictions for u exhibit slightly larger deviations from the line compared to v, especially at time-step 14. This disparity can be attributed to disturbances along the x-axis, where the larger spatial range introduces additional complexity for the GPR model.
- −
- Errors are more pronounced in regions with lower velocities, consistent with the trends identified in the relative error analysis.
- Meridional velocity ():
- −
- Predictions for v display tighter clustering around the line, reflecting higher accuracy compared to u.
- −
- This improved performance is likely due to the smaller spatial range in the v direction, which allows the model to better capture the underlying dynamics.
- Time-step variability:
- −
- Predictions at time-step 9 are more accurate than those at time-step 14. This trend aligns with earlier findings (Figure 4), where middle time steps benefit from both future and past observations.
3.1.4. Summary of Double-Gyre Model Results
- Overall performance: GPR demonstrated strong predictive accuracy, as evidenced by high values of metrics such as , D, and , which consistently approached 1 across all time steps. Error metrics (RMSE, MAE, and MBE) remained close to zero, underscoring the robustness of the model in capturing the dynamics of the velocity fields.
- Temporal trends: The GPR model performed best during the middle time steps (e.g., steps 9–14), leveraging information from both future and past observations. Slightly higher errors at the beginning and end of the time series (e.g., steps 1 and 21) reflect the limitations of one-sided information.
- Component-wise insights: The meridional velocity (v) consistently outperformed the zonal velocity (u), likely due to the perturbations in the zonal direction (x-axis) inherent to the double-gyre dynamics. This result aligns with the physical nature of the model, where the zonal direction exhibits higher variability and complexity.
- Error distribution: The relative error analysis revealed that regions with low velocity magnitudes posed the greatest challenge for GPR, with relative errors often exceeding . This limitation was most prominent near hyperbolic points and in low-velocity regions. The scatter plots further emphasized these discrepancies, highlighting larger deviations in the u component compared to v.
- Utility of predicted error (): has demonstrated its value as a reliable diagnostic tool by effectively correlating with regions of high uncertainty, significant deviations, and large relative errors. Beyond assessing prediction reliability, also highlights areas of data sparsity, which is particularly useful for refining model design and optimizing observational strategies. This dual functionality makes an essential component for improving GPR performance and identifying regions where additional data or adjustments are needed to enhance reconstruction accuracy.
- Reconstruction of velocity fields: Visual comparisons of the GPR-reconstructed and model-generated velocity fields demonstrated strong alignment, with the GPR model accurately capturing subtle dynamic features such as periodic tilts and gyre structures. Minor discrepancies were primarily localized in regions with sparse data or low velocities.
3.2. NCOM Convergence Region
3.2.1. Performance Metrics and Sampling Density Insights
- Broader region (135 km × 146 km): In the broader domain, increasing the sampling density from (20 drifters) to (200 drifters) reduces the overall relative error from approximately to . However, the improvement becomes marginal beyond a density of . This suggests that while increasing sampling density improves prediction accuracy, practical considerations such as drifter deployment feasibility must guide the selection of density levels.
- Confined region (9 km × 6 km): In the confined domain, higher densities achieve lower relative errors, approaching an optimal density of . Beyond this threshold, the performance improvement plateaus. This saturation may be attributed to inherent limitations in the GPR model’s complexity and the fidelity of the simulated drifter data. Identifying such thresholds is critical for balancing prediction accuracy with resource allocation.
- Trade-offs in broader regions: In larger domains, balancing prediction accuracy and practical constraints, such as cost and deployment logistics, is crucial. A sampling density of appears to offer an effective balance.
- Precision in confined regions: For smaller, high-priority areas, achieving densities close to maximizes accuracy while acknowledging diminishing returns beyond this threshold.
3.2.2. Velocity Field Comparisons
3.2.3. Error Analysis
- The zonal velocity (u) generally exhibits lower deviations and relative errors due to smoother flow patterns and fewer localized variations.
- The meridional velocity (v) shows higher errors, reflecting the challenges of capturing complex gradients and low-velocity regions.
- Data sparsity significantly influences error distribution, with both components experiencing larger deviations where particle coverage is limited.
- The predicted error () effectively identifies areas of high uncertainty, providing a valuable diagnostic tool for assessing the reliability of the GPR predictions.
- Zonal velocity (): deviations and relative errors are lower, benefiting from smoother gradients and large-scale flow dynamics.
- Meridional velocity (): errors are larger, particularly in regions with steep gradients and low velocities, due to the complexity of localized features.
- Data sparsity: errors are concentrated in regions with limited particle observations, underscoring the importance of adequate sampling density.
- Predicted error (): effectively identifies regions of high uncertainty, aligning well with observed deviations and relative errors.
- Low-velocity regions: relative errors significantly increase below , where the model struggles to predict both magnitude and direction accurately.
3.2.4. Summary of NCOM Model Results
- Hyperparameter configuration: The optimized hyperparameters reflect the multi-scale dynamics of the region. Larger spatial scales capture mesoscale patterns, while smaller scales highlight submesoscale variability. Correlation times reinforce the persistence of localized features and the transient nature of broader flow patterns, showcasing the adaptability of GPR in resolving multi-scale phenomena.
- Performance metrics and sampling density:
- −
- Sampling density plays a pivotal role in determining prediction accuracy.
- −
- In the broader region (135 km × 146 km), increasing sampling density from to significantly reduces relative errors. However, improvements plateau beyond this threshold, emphasizing the need for practical deployment strategies.
- −
- In the confined region (9 km × 6 km), optimal performance is achieved at approximately , where diminishing returns become evident.
- Velocity field comparisons: GPR successfully reconstructs both the magnitude and direction of the NCOM velocity fields (Figure 9). Minor discrepancies are observed in sparsely sampled areas, particularly near boundaries. The model’s robustness in replicating mesoscale and submesoscale flow structures is evident across the domain.
- Error analysis:
- −
- Grid-based analysis reveals elevated deviations and relative errors in regions with sparse observations (Figure 10).
- −
- The zonal velocity (u) exhibits lower errors due to smoother gradients, while the meridional velocity (v) shows higher deviations in areas of steep velocity gradients and low data density.
- −
- Predicted error () effectively correlates with regions of high uncertainty, offering a reliable tool for assessing prediction reliability.
- Scatter plot insights: Scatter plots (Figure 11) highlight strong agreement between GPR predictions and reference values at higher velocities. However, deviations increase in low-velocity regions, where larger predicted error values () are observed. This trend underscores the challenges of maintaining accuracy in low-velocity scenarios.
4. Discussion
4.1. Implications for Velocity Field Reconstruction and Ocean Circulation Studies
4.2. Future Directions and Enhancements
5. Conclusions
- Performance and accuracy: GPR demonstrates high accuracy in reconstructing ocean-like velocity fields, achieving overall accuracy levels exceeding 90%. This makes it a reliable tool for capturing both mesoscale and submesoscale dynamics in surface ocean studies. Furthermore, GPR’s posterior covariance matrix serves as a robust predictor of interpolation uncertainty, offering valuable insights into the reliability of reconstructed fields.
- Sampling density: Sampling density significantly influences prediction accuracy, with lower densities leading to greater errors, particularly in regions with sparse observational coverage. Our results identify an optimal sampling density of approximately seven data points per km2 per time-step, beyond which improvements plateau. This finding underscores the importance of strategic drifter deployment to balance accuracy and resource efficiency in data collection efforts.
- Velocity magnitude: GPR performs more reliably in faster-flowing regions, while accuracy diminishes in low-velocity areas (below 0.1 m/s). These errors, often exceeding 100% relative error, highlight challenges in accurately reconstructing directional flows in regions with minimal velocity variation. This limitation stems from GPR’s sensitivity to hyperparameter optimization, which tends to prioritize dynamic segments of the flow.
- Insights into temporal and spatial dependencies: The double-gyre model reveals GPR’s strength in leveraging temporal dependencies, achieving peak performance in intermediate time steps where both future and past observations are available. In contrast, the NCOM case highlights spatial dependencies, with errors concentrated in sparsely sampled or low-velocity regions. These results validate the necessity of evaluating GPR across both temporal and spatial dimensions to comprehensively understand its behavior.
6. Declaration of Generative AI and AI-Assisted Technologies in the Writing Process
Author Contributions
Funding
Institutional Review Board Statement
Informed Consent Statement
Data Availability Statement
Acknowledgments
Conflicts of Interest
References
- 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]
- Lodise, J.; Özgökmen, T.; Gonçalves, R.C.; Iskandarani, M.; Lund, B.; Horstmann, J.; Poulain, P.M.; Klymak, J.; Ryan, E.H.; Guigand, C. Investigating the formation of submesoscale structures along mesoscale fronts and estimating kinematic quantities using Lagrangian drifters. Fluids 2020, 5, 159. [Google Scholar] [CrossRef]
- Grossi, M.D.; Kubat, M.; Özgökmen, T.M. Predicting particle trajectories in oceanic flows using artificial neural networks. Ocean Model. 2020, 156, 101707. [Google Scholar] [CrossRef]
- Gupta, A.; Lermusiaux, P.F.J. Neural closure models for dynamical systems. Proc. Math. Phys. Eng. Sci. 2021, 477. [Google Scholar] [CrossRef]
- Hoteit, I.; Abualnaja, Y.; Afzal, S.; Ait-El-Fquih, B.; Akylas, T.; Antony, C.; Dawson, C.; Asfahani, K.; Brewin, R.J.; Cavaleri, L.; et al. Towards an end-to-end analysis and prediction system for weather, climate, and marine applications in the Red Sea. Bull. Am. Meteorol. Soc. 2021, 102, E99–E122. [Google Scholar] [CrossRef]
- Mariano, A.J.; Griffa, A.; Özgökmen, T.M.; Zambianchi, E. Lagrangian analysis and predictability of coastal and ocean dynamics 2000. J. Atmos. Ocean. Technol. 2002, 19, 1114–1126. [Google Scholar] [CrossRef]
- Poje, A.C.; Ozgökmen, T.M.; Lipphardt, B.L., Jr.; Haus, B.K.; Ryan, E.H.; Haza, A.C.; Jacobs, G.A.; Reniers, A.J.H.M.; Olascoaga, M.J.; Novelli, G.; et al. Submesoscale dispersion in the vicinity of the Deepwater Horizon spill. Proc. Natl. Acad. Sci. USA 2014, 111, 12693–12698. [Google Scholar] [CrossRef]
- Lumpkin, R.; Özgökmen, T.; Centurioni, L. Advances in the application of surface drifters. Ann. Rev. Mar. Sci. 2017, 9, 59–81. [Google Scholar] [CrossRef]
- Novelli, G.; Guigand, C.M.; Cousin, C.; Ryan, E.H.; Laxague, N.J.M.; Dai, H.; Haus, B.K.; Özgökmen, T.M. A biodegradable surface drifter for ocean sampling on a massive scale. J. Atmos. Ocean. Technol. 2017, 34, 2509–2532. [Google Scholar] [CrossRef]
- Dasaro, E.A.; Shcherbina, A.Y.; Klymak, J.M.; Molemaker, J.; Novelli, G.; Guigand, C.M.; Özgökmen, T.M. Ocean convergence and the dispersion of flotsam. Proc. Natl. Acad. Sci. USA 2018, 115, 1162–1167. [Google Scholar] [CrossRef]
- Griffa, A.; Piterbarg, L.I.; Özgökmen, T. Predictability of Lagrangian particle trajectories: Effects of smoothing of the underlying Eulerian flow. J. Mar. Res. 2004, 62, 1–35. [Google Scholar] [CrossRef]
- Aref, H. Stirring by chaotic advection. J. Fluid Mech. 1984, 143, 1–21. [Google Scholar] [CrossRef]
- Hernandez, F.; Le Traon, P.Y.; Morrow, R. Mapping mesoscale variability of the Azores Current using TOPEX/POSEIDON and ERS 1 altimetry, together with hydrographic and Lagrangian measurements. J. Geophys. Res. 1995, 100, 24995. [Google Scholar] [CrossRef]
- Molcard, A. Assimilation of drifter observations for the reconstruction of the Eulerian circulation field. J. Geophys. Res. 2003, 108. [Google Scholar] [CrossRef]
- Molcard, A.; Griffa, A.; Özgökmen, T.M. Lagrangian data assimilation in multilayer primitive equation ocean models. J. Atmos. Ocean. Technol. 2005, 22, 70–83. [Google Scholar] [CrossRef]
- Toner, M.; Kirwan, A., Jr.; Kantha, L.; Choi, J. Can general circulation models be assessed and their output enhanced with drifter data? J. Geophys. Res. Ocean. 2001, 106, 19563–19579. [Google Scholar] [CrossRef]
- Ide, K.; Kuznetsov, L.; Jones, C.K. Lagrangian data assimilation for point vortex systems. J. Turbul. 2002, 3. [Google Scholar] [CrossRef]
- Kuznetsov, L.; Ide, K.; Jones, C.K.R.T. A method for assimilation of Lagrangian data. Mon. Weather Rev. 2003, 131, 2247–2260. [Google Scholar] [CrossRef]
- Salman, H.; Kuznetsov, L.; Jones, C.K.R.T.; Ide, K. A method for assimilating Lagrangian data into a Shallow-Water-equation ocean model. Mon. Weather Rev. 2006, 134, 1081–1101. [Google Scholar] [CrossRef]
- Kamachi, M.; O’Brien, J.J. Continuous data assimilation of drifting buoy trajectory into an equatorial Pacific Ocean model. J. Mar. Syst. 1995, 6, 159–178. [Google Scholar] [CrossRef]
- Taillandier, V.; Griffa, A.; Molcard, A. A variational approach for the reconstruction of regional scale Eulerian velocity fields from Lagrangian data. Ocean Model. 2006, 13, 1–24. [Google Scholar] [CrossRef]
- Taillandier, V.; Griffa, A. Implementation of position assimilation for ARGO floats in a realistic Mediterranean Sea OPA model and twin experiment testing. Ocean Sci. 2006, 2, 223–236. [Google Scholar] [CrossRef]
- Salman, H. A hybrid grid/particle filter for Lagrangian data assimilation. II: Application to a model vortex flow. Q. J. R. Meteorol. Soc. 2008, 134, 1551–1565. [Google Scholar] [CrossRef]
- Salman, H. A hybrid grid/particle filter for Lagrangian data assimilation. I: Formulating the passive scalar approximation. Q. J. R. Meteorol. Soc. 2008, 134, 1539–1550. [Google Scholar] [CrossRef]
- Krause, P.; Restrepo, J.M. The diffusion kernel filter applied to Lagrangian data assimilation. Mon. Weather Rev. 2009, 137, 4386–4400. [Google Scholar] [CrossRef]
- Kennedy, M.C.; O’Hagan, A. Predicting the output from a complex computer code when fast approximations are available. Biometrika 2000, 87, 1–13. [Google Scholar] [CrossRef]
- Rasmussen, C.E.; Williams, C.K.I. Gaussian Processes for Machine Learning; MIT Press: Cambridge, MA, USA, 2006. [Google Scholar]
- Thacker, W.C.; Iskandarani, M.; Gonçalves, R.C.; Srinivasan, A.; Knio, O.M. Pragmatic aspects of uncertainty propagation: A conceptual review. Ocean Model. 2015, 95, 25–36. [Google Scholar] [CrossRef]
- Iskandarani, M.; Wang, S.; Srinivasan, A.; Thacker, W.C.; Winokur, J.; Knio, O.M. An overview of uncertainty quantification techniques with application to oceanic and oilspill simulations. J. Geophys. Res. Ocean. 2016, 121, 2789–2808. [Google Scholar] [CrossRef]
- Gonçalves, R.C.; Iskandarani, M.; Özgökmen, T.; Thacker, W.C. Reconstruction of submesoscale velocity field from surface drifters. J. Phys. Oceanogr. 2019, 49, 941–958. [Google Scholar] [CrossRef]
- Berlinghieri, R.; Trippe, B.L.; Burt, D.R.; Giordano, R.; Srinivasan, K.; Özgökmen, T.; Xia, J.; Broderick, T. Gaussian processes at the Helm(holtz): A more fluid model for ocean currents. arXiv 2023, arXiv:2302.10364. [Google Scholar]
- Walder, C.; Kim, K.I.; Schölkopf, B. Sparse multiscale Gaussian process regression. In Proceedings of the 25th International Conference on Machine Learning, Helsinki, Finland, 5–9 July 2008; pp. 1112–1119. [Google Scholar]
- Stephenson, D.; Kermode, J.R.; Lockerby, D.A. Accelerating multiscale modelling of fluids with on-the-fly Gaussian process regression. Microfluid. Nanofluidics 2018, 22, 139. [Google Scholar] [CrossRef] [PubMed]
- Li, Z.; McWilliams, J.C.; Ide, K.; Farrara, J.D. A multiscale variational data assimilation scheme: Formulation and illustration. Mon. Weather Rev. 2015, 143, 3804–3822. [Google Scholar] [CrossRef]
- Le Traon, P.Y. A method for optimal analysis of fields with spatially variable mean. J. Geophys. Res. Ocean. 1990, 95, 13543–13547. [Google Scholar] [CrossRef]
- Rohani, A.; Taki, M.; Abdollahpour, M. A novel soft computing model (Gaussian process regression with K-fold cross validation) for daily and monthly solar radiation forecasting (Part: I). Renew. Energy 2018, 115, 411–422. [Google Scholar] [CrossRef]
- Martino, L.; Laparra, V.; Camps-Valls, G. Probabilistic cross-validation estimators for Gaussian process regression. In Proceedings of the 2017 25th European Signal Processing Conference (EUSIPCO), Kos, Greece, 28 August–2 September 2017; IEEE: Piscataway, NJ, USA, 2017; pp. 823–827. [Google Scholar]
- Stone, M. Cross-validation: A review. Stat. A J. Theor. Appl. Stat. 1978, 9, 127–139. [Google Scholar]
- Jacobs, G.A.; Bartels, B.P.; Bogucki, D.J.; Beron-Vera, F.J.; Chen, S.S.; Coelho, E.F.; Curcic, M.; Griffa, A.; Gough, M.; Haus, B.K.; et al. Data assimilation considerations for improved ocean predictability during the Gulf of Mexico Grand Lagrangian Deployment (GLAD). Ocean Model. 2014, 83, 98–117. [Google Scholar] [CrossRef]
- Albert, C.G.; Rath, K. Gaussian process regression for data fulfilling linear differential equations with localized sources. Entropy 2020, 22, 152. [Google Scholar] [CrossRef]
- Bretherton, F.P.; Davis, R.E.; Fandry, C. A technique for objective analysis and design of oceanographic experiments applied to MODE-73. Deep. Sea Res. Oceanogr. Abstr. 1976, 23, 559–582. [Google Scholar] [CrossRef]
- Byrd, R.H.; Lu, P.; Nocedal, J.; Zhu, C. A limited memory algorithm for bound constrained optimization. SIAM J. Sci. Comput. 1995, 16, 1190–1208. [Google Scholar] [CrossRef]
- Willmott, C.J. On the validation of models. Phys. Geogr. 1981, 2, 184–194. [Google Scholar] [CrossRef]
- Greenwood, D.; Neeteson, J.; Draycott, A. Response of potatoes to N fertilizer: Dynamic model. Plant Soil 1985, 85, 185–203. [Google Scholar] [CrossRef]
- Shadden, S.C.; Lekien, F.; Marsden, J.E. Definition and properties of Lagrangian coherent structures from finite-time Lyapunov exponents in two-dimensional aperiodic flows. Physica D 2005, 212, 271–304. [Google Scholar] [CrossRef]
- Froyland, G.; Padberg-Gehle, K. Almost-invariant and finite-time coherent sets: Directionality, duration, and diffusion. In Ergodic Theory, Open Dynamics, and Coherent Structures; Springer Proceedings in Mathematics & Statistics; Springer: New York, NY, USA, 2014; pp. 171–216. [Google Scholar]
- Martin, P.J. Description of the navy coastal ocean model version 1.0. NRL Rep. NRL/FR/7322-00 2000, 9962, 42. [Google Scholar]
- Kamath, A.; Vargas-Hernández, R.A.; Krems, R.V.; Carrington, T.; Manzhos, S. Neural networks vs Gaussian process regression for representing potential energy surfaces: A comparative study of fit quality and vibrational spectrum accuracy. J. Chem. Phys. 2018, 148, 241702. [Google Scholar] [CrossRef]
- Ma, Y.; He, Y.; Wang, L.; Zhang, J. Probabilistic reconstruction for spatiotemporal sensor data integrated with Gaussian process regression. Probabilistic Eng. Mech. 2022, 69, 103264. [Google Scholar] [CrossRef]
- Camps-Valls, G.; Martino, L.; Svendsen, D.H.; Campos-Taberner, M.; Muñoz-Marí, J.; Laparra, V.; Luengo, D.; García-Haro, F.J. Physics-aware Gaussian processes in remote sensing. Appl. Soft Comput. 2018, 68, 69–82. [Google Scholar] [CrossRef]
- Zhou, J.; Jia, L.; Menenti, M.; Gorte, B. On the performance of remote sensing time series reconstruction methods–A spatial comparison. Remote Sens. Environ. 2016, 187, 367–384. [Google Scholar] [CrossRef]
- Li, T.; Biferale, L.; Bonaccorso, F.; Buzzicotti, M.; Centurioni, L. Stochastic Reconstruction of Gappy Lagrangian Turbulent Signals by Conditional Diffusion Models. arXiv 2024, arXiv:2410.23971. [Google Scholar]
- Bang, C.; Altaher, A.S.; Zhuang, H.; Altaher, A.; Srinivasan, A.; Cherubin, L. Physics-Informed Neural Networks to Reconstruct Surface Velocity Field from Drifter Data. Authorea Preprints 2024.
- Aravind, H.; Özgökmen, T.M.; Allshouse, M.R. Lagrangian analysis of submesoscale flows from sparse data using Gaussian Process Regression for field reconstruction. Ocean Model. 2025, 193, 102458. [Google Scholar] [CrossRef]
- Molinari, R.; Kirwan, A.D. Calculations of differential kinematic properties from Lagrangian observations in the western Caribbean Sea. J. Phys. Oceanogr. 1975, 5, 483–491. [Google Scholar] [CrossRef]
- Berta, M.; Griffa, A.; Özgökmen, T.M.; Poje, A.C. Submesoscale evolution of surface drifter triads in the Gulf of Mexico. Geophys. Res. Lett. 2016, 43, 11–751. [Google Scholar] [CrossRef]
- Mao, M.; Xia, M. Modeling blue crab (Callinectes sapidus) larval transport and recruitment dynamics in a shallow lagoon-inlet-coastal ocean system. J. Geophys. Res. Ocean. 2024, 129, e2023JC020785. [Google Scholar] [CrossRef]
- Stohl, A.; Wotawa, G.; Seibert, P.; Kromp-Kolb, H. Interpolation errors in wind fields as a function of spatial and temporal resolution and their impact on different types of kinematic trajectories. J. Appl. Meteorol. Climatol. 1995, 34, 2149–2165. [Google Scholar] [CrossRef]
- Walmsley, J.L.; Mailhot, J. On the numerical accuracy of trajectory models for long-range transport of atmospheric pollutants. Atmosphere-Ocean 1983, 21, 14–39. [Google Scholar] [CrossRef]
- Srinivasan, A.; Chin, T.; Chassignet, E.; Iskandarani, M.; Groves, N. A statistical interpolation code for ocean analysis and forecasting. J. Atmos. Ocean. Technol. 2022, 39, 367–386. [Google Scholar] [CrossRef]
- Mao, M.; Xia, M. Particle dynamics in the nearshore of Lake Michigan revealed by an observation-modeling system. J. Geophys. Res. Ocean. 2020, 125, e2019JC015765. [Google Scholar] [CrossRef]
- Fitzenreiter, K.; Mao, M.; Xia, M. Characteristics of surface currents in a shallow lagoon–inlet–coastal ocean system revealed by surface drifter observations. Estuaries Coasts 2022, 45, 2327–2344. [Google Scholar] [CrossRef]
- Ho, A.; Citrin, J.; Auriemma, F.; Bourdelle, C.; Casson, F.J.; Kim, H.T.; Manas, P.; Szepesi, G.; Weisen, H.; Contributors, J. Application of Gaussian process regression to plasma turbulent transport model validation via integrated modelling. Nucl. Fusion 2019, 59, 056007. [Google Scholar] [CrossRef]
- Zhang, Z.J.; Duraisamy, K. Machine learning methods for data-driven turbulence modeling. In Proceedings of the 22nd AIAA Computational Fluid Dynamics Conference, Dallas, TX, USA, 22–26 June 2015; p. 2460. [Google Scholar]
- Quinonero-Candela, J.; Rasmussen, C.E. A unifying view of sparse approximate Gaussian process regression. J. Mach. Learn. Res. 2005, 6, 1939–1959. [Google Scholar]
- Smola, A.; Bartlett, P. Sparse greedy Gaussian process regression. Adv. Neural Inf. Process. Syst. 2000, 13. [Google Scholar]
- Mohammadi, H.; Challenor, P.; Williamson, D.; Goodfellow, M. Cross-Validation–based Adaptive Sampling for Gaussian Process Models. SIAM/ASA J. Uncertain. Quantif. 2022, 10, 294–316. [Google Scholar] [CrossRef]
1. | n particles are randomly distributed in the research region. |
2. | The initial condition (zonal velocity ) and meridional velocity )) of the n particles are obtained from the NCOM velocity field through 16 points of nearest-neighbor interpolation. |
3. | The particles travel with the condition for one time-step , and the locations are then updated as . |
4. | Since the velocity fields do not change with time, only spatial interpolation is needed to update the and , based on the NCOM velocity fields and new locations. |
5. | Repeat steps 3 and 4 to complete the entire period. |
Hyperparameters | u | v |
---|---|---|
1.18 × 10−20 | 9.75 × 10−5 | |
0.161 | 0.152 | |
(steps) | 13.32 | 15.48 |
0.674 | 0.928 | |
0.779 | 1.01 | |
0.000634 | 0.00196 | |
(steps) | 114.52 | 247.52 |
0.00933 | 0.0127 | |
0.00546 | 0.931 |
Hyperparameters | u | v |
---|---|---|
(m s−1) | 0.000133 | 0.000356 |
(m s−1) | 0.0634 | 0.312 |
559 | 1147 | |
23.7 | 16.2 | |
12.1 | 43.9 | |
(m s−1) | 0.0016 | 0.00554 |
336 | 470 | |
3.65 | 3.11 | |
3.16 | 4.72 |
Num of Drifters | Density (n/km2) | D | |||||
---|---|---|---|---|---|---|---|
1 | 0.0012 | −0.55 | 0.49 | 0.11 | 0.25 | 0.38 | 2.0 × 10−5 |
2 | 0.0024 | −0.56 | 0.49 | 0.087 | 0.25 | 0.38 | 2.0 × 10−5 |
5 | 0.0061 | −0.43 | 0.52 | 0.13 | 0.23 | 0.37 | 1.9 × 10−5 |
10 | 0.012 | −0.30 | 0.56 | 0.18 | 0.21 | 0.35 | 1.7 × 10−5 |
20 | 0.024 | 0.82 | 0.95 | 0.84 | 0.039 | 0.13 | 4.9 × 10−6 |
30 | 0.037 | 0.85 | 0.96 | 0.85 | 0.017 | 0.12 | 4.5 × 10−6 |
40 | 0.049 | 0.94 | 0.98 | 0.94 | −0.012 | 0.078 | 2.85 × 10−6 |
50 | 0.061 | 0.94 | 0.99 | 0.95 | −0.0037 | 0.072 | 2.66 × 10−6 |
100 | 0.12 | 0.98 | 0.99 | 0.98 | 1.2 × 10−4 | 0.04 | 1.3 × 10−6 |
200 | 0.24 | 0.98 | 0.99 | 0.98 | −0.0013 | 0.038 | 1.1 × 10−6 |
Num of Drifters | Density (n/km2) | D | |||||
---|---|---|---|---|---|---|---|
1 | 0.0012 | −4.6 | 0.46 | 0.078 | 0.63 | 0.83 | 4.5 × 10−5 |
2 | 0.0024 | −2.7 | 0.59 | 0.26 | 0.34 | 0.67 | 3.6 × 10−5 |
5 | 0.0061 | −1.8 | 0.62 | 0.25 | 0.30 | 0.58 | 2.8 × 10−5 |
10 | 0.012 | 0.39 | 0.85 | 0.64 | 0.16 | 0.27 | 1.3 × 10−5 |
20 | 0.024 | 0.68 | 0.92 | 0.71 | 0.019 | 0.20 | 7.8 × 10−6 |
30 | 0.037 | 0.83 | 0.95 | 0.85 | 0.037 | 0.14 | 5.0 × 10−6 |
40 | 0.049 | 0.85 | 0.96 | 0.86 | 0.028 | 0.13 | 4.8 × 10−6 |
50 | 0.061 | 0.86 | 0.96 | 0.86 | 0.023 | 0.13 | 4.5 × 10−6 |
100 | 0.12 | 0.96 | 0.99 | 0.96 | 0.024 | 0.075 | 2.3 × 10−6 |
200 | 0.24 | 0.98 | 0.99 | 0.98 | 0.0059 | 0.052 | 1.5 × 10−6 |
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
Xia, J.; Iskandarani, M.; Gonçalves, R.C.; Özgökmen, T. Error Quantification of Gaussian Process Regression for Extracting Eulerian Velocity Fields from Ocean Drifters. J. Mar. Sci. Eng. 2025, 13, 431. https://doi.org/10.3390/jmse13030431
Xia J, Iskandarani M, Gonçalves RC, Özgökmen T. Error Quantification of Gaussian Process Regression for Extracting Eulerian Velocity Fields from Ocean Drifters. Journal of Marine Science and Engineering. 2025; 13(3):431. https://doi.org/10.3390/jmse13030431
Chicago/Turabian StyleXia, Junfei, Mohamed Iskandarani, Rafael C. Gonçalves, and Tamay Özgökmen. 2025. "Error Quantification of Gaussian Process Regression for Extracting Eulerian Velocity Fields from Ocean Drifters" Journal of Marine Science and Engineering 13, no. 3: 431. https://doi.org/10.3390/jmse13030431
APA StyleXia, J., Iskandarani, M., Gonçalves, R. C., & Özgökmen, T. (2025). Error Quantification of Gaussian Process Regression for Extracting Eulerian Velocity Fields from Ocean Drifters. Journal of Marine Science and Engineering, 13(3), 431. https://doi.org/10.3390/jmse13030431