Skip to main content
Log in

Three-dimensional time-resolved Lagrangian flow field reconstruction based on constrained least squares and stable radial basis function

  • Research Article
  • Published:
Experiments in Fluids Aims and scope Submit manuscript

Abstract

The three-dimensional time-resolved Lagrangian particle tracking (3D TR-LPT) technique has recently advanced flow diagnostics by providing high spatiotemporal resolution measurements under the Lagrangian framework. To fully exploit its potential, accurate and robust data processing algorithms are needed. These algorithms are responsible for reconstructing particle trajectories, velocities, and differential quantities (e.g., pressure gradients, strain- and rotation-rate tensors, and coherent structures) from raw LPT data. In this paper, we propose a 3D divergence-free Lagrangian reconstruction method, where three foundation algorithms—constrained least squares (CLS), stable radial basis function (RBF-QR), and partition-of-unity method (PUM)—are integrated into one comprehensive reconstruction strategy. Our method, named CLS-RBF PUM, is able to (1) directly reconstruct flow fields at scattered data points, avoiding Lagrangian-to-Eulerian data conversions; (2) assimilate the flow diagnostics in Lagrangian and Eulerian descriptions to achieve high-accuracy reconstruction; (3) process large-scale LPT datasets with more than hundreds of thousand particles in two dimensions (2D) or 3D; (4) enable inter-frame and inter-particle interpolation while imposing physical constraints (e.g., divergence-free for incompressible flows) at arbitrary time and locations. Validation based on synthetic and experimental LPT data confirm that our method can achieve the above advantages with accuracy and robustness.

Graphical abstract

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11

Similar content being viewed by others

Availability of data and materials

Datasets and codes are available from the corresponding author, Z.P. on resaonable request.

Notes

  1. In the current experimental data, the maximum number of particles that existed in all 11 frames was about 7,100.

  2. The number of frames is set to 11 throughout the paper for the following reasons: (1) directly using more frames without the PUM in time is computationally expensive; (2) having fewer frames might be insufficient to resolve complex trajectories.

References

  • Alberini F, Liu L, Stitt E, Simmons M (2017) Comparison between 3-D-PTV and 2-D-PIV for determination of hydrodynamics of complex fluids in a stirred vessel. Chem Eng Sci 171:189–203

    Google Scholar 

  • Babuška I, Melenk JM (1997) The partition of unity method. Int J Numer Methods Eng 40(4):727–758

    MathSciNet  Google Scholar 

  • Biferale L, Bonaccorso F, Buzzicotti M, Calascibetta C (2023) TURB-Lagr. A database of 3d Lagrangian trajectories in homogeneous and isotropic turbulence. arXiv preprint arXiv:2303.08662

  • Biwole PH, Yan W, Zhang Y, Roux J-J (2009) A complete 3D particle tracking algorithm and its applications to the indoor airflow study. Meas Sci Technol 20(11):115403

    Google Scholar 

  • Bobrov M, Hrebtov M, Ivashchenko V, Mullyadzhanov R, Seredkin A, Tokarev M, Zaripov D, Dulin V, Markovich D (2021) Pressure evaluation from Lagrangian particle tracking data using a grid-free least-squares method. Meas Sci Technol 32(8):084014

    Google Scholar 

  • Bowker KA (1988) Albert Einstein and meandering rivers. Earth Sci Hist 7(1):45

    Google Scholar 

  • Carr JC, Beatson RK, Cherrie JB, Mitchell TJ, Fright WR, McCallum BC, Evans TR (2001) Reconstruction and representation of 3D objects with radial basis functions. In: Proceedings of the 28th annual conference on Computer graphics and interactive techniques, pp 67–76

  • Cierpka C, Lütke B, Kähler CJ (2013) Higher order multi-frame particle tracking velocimetry. Exp Fluids 54:1–12

    Google Scholar 

  • Drake KP, Fuselier EJ, Wright GB (2022) Implicit surface reconstruction with a curl-free radial basis function partition of unity method. SIAM J Sci Comput 44(5):A3018–A3040

    MathSciNet  Google Scholar 

  • Einstein A (1926) Die Ursache der Mäanderbildung der Flußläufe und des sogenannten Baerschen Gesetzes. Naturwissenschaften 14(11):223–224

    Google Scholar 

  • Fasshauer GE (2007) Meshfree approximation methods with MATLAB, vol 6. World Scientific, Singapore

    Google Scholar 

  • Ferrari S, Rossi L (2008) Particle tracking velocimetry and accelerometry (PTVA) measurements applied to quasi-two-dimensional multi-scale flows. Exp Fluids 44:873–886

    Google Scholar 

  • Fornberg B, Piret C (2008) A stable algorithm for flat radial basis functions on a sphere. SIAM J Sci Comput 30(1):60–80

    MathSciNet  Google Scholar 

  • Fornberg B, Wright G (2004) Stable computation of multiquadric interpolants for all values of the shape parameter. Comput Math Appl 48(5–6):853–867

    MathSciNet  Google Scholar 

  • Fornberg B, Larsson E, Flyer N (2011) Stable computations with Gaussian radial basis functions. SIAM J Sci Comput 33(2):869–892

    MathSciNet  Google Scholar 

  • Fornberg B, Lehto E, Powell C (2013) Stable calculation of Gaussian-based RBF-FD stencils. Comput Math Appl 65(4):627–637

    MathSciNet  Google Scholar 

  • Fu S, Biwole PH, Mathis C (2015) Particle tracking velocimetry for indoor airflow field: a review. Build Environ 87:34–44

    Google Scholar 

  • Gautschi W (2011) Numerical analysis. Springer, Berlin

    Google Scholar 

  • Gesemann S, Huhn F, Schanz D, Schröder A (2016) From noisy particle tracks to velocity, acceleration and pressure fields using B-splines and penalties. In: 18th international symposium on applications of laser and imaging techniques to fluid mechanics, Lisbon, Portugal, vol 4

  • Haller G (2005) An objective definition of a vortex. J Fluid Mech 525:1–26

    MathSciNet  Google Scholar 

  • Huang G-B, Chen L, Siew CK et al (2006) Universal approximation using incremental constructive feedforward networks with random hidden nodes. IEEE Trans Neural Netw 17(4):879–892

    Google Scholar 

  • Hunt JC, Wray AA, Moin P (1988) Eddies, streams, and convergence zones in turbulent flows. In: Proceedings of the 1988 summer program studying turbulence using numerical simulation databases, vol 2

  • Jeon YJ, Müller M, Michaelis D (2022) Fine scale reconstruction (VIC#) by implementing additional constraints and coarse-grid approximation into VIC+. Exp Fluids 63(4):70

    Google Scholar 

  • Jeronimo MD, Zhang K, Rival DE (2019) Direct Lagrangian measurements of particle residence time. Exp Fluids 60:1–11

    Google Scholar 

  • Keerthi SS, Lin C-J (2003) Asymptotic behaviors of support vector machines with Gaussian kernel. Neural Comput 15(7):1667–1689

    Google Scholar 

  • Khojasteh AR, Laizet S, Heitz D, Yang Y (2022) Lagrangian and Eulerian dataset of the wake downstream of a smooth cylinder at a Reynolds number equal to 3900. Data Brief 40:107725

    Google Scholar 

  • Laizet S, Lamballais E (2009) High-order compact schemes for incompressible flows: a simple and efficient method with quasi-spectral accuracy. J Comput Phys 228(16):5989–6015

    MathSciNet  Google Scholar 

  • Larsson E (2023) Private communication

  • Larsson E, Lehto E, Heryudono A, Fornberg B (2013) Stable computation of differentiation matrices and scattered node stencils based on Gaussian radial basis functions. SIAM J Sci Comput 35(4):A2096–A2119

    MathSciNet  Google Scholar 

  • Larsson E, Shcherbakov V, Heryudono A (2017) A least squares radial basis function partition of unity method for solving PDEs. SIAM J Sci Comput 39(6):A2538–A2563

    MathSciNet  Google Scholar 

  • Li L (2023) Lagrangian flow field reconstruction based on constrained stable radial basis function. Master’s thesis, University of Waterloo

  • Li L, Pan Z (2023) Three dimensional divergence-free Lagrangian reconstruction based on constrained least squares and stable radial basis function. In: 15th international symposium on particle image velocimetry, vol 1

  • Li L, Sakib N, Pan Z (2022) Robust strain/rotation-rate tensor reconstruction based on least squares RBF-QR for 3D Lagrangian particle tracking. In: Fluids engineering division summer meeting, volume 85833, page V001T02A009. American Society of Mechanical Engineers

  • Lüthi B (2002) Some aspects of strain, vorticity and material element dynamics as measured with 3D particle tracking velocimetry in a turbulent flow. PhD thesis. ETH Zurich

  • Macêdo I, Gois JP, Velho L (2011) Hermite radial basis functions implicits. In: Computer graphics forum, vol 30. Wiley Online Library, pp 27–42

  • Machicoane N, López-Caballero M, Bourgoin M, Aliseda A, Volk R (2017) A multi-time-step noise reduction method for measuring velocity statistics from particle tracking velocimetry. Meas Sci Technol 28(10):107002

    Google Scholar 

  • Malik NA, Dracos T (1993) Lagrangian PTV in 3D flows. Appl Sci Res 51:161–166

    Google Scholar 

  • Melenk JM, Babuška I (1996) The partition of unity finite element method: basic theory and applications. Comput Methods Appl Mech Eng 139(1–4):289–314

    MathSciNet  Google Scholar 

  • Onu K, Huhn F, Haller G (2015) LCS Tool: a computational platform for Lagrangian coherent structures. J Comput Sci 7:26–36

    Google Scholar 

  • Ouellette NT, Xu H, Bodenschatz E (2006) A quantitative study of three-dimensional Lagrangian particle tracking algorithms. Exp Fluids 40:301–313

    Google Scholar 

  • Peacock T, Haller G (2013) Lagrangian coherent structures: the hidden skeleton of fluid flows. Phys Today 66(2):41–47

    Google Scholar 

  • Peng J, Dabiri J (2009) Transport of inertial particles by Lagrangian coherent structures: application to predator–prey interaction in jellyfish feeding. J Fluid Mech 623:75–84

    Google Scholar 

  • Romano M, Alberini F, Liu L, Simmons M, Stitt E (2021) Lagrangian investigations of a stirred tank fluid flow using 3d-PTV. Chem Eng Res Des 172:71–83

    Google Scholar 

  • Rosi GA, Walker AM, Rival DE (2015) Lagrangian coherent structure identification using a Voronoi tessellation-based networking algorithm. Exp Fluids 56:1–14

    Google Scholar 

  • Sakib MN (2022) Pressure from PIV for an oscillating internal flow. Ph.D. thesis, Utah State University

  • Scarano F, Riethmuller ML (2000) Advances in iterative multigrid PIV image processing. Exp Fluids 29(Suppl 1):S051–S060

    Google Scholar 

  • Schanz D, Gesemann S, Schröder A (2016) Shake-the-box: Lagrangian particle tracking at high particle image densities. Exp Fluids 57(5):1–27

    Google Scholar 

  • Schneiders JF, Scarano F (2016) Dense velocity reconstruction from tomographic PTV with material derivatives. Exp Fluids 57:1–22

    Google Scholar 

  • Schneiders JF, Scarano F, Elsinga GE (2017) Resolving vorticity and dissipation in a turbulent boundary layer by tomographic PTV and VIC+. Exp Fluids 58:1–14

    Google Scholar 

  • Shepard D (1968) A two-dimensional interpolation function for irregularly-spaced data. In: Proceedings of the 1968 23rd ACM national conference, pp 517–524

  • Sperotto P, Pieraccini S, Mendez MA (2022) A meshless method to compute pressure fields from image velocimetry. Meas Sci Technol 33(9):094005

    Google Scholar 

  • Takehara K, Etoh T (2017) Direct evaluation method of vorticity combined with PTV. In: Selected papers from the 31st international congress on high-speed imaging and photonics, vol 10328. SPIE, pp 193–199

  • Tan S, Salibindla A, Masuk AUM, Ni R (2020) Introducing OpenLPT: new method of removing ghost particles and high-concentration particle shadow tracking. Exp Fluids 61:1–16

    Google Scholar 

  • Taylor GI, Green AE (1937) Mechanism of the production of small eddies from large ones. Proc R Soc Lond Ser A Math Phys Sci 158(895):499–521

    Google Scholar 

  • Wendland H (1995) Piecewise polynomial, positive definite and compactly supported radial functions of minimal degree. Adv Comput Math 4:389–396

    MathSciNet  Google Scholar 

  • Wilson MM, Peng J, Dabiri JO, Eldredge JD (2009) Lagrangian coherent structures in low Reynolds number swimming. J Phys Condens Matter 21(20):204105

    Google Scholar 

  • Wright GB, Fornberg B (2017) Stable computations with flat radial basis functions using vector-valued rational approximations. J Comput Phys 331:137–156

    MathSciNet  Google Scholar 

  • Zhang K, Rival DE (2020) Direct Lagrangian method to characterize entrainment dynamics using particle residence time: a case study on a laminar separation bubble. Exp Fluids 61(12):243

    Google Scholar 

Download references

Acknowledgements

We thank Prof. Elizabeth Larsson at Uppsala University for discussing the RBF-QR and sharing codes with us. We are also thankful to Dr. Md Nazmus Sakib at Utah State University for conducting experiments and providing data and timely support. L.L. thanks for the support from the Natural Sciences and Engineering Research Council of Canada (NSERC) and the Digital Research Alliance of Canada. We thank an anonymous reviewer 2 for thorough comments and constructive suggestions.

Funding

This research is partially supported by the Natural Sciences and Engineering Research Council of Canada (NSERC), 2020-04486.

Author information

Authors and Affiliations

Authors

Contributions

Z.P. conceived the research. L.L. and Z.P. programmed the algorithm and performed the computations. All authors discussed the results and contributed to the final manuscript.

Corresponding author

Correspondence to Zhao Pan.

Ethics declarations

Ethical approval

Not applicable.

Conflict of interest

The authors declare that they have no conflict of interest.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendices

A Differential quantity

1.1 A.1 Strain-rate and rotation-rate tensor

Strain-rate and rotation-rate tensors are kinematics quantities that describe the rate of change of a fluid parcel regarding deformation and rotation, respectively. The strain-rate tensor \({{\varvec{S}}}\) and rotation-rate tensor \({{\varvec{R}}}\) are the symmetric and anti-symmetric parts of the velocity gradient \(\nabla {{\textbf {U}}}\), respectively: \(\nabla {{\textbf {U}}} =\frac{1}{2}( U _{i,j}+ U _{j,i})+\frac{1}{2}( U _{i,j}- U _{j,i})\), \({{\varvec{S}}} = S_{ij} =\frac{1}{2}( U _{i,j}+ U _{j,i})\), and \({{\varvec{R}}} = R_{ij} =\frac{1}{2}( U _{i,j}- U _{j,i})\). Once all components of \(U _{i,j}\) are calculated in Step 4, \({{\varvec{S}}}\) and \({{\varvec{R}}}\) can be evaluated.

1.2 A.2 Coherent structure based on Q-criterion

The Q-criterion is a vortex identification method first proposed by Hunt et al. (1988) to identify vortical structures in the flow. The Q-criterion is given by (Hunt et al. 1988; Haller 2005):

$$\begin{aligned} Q=\frac{1}{2}(\Vert {{\varvec{R}}}^2 \Vert - \Vert {\varvec{S}}^2 \Vert )>0 \end{aligned}$$
(15)

where \({\varvec{R}}\) is the rotation-rate tensor, and \({\varvec{S}}\) is the strain-rate tensor. After \({{\varvec{S}}}\) and \({{\varvec{R}}}\) are calculated, vortices can be found using Eq. (15). Other criteria can be achieved similarly.

B Synthetic and experimental data

1.1 B.1 Taylor–Green vortex (TGV)

To generate the TGV synthetic data for validations, we adopted a velocity field:

$$\begin{aligned} \begin{aligned} u&= \alpha _1 \cos {(\omega (x-d_x))}\sin {(\omega (y-d_y))}\sin {(\omega (z-d_z))} \\ v&= \alpha _2 \sin {(\omega (x-d_x))}\cos {(\omega (y-d_y))}\sin {(\omega (z-d_z))} \\ w&= \alpha _3 \sin {(\omega (x-d_x))}\sin {(\omega (y-d_y))}\cos {(\omega (z-d_z))} \end{aligned}, \end{aligned}$$
(16)

where \(\alpha _1=\alpha _2=0.5\), \(\alpha _3=-1\), and \(\omega =2 \pi\); \(d_x=d_y=d_z=0.25\) were the offsets of the TGV flow structure in the x, y and z directions, respectively. In the first snapshot, 20,000 particles were randomly placed in the domain \(\Omega\) of \(x \times y \times z \in [0,1]\times [0,1]\times [0,1]\). The particle locations in the next snapshot were calculated by the forward Euler method: \({\varvec{x}}_{\kappa +1}={\varvec{x}}_{\kappa }+\delta t {{\textbf {U}}}_{\kappa }\), where \({\varvec{x}}_{\kappa }=({x}_{\kappa },{y}_{\kappa },{z}_{\kappa })\) and \({\varvec{x}}_{\kappa +1}\) were the particle locations in the \(\kappa\)-th and \((\kappa +1)\)-th snapshots, respectively; \({{\textbf {U}}}_{\kappa }=(u_{\kappa },v_{\kappa },w_{\kappa })\) was the velocity calculated by Eq. (16) in the \(\kappa\)-th snapshot; \(\delta t\) was the time interval between snapshots and here we chose \(\delta t=5 \times 10^{-5}\). After the integration, particle trajectories with \(2 \times 10^{4}\) snapshots were down-sampled to 11 framesFootnote 2 to generate the ground truth of the raw LPT data. Artificial noise (see Sect. 4 for details) was added to the particle spatial coordinates of the TGV-based ground truth.

1.2 B.2 Turbulence wake flow (TWF)

To generate the synthetic TWF data, we added artificial noise to an available Lagrangian dataset. Details about the dataset can be found in Khojasteh et al. (2022) and we briefly summarize it here. The dataset are created based on a DNS. The DNS was conducted using the Incompact3D solver (Laizet and Lamballais 2009) to simulate a flow with a free stream velocity U passing a cylinder with a diameter D at a Reynolds number of \(Re=3,900\).

In the Lagrangian datasets, about 200,000 particles were scattered in the domain. The particle velocities were calculated by tri-linear interpolations using the nearest eight nodal data in space. The particle trajectories were integrated by a fourth-order Runge–Kutta method. The synthetic data had 105,000 particles that were down-sampled from the Lagrangian dataset. These particles were distributed in a wake region, located 0.5D downstream of the cylinder, with the dimensions of \(x \times y \times z \in [6D,8D]\times [3D,5D]\times [2D,4D]\). The center-line of the cylinder located at \((x,y)=(5D,4D)\) is parallel to the z axis. A total of 11 frames were selected from 350 DNS frames and these 11 frames had a time interval \(\Delta t=0.00375U/D\). The noise was defined the same way as that used in the TGV validation, with \(\zeta =0.1\%\) to represent a typical noise in LPT data.

1.3 B.3 Low-speed pulsing jet

The experimental facility consisted of a hexagonal water tank, a cylindrical piston, an impingement plate, and optical equipment. The synthetic pulsing jet was generated by driving the piston using an electromagnetic shaker, and pushing the water through a circular orifice until impinged on the plate. A dual cavity high-speed laser illuminated the measurement area with the dimensions of \(60 ~\text {mm} \times 57 ~\text {mm} \times 20~ \text {mm}\). Four high-speed cameras recorded the jet flow and provided time-resolved LPT images (see Sakib (2022) for details).

The raw experimental data were acquired using a volumetric system and commercial software LaVision DaVis 10 (Göttingen, Germany). The LPT module of DaVis 10 is based on the shake-the-box (Schanz et al. 2016) algorithm. The velocity data were post-processed by Vortex-In-Cell# (VIC#) algorithm (Jeon et al. 2022) to obtain velocity gradients on a structured mesh.

C Baseline algorithms

1.1 C.1 Finite difference methods

We use the velocity component u calculation as an example. In a given frame, the first- and second-order finite difference methods (FDMs) calculate velocities by the forward Euler method (i.e., Eq. 17a) and central difference method (i.e., Eq. 17b), respectively:

$$\begin{aligned} \tilde{u}(t_\kappa )&= \frac{\hat{x}_{\kappa +1}-\hat{x}_{\kappa }}{\Delta {t}}, \end{aligned}$$
(17a)
$$\begin{aligned} \tilde{u}(t_\kappa )&= \frac{\hat{x}_{\kappa +1}-\hat{x}_{\kappa -1}}{2\Delta {t}}, \end{aligned}$$
(17b)

where \(\hat{x}_{\kappa }\) is the particle coordinate along the x direction in the \(\kappa\)-th frame from LPT experiments, \(\Delta {t}\) is the time interval between two consecutive frames, and \(t_\kappa\) is the time instant in the \(\kappa\)-th frame. The velocities of the first and last frames are calculated by the first-order finite difference method. Since the FDMs only evaluate particle velocities, the particle locations that are ‘output’ from the FDMs are assumed to be the same as those in the synthetic data, e.g., \(\tilde{x}(t_\kappa )=\hat{x}(t_\kappa )\).

1.2 C.2 Polynomial regression

We use trajectory and velocity reconstruction in the x direction as an example. First, the trajectory polynomial model function \(\tilde{x}(t)\) is given by \(\tilde{x}(t)=\sum ^m_{j=0}{p_{m,j} t^j}\) where \(p_{m,j}\) is the polynomial coefficient, m is the order of polynomials, and t is the time. For instance, the trajectory and velocity model functions of the 2nd polynomial (2nd POLY) are

$$\begin{aligned} \begin{aligned} \tilde{x}(t)=p_{2,2} t^2+ p_{2,1}t+p_{2,0} \\ \tilde{u}(t)=\dot{\tilde{x}}(t)=2p_{2,2} t+ p_{2,1} \end{aligned}. \end{aligned}$$

The trajectory and velocity approximation functions of the other polynomials are similar to those in the x direction of the 2nd POLY.

Next, we calculate the polynomial coefficient \(p_{m,j}\). A residual \(\mathcal {R}\) between measured particle locations \(\hat{x}_{\kappa }\) and the polynomial model function \(\tilde{x}({t}_{\kappa })\) is minimized \(\min ~ \mathcal {R}=\sum ^{N_{\text {trj}}}_{\kappa =1}\Vert \tilde{x}({t}_{\kappa })-\hat{x}_{\kappa } \Vert ^2,\) where \({t}_{\kappa }\) is the time instant in the \(\kappa\)-th frame. Setting the gradient of the residual \(\mathcal {R}\) regarding \(p_{m,j}\) to zero (i.e., \(\partial {\mathcal {R}}/\partial {p_{m,j}} =0\)) and the polynomial coefficient \(p_{m,j}\) can be explicitly solved. The polynomial coefficients in the y and z directions are the same as that in the x direction. Once all the coefficients are retrieved, the polynomial model functions can be constructed and used to approximate particle trajectories, velocities, and accelerations.

1.3 C.3 Cubic B-splines

The cubic B-spline algorithm is based on the MATLAB built-in functions from the Curve Fitting Toolbox™. We adopt a function spap2(piece,k,x,y) with piece=2 and k=4 to create a cubic spline with two pieces joined together, bypass setting knots. x and y are the given data points and their values, respectively. Then we use a function fnval(f,xe), in which f=spap2(piece,k,xe,y) is the spline function calculated above, and xe is the evaluation points. To calculate first-order derivatives for velocity as an example, we use fnder(f,d), where d=1.

D Spatial and temporal frequency response

We tested the spatial and temporal frequency response of our method using TGV and HIT. The TGV has varying predefined frequencies based on its analytical expressions (i.e., Eq. 16). It is used for both spatial and temporal frequency response analysis. The Lagrangian data based on the HIT have intrinsically varying frequencies along particle trajectories and we use them to test the temporal response.

For the temporal analysis based on the TGV, the velocity amplitudes \(\alpha _1\), \(\alpha _2\), and \(\alpha _3\) in Eq. (16) are varied from 2 to 32 with an increment of 2, corresponding to the non-dimensional temporal frequency \(f'_t\) of the flow varying from 1 to 16. Physically, \(f'_t\) means the number of periodic circulations one particle in the outer layer of the vortex cell can finish within a unit of time. These TGV data consist of 33 frames with a time interval \(\Delta t = 1/32\). For the spatial analysis based on the TGV, \(\omega\) in Eq. (16) is varied from \(\pi\) to \(16 \pi\) with an increment of \(\pi\), in turn, the non-dimensional spatial frequency \(f'_s\) of the flow varies from 0.5 to 8. The frequency response is defined as the ratio of the amplitudes of the reconstructed particle spatial coordinates for temporal analysis, or of the reconstructed particle velocities for spatial analysis, to that of the input data in the frequency space. The frequency spectra are calculated by the fast Fourier transform. The normalized frequency \(f^{*}\) is the spatial or temporal frequency of the flow divided by the Nyquist sampling frequency of the data. The parameters used in the frequency response analysis are listed in Table 2.

The TGV-based frequency response is presented in Fig. 12a and b. With increasing \(f^{*}_t\) or \(f^{*}_s\), the frequency response decreases from unity to approximately zero, with a cutoff frequency varying from about 0.2 to 0.4 for \(f_{s}^*\) and \(f_{t}^*\). This indicates a low-pass filter feature of our method. When the oversampling ratio for space (\(\beta _1\) and \(\beta _2\)) and time (\(\beta _0\)) decreases, the cutoff frequency increases. This observation is expected, as using small oversampling ratios recovers high-frequency components while applying larger ones provides smooth reconstruction.

For the frequency response based on the HIT flow, we use synthetic data with artificial noise to examine how our method attenuates noise of different levels. The synthetic HIT data are generated by adding the artificial noise to the ground truth. Three artificial noise levels \(\zeta =0.1\%\), \(\zeta =1.0\%\), and \(\zeta =5.0\%\) are used and they are defined the same as that in the validation (see Sect. 4). The ground truth is from a DNS of an incompressible flow with a Reynolds number \(Re_{\lambda } \approx 130\) based on the Taylor scale (details about the HIT can be found in Biferale et al. (2023)). The non-dimensional frequency of this HIT flow is \(f'_t=U_{\text {rms}}/L=0.275\), where \(U_{\text {rms}}\) is the root mean squares of the magnitudes of the particle velocities in all frames; L is the length of the cubic computational domain. The \(f'_t\) can be considered a metric describing the characteristic spatial frequency of the particle trajectory in this flow. The synthetic HIT data are sampled to 65 frames with a time interval \(\Delta t = 0.00225\) between two consecutive frames. Therefore, the corresponding temporal frequency f based on sampling varies from 0 to about 215.28, with an increment of about 6.94. We assess the frequency response that is defined as the ratio of the amplitudes of the reconstructed particle spatial coordinates to that of the input noisy data in the frequency space.

The temporal frequency response using the HIT is shown in Fig. 12c. As illustrated in the figure, when the noise is high (e.g., \(\zeta =5.0\%\)), the high-frequency components (mainly noise) in the input data are sufficiently suppressed, while the flow frequency ones (\(f^*_t<0.2\), which is dominated by true flow structures) are preserved. When the noise level decreases (e.g., \(\zeta =1.0\%\) and \(0.1\%\)), our method recovers the particle trajectories at higher frequencies.

Table 2 Parameters for frequency response analysis
Fig. 12
figure 12

The temporal and spatial frequency response of the CLS-RBF PUM method. Temporal (a) and spatial (b) frequency response based on the TGV data; c temporal frequency response based on the synthetic HIT data. c blue, green, and scarlet lines represent the frequency response based on the HIT data with 0.1%, 1.0%, and 5.0% noise, respectively

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Li, L., Pan, Z. Three-dimensional time-resolved Lagrangian flow field reconstruction based on constrained least squares and stable radial basis function. Exp Fluids 65, 57 (2024). https://doi.org/10.1007/s00348-024-03788-y

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00348-024-03788-y

Navigation