Brought to you by:
Paper

A parametric study of 3D PTV algorithms based on a two-view collimated imaging model

, and

Published 28 December 2022 © 2022 IOP Publishing Ltd
, , Citation Q Wang et al 2023 Meas. Sci. Technol. 34 035302 DOI 10.1088/1361-6501/acab1f

0957-0233/34/3/035302

Abstract

Volumetric Lagrangian measurements of droplet or turbulent flow using particle tracking methods have attracted intensive attention recently. The performance of three-dimensional particle tracking velocimetry (3D PTV) is highly reliant on the algorithms. Most existing 3D PTV algorithms are developed for multi-view systems, which cannot be applied directly to two-view systems due to the insufficient geometry constraints. In the current study, three different 3D PTV algorithms applicable for two-view systems are investigated parametrically using synthetic data. The imaging model is established on a two-view collimated shadowgraph imaging setup, which features a high framing rate, large test volume and long depth focus. The performances of the three algorithms are tested under different image particle densities and displacement–spacing ratios. The correctness of 3D reconstruction and tracking, as well as the number of ghost particles, are obtained and compared comprehensively. The results indicate that significant improvement is achieved through the dedicated designed algorithms. The comparative study reveals the potential of each algorithm with extremely limited geometry constraints in two-view systems, which may serve as guidance for choosing appropriate algorithms under different test conditions.

Export citation and abstract BibTeX RIS

1. Introduction

Three-dimensional particle tracking velocimetry (3D PTV) has attracted great attention in recent years. In contrast to particle image velocimetry (PIV) methods, particle tracking velocimetry (PTV) techniques trace individual particles in the flow and can depict the Lagrangian velocity field [1]. A great challenge for 3D PTV techniques lies in how to calculate the 3D velocity from 2D image data using parallax algorithms. In order to eliminate the ambiguity in particle reconstruction and tracking under dense particle conditions, most 3D PTV systems are established on multi-view setups, normally three or four. Improvement in PTV techniques is mostly reliant on the development of advanced algorithms. In the past decades, numerous algorithms have been proposed, which can be divided into four categories.

Category I is to track the path of each individual particle. All the possible tracks are resolved based on the motion model among multiple neighboring frames. Then a certain cost function is introduced to figure out the most possible path. In 1991, Hassan and Canaan [2] proposed a four-frame particle tracking algorithm on 2D images, which used the variance summation of both particle displacement and acceleration. In 1993, Malik et al [3] developed two algorithms, namely three-frame minimum acceleration (3MA) and four-frame minimum change in acceleration (4MA). The former utilized the minimum acceleration of three frames to search possible tracks, while the acceleration was adopted as a cost function. The latter algorithm employed four frames to search the candidates, in which the cost function was composed of the acceleration variation. In 2006, Ouellette et al [4] developed a four-frame best estimation (4BE) algorithm. The particle position at the fourth frame was estimated using the previous three-frame trajectory, assuming acceleration was the same. The searching region was determined based on the prediction, while the difference between the candidate and prediction positions was used as the cost function. In 2008, Li et al [5] proposed a regression-based multi-frame (RMT) algorithm, which searched the possible tracks using five frames, and the root-mean-square of particle velocities served as the cost function. With the RMT algorithm, significant progress toward real-time measurement has been made using parallel computing, field-programmable gate array cameras and graphics processing unit acceleration [68]. Practical tests were conducted for large-scale airflow measurement under low particle density conditions with long-time recording [7]. The initialization tracking in the aforementioned algorithms is through selecting nearest particles, which performs well when the flow field is relatively uniform and the particle displacement in two frames is much smaller than the distances between particles in the same frame. However, when the flow field has obvious randomness or the particle displacement is comparable to the neighboring particle distances, the correctness of the initialization deteriorates dramatically. In 2019, Clark et al [9] proposed a four-frame best estimation algorithm with enhanced initialization, which could flexibly modify the shape and size of the initialization region. Under this strategy, more possible tracks were identified. The correctness of 3D particle tracking was improved under several different flow conditions. Recently, Schanz et al proposed an algorithm named 'shake-the-box' (STB), which promoted a particle density up to 0.125 ppp for imaginary particles [10]. In the STB algorithm, temporal information was introduced to predict the particle distribution in the subsequent image frame. The local residuals between the predicted particles and detected particles were minimized. Based on the predicted positions, a shaking process was conducted to fit the actual particle positions by minimizing the local residuals iteratively. The synthesized utilization of both spatial and temporal information enabled a significant improvement in performance, which makes it a milestone for PTV techniques. The STB algorithm has achieved a particle density around 0.05 ppp in real applications using tomographic PTV [11, 12] and robotic PTV [13] systems.

Category II is based on the global information of all particles to resolve possible trajectories. The particle tracking process is transformed to an optimization problem, in which the square sum of the displacements of all the possible matching particles between two frames is used as the cost function. In 1993, Ohyama et al [14] employed a genetic algorithm (GA) to optimize the particle trajectory, which included the generation of an initial population, selection and reproduction, sequencing, crossover, mutation, etc. The iteration continued up to a termination condition, which gave out the final particle tracking results. In 2007, Takagi [15] used an ant colony optimization (ACO) algorithm to solve the particle tracking problem. This method mimics an ant searching for food to find the optimization solution. If an ant confirms a specific path, then it produces a pheromone, which leads others straight to the right place. Thus, the ants can select the path based on the number of pheromones, which finally converges to the best path. Moreover, Ohmi [16] applied the self-organizing map algorithm (SOM) to optimize the particle tracking problem in 2008. In 2014, Abbasi Hoseini et al [17] further improved the cost function of the SOM method, which included not only the particle displacement but also the rotation angle. Besides the aforementioned artificial intelligence techniques, discrete tomography has also been explored for 3D particle tracking. Discrete tomography was originally developed for reconstructing crystalline objects from high-resolution transmission electron microscopy data [18], which showed better performance than conventional computer tomography in reconstructing pixel-sized objects using projection data from limited directions. In 2015, Alpers et al [19] proposed a dynamic discrete tomography (DDT) algorithm for two-view 3D PTV, which utilized the previous time step solution. The algorithm was tested using synthetic data from two perpendicular view angles, which achieved 98% correctness for 500 particles over 50 frames, with an image particle density at 0.0016 ppp.

Category III is based on using the neighboring particle information to track the particle trajectories, which is similar to the cross-correlation algorithm widely adopted in the PIV technique. The difference lies in that the PIV method resolves an average velocity in the interrogation window based on the overall particle group similarity, whereas the PTV algorithm is conducted on individual particles. In 1989, Uemura et al [20] proposed a binary image cross-correlation (BICC) algorithm, which binarized the particle image first and then tracked the particles using the cross-correlation method. In 1993, Mark [21] proposed a fuzzy logic (FL) algorithm, which employed the quasi-rigid model among neighboring particles to track the particles. Later, more algorithms based on the quasi-rigid assumption were developed, including the spring model (SPG) [22], original relaxation (ORX) [23], new relaxation (NRX) [24], polar coordinate system similarity (PCSS) [25], non-iterative approach (NIA) [26], etc.

Category IV is the combination of categories II and III, which attempts to track the global particle motion based on the neighboring particle information. Ohmi and Panday [27] established a relaxation genetic algorithm (RGA), which combined the relaxation algorithm and GA. Later, Ohmi et al [28] proposed a relaxation ACO (RACO) algorithm, which improved the tracking correctness under dense particle conditions. A summary of the aforementioned PTV algorithms can be referred to in figure 1.

Figure 1.

Figure 1. Classification of existing particle tracking algorithms in PTV.

Standard image High-resolution image

It should be noted that most of the PTV algorithms are established on multi-view setups, normally three or four. The multi-view intersections can restrict the candidate searching region to a very small area, which helps to eliminate 'ghost particles' and improve the correctness, especially under dense particle conditions. However, there are also some occasions with limited optical access or a lack of sufficient multi-view equipment. The two-view system can then offer a more compact and economic option. The two-view system can only guarantee candidate searching along the corresponding epipolar line, which is much more challenging than the multi-view systems. Although the two-view system has received less attention than multi-view systems, the algorithms have made significant progress in recent years. In 2005, the two-view system was applied to measure the air velocity within an aircraft cabin mockup, using a particle streak velocimetry algorithm [29, 30]. In 2006, Ohmi and Sapkota [31] applied a cellular neural network for two-view 3D PTV, which achieved considerable correctness of tracking (∼80%) and reconstruction (∼90%) up to 0.025 ppp. In 2011, the ACO method was employed for particle matching in two views, while reasonable results were obtained around 0.005 ppp [32]. In 2016, the GA proposed by Panday [33] further improved the performance around 0.025 ppp. In 2020, Wang et al [34] proposed an image-based 3D PTV algorithm using event information in 2D images, which was tested in the range of 0.006–0.03 ppp using synthetic data. In 2021, Wu et al [35] proposed a temporal–spatial 3D PTV algorithm for two-view systems, in which the 2D and 3D spatial information and the temporal predictions were strongly coupled. The synthetic data tests showed a correctness over 99% around 0.03 ppp. The algorithm has been further improved and tested by practical experiments [36]. The results indicated that more reasonable particle trajectories could be identified and the trajectory interruptions under dense particle conditions were greatly reduced.

Most of the PTV techniques use direct imaging with pinhole model cameras, in which the in-depth distance is limited due to the defocusing problem. In 2011, Wang and Zhang [37] proposed a two-view shadowgraph system, which could resolve high-quality particle images with little restriction on the in-depth dimension. The shadowgraph system can also avoid the contradictory between fast framing rate and large test volume. However, the system was based on two converging beams, which could not fully utilize the available volume. Recently, a two-view collimated shadowgraph system was proposed by Zhu et al [38], which enlarged the test volume by more than eight times with the same shadowgraph facility. A robust and flexible calibration algorithm for the collimated system was established, which enabled the following 3D particle tracking process. In the current study, three different 3D PTV algorithms are tested and compared using synthetic data, based on the camera model of a two-view collimated shadowgraph system. The main objective is to reveal the potential of each algorithm with limited geometry constraints. The parametric test results can serve as guidance for algorithm choosing in practical experiments, which may also trigger further development of smarter algorithms.

2. Camera model and synthetic particles

2.1. Camera model

The camera model is established on the basis of a two-view collimated shadowgraph system [38]. As shown in figure 2, the system consists of two sets of conventional Z-type shadowgraph configurations, sharing an ordinary halogen tungsten lamp. Four identical parabolic mirrors, with 30 cm in diameter and 300 cm in focal length, are used to form two collimated beams which intersect in the test region. Two synchronized high-speed cameras (Photron SA-Z and Photron Mini AX100) are used to record the images.

Figure 2.

Figure 2. Experimental setup of the two-view collimated shadowgraph system.

Standard image High-resolution image

The optical geometry of the two-view collimated shadowgraph system is shown in figure 3. The projection model of each camera includes two parts: one part is the parallel projection in the collimated light path; the other is a pinhole model in the converging region, which corresponds to the converging beams projected onto a charge coupled device (CCD). In general, the pinhole model can be simplified as a center point and an image plane. In the following, subscripts l and r are used to denote elements related to the left and right camera, respectively. Point $({x_w},{y_w},{z_w})$ in the world coordinates is first projected on the parallel projection plane, which then forms a ray by connecting the center point $C$ and is projected on the image plane as $({u_l},{v_l})$ and $({u_r},{v_r})$. As derived in [38], the relationship between the object point $({x_w},{y_w},{z_w})$ in the world coordinate system and the corresponding points in the left and the right image coordinate systems $({u_l},{v_l})$ and $({u_r},{v_r})$ can be described as

Equation (1)

Figure 3.

Figure 3. Schematic illustration of the camera model in a two-view collimated shadowgraph system [38].

Standard image High-resolution image

In equation (1), the intrinsic matrix of each camera (${A_l}$,${A_r}$) and the extrinsic matrix between two cameras (${E_{rl}}$) can be resolved through a camera calibration process.

Once the camera model is established, the synthetic particle images can be generated from 3D particle coordinates using the projection principle described by equation (1). In the current study, the projection is conducted using an idealized camera model, in which the camera calibration error and the particle identification error are not involved. Two sets of ideal camera calibration parameters are used in the current study, which are listed in table 1. The view angles in the two groups of camera parameters are both set at 30°.

Table 1. Set values of the camera parameters.

Camera parameterSet 1Set 2
${A_l}$ $\left[ {\begin{array}{*{20}{c}} {100}&0&0 \\ 0&{100}&0 \\ 0&0&1 \end{array}} \right]$ $\left[ {\begin{array}{*{20}{c}} {63.662}&0&{200} \\ 0&{63.662}&{200} \\ 0&0&1 \end{array}} \right]$
${A_r}$ $\left[ {\begin{array}{*{20}{c}} {100}&0&0 \\ 0&{100}&0 \\ 0&0&1 \end{array}} \right]$ $\left[ {\begin{array}{*{20}{c}} {63.662}&0&{200} \\ 0&{63.662}&{200} \\ 0&0&1 \end{array}} \right]$
${E_{rl}}$ $\left[ {\begin{array}{*{20}{c}} 1&0&0&0 \\ 0&{0.5}&{0.866}&0 \\ 0&0&0&1 \end{array}} \right]$ $\left[ {\begin{array}{*{20}{c}} {0.9804}&{ - 0.007}&{0.1986}&{ - 3.1416} \\ { - 0.0048}&{0.9844}&{0.1761}&{ - 3.1416} \\ 0&0&0&1 \end{array}} \right]$

2.2. Synthetic particles

The synthetic particles are generated using the direct numerical simulation (DNS) results of homogeneous isotropic turbulence from the Johns Hopkins University Turbulence Databases [39]. The calculation domain has a size of ${\text{2}}\pi \; \times \;{\text{2}}\pi \; \times \;{\text{2}}\pi $, at a flow condition of ${\operatorname{Re} _\lambda } = 418$. The particles are distributed in the flow field randomly. An example is shown in figure 4, which presents particle trajectories with 2000 particles in 50 frames with a time interval of 0.005 s.

Figure 4.

Figure 4. 3D particle trajectories with 2000 particles in 50 frames.

Standard image High-resolution image

Based on the 3D particle positions, the synthetic particle images can be generated using the camera calibration parameters shown in table 1. In the current study, three different algorithms are tested parametrically, including the DDT, 4BE and spatial-temporal four frame best estimate (ST-4BE) algorithms. The synthetic particle images for testing the DDT and 4BE algorithms are generated using Set 1 camera parameters, with an image resolution of 500 × 500 pixels. The images for ST-4BE algorithm employ Set 2 camera parameters, with an image resolution of 400 × 400 pixels. Images with different image particle densities are generated. Two examples are shown in figure 5, with 0.01 ppp and 0.025 ppp respectively. It can be seen that the particle number increases with an increase in image particle density. The dense particle conditions are more challenging for particle identification, stereo pairing and tracking in 3D space.

Figure 5.

Figure 5. Synthetic shadowgraph images at different particle image densities: (a) 0.01 ppp; (b) 0.025 ppp.

Standard image High-resolution image

Besides the image particle density, the displacement–spacing ratio $\xi $ is also an important parameter affecting the particle tracking performance. The dimensionless parameter $\xi $ is employed here, which is defined as [9]:

Equation (2)

Equation (3)

In equations (2) and (3), $\Delta r$ is the average particle displacement between two frames; ${\Delta _0}$ is the average spacing between particles in the same frame, which can be estimated using the total particle number ${N_{{\text{all}}}}$ and the total test volume $V$. The displacement–spacing ratio is related to both the image particle density and the time interval. With the same time interval, the displacement ratio increases at larger image particle densities. With the same image particle density, the displacement–spacing ratio $\xi $ can be changed by varying the time interval between two frames. As shown in figure 6, for 1600 particles in 50 frames, the particle trajectories show dramatic differences at the two displacement ratios 0.0297 and 0.1459. With increasing $\xi $, the lengths of the particle trajectories are increased. Different trajectories tangle together, which makes the tracking process more difficult.

Figure 6.

Figure 6. Trajectories of 1600 particles in 50 frames under different displacement-spacing ratios (a) $\xi $ = 0.0297; (b) $\xi $ = 0.1459.

Standard image High-resolution image

The summary of the test conditions for the three algorithms is listed in table 2. As will be shown in sections 35, the performances of the three algorithms deteriorate with increasing image particle density and displacement–spacing ratio. In the present study, the test is terminated when the correct ratio is below 60%. As shown in table 2, the test range of the displacement–spacing ratio does not significantly overlap for the three algorithms, which is mainly due to their performance differences.

Table 2. Summary of test conditions.

AlgorithmImage particle density (ppp) $\xi $ AlgorithmImage particle density (ppp) $\xi $
DDT0.0050.0094 0.0050.0236
 0.010.0118 0.0050.1158
 0.0150.0136 0.00630.0563
 0.020.0149 0.010.0297
 0.0250.0161 0.010.1459
    0.010.2861
   ST-4BE0.0180.0362
4BE0.0050.0108 0.0250.0404
 0.010.0136 0.0250.1980
 0.0150.0156 0.0380.0465
 0.020.0171 0.050.0508
 0.0250.0185 0.050.2495
 0.030.0196 0.0750.0582
 0.0350.0207 0.0880.0615
 0.040.0216 0.10.0641

3. DDT

The DDT algorithm belongs to Category II, which tracks the particles from a global view. In the current study, the performance of the DDT algorithm has been parametrically investigated, to reveal its potential for two-view 3D PTV applications.

3.1. DDT algorithm

For two-view systems, the particles at each time step are recorded from two different views. The 3D particle coordinates can be obtained using the imaging coordinates and camera calibration parameters. Figure 7 shows a two-view imaging example with parallel beams, while the following description is also applicable to the pinhole model. For each particle, there are two projecting lines passing through it. The intersections of the projecting lines are called candidate points; the whole set is called a candidate grid. It is noted that the candidate grid may contain additional points without real particles. It is also possible for multiple particles to lie on a projecting line. For n particles, there are at most n projecting lines for each projection direction, which corresponds to a maximum of n2 grid points. With the camera calibration parameters, the candidate grid can be easily resolved from the projecting lines.

Figure 7.

Figure 7. Schematic of 3D reconstruction of the DDT algorithm.

Standard image High-resolution image

Assume that at a specific time step t, the candidate grid ${G^{(t)}}$ contains $l(t)$ candidate points. At each candidate point $g_i^{(t)}$, a discretized variable $\xi _i^{(t)}$ is designated. If there is a particle then $\xi _i^{(i)} = 1$; otherwise $\xi _i^{(i)} = 0$. The 3D reconstruction process is transformed to resolve the solution ${\underline x ^{(t)}}: = {(\underline \xi _1^{(t)},\ldots,\underline \xi _{l(t)}^{(t)})^T} \in {\{ 0,1\} ^{l(t)}}$ constrained by the linear inequalities:

Equation (4)

where ${b^{(t)}}: = {(1,\ldots,1)^{\text{T}}} \in {\left\{ 1 \right\}^{k(t)}}$ represents the imaging data from two views; $k(t)$ denotes the total number of detected particles; and ${A^{(t)}}$ represents the particle imaging effect matrix. If the candidate point $g_i^{(t)}$ has contributions to the particle image $b_j^{(t)}$, $A_{ji}^{(t)} = 1$; otherwise, $A_{ji}^{(t)} = 0$. Here, we use ${x^{(t)}}$ and ${\underline x ^{(t)}}$ to denote the variable and specific solutions, respectively.

For 3D particle tracking, it does not only require the 3D reconstruction solution based on equation (4) to be resolved, but also for the particle position ${x^{(t - 1)}}$ at time step $t - 1$ to correspond to the following position ${x^{(t)}}$ at time t. Here a vector ${w^{(t)}} = {(\omega _1^{(t - 1,t)},\ldots,\omega _{l(t)}^{(t - 1,t)})^{\text{T}}}$ is defined to denote the weights associated with each grid point. Then the tracking from time step $t - 1$ to t can be described by the following discrete optimization problem:

Equation (5)

In equation (5), ${w^{(t)}} \bullet {x^{(t)}}$ represents the scalar product between weight parameter vector ${w^{(t)}}$ and solution vector ${x^{(t)}}$.

Here the weight parameter vector ${w^{(t)}}$ is defined as

Equation (6)

where ${\text{dist}}(g_i^{(t)},g_j^{(t - 1)})$ denotes the distance of a certain particle from the previous time step t – 1 to the next frame at time t. The initialization can be set as ${w^{(0)}}: = 0$, if there is no prior information available. When the framing rate reaches the kHz or MHz order, the distance between neighboring frames can be considered as 'short' for most test conditions. Then the distance can be expressed as the Euclidian distance

Equation (7)

It should be noted that the solutions for a finite number of projections have the possibility of non-uniqueness. The ambiguities can be improved by introducing extra constraints, while the particle position in the space domain and the maximum velocity value are employed in the current study.

3.2. DDT test tracking

In the present study, four parameters are defined to quantitatively evaluate the algorithm performance, which are listed below:

Equation (8)

${N_t}$ represents the particle number being correctly reconstructed at every frame. ${N_{\text{g}}}$ is the number of ghost particles at each frame. ${N_{t({\text{trac}})}}$ is the correctly tracked trajectories at each frame, while ${N_{{\text{g}}({\text{trac}})}}$ represents the wrong ones. ${N_{{\text{all}}}}$ is the total number of correct particles. From the definition in equation (8), ${F_t}$ and ${F_{\text{g}}}$ are used to evaluate the 3D reconstruction correctness, while ${F_{t({\text{trac}})}}$ and ${F_{{\text{g}}({\text{trac}})}}$ are employed for 3D tracking.

3.2.1. Tracking results at different frames.

The DDT algorithm is first tested using 1000 particles, with an image particle density of 0.005 ppp and a displacement–spacing ratio of 0.0094. The 3D reconstruction and tracking results in 100 frames are shown in figure 8. The curves indicate that the variations of ${F_{t({\text{trac}})}}$ and ${F_{{\text{g}}({\text{trac}})}}$ with frame number are in accordance with ${F_t}$ and ${F_{\text{g}}}$ respectively. The correctness ${F_t}$ decreases from 100% in the first frame to 83% in the tenth frame. Then it increases with the frame number, reaching 99% in the 40th frame and remaining there until the final frame. Although the first frame has the highest correctness, it also involves the most ghost particles, as high as 158%. The number of ghost particles decreases dramatically with the frame number, which is reduced to only 1% in the 40th frame. This variation is mainly due to the lack of prior information in the first frame, where it is assumed that all the candidate points have particles. With the processing of the DDT program, the ghost particles are excluded gradually, which also includes some real particles. After 40 frames, the constraint is more reliable, which helps to track the lost particles again.

Figure 8.

Figure 8. (a) 3D reconstruction and (b) particle tracking results with frame number using the DDT algorithm (0.005 ppp, $\xi $ = 0.0094).

Standard image High-resolution image

Considering that the poor performance in the first 40 frames is mainly due to the lack of constraints from the time domain, an iteration strategy is introduced to improve the DDT algorithm. The iteration is conducted from forth-to-back for odd frame numbers using time steps $t - 1$ and t, and back-to-forth for even frame numbers using $t + 1$ and t. The results after two iteration operations are shown in figure 9. It can be seen that the correctness of the DDT algorithm is greatly improved by the iteration. The correctness of ${F_t}$ and ${F_{t({\text{trac}})}}$ is higher than 99% from the first to the end frame, where the ratio of ghost particles is kept at less than 1%.

Figure 9.

Figure 9. (a) 3D reconstruction and (b) particle tracking results with frame number using the DDT algorithm with iteration (0.005 ppp, $\xi $ = 0.0094).

Standard image High-resolution image

3.2.2. Tracking results at different image particle densities.

Based on the iteration DDT algorithm, the performances at different image particle densities from 0.005 ppp to 0.025 ppp are tested, with a constant time interval of 0.005 s. The 3D reconstruction and tracking results are shown in figure 10. It can be seen that the variations of ${F_{t({\text{trac}})}}$ and ${F_{{\text{g}}({\text{trac}})}}$ are very similar with ${F_t}$ and ${F_{\text{g}}}$. The correctness is over 99% at 0.005 ppp, and decreases with increasing image particle density. At 0.01 ppp, the correctness is around 94%, while it drops below 90% from 0.015 ppp. At 0.025 ppp, the correctness is only about 60%. The number of ghost particles increases with the image particle density, in a mirroring curve with respect to the correctness. The variations indicate a strong relevance between ghost particles and wrong detections of candidate points. With a dramatic increase in ghost particles, the correctness decreases accordingly. The quantitative analysis indicates that it is a great challenge for the global DDT algorithm to deal with dense particles.

Figure 10.

Figure 10. Performance of the DDT algorithm at different particle image densities.

Standard image High-resolution image

3.2.3. Tracking results at different displacement–spacing ratios.

The 3D reconstruction and tracking results under different displacement–spacing ratios are shown in figure 11. The displacement–spacing ratio is in a range from 0.0094 to 0.0161. It can be seen that the correctness of ${F_t}$ and ${F_{t({\text{trac}})}}$ is around 99% when $\xi $ is 0.0094. The correctness can stay as high as 92% when $\xi $ is below 0.012. With further increasing of the displacement ratio, the correctness decreases dramatically, dropping to 60% at $\xi = 0.016$. The ghost particles and ghost velocity vectors increase with $\xi $ accordingly. The results indicate that the algorithm performance is very sensitive to the displacement–spacing ratio, which should be kept small enough for practical applications.

Figure 11.

Figure 11. Performance of the DDT algorithm at different displacement–spacing ratios.

Standard image High-resolution image

3.2.4. Tracking performance summary for DDT algorithm.

The parametric analysis indicates that the current DDT algorithm can be applied for relatively sparse particles, where the particle image density is no more than 0.01 ppp with a displacement–spacing ratio less than 0.01. It should be pointed out that with increasing of particle density, the candidate points increase dramatically. For the test condition at 0.025 ppp, the number of candidate points is as large as $8 \times {10^4}$, while the number of correct points is only 6250.

4. Four-frame best estimate (4BE)

The 4BE algorithm belongs to Category I, which tracks the individual particle paths. The four-frame best estimation (4BE) algorithm is widely applied for both two-dimensional particle tracking velocimetry (2D PTV) and 3D PTV measurements, featuring low calculation cost and available open source code [9]. Thus it is worthy of exploring its application under two-view conditions. In the following subsections, the 4BE algorithm is modified and applied to two-view systems.

4.1. Four-frame best estimate (4BE) algorithm

As aforementioned, the multi-view system can help to restrict the candidate points to a very small region. The 3D coordinates of the particles are reconstructed first, which are then tracked in time sequences. Since the candidate points are quite limited, it is easy to exclude the ghost particles using certain constraints. In two-view systems, the candidate points can only be confined to those near an epipolar line, which makes it quite difficult to exclude the large number of ghost particles in the 3D reconstruction process. The strategy of the 4BE algorithm in the present study is shown in figure 12. The 3D reconstruction is based on a searching region near the epipolar line, where a small number of ghost particles are excluded if they are out of the measurement volume. Then most of the ghost particles are identified and removed in the tracking process in space.

Figure 12.

Figure 12. Schematic of the 4BE algorithm.

Standard image High-resolution image

In the current 4BE algorithm, the 3D coordinates in four frames, $n,n + 1,n + 2,n + 3$, are used to track the particles. Firstly, it needs to initialize the track by matching the particle $x_i^n$ to its position in next frame $x_i^{n + 1}$. In frame n+ 1, a searching region is defined to exclude ghost particles, which is set as 1.1 times the maximum particle displacement. All the particles in the searching region are considered as candidate points. The consideration is to involve all the possible points but with affordable calculation cost.

After the initialization, the two positions at frames n and n + 1 can form a track. Then the particle position $\widetilde x_i^{\,n + 2}$ at frame n + 2 can be predicted using the following equations:

Equation (9)

Equation (10)

In equations (9) and (10), $\Delta t$ is the time interval between two frames. $\widetilde v_i^{\,n + 1}$ is the particle velocity, which can be calculated using equation (10), based on a constant velocity assumption.

Then the searching area for frame n + 2 is set based on the predicted positions. The searching region is set at 1.1 times the maximum possible displacement to exclude the ghost particles. If there is no particle detected in the searching region, then this track will be discarded. If there is one or more particles detected, all the tracks are kept for the following procedures. The predictions at frame n + 3 are based on the previous three frames n, n + 1 and n + 2, using both the constant velocity and acceleration assumptions, as indicated by equations (11)–(13).

Equation (11)

Equation (12)

Equation (13)

In equation (13), $\widetilde a_i^{\,n + 1}$ is the acceleration at frame n + 1. After the prediction, a searching region for frame n + 3 is selected, where the candidate points are tracked. If there are no particles detected, this track is discarded. If one or more particles are found, then multi-tracks remain.

Up to now, for any particles in frame n, there might be multiple candidate tracks in the following three frames. In order to resolve the most possible track, a cost function $\phi _{ij}^n$ is introduced to optimize the tracking process, which is defined as

Equation (14)

In equations (11) and (14), $x_j^{n + 3}$ is the detected particle position at frame n + 3; $\widetilde x_i^{\,n + 3}$ is the predicted particle position. With the cost function optimization, the minimum distance between the detected and predicted particles is selected as the best track. In the current algorithm, only the trajectories longer than ten time steps are kept, which helps to exclude the large number of ghost particles in two-view systems.

4.2. Four-frame best estimate (4BE) test tracks

4.2.1. Tracking results at different frames.

The ratios of correctly reconstructed and tracked particles at different frames are presented in figure 13. The image particle density is 0.005 ppp, with a displacement–spacing ratio $\xi $ of 0.0108. It can be seen that at relatively small image particle densities, the correctness of both ${F_t}$ and ${F_{t({\text{trac}})}}$ is kept very close to 100%. The ratio of ghost particles ${F_{\text{g}}}$ is about 7.5% in the first frame, which then decreases gradually and reaches 3.8% in the 49th frame. Due to the lack of sufficient prior knowledge at the very beginning, some ghost particles cannot be excluded. With the progression of particle tracking, the constraint from the time domain begins to play a role, which helps to remove more ghost particles. As can be seen from figure 13(b), the trends of ${F_{{\text{g}}({\text{trac}})}}$ are very similar to ${F_{\text{g}}}$, with differences of less than 1.5%.

Figure 13.

Figure 13. (a) 3D reconstruction and (b) particle tracking results with frame number using the 4BE algorithm (0.005 ppp, $\xi $ = 0.0108).

Standard image High-resolution image

4.2.2. Tracking results at different image particle densities.

The 4BE algorithm is tested with different image particle densities in the range of 0.005 ppp–0.04 ppp, with a constant time interval of 0.005 s. The results are shown in figure 14, where the DDT test results are also presented for comparison. It can be seen that the ratio of ${F_t}$ is kept at more than 99% up to 0.04 ppp. However, the ghost particles increase dramatically with an increase in image particle density, which rises from 4.5% to 35.0% in the test range. The variations of ${F_{t({\text{trac}})}}$ and ${F_{{\text{g}}({\text{trac}})}}$ are very similar to those of ${F_t}$ and ${F_{\text{g}}}$. The comparison with the DDT algorithm indicates that the 4BE algorithm has better performance when the image particle density is larger than 0.01 ppp.

Figure 14.

Figure 14. (a) 3D reconstruction and (b) particle tracking results of the 4BE algorithm at different particle image densities.

Standard image High-resolution image

4.2.3. Tracking results at different displacement–spacing ratios.

The 4BE algorithm is tested under different displacement–spacing ratios from 0.0108 to 0.0216. The results are presented in figure 15. It can be seen that both ${F_t}$ and ${F_{t({\text{trac}})}}$ remain higher than 98% for all the test cases. However, the ratio of ghost particles increases with the displacement–spacing ratio, which increases from to 7% to 36% in the test range. Compared with the DDT algorithm, the 4BE algorithm has much higher correctness and fewer ghost particles when the displacement–spacing ratio is larger than 0.012.

Figure 15.

Figure 15. (a) 3D reconstruction and (b) particle tracking results of the 4BE algorithm under different displacement–spacing ratios.

Standard image High-resolution image

4.2.4. Tracking performance summary for 4BE algorithm.

The results in section 4 indicate that the 4BE algorithm has an overall better performance than the DDT algorithm. The correctness can be kept higher than 98% in a relatively wide range. However, it is also noticed that the ghost particle ratio increases significantly with the image particle density and displacement–spacing ratio in the test range. The ghost particles induce wrong trajectories and vectors. In order to improve the performance, some extra efforts are required to exclude these ghost particles.

5. ST-4BE

The DDT and 4BE algorithms conduct the 3D coordinate reconstruction first. Then the 3D particle tracking process is performed. However, in two-view systems, a large number of ghost particles are generated during the 3D reconstruction process, which is a key factor limiting the performance under dense particle conditions. In the current study, a recently proposed ST-4BE algorithm [35] for the two-view system is tested.

5.1. ST-4BE algorithm

The strategy of the ST-4BE algorithm is shown in figure 16, in which the reconstruction and tracking processes are strongly coupled. The cost function involves the constraints from both the space and time domains. The constraints from epipolar geometry, predictions in 2D images and 3D space, and reprojection from space to image coordinates are utilized in a comprehensive way. The algorithm is explained in detail below.

Figure 16.

Figure 16. Schematic of the proposed ST-4BE algorithm [35].

Standard image High-resolution image

For an image at time n, the epipolar constraint is first applied to find all the possible stereo pairing candidates $x_1^n$ and $x_2^n$ in two images, based on which the 3D coordinates ${x^n}$ are reconstructed using the calibration parameters. For the particle image in the next frame n + 1, specific searching regions are selected based on the positions of $x_1^n$ and $x_2^n$, where the stereo pairing candidates $x_1^{n + 1}$ and $x_2^{n + 1}$ are found by epipolar constraints. Then the 3D coordinates ${x^{n + 1}}$ at frame n + 1 can be obtained. The size of the searching region is set at 1.1 times the maximum possible displacement.

After initialization, the 3D coordinates ${\widetilde x^{\,n + 2}}$ at frame n + 2 can be predicted using equations (9) and (10), which are re-projected back to two-view images with coordinates $\widetilde x_1^{\,n + 2}$ and $\widetilde x_2^{\,n + 2}$. Then the searching region is selected based on $\widetilde x_1^{\,n + 2}$ and $\widetilde x_2^{\,n + 2}$, where the stereo pairing is conducted to tackle corresponding particle coordinates $x_1^{n + 2}$ and $x_2^{n + 2}$. The 3D coordinates at frame n + 2 can then be obtained. This procedure is repeated to resolve the 3D coordinates at frame n + 3.

Up to now, all the possible tracks are established. Then a cost function is established to resolve the best particle trajectories. Several cost functions have been attempted in [35], among which the following one is selected,

Equation (15)

In equation (15), $i = 1,2$ represents the two views; $x_i^{n + 3}$ is the re-projected 2D image coordinates from reconstructed particles at frame $n + 3$; $\widetilde x_i^{\,n + 3}$ is the predicted 2D image coordinates at frame $n + 3$. Only the trajectories with minimum cost function value are remained; the others are discarded.

5.2. ST-4BE test tracks

5.2.1. Tracking results at different frames.

The 3D particle reconstruction and tracking results of ST-4BE are shown in figure 17, in which the results of the 4BE algorithm are also presented for comparison. The image particle density is 0.005 ppp, with a displacement–spacing ratio $\xi $ of 0.0236. It can be seen that the ${F_t}$ and ${F_{t({\text{trac}})}}$ of the ST-4BE algorithm keep a high correctness over 99% for all the frames. The two lines coincide with the 4BE results. However, the two algorithms resolve dramatically different ratios of ghost particles. As shown in figure 17, the resolved ghost particle ratio by the ST-4BE algorithm is always less than 1%, which is much less than that of the 4BE algorithm. The comparison indicates that the comprehensive application of multi-constraints in both time and space of the ST-4BE algorithm excludes the ghost particles effectively.

Figure 17.

Figure 17. (a) 3D reconstruction and (b) particle tracking results with frame number using the ST-4BE algorithm (0.005 ppp, $\xi $ = 0.0236).

Standard image High-resolution image

5.2.2. Tracking results at different image particle densities.

The performance of ST-4BE is investigated at different image particle densities, ranging from 0.005 ppp to 0.1 ppp, with a constant time interval of 0.005 s. As shown in figure 18, the results are compared with the other two algorithms DDT and 4BE. The comparison shows that the ST-4BE algorithm makes significant improvements. The correctness of ${F_t}$ and ${F_{t({\text{trac}})}}$ remains higher than 99% until 0.1 ppp, while the ghost particle ratio of ${F_{\text{g}}}$ and ${F_{{\text{g}}({\text{trac}})}}$ is less than 1% in the whole test range.

Figure 18.

Figure 18. (a) 3D reconstruction and (b) particle tracking results of the ST-4BE algorithm at different particle image densities.

Standard image High-resolution image

5.2.3. Tracking results at different displacement–spacing ratios.

The ST-4BE algorithm is tested at different displacement–spacing ratios, which vary from 0.0236 to 0.2861, by varying both the particle image density and time interval. It can be seen from figure 19 that the correctness is higher than 92.5% at $\xi = 0.2$, and decreases dramatically with a further increase in the displacement–spacing ratio. The correctness drops below 58% at $\xi = 0.2861$. The correctness of ST-4BE and 4BE is similar when $\xi $ is smaller than 0.03, which is much higher than the DDT algorithm. However, the ST-4BE shows far fewer ghost particles than the 4BE algorithm, remaining below 5% in the wide test range.

Figure 19.

Figure 19. (a) 3D reconstruction and (b) particle tracking results of the ST-4BE algorithm under different displacement–spacing ratios.

Standard image High-resolution image

5.2.4. Tracking performance summary for ST-4BE algorithm.

The ST-4BE algorithm makes use of the constraints from space and time through a dedicated design. The comparison indicates that the ST-4BE algorithm shows the best performance among the three algorithms, which extends the range of image particle densities up to 0.1 ppp. This improves the spatial resolution to ten times finer than the 4BE and DDT algorithms. The ST-4BE algorithm also greatly extends the dynamic range, which resolves reasonable correctness when the displacement–spacing ratio is less than 0.2. Considering the ghost particle ratio, the applicable dynamic range is about one order larger than those of the 4BE and DDT algorithms. The significant improvement enables the two-view system to be applicable to more challenging test conditions, which represents remarkable progress for 3D PTV techniques.

6. Conclusions

In the current study, three different two-view 3D PTV algorithms are tested and compared using synthetic particle images, which are generated with an idealized shadowgraph imaging model in collimated beams and a DNS homogeneous turbulence database. For the global DDT algorithm, the iteration process is introduced to improve the performance at initial stages. The test results indicate that DDT can perform well when the particle density is less than 0.01 ppp, with a displacement–spacing ratio less than 0.01. The correctness drops to only 60% at 0.025 ppp. With increasing particle numbers, a huge number of ghost particles are generated, which induce a heavy calculation burden and dramatic decrease in correctness. The 4BE algorithm based on local searching shows a better performance than DDT, with higher correctness and fewer ghost particles. It can track more than 99% of the correct particles at 0.04 ppp, with 36% ghost particles. In a recently proposed ST-4BE algorithm, comprehensive constraints from both space and time domains are employed to exclude the ghost particles. The ST-4BE algorithm shows far superior performance to the other two algorithms. The correctness is kept around 99% at 0.1 ppp, while the applied displacement–spacing ratio is extended to 0.2. The enhanced constraints effectively exclude the large number of ghost particles, which are kept under 5% in the wide test range. The performances of all three algorithms deteriorate with increasing displacement–spacing ratio. Thus a high enough framing rate should be guaranteed for practical applications, which can be selected according to the parametric test results.

Comparative studies have indicated that the large number of ghost particles in two-view systems is a major challenge for the 3D reconstruction and tracking process, far more than in multi-view systems. The local searching algorithms such as 4BE and ST-4BE show better potential than the global algorithm DDT. The main reason lies in that the local searching algorithm can utilize more information in time and space domain, which helps to exclude the large number of ghost particles more effectively. For the global DDT algorithm, the exponential increase in ghost particles with image particle density limits its application under dense conditions. The current study reveals the potential of the two-view collimated system for 3D particle tracking; the parametric test results can serve as guidance for algorithm choosing under different test conditions. It needs to be pointed out that the camera calibration and particle identification errors are not involved at present, which deteriorates the performance for all the algorithms to some extent. Practical experimental applications of the 3D PTV algorithms will be conducted in our future studies.

Acknowledgments

The financial support by the National Natural Science Foundation of China (Grant Nos. 51976121 and 52011530187) is gratefully acknowledged.

Data availability statement

No new data were created or analyzed in this study.

Please wait… references are loading.
10.1088/1361-6501/acab1f