Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Deep learning-based size prediction for optical trapped nanoparticles and extracellular vesicles from limited bandwidth camera detection

Open Access Open Access

Abstract

Due to its ability to record position, intensity, and intensity distribution information, camera-based monitoring of nanoparticles in optical traps can enable multi-parametric morpho-optical characterization at the single-particle level. However, blurring due to the relatively long (10s of microsecond) integration times and aliasing from the resulting limited temporal bandwidth affect the detected particle position when considering nanoparticles in traps with strong stiffness, leading to inaccurate size predictions. Here, we propose a ResNet-based method for accurate size characterization of trapped nanoparticles, which is trained by considering only simulated time series data of nanoparticles’ constrained Brownian motion. Experiments prove the method outperforms state-of-art sizing algorithms such as adjusted Lorentzian fitting or CNN-based networks on both standard nanoparticles and extracellular vesicles (EVs), as well as maintains good accuracy even when measurement times are relatively short (<1s per particle). On samples of clinical EVs, our network demonstrates a well-generalized ability to accurately determine the EV size distribution, as confirmed by comparison with gold-standard nanoparticle tracking analysis (NTA). Furthermore, by combining the sizing network with still frame images from high-speed video, the camera-based optical tweezers have the unique capacity to quantify both the size and refractive index of bio-nanoparticles at the single-particle level. These experiments prove the proposed sizing network as an ideal path for predicting the morphological heterogeneity of bio-nanoparticles in optical potential trapping-related measurements.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Due to its non-invasive property and frequently label-free nature, harmonic optical potential related techniques such as optical tweezers and optical waveguides have become increasingly attractive in numerous biological applications ranging from the manipulation of living cells [1] and individual virus/extracellular vesicles [2,3], to real-time observation of the dynamic processes of trapped particles by accurately measuring their precise displacement or piconewton trapping force. Furthermore, through high-speed observations of the diffusive Brownian motion of trapped particles and calculating their stiffness and time evolution of the mean square displacement (MSD), morphological properties below the optical diffraction limit such as the hydrodynamic diameter [4], as well as related physical quantities such as the refractive index [5,6] could be characterized. In addition, joint morpho-chemical information provided by combining laser tweezers with other optical detection modes (such as Raman [7], fluorescence [8], and others) are capable of achieving precise, quantitative chemical characterization of heterogeneous nanoparticles at the individual-particle level. All the techniques above require a precise and high-bandwidth measurement of the trapped particle positions.

The position of trapped particles has traditionally been measured using either position sensitive detectors [9] (such as quadrant photodiode (QPD) and position sensitive diode (PSD), etc.) or sensitive high-speed video cameras (such as sCMOS) [10] with particle tracking algorithms. Compared to the former that only provide only location information from a single detection spot, the spatially resolved high-speed camera offers a more comprehensive solution by observing the constrained Brownian motion process of single/multiple trapped nanoparticles simultaneously. Moreover, once their sizes were determined, refractive index distributions could also be extracted by inverting the scattering intensity recorded in images based on homogeneous or core-shell sphere Mie scattering models [11]. However, common high-speed cameras only offer the measurement rates up to a maximum of a few kHz, which leads to a conflict. Namely, strong stiffness traps are typically required to stably trap low index contrast, nanometer-sized particles such as liposomes or extracellular vesicles, yet the stronger the trap stiffness, the faster the instrument measurement speed must be to ensure that the measurement captures primarily ballistic motion. It has been proven that few kHz acquisition rates do not meet the full sampling measurement conditions of the diffuse Brownian motion for stable trapped particles under strong stiffness [4,12,13], which leads to overestimation of particle size due to inaccurate MSD measurement, and ultimately affecting the characterization accuracy of related physical properties such as refractive index [14].

Various methods have been explored for solving the sizing distortion problem in camera-based measurements, which can be mainly divided into three approaches. First, since the frequency of overdamped oscillatory motion is inversely proportional to the trapping stiffness, relatively weak stiffness traps could be applied for trapping and observing the diffusive Brownian motion of the nanoparticle, such that full-sampling lies within the acquisition bandwidth of camera [15,16]. However, this might sacrifice the trapping stability and scattering intensity of nanoparticles, and may be incompatible with other measurement modes demanding a strong laser (e.g. Raman spectroscopy). A second approach is to establish the calibrated constrained Brownian motion model by considering both the blurring and aliasing effect caused by the limited bandwidth of the camera measurement, and obtain a corrected size by fitting the experimental measurements to an adjusted model [17]. However, as the constrained model is, in general, dependent on particle size, the corrected fitting quality relies on the deviation between a guessed value for the particles and their actual size. This leads to a lack of precision when applying such models for the characterization of bio-nanoparticles with highly unknown and heterogeneous characteristics, which is typically the case for biological nanoparticles such as extracellular vesicles. Finally, deep learning algorithms such as the convolutional neural network (CNN) provide an efficient technical means to solve the sizing regression problem. It has been reported that by using a CNN, physical properties such as particle size, refractive index [18] or even fluid viscosity around particles [19] can be successfully predicted from time resolved displacement/force series of polystyrene and silica nanoparticles. Despite to improve the diffusion coefficient and size calculation accuracy of nanoparticles under free Brownian motion [20,21], as well as the autocorrelation function accuracy [22] under constrained Brownian motion, when considering size prediction under constrained Brownian motion, deep learning has been applied only to fully simulated datasets, and its exploration on experimental measurements (particularly for biological nanoparticles) remains to be demonstrated.

To address the above gaps, here we describe a residual neural network (ResNet) to characterize the size of optically trapped nanoparticles based on temporally under-sampled position measurements acquired by a camera with a limited detection bandwidth. Our proposed network was trained using simulated trapped nanoparticles position series under strong stiffness traps. We experimentally demonstrate that our simulated data-driven network enables size prediction of standard nanoparticles as well as extracellular vesicles (EVs), resulting in more accurate size prediction results than other algorithms such as adjusted Lorentzian fitting method or CNN-based networks. We further explored the effect of the time series length on the networks performance and demonstrate that accurate particle sizes could be predicted even with short data lengths, providing substantial improvements in measurement speeds for real experiments. Lastly, by expanding the network to human platelet-free plasma (PFP)-derived vesicles, we found the predicted sizing results are in close agreement with gold-standard nanoparticle tracking analysis (NTA) characterization. To highlight the unique capacity of high-speed camera based optical tweezers measurements compared with traditional QPD measurements, we further quantify the refractive index of vesicles separated by different purification procedures by the combination of our proposed network and individual scattering images from the video series. Overall, this study offers a promising ResNet-based sizing method for studying the morphological characterization of optical trapped nanoparticles on single-particle level, obtaining more accurate results than other state-of-art algorithms on experimental measured datasets such as standard nanoparticles and EVs, and can be generalized to many other optical potential trapping measurements.

2. Materials and methods

2.1 Laboratory-built optical tweezers system and its data processing

A laboratory-built optical tweezers system was developed previously [13], with its detailed optical configuration description in the Fig. 1(a). Briefly, a 785 nm single-mode continuous wave Gaussian laser (TA Pro, Toptica, Germany, output power of 300 mW in front of objective) is used to create a tightly confined laser trap via a high NA water immersion objective (60x / NA = 1.27, Nikon Plan Apo SR IR), and a gradient attenuator was applied for power reduction during trapping measurement. Elastically scattered light is collected and directed towards a high-speed sCMOS detector (zyla 4.2, Andor, Belfast, UK) placed in a conjugate plane to the sample plane. An imaging lens (f = 500 mm) is coupled with the camera and the pixel to micron conversion factor of the sCMOS image is 0.0443 micron/pixel. The region of interest (ROI) was restricted to image the motion of trapped individual nanoparticle, which was selected to 20 × 60 pixels (corresponding to 0.89 µm × 2.66 µm in the sample plane), allowing images to be taken at 6500 Hz.

 figure: Fig. 1.

Fig. 1. Optical tweezers measurement system and data preprocessing. (a) Schematic of the optical tweezers system setup. A 785 nm laser (red) is used to trap nanoparticles and excite elastic and Raman scattering. A bright-field imaging path is shown in green, while an sCMOS camera collects the backward elastic scattering of trapped nanoparticles. (b) Acquired high-speed image of a weak-stiffness trapped single PS 200 nm nanobead with an overlaid trajectory recorded across ∼1s. The extracted 2D and 1D displacement are shown in (c), and then converted into variance vs. time-lag plots for size extraction as shown in (d). Black and Blue dots represent for measurement under weak and strong stiffness, respectively.

Download Full Size | PDF

To simplify the network training process, preprocessing, including extraction of 1D particle position from 2D images, is necessary. Firstly, we determine the sub-pixel level centroid of each particle in each frame and then re-link the individual nanoparticles between frames. Figure 1(b) shows an example of extracting 2D displacement from high-speed images of a weak-stiffness trapped individual 200 nm polystyrene (PS) nanoparticle. Next, the obtained time stack of 2D position was further reduced into a 1D position series along the direction of weakest trap stiffness (the blue-line axis shown in Fig. 1(c)) and use it as an input to a pre-trained network. As the particle’s motion is least constrained in this direction, it yields the particle motion characteristics that most closely approaches free Brownian motion. Comparison of size predictions along both axes are shown in Figure S1 in Supplement 1. To validate the performance of size prediction neural network on experimental data, size information was extracted from the same nanoparticle measured with reduced trap stiffness (10%∼20% of the above strong stiffness) and used as ground truth of training dataset. Generally, the relationship between time lag $\Delta t$ and position autocorrelation time ${\tau}$ were used determine the dynamic regime of trapped nanoparticles, the latter is written as,

$$\tau = {\raise0.7ex\hbox{${3\pi \eta d}$} \!\mathord{\left/ {\vphantom {{3\pi \eta d} K}}\right.}\!\lower0.7ex\hbox{$K$}}$$
Here $\eta $, d and K represent the viscosity of the solution, particle diameter and trapping stiffness. The time-scale condition $\Delta t < \mathrm{\tau }$ holds when using weak-stiffness trapping and short time lag [4], it allows approximating the particles’ motion as occurring in the diffusive-regime where the variance of displacement is equal to:
$${\sigma ^2}({\Delta t} )\approx 2D\Delta t$$
Therefore, the size-dependent diffusion coefficient D could be extracted by linear fitting between the variance of displacement to the time lag (as shown in Fig. 1(c)), which was described in the Stokes-Einstein equation:
$$D = \frac{{{k_B}T}}{{3\pi \eta d}}$$
Where ${k_B}$ is the Boltzmann coefficient, T is the absolute temperature. Variance would finally approach to a stiffness-determined constant $\frac{{{\textrm{k}_b}T}}{K}$ as $\Delta t$ increased to $\Delta t \gg \mathrm{\tau }$. We notice from Fig. 1(d) that size obtained from strong stiffness (25pN/µm) optical tweezers is severely underestimated due to capturing primarily non-diffusive motion, where the smallest time lag ($\Delta t$=1/6500${\approx} $0.15 ms) of our system is larger than $\mathrm{\tau }$ (around 0.07 ms). However, size obtained from weak stiffness (2.5 pN/µm) is accurate since diffusive motion could be monitored with the first three time lags (3$\Delta t$<$\mathrm{\tau }$).

2.2 Sample preparation and data processing of nanoparticles

NIST standard nanoparticles including 100 nm and 200 nm polystyrene (PS, Thermo Fisher, Massachusetts, USA) and 200 nm silica (Nanocomposite, San Diego, USA) nanoparticles were diluted with ultrapure water, then sonicated for 5 min and filtered using polycarbonate filters (Millipore) with 220 nm pore size before use. Additionally, EVs extracted from HN4 cell supernatant and human PFP were prepared by size exclusion chromatography (SEC) and ultracentrifugation, respectively. The detailed preparation steps of the EV samples are shown in the Supplement 1.

2.3 Generation of simulated time tracks of trapped nanoparticles

Considering the challenge of obtaining large datasets by single-beam optical tweezers with limited throughput, here we chose to instead train our network using purely simulated data. Based on the Monte-Carlo method [23], we simulated a large dataset of one-dimensional constrained Brownian motion of trapped nanoparticles to utilize as a training dataset. Since the nanoparticle moves during the recording of its position, its measured position is significantly affected by motion blur when the temporal camera bandwidth of the measurement is limited. Here we incorporate this blurring effect and subsequent positioning error into our simulations, with the simulated time series of individual trapped nanoparticle positions P(t) could be written as:

$$\textrm{P}(\textrm{t} )= 1/10(\mathop \sum \nolimits_{i = 1}^{10} {P_s}({\textrm{t} + \mathrm{\delta }{\textrm{t}_i}} )) + \varepsilon _G^2$$
Following Mortensen, K. et al [24], Where $\varepsilon _G^2$ represents frequency-independent position detection error from particle tracking process as well as the mechanical vibration of sample stage, with its amplitude depending on experimental measurements of a fixed nanoparticle on coverslip, as described by Ando. J. et al [25] (shown in Fig. 2(a)). ${P_s}(\textrm{t} )$ and $\mathrm{\delta }{\textrm{t}_i}$ represent the simulated 65kHz (where Δt<τ) sampled position series and sub-intervals during time lapse, the first item in Eq. (4) represents the motion blur by time-averaging the position across 10 sub-intervals. Theoretically, ${P_s}(\textrm{t} )$ could be described by the position probability density function, its normalized solution is a Gaussian function with a zero mean and variance given by Smoluchovwski equation [26]:
$${\sigma ^2}({\Delta t} )= \frac{{{\textrm{k}_b}T}}{K}\left( {1 - \exp \left( { - \frac{{2KD\Delta t}}{{{\textrm{k}_b}T}}} \right)} \right)$$

 figure: Fig. 2.

Fig. 2. Establishment of a simulated training dataset and the size prediction neural network. (a) The simulated training dataset under strong-stiffness (right) was established based on the Monte-Carlo method, including the camera’s blurring effect by time-averaging of position in each 10 frames (left), as well as the positioning error measured by tracking an individual 500 nm nanoparticle fixed on coverslip (middle). (b) The ResNet-50 architecture for particle size prediction. During training, the input to the training is the time series displacement under strong stiffness and the “ground truth” size information. The numbers (×3, × 4, × 6) under each residual block represents the number of times the three convolutional layers are repeated.

Download Full Size | PDF

Since size of nanoparticles in this work are much smaller than laser wavelength, stiffness K is primarily determined by the gradient force when ignoring the extremely weak scattering force (Taking PS200nm nanoparticle as an example, the calculated ${F_{sca}} = 6.1 \times {10^{ - 4}}$pN and ${F_{gra}}$=0.43pN respectively, ${F_{sca}}$ is ∼3 orders of magnitude smaller than ${F_{gra}}$), which can be written based on dipole approximation [27]:

$${F_{grad}} = \; \frac{{2\pi {d^3}}}{{{c_1}}}\left( {\frac{{{m^2} - 1}}{{{m^2} + 2}}} \right)\frac{{\partial I}}{{\partial r}}$$
Here m represents the ratio of the trapped nanoparticle's index to the medium's index, and ${c_1}$ is a constant. Assuming a Gaussian beam intensity distribution, the radial gradient of laser intensity $\frac{{\partial I}}{{\partial r}}$ could be written as:
$$\frac{{\partial I}}{{\partial r}} = \frac{{ - 4A}}{{{w^2}}}\exp \left( {\frac{{ - 2{r^2}}}{{{w^2}}}} \right)\ast r$$
Where w and A represent the beam waist ($\mathrm{\mu}\textrm{m}$) and intensity of the Gaussian beam (W/$\mathrm{\mu }{\textrm{m}^2})$, respectively. The intensity gradient could be simplified as a linear function of term r near the trap center (r→0). Thus, according to the simplified linear relationship ${F_{grad}} = Kr$, the stiffness can be rewritten as:
$$K = \; {c_2}\ast {d^3}\ast \left( {\frac{{{m^2} - 1}}{{{m^2} + 2}}} \right)$$
Where ${c_2} = \; - \frac{{8\pi A}}{{{c_1}\ast {w^2}}}$ is a beam-related constant. We can tell from Eq.8 that the stiffness K only depends on the size d and refractive index m of nanoparticles once the configuration of the laser trapping system is fixed, in which the constant ${c_2}$ could be determined by experimental measurement and downstream thermal motion analysis [28]. Taking individual PS 200 nm nanoparticle as example, measured optical stiffness as well as its related constant ${c_2}$ were determined to be 25 pN/µm and 0.2 µN/µm4, respectively. The beam in this configuration enabled stable trapping and strong scattering excitation, as shown in Supplement 1 Figure S2. Simulated stiffness of silica and extracellular vesicles could also be calculated according to their refractive index range reported in the literature [29,30]. Hence, the essence of the simulation is performing theoretical constrained Brownian profiles with user-defined particle size and refractive index distributions as input. The generation process of the simulation data is described in Fig. 2(a).

After following the above steps, a simulated constrained Brownian motion time series data was generated as shown in Fig. 2(a). In our case, a 20000-series-large database was constructed to perform training of each ResNet network, each time series were composed initially of 6000 frames (corresponding to an experimental measurement about 1 second in length). The a priori known size of nanoparticles and their corresponding simulated time series of strongly-constrained Brownian motion are regraded as the output and input for the neural network training, respectively.

2.4 ResNet-based size prediction network for strongly trapped nanoparticles

We propose a single-output one dimensional ResNet50-based network [31] for size prediction, as shown in Fig. 2(b). Briefly, our proposed network is composed of five main stages (1×conv at the beginning, 16 × 3 conv in the residual blocks, and 1×fully connected layer at the end, making up the 50 layers), with shortcut connections between the input and output of every residual block. A convolutional block is added in the shortcut path such that input and output have the same dimension. The shortcut or skip connections result in great improvement in both performance and training time in deeper networks, and help to mitigate the diminishing performance observed in deep neural networks by allowing the gradient information to pass through the layers. Briefly, the initial convolution layer has 64 filters, with the kernel size of 5, which is followed by the batch normalization, ReLU activation layer, and a max pooling layer of size 1 × 3, with a stride of 2. Next are the different groups of residual blocks with default filters and kernel sizes as indicated by the different colors in Fig. 2(b). The loss function is the mean squared error loss, with L2 regularization and dropout [32] applied to avoid overfitting.

We apportioned 90% of the simulated constrained Brownian motion of nanoparticles for the network training and the remaining 10% for validation, and employed a 10-fold cross-validation method to validate our model. The networks are trained using the Adam optimizer [33] for 50 epochs. The batch size equal to 32 was found to be empirically optimal and learning rate is set to 5e-5, The input to the network is the strong stiffness time series, and the ground truth size information. The network is constructed using the Tensorflow framework and Keras library, and training is accelerated with NVIDIA Geforce RTX-2080 GPU 8 GB RAM. Further, the network performance was evaluated by the root-mean-squared error of size prediction (RMSEP) on both simulated and experimental nanoparticles and extracellular vesicles datasets. An example of the training-testing loss curve during the network training is shown in Supplement 1 Figure S3.

3. Results and discussion

3.1 Simulation study and experimental validation of the size characterization network

Following training, simulated data of constrained Brownian motion from particles with different size and refractive index distributions (including 100 nm and 200 nm polystyrene (PS) standard nanobeads, 200 nm silica standard nanobeads and extracellular vesicles (EVs)) were generated as described above and used to evaluate the performance of our size prediction deep network. For each of these three different particle types studied, the validation dataset consists of 2000 time series examples that were generated independently of the data used for the training.

The performance for predicting size of different nanoparticles is shown in Fig. 3. As the camera bandwidth is not high enough to accurately sample diffusive motion, directly applying the Stokes-Einstein equation to the raw camera data yields very poor sizing results. These can be markedly improved by correcting the data and fitting with the adjusted Lorentzian fitting [17] (detail shown in Supplement 1 Figure S4). However, even better results can be obtained by both CNN-based deep learning (trained by the same simulated data, detail shown in Supplement 1 Figure S5), as well as our proposed ResNet, which has the optimum performance. It’s clearly shown that the ResNet achieved accurate precision and maintained a good degree of linear agreement in predicting the particle size across a wide range of sizes and nanoparticle types. The detailed root mean squared error of prediction (RMSEP) and mean absolute percentage error (MAPE) values are summarized in Table 1, including results from multiple different lengths of input time series. Following simulation validation, measurement of nanoparticles with the same sizes and materials as the simulation data were used to validate whether our proposed network training based on simulated datasets could generate accurate predictions from experimental data. It can be seen from Fig. 3(b) and Table 1 that experimental prediction error is slightly larger than simulated data in all cases, which points to the existence of data bias between simulated and measured data. This might be caused by the environmental flicker noise caused by instability of the laser source. Interestingly, we noticed from Fig. 3(b) that the ResNet performance on EVs seems to be degraded by the addition of a deterministic offset of the sizing predictions, which could be reduced by a simple linear regression (as shown in the Supplement 1 Figure S6 a). Therefore, despite the slightly increased error values, the size predictions are all highly accurate, with MAPE < 5% for both standard nanobeads and complex, highly heterogeneous EVs (shown in the second and fifth column in Table 1), which are close to the simulated results. Echoing the performance on experimental data, adjusted Lorentzian fitting and CNN-based deep networks showed worse performance. Detailed results of their fitting performance are shown in the Supplement 1.

 figure: Fig. 3.

Fig. 3. Single-particle level ResNet size prediction performance on both simulated (a) and experimental (b) datasets of polystyrene (PS) and silica standard nanobeads, and extracellular vesicles (EVs). Red lines represent lines of perfect agreement. (c) Averaged RMSEP for raw R-NTA, adjusted Lorentzian fitting, CNN and our ResNet sizing network. (error bars = mean ± STD).

Download Full Size | PDF

Tables Icon

Table 1. RMSEP and MAPE of the particle size predictions by the ResNet on the simulated and experimental datasets considering the different time series length.a

3.2 Exploring the effect of time series length on sizing performance

To further validate our ResNet prediction network, we evaluate its performance on reduced data lengths. To match the reduced input dimension, here we retrained network by using the same 20000 time series with either 3000 or 1000 frames extracted from the original full 6000-frame data sequences used in the original training. Figure 4(a) and (b) shows the effect of time series length on the ResNet prediction error using the experimental datasets. Meanwhile, linear regression model established on full frames data is extended to short-data-length conditions, the RMSEP versus time series length can be depicted in Fig. 4(c), with their specific values reported in Table 1. Traditionally, a certain number of frames (∼1s acquisition time) is required to obtain robust size predictions of nanoparticles under the diffusive-regime of constrained Brownian motion, the prediction accuracy decreases significantly with fewer data points (as shown in the Supplement 1 Figure S7). It can be seen from Fig. 4(c) that the EVs size estimation errors of our ResNet network have increased by 0.6 and 1.4 times when using only 1/2 or 1/6th of the full frames, respectively, indicating that the network is difficult to maintain performance after reducing the number of training data points. However, despite this performance degradation, our purposed network output still obtains substantially more accurate results compared to adjusted Lorentzian analysis and CNN, as well as the weak-stiffness trapping measurement. Furthermore, shortening the data lengths from, say, 6000 to 1000 also may allow us to increase the amount of training data, as now the original 20000 simulated tracks with 6000- frames each could be converted to 120000 tracks with 1000- frames each. These vectors could then be independently used as training data. The experimental results indicate that the RMSEP/ MAPE of EVs size prediction could be modestly improved from 18.33/8.6% to 17.2/6.5% respectively by increasing the amount of training data in this fashion, with the detail shown in Supplement 1s Figure S8.

 figure: Fig. 4.

Fig. 4. (a-b) Effect of time series length on ResNet prediction performance using experimental datasets of polystyrene (PS), silica standard nanobeads, and EVs using half and 1/6th of the full 6000 frames. (c) RMSEP versus time series length on EVs experimental data using different sizing methods. (error bars = mean ± STD).

Download Full Size | PDF

3.3 Size and refractive index characterization of human PFP-derived Vesicles

To validate the generalization ability, our proposed network is further applied to the analysis of vesicles derived from identical human PFP samples, but prepared with different purification methods (size exclusion chromatography, SEC and ultracentrifugation, UC). SEC and UC are both common purification methods for EVs [34], yet yield purified populations with slightly different size and refractive index distributions due to SEC selecting EVs based on size, and UC selecting based on density, as well as the “gentler” SEC purification limiting formation of particle aggregates. Further, a highlight of the high-speed camera-based optical tweezers measurement is its ability to extract multi-parametric information, such as refractive index of nanoparticles by fusing quantitative size and scattering images together. Similar to our previous work [13], by making the simplifying assumption that vesicles are structurally homogenous nanoparticles, Mie theory could be employed to determine the refractive index of each single vesicle, as details described in Figure S9 in the Supplement 1. Size distributions of SEC and UC-derived vesicles predicted by our proposed ResNet network were shown in Fig. 5(a), demonstrating extremely good agreement with results from gold-standard NTA measurements. The detailed average sizes are summarized in Table S2, we note that the UC derived vesicles have a wider distribution with larger mean sizes compared with SEC, which may be due to the aggregation effect caused by UC preparation [35]. In addition, the average refractive index analysis results of EVs yield distributions generally in the recognized EVs distribution range of 1.4-1.42. However, there are still around 20% of vesicles with refractive index above 1.45. These vesicles are likely not EVs but residual lipoprotein contaminants (such as very-low density lipoprotein, VLDL particles), which generally have higher refractive index than EVs [36] and are commonly observed in SEC purified PFP samples [37]. In contrast, these high refractive index vesicles are rarely present in UC preparation samples, which provides higher purification efficiency.

 figure: Fig. 5.

Fig. 5. Morpho-chemical characterization of PFP-derived EVs (total of 66 and 50 for SEC and UC respectively) based on ResNet size predictions. (a) Histogram of particle sizes predicted by ResNet (bins) and NTA (curves) for SEC (blue) and ultracentrifugation methods (black), respectively. (b) RI distributions vs. size, determined by the joint combination of scattering images and ResNet size predictions. Inset shows the RI distribution of SEC (blue) and UC (black) purified vesicles.

Download Full Size | PDF

4. Conclusions

In this paper, we have presented a ResNet-regression-enabled size prediction model based on imaging measurements of nanoparticles in optical tweezers, with results showing that the network accurately outputs the sizes of both simulated and experimental nanoparticles under strong trapping forces, which is challenging or impossible to achieve using standard solutions such as via a constrained Brownian model. The strong-stiffness trap regime is particularly important for measurements of small nanoparticles with low refractive-index contrast such as EVs. However, physics-based models do not perform well in these regimes, and data-driven methods such as deep learning offer a more promising approach for these complex non-linear regression problems. To demonstrate the utility of our method in these difficult situations, we explored the performance of our proposed network on experimentally measured bio-nanoparticles such as EVs, showing that our method substantially outperforms other state-of-art algorithms such as CNN and adjusted Lorentzian fitting methods.

Importantly, the training dataset was generated completely using simulated data, which enables the network to be rapidly and easily trained for different experimental settings and expected particle size and refractive index distributions. Additionally, experiments on EVs show that the ResNet network possesses a good generalization capability, where vesicles from different biological sources and from different preparation protocols were accurately predicted using a single trained network. Importantly, considering the sensitive sCMOS detection of weakly scattered light as well as the combination of size prediction network with scattering images measured by camera, optical tweezers measurement enables the ability to determine the refractive index of individual trapped vesicles according to Mie theory, which is challenging for methods based on position-sensitive sensors. Experimental results indicate that >100 plasma derived vesicles predicted by ResNet show strong agreement with measurements determined from gold-standard NTA, yielding both size and refractive index distributions in the range of 80-250 nm and 1.35-1.59, respectively. Considering the differences in physical characteristics between impurities (lipoproteins etc.) and EVs in PFP derived samples, simulated training datasets with larger sample numbers and more widely varying physical parameters could be employed in the future to reduce generalization error and obtain more accurate and reliable size predictions.

The essence of this work is aimed at solving the under-sampling problem in camera-based particle tracking through a ResNet network, which confirms the superiority of deep learning in solving complex nonlinear fitting issues. Wide potential exists for our network to be integrated with various optical tweezers-related characterization techniques, such as laser tweezer Raman spectroscopy (LTRS), etc. Our network could also possibly be applied for simultaneous multiplex sizing when using a high-speed camera and an integrated multi-focal tweezers system. Ultimately, then, we anticipate that further improvements in our proposed methods will find wider applications in any area where characterization of small, low-RI-contrast nanoparticles plays an important role, especially in EVs, liposomal nanomedicines etc.

Funding

National Natural Science Foundation of China (61775208); Anhui Province Key Research and Development Project (202003a07020020); Natural Science Foundation of Anhui Province (2208085QF220).

Disclosures

The authors declare no conflicts of interest related to this work.

Data Availability

The codes and simulated data for the implementation of our size prediction neural network presented in this paper are available in Ref. [38].

Supplemental document

See Supplement 1 for supporting content.

Reference

1. M.-C. Zhong, X.-B. Wei, J.-H. Zhou, Z.-Q. Wang, and Y.-M. Li, “Trapping red blood cells in living animals using optical tweezers,” Nat. Commun. 4(1), 1768 (2013). [CrossRef]  

2. Z. J. Smith, C. Lee, T. Rojalin, R. P. Carney, S. Hazari, A. Knudson, K. Lam, H. Saari, E. L. Ibañez, T. Viitala, T. Laaksonen, M. Yliperttula, and S. Wachsmann-Hogiu, “Single exosome study reveals subpopulations distributed among cell lines with variability related to membrane content,” J. Extracell. Vesicles 4(1), 28533 (2015). [CrossRef]  

3. Y. Pang, H. Song, J. H. Kim, X. Hou, and W. Cheng, “Optical trapping of individual human immunodeficiency viruses in culture fluid reveals heterogeneity with single-molecule resolution,” Nat. Nanotechnol. 9(8), 624–630 (2014). [CrossRef]  

4. B. Lukić, S. Jeney, C. Tischer, A. J. Kulik, L. Forró, and E. L. Florin, “Direct Observation of Nondiffusive Motion of a Brownian Particle,” Phys. Rev. Lett. 95(16), 160601 (2005). [CrossRef]  

5. Y. Pang, H. Song, and W. Cheng, “Using optical trap to measure the refractive index of a single animal virus in culture fluid with high precision,” Biomed. Opt. Express 7(5), 1672–1689 (2016). [CrossRef]  

6. R. H. Shepherd, M. D. King, A. A. Marks, N. Brough, and A. D. Ward, “Determination of the refractive index of insoluble organic extracts from atmospheric aerosol over the visible wavelength range using optical tweezers,” Atmos. Chem. Phys. 18(8), 5235–5252 (2018). [CrossRef]  

7. Y. Dai, Y. Yu, X. Wang, Z. Jiang, Y. Chen, K. Chu, and Z. J. Smith, “Hybrid principal component analysis denoising enables rapid, label-free morpho-chemical quantification of individual nanoliposomes,” Anal. Chem. 94(41), 14232–14241 (2022). [CrossRef]  

8. P. M. Bendix and L. B. Oddershede, “Expanding the optical trapping range of lipid vesicles to the nanoscale,” Nano Lett. 11(12), 5431–5437 (2011). [CrossRef]  

9. K. C. Neuman and S. M. Block, “Optical trapping,” Rev. Sci. Instrum. 75(9), 2787–2809 (2004). [CrossRef]  

10. S. Keen, J. Leach, G. Gibson, and M. J. Padgett, “Comparison of a high-speed camera and a quadrant detector for measuring displacements in optical tweezers,” J. Opt. A: Pure Appl. Opt. 9(8), S264–S266 (2007). [CrossRef]  

11. C. F. Bohren and D. R. Huffman, Absorption and Scattering by a Sphere, 1998.

12. K. Zembrzycki, S. Pawłowska, F. Pierini, and T. A. Kowalewski, Journal 15, (2023).

13. Y. Dai, S. Bai, C. Hu, K. Chu, B. Shen, and Z. J. Smith, “Combined morpho-chemical profiling of individual extracellular vesicles and functional nanoparticles without labels,” Anal. Chem. 92(7), 5585–5594 (2020). [CrossRef]  

14. E. van der Pol, F. A. W. Coumans, A. Sturk, R. Nieuwland, and T. G. van Leeuwen, “Refractive Index Determination of Nanoparticles in Suspension Using Nanoparticle Tracking Analysis,” Nano Lett. 14(11), 6195–6201 (2014). [CrossRef]  

15. D. O’Dell, P. Schein, and D. Erickson, “Simultaneous Characterization of Nanoparticle Size and Particle-Surface Interactions with Three-Dimensional Nanophotonic Force Microscopy,” Phys. Rev. Appl. 6(3), 034010 (2016). [CrossRef]  

16. P. Schein, D. O’Dell, and D. Erickson, “Orthogonal Nanoparticle Size, Polydispersity, and Stability Characterization with Near-Field Optical Trapping and Light Scattering,” ACS Photonics 4(1), 106–113 (2017). [CrossRef]  

17. A. van der Horst and N. R. Forde, “Power spectral analysis for optical trap stiffness calibration from high-speed camera position detection with limited bandwidth,” Opt. Express 18(8), 7670–7677 (2010). [CrossRef]  

18. H. W. H. Lachlan, R. M. Lauren, W. S. Oscar, A. N. Timo, R.-D. Halina, and C. D. L. Isaac, “Predicting particle properties in optical traps with machine learning,” Proc. SPIE 11469, 114691Z (2020). [CrossRef]  

19. G. S. Matthew, R. Jack, F. Eky, R. Jorge, O. M. Helen, B. M. Andrew, M. G. Graham, F. Daniele, and T. Manlio, “Machine learning opens a doorway for microrheology with optical tweezers in living systems,” arXiv, arXiv:2211.09689 (2022). [CrossRef]  

20. J. M. Newby, A. M. Schaefer, P. T. Lee, M. G. Forest, and S. K. Lai, “Convolutional neural networks automate detection for tracking of submicron-scale particles in 2D and 3D,” Proc. Natl. Acad. Sci. 115(36), 9026–9031 (2018). [CrossRef]  

21. Z. Wang, G. Chen, S. Wang, and X. Su, “Three-dimensional deep regression-based light scattering imaging system for nanoscale exosome analysis,” Biomed. Opt. Express 14(5), 2055–2067 (2023). [CrossRef]  

22. S. Helgadottir, A. Argun, and G. Volpe, “Digital video microscopy enhanced by deep learning,” Optica 6(4), 506–513 (2019). [CrossRef]  

23. S.-H. Xu, Y.-M. Li, L.-R. Lou, and Z.-W. Sun, “Computer simulation of the collision frequency of two particles in optical tweezers,” Chin. Phys. 14(2), 382–385 (2005). [CrossRef]  

24. K. I. Mortensen, H. Flyvbjerg, and J. N. Pedersen, “Confined Brownian Motion Tracked With Motion Blur: Estimating Diffusion Coefficient and Size of Confining Space,” Front. Phys. 8, 583202 (2021). [CrossRef]  

25. J. Ando, A. Nakamura, A. Visootsat, M. Yamamoto, C. Song, K. Murata, and R. Iino, “Single-Nanoparticle Tracking with Angstrom Localization Precision and Microsecond Time Resolution,” Biophys. J. 115(12), 2413–2427 (2018). [CrossRef]  

26. M. Lindner, G. Nir, A. Vivante, I. T. Young, and Y. Garini, “Dynamic analysis of a diffusing particle in a trapping potential,” Phys. Rev. E 87(2), 022716 (2013). [CrossRef]  

27. Y. Harada and T. Asakura, “Radiation forces on a dielectric sphere in the Rayleigh scattering regime,” Opt. Commun. 124(5-6), 529–541 (1996). [CrossRef]  

28. E. L. Florin, A. Pralle, E. H. K. Stelzer, and J. K. H. Hörber, “Photonic force microscope calibration by thermal noise analysis,” Appl. Phys. A 66(7), S75–S78 (1998). [CrossRef]  

29. I. H. Malitson, “Interspecimen Comparison of the Refractive Index of Fused Silica,” J. Opt. Soc. Am. 55(10), 1205–1209 (1965). [CrossRef]  

30. Y. Tian, C. Xue, W. Zhang, C. Chen, L. Ma, Q. Niu, L. Wu, and X. Yan, “Refractive Index Determination of Individual Viruses and Small Extracellular Vesicles in Aqueous Media Using Nano-Flow Cytometry,” Anal. Chem. 94(41), 14299–14307 (2022). [CrossRef]  

31. K. He, X. Zhang, S. Ren, and J. Sun, “Deep Residual Learning for Image Recognition,” arXiv, arXiv:1512.03385 (2015). [CrossRef]  

32. G. E. Hinton, N. Srivastava, A. Krizhevsky, I. Sutskever, and R. R. Salakhutdinov, “Improving neural networks by preventing co-adaptation of feature detectors,” arXiv, arXiv:1207.0580 (2012). [CrossRef]  

33. D. P. Kingma and J. Ba, “Adam: A Method for Stochastic Optimization,” arXiv, arXiv:1412.6980 (2017). [CrossRef]  

34. K. Brennan, K. Martin, S. P. FitzGerald, J. O’Sullivan, Y. Wu, A. Blanco, C. Richardson, and M. M. Mc Gee, “A comparison of methods for the isolation and separation of extracellular vesicles from protein and lipid particles in human serum,” Sci. Rep. 10(1), 1039 (2020). [CrossRef]  

35. R. Linares, S. Tan, C. Gounou, N. Arraud, and A. R. Brisson, “High-speed centrifugation induces aggregation of extracellular vesicles,” J. Extracell. Vesicles 4(1), 29509 (2015). [CrossRef]  

36. D. K. Anna, B. Martin, R. Marie, S. B. Andreas, S. Vahid, and D. Jan Van, “Interferometric nanoparticle tracking analysis enables label-free discrimination of extracellular vesicles from large lipoproteins,” bioRxiv, 2022.2011.2011.515605 (2022). [CrossRef]  

37. L. de Rond, S. F. W. M. Libregts, L. G. Rikkert, C. M. Hau, E. van der Pol, R. Nieuwland, T. G. van Leeuwen, and F. A. W. Coumans, “Refractive index to evaluate staining specificity of extracellular vesicles by flow cytometry,” J. Extracell. Vesicles 8(1), 1643671 (2019). [CrossRef]  

38. D. Boateng, K. Chu, Z. J. Smith, J. Du, and Y. Dai, “Deep learning-based size prediction for optical trapped nanoparticles and extracellular vesicles from limited bandwidth camera detection: code,” Github, 2023, https://github.com/derrick756/OpticalTweezerNet.

Supplementary Material (1)

NameDescription
Supplement 1       1. 1. Configuration of Optical Tweezers System, etc

Data Availability

The codes and simulated data for the implementation of our size prediction neural network presented in this paper are available in Ref. [38].

38. D. Boateng, K. Chu, Z. J. Smith, J. Du, and Y. Dai, “Deep learning-based size prediction for optical trapped nanoparticles and extracellular vesicles from limited bandwidth camera detection: code,” Github, 2023, https://github.com/derrick756/OpticalTweezerNet.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (5)

Fig. 1.
Fig. 1. Optical tweezers measurement system and data preprocessing. (a) Schematic of the optical tweezers system setup. A 785 nm laser (red) is used to trap nanoparticles and excite elastic and Raman scattering. A bright-field imaging path is shown in green, while an sCMOS camera collects the backward elastic scattering of trapped nanoparticles. (b) Acquired high-speed image of a weak-stiffness trapped single PS 200 nm nanobead with an overlaid trajectory recorded across ∼1s. The extracted 2D and 1D displacement are shown in (c), and then converted into variance vs. time-lag plots for size extraction as shown in (d). Black and Blue dots represent for measurement under weak and strong stiffness, respectively.
Fig. 2.
Fig. 2. Establishment of a simulated training dataset and the size prediction neural network. (a) The simulated training dataset under strong-stiffness (right) was established based on the Monte-Carlo method, including the camera’s blurring effect by time-averaging of position in each 10 frames (left), as well as the positioning error measured by tracking an individual 500 nm nanoparticle fixed on coverslip (middle). (b) The ResNet-50 architecture for particle size prediction. During training, the input to the training is the time series displacement under strong stiffness and the “ground truth” size information. The numbers (×3, × 4, × 6) under each residual block represents the number of times the three convolutional layers are repeated.
Fig. 3.
Fig. 3. Single-particle level ResNet size prediction performance on both simulated (a) and experimental (b) datasets of polystyrene (PS) and silica standard nanobeads, and extracellular vesicles (EVs). Red lines represent lines of perfect agreement. (c) Averaged RMSEP for raw R-NTA, adjusted Lorentzian fitting, CNN and our ResNet sizing network. (error bars = mean ± STD).
Fig. 4.
Fig. 4. (a-b) Effect of time series length on ResNet prediction performance using experimental datasets of polystyrene (PS), silica standard nanobeads, and EVs using half and 1/6th of the full 6000 frames. (c) RMSEP versus time series length on EVs experimental data using different sizing methods. (error bars = mean ± STD).
Fig. 5.
Fig. 5. Morpho-chemical characterization of PFP-derived EVs (total of 66 and 50 for SEC and UC respectively) based on ResNet size predictions. (a) Histogram of particle sizes predicted by ResNet (bins) and NTA (curves) for SEC (blue) and ultracentrifugation methods (black), respectively. (b) RI distributions vs. size, determined by the joint combination of scattering images and ResNet size predictions. Inset shows the RI distribution of SEC (blue) and UC (black) purified vesicles.

Tables (1)

Tables Icon

Table 1. RMSEP and MAPE of the particle size predictions by the ResNet on the simulated and experimental datasets considering the different time series length.a

Equations (8)

Equations on this page are rendered with MathJax. Learn more.

τ = 3 π η d / 3 π η d K K
σ 2 ( Δ t ) 2 D Δ t
D = k B T 3 π η d
P ( t ) = 1 / 10 ( i = 1 10 P s ( t + δ t i ) ) + ε G 2
σ 2 ( Δ t ) = k b T K ( 1 exp ( 2 K D Δ t k b T ) )
F g r a d = 2 π d 3 c 1 ( m 2 1 m 2 + 2 ) I r
I r = 4 A w 2 exp ( 2 r 2 w 2 ) r
K = c 2 d 3 ( m 2 1 m 2 + 2 )
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.