
Artificial Intelligence for Accelerating Time Integrations in Multiscale Modeling
Changnian Han
Peng Zhang
Danny Bluestein
Guojing Cong
Yuefan Deng
Corresponding Author: Yuefan Deng, Ph. D, Stony Brook University, Stony Brook, NY 11794, USA, Department of Applied Mathematics and Statistics,Yuefan.Deng@StonyBrook.edu
Issue date 2021 Feb 15.
Abstract
We developed a novel data-driven Artificial Intelligence-enhanced Adaptive Time Stepping algorithm (AI-ATS) that can adapt timestep sizes to underlying biophysical dynamics. We demonstrated its values in solving a complex biophysical problem, at multiple spatiotemporal scales, that describes platelet dynamics in shear blood flow. In order to achieve a significant speedup of this computationally demanding problem, we integrated a framework of novel AI algorithms into the solution of the platelet dynamics equations. Our framework involves recurrent neural network-based autoencoders by the Long Short-Term Memory and the Gated Recurrent Units as the first step for memorizing the dynamic states in long-term dependencies for the input time series, followed by two fully-connected neural networks to optimize timestep sizes and step jumps. The computational efficiency of our AI-ATS is underscored by assessing the accuracy and speed of a multiscale simulation of the platelet with the standard time stepping algorithm (STS). By adapting the timestep size, our AI-ATS guides the omission of multiple redundant time steps without sacrificing significant accuracy of the dynamics. Compared to the STS, our AI-ATS achieved a reduction of 40% unnecessary calculations while bounding the errors of mechanical and thermodynamic properties to 3%.
Keywords: Adaptive time stepping, Artificial intelligence, Multiscale modeling, Platelet dynamics
1. Introduction
Multiscale modeling (MSM), a method that multiple models at different scales are considered simultaneously, is a commonly used and rapidly evolving approach to study large and complex systems in biology, chemistry, material science, and fluid dynamics as well as biomedical engineering, involving spatial scales ranging from nanometer to millimeter and temporal scales from picosecond to millisecond [1]. To simulate a system in these areas often requires the hybridization of models at various scales and the processes governed by different laws of physics. These multiple scales have been manipulated intuitively by insights of the dynamics and, for temporal scales, the conventional standard time stepping algorithms (STS) use the smallest single timestep size for capturing sufficient details at the finest scale. This STS wastes substantial computing resources when high temporal resolutions are unnecessary. Since simulation at all scales proceeds concurrently, the has to be adjusted seamlessly to small and large for high- and low-frequency dynamics, respectively.
All-atom molecular dynamics (MD) simulation can capture sufficient details on spatiotemporal scales, but they must be the finest for all biomolecular states and require prohibitive, and unnecessary for many scales, computations [2]. By representing a cluster of atoms as one effective particle, MSM reduces computing time with potential loss of simulation accuracy. We developed a multiscale particle-based biomechanical model to study the platelets’ properties in thrombosis initial formation under flow-induced stresses at molecular scales [3]. The hydrodynamics of blood flow is simulated using dissipative particle dynamics (DPD) at a mesoscopic scale ranging from micrometer to millimeter in space and microsecond in time. The platelet particles are modeled by coarse-grained molecular dynamics (CGMD) at a microscopic scale with the timescale in picosecond to nanosecond, including bilayer membrane, gel-like cytoplasm, and elastic cytoskeleton [4,5]. STS uses the smallest during the entire simulation for all temporal scales to capture sufficient details, leading to a waste of resources for unnecessary computations on low-frequency dynamics. A multiple time stepping algorithm (MTS) was introduced to significantly speed up simulations by splitting components in modeling platelet rotation and each component is modeled with a specific [6]. Scalability and other performance analysis of the platelet-fluid model using MTS were carried out on three supercomputer architectures [7]. Recently, a more complex model of inter-platelet interaction was investigated by platelet recruitment aggregation within vitro experiments [8]. Platelets experience dynamic changes in time scale during various events in one simulation. The lack of adaptivity to time scales in time stepping algorithms introduces great complexity in the algorithm implementation. Advancing to study large-scale, long-time multi-platelet adhesion and activation in microchannels [8–10], we face prohibitive complexity for examining the platelet dynamics by the associated optimal temporal scales and, like many other ventures, this project certainly can benefit from artificial intelligence.
Developing such an AI-enhanced Adaptive Time Stepping algorithm (AI-ATS) to efficiently integrate the dynamics is the core of this study. Coupling large experimental data sets with the extracted dynamical features permits smart selection of modeling parameters, resulting in substantial savings in simulation time without visible loss of accuracy. We introduce the AI-ATS to intelligently adapt’s to underlying biophysical states for improving MSM. The AI-ATS framework constitutes neural networks (NNs) to guide online determinations of’s through training and inferencing. We use two state-of-the-art recurrent neural network (RNN) cells to analyze the time series of platelet dynamics and represent them by latent features, which help predict components’’s for the subsequent time stepping. To demonstrate the performance of our AI-ATS, we conduct numerical simulations with various’s and characterize the system dynamics by representative measurements of mechanics and thermodynamics. A NN-based solution framework is trained to output and step jump, indicating the duration before invoking the next inferencing. The dynamics simulation proceeds with the new for duration. Through benchmarking, we further compare our AI-ATS with STS in terms of accuracy and speed. Overall, our algorithm reduces computing time by ~40% while bounding the numerical errors over STS within 3%.
The rest of the manuscript is organized as follows:Sec. 2 reviews related work about MSM methods and time stepping algorithms by focusing on revealing their limitations to MSM simulation.Sec. 3 presents an overview of our previous MSM of platelets in shear blood flow and the detailed framework of our AI-ATS.Sec. 4 describes our MSM platelet simulation setups and summarizes the performance of our AI-ATS in accuracy and speed.Sec. 5 discusses the impacts, and the future extensions, of our AI-ATS.
2. Related Work
MSM can be classified into serial approaches and concurrent approaches [1,11] based on the ways of information passing between different scales. In the serial approaches, some constitutive relations in the macroscale model are precomputed from microscale models. Different scales are treated separately and these approaches, requiring less parameter passing, are suitable for weakly coupled problems such as peptides and proteins dynamics in the liquid solvent by multiscale coarse-graining methods [12]. For strongly coupled systems requiring computations of inter-particle forces at different scales on-the-fly, concurrent approaches are needed. However, time integrations at large scales produce many redundant computations leading to a significant waste. To address this issue, many MTS’s were introduced, including splitting spatial scales and matching’s to time integrations. Tuckerman et al. [13] introduced a reference system propagator algorithm (RESPA) to deal with systems consisting of mixtures of light and heavy particles. The algorithm splits the model by masses, treats fast motion of light particles as a reference system, and integrates the reference system for small’s and then the whole system for one large. Afterward, Tuckerman et al. [14,15] applied RESPA to systems with short and long-range forces by splitting forces based on the interaction ranges and achieved 10–20 times acceleration in a Lennard-Jones model of 864 molecules over STS algorithms. The same group improved the method by deriving reversible RESPA (r-RESPA) using the Trotter factorization of the Liouville propagator to solve the numerical instabilities caused by not reversible in time of reference system methods [16]. Humphreys et al. [17] generalized r-RESPA to macromolecular simulations with more complex potentials by force splitting and reduced computing time in 4–5 folds for a Crambin simulation of 655 particles. Due to the effectiveness and efficiency of the r-RESPA method, it is widely applied to various simulations [18] and detailed particle or force-based splitting strategies were illustrated through examples [19]. By extending the usage, Zhou et al. [20] combined the particle-particle particle-mesh Ewald method with r-RESPA and provided a new P3ME/RESPA method handling short and long-range electrostatic interactions more efficiently, 6–8 times faster than the Verlet algorithm for solvated protein simulations. Han et al. [21] derived a closed-form time step dependent Hamiltonian and analytically studied force splitting by Ewald summation and r-RESPA and numerically analyzed error and timing results for electrostatic potential computation. The multiscale separation is alleviated by MTS, but the maximum outer is limited by resonance phenomena [22]. Minary et al. [23] developed a set of resonance-free equations of motion permitting very large in MD simulations. Morrone et al. [24] designed a simple colored noise thermostatting scheme to break through the resonance limitation while preserving sufficient accuracy. Leimkuhler et al. [25] integrated stochastic isokinetic equations of motion with MTS (Stochastic-Iso-NH-RESPA) allowing as large as 100 fs for slow motions in a fixed-charge flexible water model. Margul et al. [26] presented a stochastic isokinetic algorithm using a polarizable model and achieved 10–20 times speedup in MD simulations of the AMOEBA water model. To sum up, MTS significantly accelerates the simulation without losing accuracy by computing the slow motion of heavy particles or long-range forces at the end of a large time step, a multiple of small’s fast motion or short-range forces.
In our study on MSM for platelets in shear blood flow by MTS, we split the forces based on both the component and the interaction range [6]. Regarding components, the system is decomposed into flow and platelet subsystems. Interaction forces within the platelet subsystem are considered as bounded and nonbounded forces by interaction ranges. Along with fluid-platelet interfacing, a four-level MTS for a model of platelet rotating freely in Couette flow was introduced with temporal scales ranging from micro- to femto-scales for corresponding spatial scales covering millimeter down to angstrom. Besides, the complex phenomena involved in thrombosis formation include platelet activation, aggregation, and adhesion. Current time stepping algorithms assess’s based on intuitions and adjustment of them on-the-fly greatly increases the implementation complexity. These issues, extracting physics-informed features from noisy data at runtime, prompt us to propose an ATS that intelligently adjusts’s. The ATS investigates the biophysical dynamics along the time and split states and assign optimal’s to each of them, in which way it further improves the computing efficiency by eliminating redundant computation carried out by the conventional algorithms.
3. AI-Enhanced Adaptive Time Stepping
For our prior MSM framework [3], we propose a novel ATS schema to enable a data-driven solution that can learn the optimal from the fluid-platelet interaction during platelet’s translation and rotation. For this schema, the implementation details including the learning architecture, the training and inferencing workflow are presented. Naturally, the proposed schema and its implementation are generalizable to a more complex fluid-structure interaction system of multiphysics involving fluid and solid mechanics at multiple scales.
3.1. Overview
The current study of fluid-platelet interactions follows our prior MSM framework [3–5,8–10] by incorporating the DPD method for the macroscopic transport of blood plasma in the vessel, governed by
(1) |
(2) |
where,, are the conservative, dissipative, random forces acting on particle by particle respectively, and are external forces exerted on each particle [27]. is the inter-particle distance, is the relative velocity, is the unit vector in the direction, is a random variable following normal distribution, and parameters,, control conservative, dissipative and random forces strength [27]. Español and Warren [28] introduced and, where is the Boltzmann constant and is the temperature of the system.
The classical DPD scheme, favored for its simple formulation, suffers from limitations in the thermodynamics behavior, too simplistic friction forces, lack of scaling, and sustaining temperature gradient [29]. The many-body DPD (MDPD) method [30] relaxes some of these limitations in allowing general equations of state by modulating the thermodynamics behavior at the interaction level between particles. The energy-conserving DPD (EDPD) model [31] extends the DPD model to non-isothermal situations for sustaining temperature gradients. The fluid particle model (FPM) [32] addresses the simplistic friction issue by incorporating shear forces between dissipative particles. By employing discretization of the Navier-Strokes equations in a Lagrangian moving Voronoi gird and thermal fluctuations, the smoothed DPD (SDPD) model [33] encapsulates all benefits of the three models discussed above.
The CGMD method for the dynamics of individual platelet in response to extracellular hydrodynamic stresses governed by
(3) |
The last two terms are bonded interactions in which and are the force constants and and are the equilibrium distance and angle, existing in the membrane and cytoskeleton. The Lennard-Jones (LJ) potential describes nonbonded interactions between any particle pairs within a cutoff range defined as
(4) |
where is the depth of the potential well and indicates the distance at which the inter-particle potential is zero. The disparate spatial scales between DPD and CGMD are handled by a hybrid force field [5] which establishes a functional spatial interface between the platelet membrane and the surrounding viscous fluid, governed by
(5) |
The platelet moves continuously according to the fluid-induced instantaneous stresses and, conversely, the flow’s state varies by the platelet’s motion.
In this MSM system, the flow-driven dynamics of individual platelet whose motion can be modeled as a rigid body in circumstances like flipping platelets in the free flow [5] and marginated platelets before adhesion to blood vessel walls [8], described as
(6) |
where is the mass, is the center-of-mass (COM) velocity, is the angular velocity, is the moment of inertia, of the platelet. Among them, we choose and rotational and translational kinetic energies, and, to demonstrate the platelet’s rotation and translation. The fluid state influencing the platelet is measured by the surrounding flow rate, the flow speed at the platelet’s COM. As the platelets flip slow (fast), the system evolves at large (small), adaptively. Algorithmically, the should change along with to facilitate capture of adequate dynamics at a low cost.
3.2. Data-Driven Time Stepping
Time stepping algorithms use time discretization to integrate the governing equations with a timestep size such as the scheme of the velocity Verlet algorithm [6]
(7) |
With initial conditions, the simulator iterates to time by computing the velocity at a half step and the position at full step, and then completing the calculation of the second half step velocity. In STS, all particles share a constant throughout the entire simulation, a conservative approach allowing capture of the fastest motion. In MTS, the system is decomposed into sub-systems according to spatial scales of system constituents or interaction force types. Then, small (large)’s are prescribed for sub-systems at bottom (top) scales. We consider finding as a function that maps system states to’s as subject to particles’ physical states. STS maps everything to a fixed conservative as regardless of the discrepancy between the system time scales. Comparing to STS, MTS considers multiple spatial scales and the mapping function follows, where is a vector of characteristic’s for component sub-systems [6,7]. To further improve simulation efficiency, a regulation on for the time integrator is derived to accommodate the constantly varying system dynamics.
The system state evolves with particles’ dynamics and the’s should synchronize itself with the underlying dynamics by mapping states to a variable as, where is an adaption at time bounded by the at bottom scales and at top scales. Intuitively, the adaptive can be computed by following the particle motion within a safe range. A data-driven ATS detects system dynamics and adjusts automatically. At the end of each integration step, ATS reassesses the most recent system states and alters to match the states when necessary.
We design the workflow for ATS to handle large streaming data generated during the simulation. ATS maintains the velocity Verlet algorithm as demonstrated inEq. (7). Next, instantaneous platelet and fluid properties are calculated and collected as raw data frequently from the simulator. Parallel to the simulator, ATS analyzes the raw data to predict an optimal for the near-future integrations. As a feedback, the ATS outputs are transferred back to the simulator to update the internal, and the simulation moves on with the new. Then, ATS waits for the time trigger for the next inference by repeating the same procedure above. Apparently, the frequency of ATS arbitration adapts to the state changes for an optimal examination of the system. For this optimization, a variable, step jump, is introduced to indicate the number of consecutive integration steps that can be skipped without computing. As depicted inFig. 1, ATS produces a pair of, where is a feedback to updating the dynamic trigger of invoking ATS.
Fig. 1.
AI-ATS framework showing (1) the collection of the raw data sequences from the simulator, (2) MA and WT filters for denoising these raw data, (3) the RNN-based AEs for latent feature extraction from the denoised data, and (4) the two FCNs for predicting and, the simulation timestep size and the duration before the next inference, respectively.
Because of the high-frequency oscillation, the dynamics requires pre-processing to better monitor the core motion. A denoising technique compresses the high-frequency signals along the time in each measurement separately as the first step in ATS. A long-term analysis is ideal to capture the present state and predict the trend. Therefore, we formulate denoised data points in a certain period in time series format to maintain the continuity of the data, and one sequence of each measurement is treated as an entry. To study the pattern in the large data sets, we integrate AI techniques with the data-driven ATS to predict and.
3.3. Learning Architecture
AI-ATS aims to predict the next, from the historical system states, by a deep learning framework that integrates RNN-based AEs and FCNs. The framework (Fig. 1) is composed of three components: 2-stage denoising by moving average (MA) and wavelet transform (WT) filters to clean high-frequency noise in raw data, two RNN-based AEs to compress and extract latent features, and two FCNs to predict and.
The platelet, modeled from the finest molecular scales, allows dynamic interactions with the surrounding flow and, for the first time, this study depicts a high-fidelity flow-induced stress mapping as the basis for accurate prediction of platelet activation and aggregation [8,10,34]. On the other hand, this made it challenging to extract the long-term patterns of platelet’s motion mixed by intense high-frequency oscillations on the membranes. Algorithmically in MSM, a platelet object is described as a collection of in-homogenous particles (membrane, receptors, cytoplasm, and cytoskeleton) interacted by a variety of governing functions. Among these massive particles’ interactions, our special interest is the large-scale motion and energy of the platelet subsystem under external forces. As an example,Fig. 1 depicts a snapshot of the raw data for the platelet rotational energy at which the features of interest are buried with the noisy raw data. To capture physics-informed features of a platelet dynamics in MSM (Sec. 3.1), we develop a 2-stage data denoising approach. Initially, the raw data is sampled directly from the MSM simulator with the same frequency as and stored as a time series. Stage-1: a MA filter is used for regulating time series of sampled raw data. It takes the samples of a time window of a fixed size and outputs the time-averaged values at the time point. The stride of MA is fixed as. After being processed by the MA filter, the unevenly sampled raw data are converted to an evenly spaced and filtered time series. Stage-2: a WT filter, when applied to the Stage-1 outputs for the noise removal, decomposes the input time series as the low and high-frequency components by [35]
(8) |
where
(9) |
The is the input data, and are the approximate and detail coefficients representing the orthogonal relationship between wavelet at resolution level and, in which and. The smoothed approximation is described by a function, translating and dilating of a scaling function, and the fine-scale details are generated by an orthonormal basis function, scaling of a mother wavelet [35]. By cleaning coefficients with thresholds at different resolutions, a smooth estimated signal is reconstructed as inverting the transform. The hard and soft thresholds are adopted, governed by operators [36]
(10) |
where the function extracts the signum, is the threshold at resolution level, for which we used a universal threshold [37] where is the length of the signal, and is an indicator function. The noise level is estimated from the finest scale wavelet coefficient defined as. By examining the snapshots showing passes through MA and WT filters inFig. 1, we find that the high-frequency oscillation movements are sufficiently suppressed and meanwhile informative dynamics are attained for being processed by the Autoencoders (AEs) to extract physics-informed learnable features.
The AEs transform these large amounts of filtered high-dimensional inputs further into a coded representation, a smaller amount of lower-dimensional latent features which, by design, preserve sufficient properties of original inputs. To achieve this, we employed RNN cells in AEs to process those time series inputs as RNNs “memorize” the relevant long-term information inside their hidden states throughout the processing of the time series. In this study, we used two state-of-the-art RNNs: Long Short-Term Memory (LSTM) [38] and Gated Recurrent Unit (GRU) [39] (Fig. 1). LSTM is developed to regulate memories by mechanisms called gates to better analyze long sequences obeying
(11) |
where the subscript indicates the time step, is the sigmoid function, and is the customized activation. The cell state, acting as the internal memory of the network, and the hidden state outputs the current state of the cell. GRU is a variant of LSTM and governed by
(12) |
where the hidden state acts as the internal memory. Update gate and reset gate are similar to and in the LSTM cell. We construct two AEs by LSTM and GRU respectively, and the training processes are discussed inSec. 3.4. The resultant latent features obtained from AEs are inputs of following FCNs.
In the last step, two FCNs using perceptron neurons are constructed for predicting and independently (Fig. 1). The input vector for both FCNs consists of latent features and most recently used’s. In prediction, a discrete value is outputted as a candidate of adaptive. In prediction, the integer used to measure the duration before the next inference, we select among a group of predefined categorical values by their predicted probability and this predication is a classification problem. At a time, a pair of and are provided as feedbacks to the simulator and the AI-ATS program: the is used in the integrator of the simulator and the is used by the AI-ATS scheme as a time trigger for the next inference.
3.4. Training Method
Training samples cover the duration of the platelet rotation of π in each of four typical motions, to be discussed inSec. 4.1, of platelets rotating and translating under shear blood flow. In the training and inferencing, we follow the same dimensionless unit system as in MSM. The conversion between the dimensionless unit and the SI unit can be found inTable 2 of [8]. During the data production process, six physics measurements (Sec. 3.1) are recorded as time series. In the MA filter, the sampling step size follows, and the moving window size and the moving stride are both 0.02. The in each time series covers 100–170 consecutive data points. One training sample is produced by averaging the time series within for all six measurements. In this study, we collected 400 samples for each of these four motions, resulting in a total of 1,600 samples prepared in the training database. In the training processes, 90% of these samples are used for training while the remaining 10% for validation.
Table 2.
The absolute and relative s in simulations under shear stress 40 dyne/cm2
Measures | Absolute Error | Relative Error | |||||||
---|---|---|---|---|---|---|---|---|---|
ATS | AI-ATS | SI Unit | ATS | AI-ATS | |||||
R.40 | RT.40 | R.40 | RT.40 | R.40 | RT.40 | R.40 | RT.40 | ||
0.02 | 0.01 | 0.02 | 0.03 | rad/ms | 1.45% | 1.10% | 1.32% | 2.58% | |
0.50 | 1.14 | 1.05 | 0.92 | ×10-9 pJ | 1.06% | 2.62% | 2.59% | 2.63% | |
2.47 | 43.62 | 2.30 | 44.81 | ×10-9 pJ | 0.00% | 0.01% | 0.00% | 0.03% | |
0.13 | 0.33 | 0.12 | 0.34 | mm/s | 0.33% | 0.15% | 0.07% | 0.45% |
The labeling process of finding the optimal and is based on a trial-and-error approach to determine the most aggressive and values that allow platelet-fluid dynamic interactions in our MSM. The search space for and are based on our prior knowledge as in [5,6]. At the trial, possible values are examined for performance tests; at the error, particle-based dynamics breaks the constraints discussed inSec. 3.2, determining the and.
In the training processes, AEs and FCNs are trained iteratively: the training of AEs is performed first for clearer physics-informed latent features, followed by the training of FCNs for the better approximation to labels. In the implementation, an AE is constructed as a 2-layer encoder of 16 and 4 units respectively and a 1-layer decoder of 16 units. Two types of AEs are implemented using LSTM and GRU cells (Fig. 1). In the LSTM-AE, mish and tanh are selected as the activation functions in the encoder and mish in the decoder, respectively. In the GRU-AE, the’s in the first and last layers are replaced by swish. Activation functions used in the model are illustrated inAppendix A. Each training sample is formatted as a 2-D matrix of dimensions 50×6 (steps×features). The mean square error (MSE) is employed as the training loss function.
Fig. 2 illustrates the performance of feature engineering processes by showing the linear correlation among the raw data, filtered data, latent features, and the target’s. When correlation is less than 0.3, most of the raw data cease to impact because particles’ high-frequency vibrations are mistreated in the presence of the generation motions. The data processed by the MA and WT filters exploit the hidden behaviors of measurements to amplify correlation. For example, those rotation-relevant measurements achieved more than 0.5. RNN-based AEs map data into a latent space where each projection combines partial original measurements and forms the most representative new features, boosting up the correlation to above 0.8. The correlation enhancement, aided by the feature engineering, boosts the ultimate determination of. The scatter plot (b) inFig. 3 uses the raw data (and the filtered and the latent feature) as an example to show a gradual improvement of the feature engineering performance. It is intuitively evident that our data denoising and feature extracting methods can effectively extract physics-informed features out of the noisy raw data with high-frequency motions.
Fig. 2.
Linear correlation among raw data, denoised data, latent features, and’s. In (a), the raw data (red) and the denoised data (green) from MA and WT filters, as well as the eight latent features (blue) extracted by RNN-based AEs. The correlation matrix shows linear relationship among all variables in R.30. The lower three rows are for RT.30, R.50, and RT.50. The insert (b) compares the raw data, the denoised data, and the latent feature, an example demonstrating the effectiveness of the feature engineering process.
Fig. 3.
AI-ATS inferencing workflow. The MSM simulator carries out the velocity Verlet algorithm. Required physics measurements computed in the MSM simulator are exported to the external datastore. AI-ATS is invoked starting to import MSM raw data and executing the inference. The outputs, and, are transferred back to the MSM simulator through the other external datastore. The simulator updates its internal parameters and continues to the next step.
Regarding FCNs, the input vector is composed of eight latent features obtained from AEs and five historical’s evenly selected from the most recent time window of 1 with interval 0.2. In the regression FCN for prediction, we construct a 4-layer NN of {32, 16, 8, 1} neurons in each layer respectively. Rectified linear units (ReLU) (All activation functions used in the NN model are depicted inFig. A.1.) are used as activation functions for all neurons, and the MSE is the loss function. We introduce a threshold to restrict the regulation strength on to prevent sharp changes by
(13) |
where controls the maximum deviation in one step, and the indicator function as the deviation between currently used and predicted reaches a threshold. In the classification FCN for prediction, the same NN architecture is applied in which the activation functions are swish and the last layer is replaced by a softmax layer. The target categories of are prescribed as {1,2,4} where each category represents a time duration of. For training purposes, target categories are transformed by one-hot encoding, using a binary value to indicate the presence of a sample at each category. The categorical cross-entropy is used as the loss function, defined as, where is the number of categories, is the ground truth (0 or 1), and is the predicted probability falls into the category.
The inference model is trained before deploying to test cases. In performance testing, the pre-trained model is used to collaborate with the simulator for determination. Relative to the underlying MSM time, the inferencing time is insignificantly small. The performance gain of AI-ATS highly depends on dynamics being modeled. Our current efforts focus on proving feasibility while showing the performance for a moderately steady motion
3.5. Inferencing Workflow
The inferencing method integrates the AI-ATS with the MSM simulator. The MSM simulator carries out time integration and passes simulation data to the AI-ATS program who analyzes data and predicts and for the simulator. The inferencing workflow is depicted inFig. 3. In the MSM simulator, per-particle properties,, and are updated in each time step following the velocity Verlet algorithm asEq. (7). Then, platelet and at the current time step are calculated based on in-memory per-particle data. Together with and, totally six physics measurements showing platelet and flow instantaneous states are exported from the simulator to the external datastore. Once the AI-ATS program is invoked by the trigger emitted from the simulator, it begins to import MSM raw data from the datastore and processes them through the pipeline consisting of 2-stage denoising filters, AEs, and FCNs. The inference outputs, and, are transferred back to the MSM simulator through the other external datastore. Meanwhile, the simulator waits for the feedback from the AI-ATS program to update its internal and. After synchronizing parameters by the end of step, all simulator processors continue to the next step.
Algorithm 1 illustrates the pseudocode of the MSM simulator. The velocity Verlet algorithm is implemented in Lines 2–5, followed by the computation and exportation of six physics measurements in Lines 6–7. The timer records simulated time since last inference on and. Each time reaches, the simulator triggers the AI-ATS program. As illustrated in Algorithm 2, the AI-ATS program starts with importing raw MSM data into in time series format. The Stage-1 denoising, the MA filter, processes the data with window size and moving stride and assembles evenly spaced results in; the Stage-2 denoising, the WT filter, outputs denoised time series in. The following AEs compress and extract eight physics-informed latent features. Then, both regression and classification FCNs take and historical as inputs and give the inference on and simultaneously. A physics constraint as inEq. (13) is applied to the final examination on. The AI-ATS program is concluded with exporting outputs to the other external datastore. The MSM simulator imports and and updates its internal parameters accordingly. At the end of each time step, parameters are broadcasted to all processors to synchronize the status. We set the MA window size, the moving stride, and the maximum allowance of variation. The threshold for variation is.
The inference model is trained before deploying to test cases. In performance testing, the pre-trained model is used to collaborate with the simulator for determination. Relative to the underlying MSM time, the inferencing time is insignificantly small. The performance gain of AI-ATS highly depends on dynamics being modeled. Our current efforts focus on proving feasibility while showing the performance for moderately steady motion.
4. Numerical Simulations and Analysis
To illustrate and study AI-ATS, we conduct numerical simulations on an MSM of platelet under shear blood flow by STS, ATS, and AI-ATS. We benchmarked AI-ATS in terms of accuracy and speed. An overall performance statement of comparing AI-ATS to STS is summarized in the end.
4.1. Simulations
In the MSM, we study platelet dynamics in a variety of shear blood flow. A microchannel is simulated by a rectangular region with dimensions of 16×16×8 in as displayed inFig. 4. The periodic boundary condition is applied along- and-axes. The blood vessel walls are modeled at top and bottom boundaries in-axis and no-slip boundary condition [40] is applied. The entire system incorporated 1,091,360 fluid particles and 140,303 platelet particles. The shear blood flow is modeled by counter Couette flow by driving top and bottom walls in opposite directions along-axis. We consider two cases of platelet initial positions, near the central line–rotation motion (R) is dominant–and near the blood vessel–mixture of rotation and translation (RT). With different moving velocities of walls, flow conditions of shear stresses {30, 40, 50} in dyne/cm2 are applied to the system. A total of 18 simulations (Table 1) were conducted to cover three different algorithms, two platelet motion types, and three shear stresses. The same force field parameters in [8] are used in this work. Simulations under shear stresses 30 and 50 in dyne/cm2 are used for training purposes. Adaptive’s are selected from four empirical values of {416, 624, 832, 936} in ps. The instantaneous physics measurements are collected at the 40-ns basis from the simulator. With the trained model, AI-ATS is benchmarked in simulations under shear stresses 40 dyne/cm2 against a reference system of using STS at the smallest ps. All numerical simulations are implemented by modified LAMMPS package [41] on the IBM’s WSC Cluster consisting of IBM AC-922 nodes.
Fig. 4.
A 2-D cross-section of a typical 3-D simulation setup in domain 16×16×8 in. No-slip boundary condition is applied on top and bottom vessel walls, and periodic boundary condition on- and-axes. Two cases represent two different platelet initial conditions: near the central line (R) and near the blood vessel (RT).
Table 1.
The list, and naming convention, of 6 simulations each of which is tested by STS, ATS, and AI-ATS, for a total of 18 simulations
30 dyne/cm2 | 40 dyne/cm2 | 50 dyne/cm2 | |
---|---|---|---|
R | R.30 | R.40 | R.50 |
RT | RT.30 | RT.40 | RT.50 |
4.2. Analysis
We performed simulations of platelet rotation in R.40 and RT.40 by STS, ATS, and AI-ATS. Performance metrics are defined to quantize measurement precisions. Accuracy is investigated in terms of platelet dynamics and mechanics. Speed is analyzed in terms of the variation of’s and simulation steps.
4.2.1. Performance Metrics
The noise in our particle-based simulations results from the high-frequency particle oscillation. To reveal the major trend of dynamics, we used the same MA filter as inSec. 3.3 to clean the raw data. The accuracy analysis is carried out in terms of dynamics and mechanics. In dynamics, both platelet rotation and translation are examined. For analysis of rotation, we fit regression curves to discrete data points by the Jeffery orbits equation [42] on the angular velocity with two translation and one dilation coefficients, given by
(14) |
where
(15) |
and is the aspect ratio of the minor axis to the major axis. is a dilation coefficient and and are translation coefficients. For analysis of translation, the fact fluctuates around a value allows us to describe the motion by a mean-value trending line. The deviation between a simulation and the reference system is measured by defining the time-averaged absolute and relative errors as
(16) |
(17) |
where indicates the duration of the integration and the physics measurement as a function of time.
In mechanics, the per-particle stress is expressed by a tensor where and take on values,, to generate the tensor. Following [5], the same algorithm is used here to calculate the instantaneous stress scalar by
(18) |
where denotes the stress scalar of particle at time. To reveal the pattern of stress on the platelet membrane, we average in time and space. Following the temporal-spatial averaging technique as described in [5], we select the time-averaging of 5 µs and spatial-averaging of 0.178 µm. The averaged stress scalar is normalized by the mean stress value among all membrane particles. The per-particle stress relative deviation between two systems is defined as
(19) |
where is the normalized stress scalar.
4.2.2. Accuracy
The platelet rotation and translation results are presented and compared in the first and second rows ofFig. 5, respectively. The scatter points represent the processed raw data after MA filters. The rotation data is further analyzed by fitting to the modified Jeffery orbits equation as described inEq. (14), shown as the “fit curve” in the figure. The translation data is calculated as a mean value for each case so-called as the “trending line”. By analyzing the results (Fig. 5) from simulations involving STS, ATS, and AI-ATS, we claim (a) in R.40, all algorithms preserved the same insignificant translation of the platelet COM, implying that the platelet rotates near the channel central line; (b) in RT.40, all algorithms described the same translation velocity of the platelet COM; (c) in R.40 and RT.40, the fit curves demonstrated fairly consistent rotations (including angular velocities and rotational periods) by all algorithms.
Fig. 5.
Comparison of and in R.40 and RT.40 by STS, ATS, and AI-ATS. The scatter points are the preprocessed data by the MA filter. The fit curves in plots are generated usingEq. (14), and trending lines of mean values of preprocessed data are used in plots.
In addition,Table 2 summarizes absolute and relative errors of platelet dynamics in terms of velocities and kinetic energies calculated followingEqs. (16)-(17). In platelet rotation, both ATS and AI-ATS bound the error within 3% accuracy loss, where the deviation of and are bounded by 0.02 rad/ms and 1.05×10-9 pJ in R.40, and 0.03 rad/ms and 1.14×10-9 pJ in RT.40. In platelet translation, both ATS and AI-ATS achieve less than 1% accuracy loss, where the deviation of and are bounded by 2.47×10-9 pJ and 0.13 mm/s in R.40, and 4.48×10-8 pJ and 0.34 mm/s in RT.40.
In mechanics, we examine the per-particle virial stress as a means to reveal the spatial variation of the mechanical stress [43].Fig. 6-Fig. 7 show normalized virial stress scalar on the platelet membrane in R.40 and RT.40, respectively. Five representative orientations are selected at rotation angles of {0, π/4, π/2, 3π/4, π}, where the angle 0 represents the platelet major axis is aligned to-axis, reaching the fastest angular velocity. The results inFig. 6-Fig. 7 consistently demonstrated: (a) intuitively, all the STS, ATS, and AI-ATS showed the same stress distribution patterns and evolution processes. For example, in all snapshots, relatively larger (smaller) stress mapping all appears near the peripheral (center) areas. The high and low stresses (i.e., red and blue spots) distributions are in good agreement among all algorithms. Thus, the accuracy for hemodynamic stress mapping on platelet membranes is very high using the AI-ATS; (b) quantitatively, the histogram statistics of relative errors in stress mapping computed byEqs. (18)-(19) further quantified the negligible numeric errors by comparing AI-ATS with STS. More than 72% of particles maintain a relative error of less than 1% through the entire simulation. Overall, the error is well-bounded by 3%.
Fig. 6.
Normalized virial stress profile on platelet membrane in R.40. From top to down, each row shows the results of the same setup but using STS, ATS, or AI-ATS, respectively. Each column represents the results at the same orientation. Last row: the histogram plots statistics of per-particle relative error, which is bounded by 3% through the entire simulation.
Fig. 7.
Normalized virial stress profile on platelet membrane in RT.40. From top to down, each row shows the results of the same setup but using STS, ATS, or AI-ATS, respectively. Each column represents the results at the same orientation. Last row: the histogram plots statistics of per-particle relative error, which is bounded by 3% through the entire simulation.
4.2.3. Computational Efficiency
Simulation time is estimated as the number of integration steps. Simulation efficiency is improved by varying the values.Fig. 8 presents the values during the simulation in R.40 and RT.40 by STS, ATS, and AI-ATS. STS utilizes the smallest and fixed for accuracy purposes. ATS alters for six times by using four empirical values (as described inSec. 4.1) with a relatively large jump at each adjustment. AI-ATS regulates’s in a relatively smoother manner for more than 20 times with a variety of 12–14 values. The maximally allowed jump of at each inference is 42 ps, one-tenth of the used in STS. The results using AI-ATS consistently follows the pattern on regulation in simulations by ATS. Compared to STS, both ATS and AI-ATS obtained a reduction of 40% unnecessary simulation computations.Fig. 9 compares the overhead associated with one time inference. In ATS, the overhead in manual inference for a new includes state data downloading, visual inspection, analysis, and the MSM restart. The period of such inference is approximately two hours in wall clock time. In AI-ATS, the overhead of the inference, free of manual labor and executed on-the-fly, is negligible compared to that of the ATS. To minimize inference requests, we introduce to maximize the duration to use the current. In this case, without, each ATS simulation requires more than 400 inferences while, with, the inference requests drop to 120–130, cutting by more than 65%. The data needed for the inferences are proportional to the system complexity measured in the number of platelets while the processing overhead of the inferences, growing rapidly for the ATS with the system complex and dynamics, will change little for the AI-ATS. The efficiency results demonstrate the efficacy of our delicate AI-ATS in eliminating unnecessary computations in complex multiscale systems from a practical standpoint. Extrapolating from the training shear stress of 30 and 50 dyne/cm², we made predictions of the for an aggressive shear stress of 100 dyne/cm² for which the AI-ATS sped up the simulation by 19.3%.
Fig. 8.
Timestep sizes comparison in R.40 and RT.40. In STS (green solid), one smallest and fixed is applied. In ATS (red dot),’s are adjusted manually with four empirical values (Sec. 4.1). In AI-ATS (blue dash),’s are regulated automatically by NNs in a smoother manner.
Fig. 9.
Comparison of overhead associated with one time inference in ATS and AI-ATS. In ATS, each inference requires human involved workload including data visualization, analysis, and restart. In AI-ATS, each inference goes through the data pre-processing, AI analysis, and resume. AI-ATS reduced the overhead by 99.87%.
5. Discussions
We introduce an ATS that is capable of adjusting the time step size’s to accommodate the needs of multiple spatial scales in the biophysical dynamics. Such empirical and static ATS, incapable of handling large and complex dynamics, is enhanced by AI with our designed neural network to infer from the latest physical states. To effectively use the noisy state data, we develop a sequence of denoising filters to extract physics-informed latent features, and the FCNs to regulate’s. AI-ATS workflow is implemented by modifying the velocity Verlet algorithm with a dynamic trigger to alter. The capability and potential of our AI-ATS are demonstrated by a group of MSM simulations of studying the platelet motion in shear blood flow. The AI-ATS, benchmarked against a reference system using STS for this fairly steady flow, shows its merit in preserving physical states in terms of platelet dynamics and mechanics with over 97% accuracy and in eliminating 40% unnecessary computing. Automation of the AI-ATS remedies the complexity in ATS to predict’s by expensive manual efforts.
Problems to be modeled grow rapidly in size and complexity, for example, in the study of shear-induced platelet-mediated thrombosis formation, the system is large and the biophysical processes are complex. Incorporating various types of interactions at multiple scales including inter-cellular interactions in margination and aggregation, cell-vessel interactions in adhesion, and intra-cellular interactions in activation, adaptive’s are needed to speedup computations substantially while achieving expected simulation accuracy. AI, as the wave of the future, eliminates massive manual labor to monitor and to adjust simulations, adaptively, and enables modeling of realistic systems that are much bigger and much more complex. AI-ATS is one example to demonstrate the need and feasibility. AI-ATS gained speedup for a particular application of tracking the motions of platelets in shear flows that consist of multiple constituents of particles interacting at molecular to cellular scales. The general applicability of AI-ATS has yet to be explored, although the potential of physics-informed machine learning for accelerating complex simulations is noticeable.
Overlooking the huge gains in the overall performance of the AI-ATS over the ATS is not uncommon. The fact that ATS’s determination of requires significant manual interruptions of the underlying simulations to process several massive time series including pausing, reverting, analyzing, and restarting is usually missing in terse descriptions. The AI-ATS, on the other hand, can by-pass these steps when determining the on-the-fly. Furthermore, as we expand the simulated systems in sizes and physical complexities for more realistic applications, the overhead of ATS will increase rapidly overwhelm the underlying simulations.
Optimizing multiscale modeling of complex biomedical systems posts severe challenges. The AI-ATS introduced in this work, displaying its initial potency to address some of these challenges, can be further improved in three directions. First, to expand the automation of scale detection to spatial dimensions to intelligently and dynamically group system components of similar scales for adaptive time stepping. Second, to map various components to minimize load imbalance and communication by topology-aware task mapping techniques [34,35] on supercomputing architectures involving CPUs and GPUs complexes. Third, to analyze the detailed balance and time reversibility among the multiple’s.
This integration of MSM with AI that intelligently selects the optimal’s by examining the underlying dynamics has demonstrated great potential in improving the accuracy and speed of modeling a complex biomedical system.
Algorithms
Supplementary Material
Algorithm 1: MSM simulator. | ||
1 | repeat | |
2 | /* Verlet */ | |
3 | /* Verlet */ | |
4 | compute | /* Verlet */ |
5 | /* Verlet */ | |
6 | compute physics measurements,,, | /* measurements computation */ |
7 | export data to datastore | /* data export */ |
8 | update simulation time | /* accumulate t */ |
9 | update timer | /* accumulate l */ |
10 | if timerl = step jumpthen | /* AI-ATS trigger */ |
11 | invoke AI-ATS program | /* invoke AI-ATS */ |
12 | import from datastore | /* data import */ |
13 | update | /* update step size */ |
14 | update | /* update step jump */ |
15 | reset timer | /* reset timer l */ |
16 | end if | |
17 | synchronize on all processors | /* synchronization */ |
18 | until simulation ends |
Algorithm 2: AI-ATS. | ||
parameters:,,, | ||
1 | raw data import data from datastore | /* data import */ |
2 | averaged data MA_filter | /* denoising */ |
3 | denoised data WT_filter | /* denoising */ |
4 | latent features AE | /* feature extraction */ |
5 | regression_FCN | /* step size inference */ |
6 | classification_FCN | /* step jump inference */ |
7 | ifthen | /* physics constraint */ |
8 | set | /* finalize step size */ |
9 | end if | |
10 | export to datastore | /* data export */ |
Acknowledgement
This publication was made possible by a grant from the National Institutes of Health NHLBI 5U01 HL13105205 (PI: D. Bluestein, Co-Investigators: Y. Deng, M. J. Slepian) and a grant from the SUNY-IBM Consortium Award, IPDyna: Intelligent Platelet Dynamics, FP00004096 (PI: Y. Deng, Co-Investigator: P. Zhang). The simulations in this study were conducted on WSC Cluster at the IBM Thomas J. Watson Research Center through an IBM Faculty Award FP0002468 (PI: Y. Deng).
Footnotes
Appendix A
All activation functions used in the NN model are depicted inFig. A.1.
Fig.A.1. Activation functions used in the model.
References
- [1].Ayton GS, Noid WG, Voth GA, Multiscale Modeling of Biomolecular Systems: In Serial and in Parallel, Current Opinion in Structural Biology, 17 (2007) 192–198. [DOI] [PubMed] [Google Scholar]
- [2].Dror RO, Dirks RM, Grossman J, Xu H, Shaw DE, Biomolecular Simulation: A Computational Microscope for Molecular Biology, Annual Review of Biophysics, 41 (2012) 429–452. [DOI] [PubMed] [Google Scholar]
- [3].Zhang P, Zhang L, Slepian MJ, Deng Y, Bluestein D, A Multiscale Biomechanical Model of Platelets: Correlating with in-Vitro Results, Journal of Biomechanics, 50 (2017) 26–33. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [4].Zhang N, Zhang P, Kang W, Bluestein D, Deng Y, Parameterizing the Morse Potential for Coarse-Grained Modeling of Blood Plasma, Journal of Computational Physics, 257 (2014) 726–736. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [5].Zhang P, Gao C, Zhang N, Slepian MJ, Deng Y, Bluestein D, Multiscale Particle-Based Modeling of Flowing Platelets in Blood Plasma Using Dissipative Particle Dynamics and Coarse Grained Molecular Dynamics, Cellular and Molecular Bioengineering, 7 (2014) 552–574. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [6].Zhang P, Zhang N, Deng Y, Bluestein D, A Multiple Time Stepping Algorithm for Efficient Multiscale Modeling of Platelets Flowing in Blood Plasma, Journal of Computational Physics, 284 (2015) 668–686. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [7].Zhang P, Zhang N, Gao C, Zhang L, Gao Y, Deng Y, Bluestein D, Scalability Test of Multiscale Fluid-Platelet Model for Three Top Supercomputers, Computer Physics Communications, 204 (2016) 132–140. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [8].Gupta P, Zhang P, Sheriff J, Bluestein D, Deng Y, A Multiscale Model for Recruitment Aggregation of Platelets by Correlating with in Vitro Results, Cellular and Molecular Bioengineering, 12 (2019) 327–343. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [9].Gao C, Zhang P, Marom G, Deng Y, Bluestein D, Reducing the Effects of Compressibility in Dpd-Based Blood Flow Simulations through Severe Stenotic Microchannels, Journal of Computational Physics, 335 (2017) 812–827. [Google Scholar]
- [10].Pothapragada S, Zhang P, Sheriff J, Livelli M, Slepian MJ, Deng Y, Bluestein D, A Phenomenological Particle-Based Platelet Model for Simulating Filopodia Formation During Early Activation, International Journal for Numerical Methods in Biomedical Engineering, 31 (2015) 1–16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [11].Garcia-Cervera CJ, Ren W, Lu J, Weinan E, Sequential Multiscale Modeling Using Sparse Representation, Communications in Computational Physics, 4 (2008) 1025–1033. [Google Scholar]
- [12].Hills RD Jr, Lu L, Voth GA, Multiscale Coarse-Graining of the Protein Energy Landscape, PLOS Computational Biology, 6 (2010) 1–12. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [13].Tuckerman ME, Berne BJ, Rossi A, Molecular Dynamics Algorithm for Multiple Time Scales: Systems with Disparate Masses, The Journal of Chemical Physics, 94 (1991) 1465–1469. [Google Scholar]
- [14].Tuckerman ME, Berne BJ, Martyna GJ, Molecular Dynamics Algorithm for Multiple Time Scales: Systems with Long Range Forces, The Journal of Chemical Physics, 94 (1991) 6811–6815. [Google Scholar]
- [15].Tuckerman ME, Berne BJ, Molecular Dynamics in Systems with Multiple Time Scales: Systems with Stiff and Soft Degrees of Freedom and with Short and Long Range Forces, The Journal of Chemical Physics, 95 (1991) 8362–8364. [Google Scholar]
- [16].Tuckerman ME, Berne BJ, Martyna GJ, Reversible Multiple Time Scale Molecular Dynamics, The Journal of Chemical Physics, 97 (1992) 1990–2001. [Google Scholar]
- [17].Humphreys DD, Friesner RA, Berne BJ, A Multiple-Time-Step Molecular Dynamics Algorithm for Macromolecules, The Journal of Physical Chemistry, 98 (1994) 6885–6892. [Google Scholar]
- [18].Watanabe M, Karplus M, Simulations of Macromolecules by Multiple Time-Step Methods, The Journal of Physical Chemistry, 99 (1995) 5680–5697. [Google Scholar]
- [19].Stuart SJ, Zhou R, Berne BJ, Molecular Dynamics with Multiple Time Scales: The Selection of Efficient Reference System Propagators, The Journal of Chemical Physics, 105 (1996) 1426–1436. [Google Scholar]
- [20].Zhou R, Harder E, Xu H, Berne B, Efficient Multiple Time Step Method for Use with Ewald and Particle Mesh Ewald for Large Biomolecular Systems, The Journal of Chemical Physics, 115 (2001) 2348–2358. [Google Scholar]
- [21].Han G, Deng Y, Glimm J, Martyna G, Error and Timing Analysis of Multiple Time-Step Integration Methods for Molecular Dynamics, Computer physics communications, 176 (2007) 271–291. [Google Scholar]
- [22].Biesiadecki JJ, Skeel RD, Dangers of Multiple Time Step Methods, Journal of Computational Physics, 109 (1993) 318–328. [Google Scholar]
- [23].Minary P, Tuckerman ME, Martyna GJ, Long Time Molecular Dynamics for Enhanced Conformational Sampling in Biomolecular Systems, Phys Rev Lett, 93 (2004) 150201. [DOI] [PubMed] [Google Scholar]
- [24].Morrone JA, Markland TE, Ceriotti M, Berne B, Efficient Multiple Time Scale Molecular Dynamics: Using Colored Noise Thermostats to Stabilize Resonances, The Journal of chemical physics, 134 (2011) 014103. [DOI] [PubMed] [Google Scholar]
- [25].Leimkuhler B, Margul DT, Tuckerman ME, Stochastic, Resonance-Free Multiple Time-Step Algorithm for Molecular Dynamics with Very Large Time Steps, Molecular Physics, 111 (2013) 3579–3594. [Google Scholar]
- [26].Margul DT, Tuckerman ME, A Stochastic, Resonance-Free Multiple Time-Step Algorithm for Polarizable Models That Permits Very Large Time Steps, Journal of Chemical Theory and Computation, 12 (2016) 2170–2180. [DOI] [PubMed] [Google Scholar]
- [27].Groot RD, Warren PB, Dissipative Particle Dynamics: Bridging the Gap between Atomistic and Mesoscopic Simulation, The Journal of chemical physics, 107 (1997) 4423–4435. [Google Scholar]
- [28].Espanol P, Warren P, Statistical Mechanics of Dissipative Particle Dynamics, EPL (Europhysics Letters), 30 (1995) 191. [Google Scholar]
- [29].Espanol P, Warren PB, Perspective: Dissipative Particle Dynamics, J Chem Phys, 146 (2017) 150901. [DOI] [PubMed] [Google Scholar]
- [30].Trofimov S, Nies E, Michels M, Thermodynamic Consistency in Dissipative Particle Dynamics Simulations of Strongly Nonideal Liquids and Liquid Mixtures, The Journal of chemical physics, 117 (2002) 9383–9394. [Google Scholar]
- [31].Espanol P, Dissipative Particle Dynamics with Energy Conservation, EPL (Europhysics Letters), 40 (1997) 631. [Google Scholar]
- [32].Espanol P, Fluid Particle Model, Physical Review E, 57 (1998) 2930. [Google Scholar]
- [33].Espanol P, Revenga M, Smoothed Dissipative Particle Dynamics, Phys Rev E Stat Nonlin Soft Matter Phys, 67 (2003) 026705. [DOI] [PubMed] [Google Scholar]
- [34].Bluestein D, Soares JS, Zhang P, Gao C, Pothapragada S, Zhang N, Slepian MJ, Deng Y, Multiscale Modeling of Flow Induced Thrombogenicity with Dissipative Particle Dynamics and Molecular Dynamics, Journal of Medical Devices, 8 (2014) 024502–024503. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [35].Mallat SG, A Theory for Multiresolution Signal Decomposition: The Wavelet Representation, IEEE Transactions on Pattern Analysis and Machine Intelligence, 11 (1989) 674–693. [Google Scholar]
- [36].Donoho DL, Johnstone JM, Ideal Spatial Adaptation by Wavelet Shrinkage, Biometrika, 81 (1994) 425–455. [Google Scholar]
- [37].Donoho DL, De-Noising by Soft-Thresholding, IEEE Transactions on Information Theory, 41 (1995) 613–627. [Google Scholar]
- [38].Hochreiter S, Schmidhuber J, Long Short-Term Memory, Neural Computation, 9 (1997) 1735–1780. [DOI] [PubMed] [Google Scholar]
- [39].Cho K, Van Merriënboer B, Gulcehre C, Bahdanau D, Bougares F, Schwenk H, Bengio Y, Learning Phrase Representations Using Rnn Encoder-Decoder for Statistical Machine Translation, Conference on Empirical Methods in Natural Language Processing (EMNLP 2014), (2014) 1724–1734. [Google Scholar]
- [40].Soares JS, Gao C, Alemu Y, Slepian M, Bluestein D, Simulation of Platelets Suspension Flowing through a Stenosis Model Using a Dissipative Particle Dynamics Approach, Annals of Biomedical Engineering, 41 (2013) 2318–2333. [DOI] [PMC free article] [PubMed] [Google Scholar]
- [41].Plimpton S, Fast Parallel Algorithms for Short-Range Molecular Dynamics, Journal of computational physics, 117 (1995) 1–19. [Google Scholar]
- [42].Jeffery GB, The Motion of Ellipsoidal Particles Immersed in a Viscous Fluid, Proceedings of the Royal Society of London. Series A, Containing Papers of a Mathematical and Physical Character, 102 (1922) 161–179. [Google Scholar]
- [43].Thompson AP, Plimpton SJ, Mattson W, General Formulation of Pressure and Stress Tensor for Arbitrary Many-Body Interaction Potentials under Periodic Boundary Conditions, The Journal of Chemical Physics, 131 (2009) 154107. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.