\AuthorHeaders

Chen et al.

MetaFollower: Adaptable Personalized Autonomous Car Following

Xianda Chen
Intelligent Transportation Thrust
   Systems Hub
The Hong Kong University of Science and Technology (Guangzhou)
   Guangzhou    511400    China
[email protected]
Kehua Chen
Division of Emerging Interdisciplinary Areas (EMIA)
   Interdisciplinary Programs Office
The Hong Kong University of Science and Technology
   Hong Kong    China
[email protected]
Meixin Zhu, Ph.D., Corresponding Author
Assistant Professor
Intelligent Transportation Thrust
   Systems Hub
The Hong Kong University of Science and Technology (Guangzhou)
   Guangzhou    511400    China
Guangdong Provincial Key Lab of Integrated Communication
   Sensing and Computation for Ubiquitous Internet of Things    Guangzhou    511400    China
[email protected]
Hao (Frank) Yang, Ph.D.
Research Assistant Professor
Department of Civil and System Engineering
   Whiting School of Engineering
Johns Hopkins University
   Baltimore    MD 21218
[email protected]
Shaojie Shen, Ph.D.
Associate Professor
Department of Electronic and Computer Engineering
The Hong Kong University of Science and Technology
   Hong Kong    China
[email protected]
Xuesong Wang, Ph.D.
Professor
School of Transportation Engineering
   Tongji University    Shanghai    201804    China
[email protected]
Yinhai Wang, Ph.D.
Professor
Department of Civil and Environmental Engineering
University of Washington
   Seattle    WA 98195    USA
[email protected]

1 Abstract

Car-following (CF) modeling, a fundamental component in microscopic traffic simulation, has attracted increasing interest of researchers in the past decades. In this study, we propose an adaptable personalized car-following framework —– MetaFollower, by leveraging the power of meta-learning. Specifically, we first utilize Model-Agnostic Meta-Learning (MAML) to extract common driving knowledge from various CF events. Afterward, the pre-trained model can be fine-tuned on new drivers with only a few CF trajectories to achieve personalized CF adaptation. We additionally combine Long Short-Term Memory (LSTM) and Intelligent Driver Model (IDM) to reflect temporal heterogeneity with high interpretability. Unlike conventional adaptive cruise control (ACC) systems that rely on predefined settings and constant parameters without considering heterogeneous driving characteristics, MetaFollower can accurately capture and simulate the intricate dynamics of car-following behavior while considering the unique driving styles of individual drivers. We demonstrate the versatility and adaptability of MetaFollower by showcasing its ability to adapt to new drivers with limited training data quickly. To evaluate the performance of MetaFollower, we conduct rigorous experiments comparing it with both data-driven and physics-based models. The results reveal that our proposed framework outperforms baseline models in predicting car-following behavior with higher accuracy and safety. To the best of our knowledge, this is the first car-following model aiming to achieve fast adaptation by considering both driver and temporal heterogeneity based on meta-learning.

Keywords: Car Following, Meta Learning, Driving Style Adaptation, Autonomous Driving, Adaptive Cruise Control.

2 Introduction

Car-following (CF) is a fundamental driving behavior in traffic flow, where each vehicle follows the one in front of it at a certain distance. Accurately modeling car-following behavior is crucial for microscopic traffic simulation and plays a key role in adaptive cruise control (ACC) systems. Over the years, researchers have shown significant interest in develo** car-following models [1, 2, 3, 4, 5] to simulate and understand the dynamics of car-following.

Car-following models can be classified into three categories based on their modeling approaches [6, 7, 8]: physics-based models, data-driven models, and hybrid models. Physics-based models use heuristic rules and mathematical equations to simulate car-following behavior. Data-driven models, on the other hand, leverage large amounts of real-world driving data and machine learning techniques to extract patterns. Hybrid models combine both physics-based and data-driven approaches to take advantage of their respective strengths, aiming to improve the accuracy and interpretability of car-following models.

However, accurately capturing the heterogeneity in car-following behavior caused by different driving styles and temporal variations remains a challenge [9, 10, 11, 12, 13]. Previous research has identified several sources of heterogeneity in car-following behavior, such as age [14], gender [15], vehicle type [16], road condition [17], as well as intra-driver heterogeneity [18]. These pose challenges for accurately modeling car-following and lead to limitations on ACC systems. Most existing ACC systems utilize either a fixed time headway or distance headway setting, which lacks adaptability to individual drivers or varying driving conditions. Furthermore, develo** personalized car-following models that can quickly learn from limited data and adapt to new drivers is a tricky problem and has hardly been investigated before.

To address the aforementioned limitations, we propose MetaFollower, an adaptable personalized car-following framework that combines data-driven and physics-based models under the Model-Agnostic Meta-Learning (MAML) framework (Fig. 1). In detail, we leverage Long Short-Term Memory (LSTM) to generate parameters for the Intelligent Driver Model (IDM), incorporating temporal heterogeneity and interpretability. We then train the hybrid model with MAML to generate good initialization parameters, the pre-trained CF model can fast adapt to new drivers with only a few CF events under the MAML framework. To evaluate the performance of MetaFollower, we extracted 3050 driver-specific car-following events from the Shanghai Naturalistic Driving Study-world dataset, involving a total of 44 drivers. The results show that the proposed model outperformed both data-driven and physics-based models in terms of both accuracy and safety.

The main contributions of this study are as follows:

  • We introduce MetaFollower, the first car-following model based on meta-learning, that considers both driver and temporal heterogeneity. Our model enables fast adaptation to new drivers with limited training data.

  • We visually segment the driving behavior of different drivers, showcasing the heterogeneity in driving behavior.

  • We validate the effectiveness and superiority of our approach by conducting experiments using real-world driving data from naturalistic drivers and comparing it with baseline models.

Refer to caption
Figure 1: A Roadmap of MetaFollower

3 Related Work

3.1 Car-Following Model

3.1.1 Physics-Based Car-Following Models

Physics-based car-following models utilize fundamental principles from traffic flow theory to describe the interaction between vehicles. These models incorporate parameters such as desired time headway, acceleration, and the effect of surrounding traffic conditions. Prominent examples in this category include Gazis-Herman-Rothery (GHR) model [19], Helly’s model [20], Wiedemann model [21], Gipps Model [22], Optimal Velocity Model (OVM) [23] and Intelligent Driver Model (IDM) [24]. To gain a comprehensive understanding of conventional car-following models, Chen et al. [4], Brackstone and McDonald [25], Saifuzzaman and Zheng [26] provide valuable insights into the development and incorporation of car-following models in transportation research. These models are based on simplified assumptions and equations that may not fully capture the complexity and variability of real-world driving behavior. As a result, their accuracy may be lower compared to data-driven models.

3.1.2 Data-Driven Car-Following Models

Data-driven car-following models leverage large amounts of real-world driving data to establish models and extract patterns and rules with machine learning techniques. These models can learn from historical data and use statistical techniques to capture complex relationships between different variables, allowing for more accurate predictions and simulations compared to physics-based models. There are various data-driven car-following models available for predicting and modeling vehicle behavior in the car-following scenario. For example, Neural Networks (NN) [27, 28, 29, 30, 31] are commonly used to capture complex relationships and make predictions based on observed data. LSTM [32, 33, 34, 35], a type of recurrent neural network, is particularly suitable for handling time-series data and capturing temporal dependencies in car-following behavior. Transformer-based models, originally designed for natural language processing, can encode and decode vehicle trajectory data while considering global and local information for predicting vehicle behavior [36]. In addition, Reinforcement Learning (RL) [37, 38, 39, 40, 41, 42] algorithms can train intelligent agents to learn tracking policies by providing states and rewards. However, data-driven approaches are black-box models and have low interpretability, which hinders their application.

3.1.3 Hybrid Car-Following Models

Hybrid car-following models, also known as physics-informed deep learning (PIDL) models, represent an innovative approach that combines the strengths of mathematical car-following models and data-driven methods. By incorporating physics-based constraints, these hybrid models enhance their interpretability. For example, Yang et al. [31] developed a combination car-following model that merges machine-learning-based and kinematics-based models by optimizing weight values. This model demonstrates superior performance in terms of safety and robustness compared to individual models. Mo et al. [7] proposed a PIDL approach for car-following models, combining physics-based models with deep-learning models. This integration enhances prediction accuracy and data efficiency by incorporating fundamental traffic flow theories. Building upon this approach, Mo and Di [43] addressed uncertainty quantification in car-following behavior by integrating stochastic physics into the PIDL structure.

3.2 Driving Heterogeneity in Car-Following Behavior

Car-following behavior varies significantly among different drivers, reflecting the inherent heterogeneity in driving styles. Several studies have examined driving style heterogeneity in car-following behavior through field observations, driving simulator experiments, and data analysis. Ossen and Hoogendoorn [10] investigated the heterogeneity in car-following behavior using a large dataset of trajectory observations. It identifies significant differences in behavior among passenger car drivers and between passenger car drivers and truck drivers. This study also reveals that truck drivers exhibit a more consistent and robust car-following behavior compared to passenger car drivers. Research by Ding et al. [44] assumed that drivers share a set of driver states, and each driver has a unique driver profile that characterizes driving style. The method considers both intra-driver and inter-driver heterogeneity. Wen et al. [45] used the Waymo Open Dataset to investigate car-following behavior between human-driven vehicles (MVs) following automated vehicles (AVs) and MVs following MVs. Results show that MV-following-AV events have lower driving volatility, smaller time headways, and higher time to collision (TTC) values, and human drivers exhibit four distinct car-following styles. Huang et al. [46] presented an enhanced Fog-related Intelligent Driver Model (FIDM) that considers unobserved driver heterogeneity in fog conditions to accurately reproduce car-following behavior. The results demonstrated that as fog density decreased, unobserved driver heterogeneity increased. Kim et al. [47] calibrated a car-following model with random coefficients, which can capture the heterogeneity across drivers who respond differently to stimuli. The expectation-maximization (EM) algorithm is employed to overcome challenges related to dimensionality and empirical identification. The calibration results confirm significant variations in random coefficients among drivers, with correlations between them.

3.3 Meta-Learning

Meta-learning, also known as "learning to learn," is a subfield of machine learning that focuses on develo** algorithms and techniques that enable models to learn new tasks or adapt to new environments quickly and effectively. Li and Malik [48] introduced a meta-learning approach for learning optimization algorithms. Through this framework, the authors successfully learn to optimize black-box functions using recurrent neural networks. This innovative application of meta-learning provides insights into the automatic design of optimization algorithms. Additionally, Santoro et al. [49] proposed a memory-augmented neural network architecture capable of learning new concepts from limited examples. The model employs external memory, resembling working memory in humans, to store information for generalization to unseen tasks. This work showcases the potential of meta-learning in enabling one-shot learning capabilities. Finn et al. [50] proposed a Model-Agnostic Meta-Learning (MAML) algorithm that enables rapid adaptation of deep neural networks to new tasks. Their method demonstrates impressive performance across various domains, showcasing the potential of meta-learning techniques for efficient learning. Furthermore, Lee and Choi [51] proposed a gradient-based meta-learning framework with learned layerwise metrics. The authors demonstrate improved performance on few-shot classification tasks by explicitly learning adaptive metrics within the network. Moreover, Meta-learning has applications in various domains, including computer vision [52], natural language processing [53], robotics [54], and reinforcement learning [55]. In regard to transportation, Ye et al. [56] proposed the use of Meta Reinforcement Learning (MRL) to improve the generalization capabilities of autonomous driving agents in new environments. Specifically, the method focuses on automated lane-changing maneuvers under different traffic congestion levels. Results show that the proposed MRL approach achieves higher success rates and lower collision rates compared to the benchmark model, particularly in heavy traffic conditions that were not encountered during training. ** et al. [57] introduced CrossTReS, a framework for traffic prediction using meta-learning in transportation. It addresses the scarcity of data by utilizing cross-city transfer learning. Experimental results demonstrate that CrossTReS outperforms state-of-the-art baselines by up to 8% in real-world scenarios.

4 Preliminary Knowledge

4.1 Intelligent Driver Model (IDM)

IDM is a widely recognized physics-based car-following model that considers both distance-based and time-based driving behavior. The model takes into account several influencing factors as input. These factors include the following vehicle’s speed, the gap as well as the relative speed between the following vehicle (FV) and the lead vehicle (LV). It reproduces realistic car-following behavior, including smooth acceleration, maintaining a safe distance, and adapting to changes in speed and traffic conditions, the model expressions are:

an(t)=a0[1(vn(t)v~n)λ(S~n(t)sn)2]subscript𝑎𝑛𝑡subscript𝑎0delimited-[]1superscriptsubscript𝑣𝑛𝑡subscript~𝑣𝑛𝜆superscriptsubscript~𝑆𝑛𝑡subscript𝑠𝑛2a_{n}(t)=a_{0}\left[1-\left(\frac{v_{n}(t)}{\widetilde{v}_{n}}\right)^{\lambda% }-\left(\frac{\widetilde{S}_{n}(t)}{s_{n}}\right)^{2}\right]italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ) = italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT [ 1 - ( divide start_ARG italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ) end_ARG start_ARG over~ start_ARG italic_v end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_λ end_POSTSUPERSCRIPT - ( divide start_ARG over~ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ) end_ARG start_ARG italic_s start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] (1)
S~n(t)=S0+vn(t)T~+vn(t)Δv(t)2a0bsubscript~𝑆𝑛𝑡subscript𝑆0subscript𝑣𝑛𝑡~𝑇subscript𝑣𝑛𝑡Δ𝑣𝑡2subscript𝑎0𝑏\widetilde{S}_{n}(t)=S_{0}+v_{n}(t)\widetilde{T}+\frac{v_{n}(t)\Delta v(t)}{2% \sqrt{a_{0}b}}over~ start_ARG italic_S end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ) = italic_S start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ) over~ start_ARG italic_T end_ARG + divide start_ARG italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ) roman_Δ italic_v ( italic_t ) end_ARG start_ARG 2 square-root start_ARG italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_b end_ARG end_ARG (2)
Δv(t)=vn(t)vn1(t)Δ𝑣𝑡subscript𝑣𝑛𝑡subscript𝑣𝑛1𝑡\Delta v(t)=v_{n}(t)-v_{n-1}(t)roman_Δ italic_v ( italic_t ) = italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ) - italic_v start_POSTSUBSCRIPT italic_n - 1 end_POSTSUBSCRIPT ( italic_t ) (3)

where an(t)subscript𝑎𝑛𝑡a_{n}(t)italic_a start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ) and vn(t)subscript𝑣𝑛𝑡v_{n}(t)italic_v start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ) represent the acceleration and the velocity of the FV at time t𝑡titalic_t, respectively. Sn(t)subscript𝑆𝑛𝑡S_{n}(t)italic_S start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_t ), Δv(t)Δ𝑣𝑡\Delta v(t)roman_Δ italic_v ( italic_t ) are the spacing and relative speed between the FV and the LV. The desired maximum acceleration, comfortable deceleration, desired velocity, and desired time headway are represented by a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, b𝑏bitalic_b, v~~𝑣\widetilde{v}over~ start_ARG italic_v end_ARG, and T~~𝑇\widetilde{T}over~ start_ARG italic_T end_ARG, respectively. S0subscript𝑆0S_{0}italic_S start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the minimum safe headway and λ𝜆\lambdaitalic_λ is a constant to be calibrated.

4.2 Long Short Term Memory (LSTM) Networks

LSTM [35] is a type of recurrent neural network (RNN) that addresses the vanishing gradient problem encountered in traditional RNNs. It is particularly useful for modeling car-following behavior due to its ability to capture temporal dependencies. Compared to a traditional RNN, LSTM has two main pathways: the cell state (C𝐶Citalic_C) and the hidden state (H𝐻Hitalic_H). The cell state serves as the long-term memory, allowing important information to be preserved over time and preventing the vanishing gradient problem. The hidden state functions similarly to the hidden state in a regular RNN, capturing the short-term working memory. LSTM incorporates gates, such as the input gate, forget gate, and output gate, to control the flow of information. The input gate determines how much new information should be added to the cell state, while the forget gate controls what information should be discarded. The output gate controls which parts of the cell state should be outputted after analysis. The overall structure and flow of information in an LSTM model are illustrated in Fig. 2. This architecture enables LSTM to effectively learn and predict car-following behavior by considering both short-term and long-term dependencies.

Refer to caption
Figure 2: LSTM Network Architecture [35]

4.3 Model-Agnostic Meta-Learning (MAML)

MAML presents a novel and efficient meta-learning algorithm designed for the rapid adaptation of deep neural networks to new tasks using limited training data. The core idea of MAML is to learn an initial parameter θ𝜃\thetaitalic_θ that can be quickly adapted to new tasks using only a few gradient updates for each task. This adaptation process generates new parameters θisuperscriptsubscript𝜃𝑖\theta_{i}^{\prime}italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT specific to each task. Subsequently, the global initial parameter θ𝜃\thetaitalic_θ is further updated from all tasks. Fig. 3 below illustrates this process, where the blue branching lines represent the update directions for different tasks, and the orange main axis line represents the overall direction of model parameters. This approach can be seen as a balance between different tasks, preventing parameters from overfitting to any individual task. The dashed line indicates the adaptation process for a new task, which involves fine-tuning model parameters. The adaptation process can be achieved with a few gradient updates, allowing the model to quickly adapt to the new task. The pseudocode for the MAML is shown in algorithm 1.

Refer to caption
Figure 3: Illustration of the MAML Framework
Algorithm 1 Model-Agnostic Meta-Learning (MAML)
0:  p(T)𝑝𝑇p({T})italic_p ( italic_T ): distribution over tasks
0:  α,β𝛼𝛽\alpha,\betaitalic_α , italic_β: learning rates for inner and outer loop
0:  K𝐾Kitalic_K: number of inner loop gradient steps
1:  Initialize 𝜽𝜽\boldsymbol{\theta}bold_italic_θ: model parameters
2:  while not done do
3:     Sample batch of tasks Tip(T)similar-tosubscript𝑇𝑖𝑝𝑇{T}_{i}\sim p({T})italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∼ italic_p ( italic_T )
4:     for all Tisubscript𝑇𝑖{T}_{i}italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT do
5:        Sample Dtrainisuperscriptsubscript𝐷train𝑖D_{\text{train}}^{i}italic_D start_POSTSUBSCRIPT train end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT and Dtestisuperscriptsubscript𝐷test𝑖D_{\text{test}}^{i}italic_D start_POSTSUBSCRIPT test end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT from Tisubscript𝑇𝑖{T}_{i}italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT
6:        Compute adapted parameters 𝜽i𝜽α𝜽LTi(𝜽;Dtraini)superscriptsubscript𝜽𝑖𝜽𝛼subscript𝜽subscript𝐿subscript𝑇𝑖𝜽superscriptsubscript𝐷train𝑖\boldsymbol{\theta}_{i}^{\prime}\leftarrow\boldsymbol{\theta}-\alpha\nabla_{% \boldsymbol{\theta}}{L}_{{T}_{i}}(\boldsymbol{\theta};D_{\text{train}}^{i})bold_italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ← bold_italic_θ - italic_α ∇ start_POSTSUBSCRIPT bold_italic_θ end_POSTSUBSCRIPT italic_L start_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( bold_italic_θ ; italic_D start_POSTSUBSCRIPT train end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT )
7:        for k=1𝑘1k=1italic_k = 1 to K𝐾Kitalic_K do
8:           Update adapted parameters 𝜽i𝜽iα𝜽iLTi(𝜽i;Dtraini)superscriptsubscript𝜽𝑖superscriptsubscript𝜽𝑖𝛼subscriptsuperscriptsubscript𝜽𝑖subscript𝐿subscript𝑇𝑖superscriptsubscript𝜽𝑖superscriptsubscript𝐷train𝑖\boldsymbol{\theta}_{i}^{\prime}\leftarrow\boldsymbol{\theta}_{i}^{\prime}-% \alpha\nabla_{\boldsymbol{\theta}_{i}^{\prime}}{L}_{{T}_{i}}(\boldsymbol{% \theta}_{i}^{\prime};D_{\text{train}}^{i})bold_italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ← bold_italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT - italic_α ∇ start_POSTSUBSCRIPT bold_italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT italic_L start_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( bold_italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ; italic_D start_POSTSUBSCRIPT train end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT )
9:        end for
10:        Compute meta-objective LmetaiLTi(𝜽i;Dtesti)superscriptsubscript𝐿meta𝑖subscript𝐿subscript𝑇𝑖superscriptsubscript𝜽𝑖superscriptsubscript𝐷test𝑖{L}_{\text{meta}}^{i}\leftarrow{L}_{{T}_{i}}(\boldsymbol{\theta}_{i}^{\prime};% D_{\text{test}}^{i})italic_L start_POSTSUBSCRIPT meta end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ← italic_L start_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( bold_italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ; italic_D start_POSTSUBSCRIPT test end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT )
11:     end for
12:     Update model parameters 𝜽𝜽β𝜽Tip(T)Lmetai𝜽𝜽𝛽subscript𝜽subscriptsimilar-tosubscript𝑇𝑖𝑝𝑇superscriptsubscript𝐿meta𝑖\boldsymbol{\theta}\leftarrow\boldsymbol{\theta}-\beta\nabla_{\boldsymbol{% \theta}}\sum_{{T}_{i}\sim p({T})}{L}_{\text{meta}}^{i}bold_italic_θ ← bold_italic_θ - italic_β ∇ start_POSTSUBSCRIPT bold_italic_θ end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∼ italic_p ( italic_T ) end_POSTSUBSCRIPT italic_L start_POSTSUBSCRIPT meta end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT
13:  end while

4.4 Problem Definition

Given H𝐻Hitalic_H drivers with numerous car-following trajectories and K𝐾Kitalic_K new drivers with a few trajectories, we aim to train a model \mathcal{M}caligraphic_M on H𝐻Hitalic_H drivers to extract driver styles and achieve fast adapting to the driving styles of the K𝐾Kitalic_K new drivers. Formally, we have a dataset DHsubscript𝐷𝐻{D}_{H}italic_D start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT consisting of H𝐻Hitalic_H drivers and we treat each driver as one task under the MAML framework, where each driver i𝑖iitalic_i is associated with a set of trajectories Ti={Ti1,Ti2,,TiN}subscript𝑇𝑖superscriptsubscript𝑇𝑖1superscriptsubscript𝑇𝑖2superscriptsubscript𝑇𝑖𝑁{T}_{i}=\{T_{i}^{1},T_{i}^{2},...,T_{i}^{N}\}italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = { italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT , italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , … , italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT }, with N𝑁Nitalic_N being the number of trajectories for driver i𝑖iitalic_i. Each trajectory Tinsuperscriptsubscript𝑇𝑖𝑛T_{i}^{n}italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT consists of a time sequence of states, represented as (smn,smn,,sMn)superscriptsubscript𝑠𝑚𝑛superscriptsubscript𝑠𝑚𝑛superscriptsubscript𝑠𝑀𝑛(s_{m}^{n},s_{m}^{n},...,s_{M}^{n})( italic_s start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT , italic_s start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT , … , italic_s start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ), where M𝑀Mitalic_M is the number of states in trajectory Tinsuperscriptsubscript𝑇𝑖𝑛T_{i}^{n}italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT. We also have a smaller dataset DKsubscript𝐷𝐾{D}_{K}italic_D start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT containing trajectories from K new drivers, where each driver j𝑗jitalic_j is associated with a set of trajectories Tj={Tj1,Tj2,,TjL}subscript𝑇𝑗superscriptsubscript𝑇𝑗1superscriptsubscript𝑇𝑗2superscriptsubscript𝑇𝑗𝐿{T}_{j}=\{T_{j}^{1},T_{j}^{2},...,T_{j}^{L}\}italic_T start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = { italic_T start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT , italic_T start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , … , italic_T start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_L end_POSTSUPERSCRIPT }, with L𝐿Litalic_L being the number of trajectories for driver j𝑗jitalic_j. Our objective is to train a model \mathcal{M}caligraphic_M that can effectively extract driving styles from the trajectories in DHsubscript𝐷𝐻{D}_{H}italic_D start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT and then adapt quickly to the driving styles in the trajectories of DKsubscript𝐷𝐾{D}_{K}italic_D start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT. To achieve this, the model \mathcal{M}caligraphic_M needs to learn a map** function f𝑓fitalic_f that takes a trajectory Tinsuperscriptsubscript𝑇𝑖𝑛T_{i}^{n}italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT as input and predicts the corresponding driving behavior. Mathematically, we can represent this as:

𝕪^in=f(T;θ)superscriptsubscript^𝕪𝑖𝑛𝑓𝑇𝜃\hat{\mathbb{y}}_{i}^{n}=f(T;\theta)over^ start_ARG blackboard_y end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT = italic_f ( italic_T ; italic_θ ) (4)

where 𝕪^insuperscriptsubscript^𝕪𝑖𝑛\hat{\mathbb{y}}_{i}^{n}over^ start_ARG blackboard_y end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT is the predicted driving states of n𝑛nitalic_n trajectory of driver i𝑖iitalic_i, T𝑇Titalic_T is the trajectory, and θ𝜃\thetaitalic_θ represents the parameters of the model \mathcal{M}caligraphic_M. During training, the model \mathcal{M}caligraphic_M learns the optimal values for the parameters θ𝜃\thetaitalic_θ by minimizing the discrepancy between the predicted driving behavior and the ground truth labels in the dataset DHsubscript𝐷𝐻{D}_{H}italic_D start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT. This can be formulated as an optimization problem:

minθi=1Hn=1NL(f(Tin;θ),𝕪in)subscript𝜃superscriptsubscript𝑖1𝐻superscriptsubscript𝑛1𝑁𝐿𝑓superscriptsubscript𝑇𝑖𝑛𝜃superscriptsubscript𝕪𝑖𝑛\min_{\theta}\sum_{i=1}^{H}\sum_{n=1}^{N}L(f(T_{i}^{n};\theta),\mathbb{y}_{i}^% {n})roman_min start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_H end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_L ( italic_f ( italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ; italic_θ ) , blackboard_y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) (5)

where L𝐿Litalic_L is the loss function that measures the discrepancy between the predicted behavior f(Tin;θ)𝑓superscriptsubscript𝑇𝑖𝑛𝜃f(T_{i}^{n};\theta)italic_f ( italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ; italic_θ ) and ground truth driving behavior 𝕪insuperscriptsubscript𝕪𝑖𝑛\mathbb{y}_{i}^{n}blackboard_y start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT. Once the model \mathcal{M}caligraphic_M is trained on the dataset DHsubscript𝐷𝐻{D}_{H}italic_D start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT, it can be fine-tuned or adapted using the trajectories in DKsubscript𝐷𝐾{D}_{K}italic_D start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT. This adaptation process involves updating the parameters θ𝜃\thetaitalic_θ based on the trajectories and driving behavior of the new drivers. The objective is to minimize the discrepancy between the predicted f(Tjk;θ)𝑓superscriptsubscript𝑇𝑗𝑘𝜃f(T_{j}^{k};\theta)italic_f ( italic_T start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ; italic_θ ) and true driving behavior 𝕪jksuperscriptsubscript𝕪𝑗𝑘\mathbb{y}_{j}^{k}blackboard_y start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT for the new drivers, similar to the training phase:

minθj=1Kk=1NL(f(Tjk;θ),𝕪jk).subscript𝜃superscriptsubscript𝑗1𝐾superscriptsubscript𝑘1𝑁𝐿𝑓superscriptsubscript𝑇𝑗𝑘𝜃superscriptsubscript𝕪𝑗𝑘\min_{\theta}\sum_{j=1}^{K}\sum_{k=1}^{N}L(f(T_{j}^{k};\theta),\mathbb{y}_{j}^% {k}).roman_min start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_L ( italic_f ( italic_T start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ; italic_θ ) , blackboard_y start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ) . (6)

By effectively training and adapting the model \mathcal{M}caligraphic_M on both the dataset DHsubscript𝐷𝐻{D}_{H}italic_D start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT and the trajectories from the new drivers in DKsubscript𝐷𝐾{D}_{K}italic_D start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT, we can achieve fast adaptation to the driving styles of the K𝐾Kitalic_K new drivers.

5 Proposed Method

In this work, we propose a novel model, MetaFollower, which combines the principle of meta-learning with a PIDL model. The PIDL model incorporates the IDM model into LSTM. The MetaFollower model is trained under the MAML framework.

5.1 Driving Style Analysis

To facilitate the semantic interpretation of the driving modes, we divided three variables into different levels based on the following driver’s car-following data statistics [11]: spacing, relative speed, and acceleration. We fitted them using gamma distributions according to [11, 58] and determined the threshold for each variable from a statistical perspective, as shown in Table 1. Based on the predefined thresholds for each variable, we obtained a library of original car-following modes with a size of 75 (5 × 5 × 3 = 75). Fig. 4 illustrates the diverse driving behavior exhibited by three drivers under different car-following distances. The white color indicates a higher probability that the driver adheres to the corresponding mode, while the dark black color indicates a lower probability (close to zero). For instance, when following an LV at a long distance, driver 8 (Fig. 4(f)) and driver 10 (Fig. 4(i)) tend to approach the LV quickly, whereas driver 2 (Fig. 4(c)) prefers a slower acceleration to approach the LV. To better visually differentiate the preferred driving modes of each driver, we selected the mode with the highest probability in each distance pattern, as shown in Fig. 5. The color orange represents the "close gap" mode, purple represents the "normal gap" mode, and black represents the "long gap" mode.

Table 1: Feature Distribution and Threshold Selection
Physical quantity Threshold Semantic state
Acceleration [m/s2𝑚superscript𝑠2m/s^{2}italic_m / italic_s start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT] (-\infty, -0.39) Aggressive positive deceleration
[-0.39, -0.08) Gentle negative deceleration
[-0.08, 0.16) Kee** acceleration
[0.16, 0.46) Gentle positive acceleration
(0.46, +\infty) Aggressive positive acceleration
Relative speed [m/s𝑚𝑠m/sitalic_m / italic_s] (-\infty, -0.82) Aggressive negative relative speed
[-0.82, -0.21) Gentle negative relative speed
[-0.21, 0.28) Kee** relative speed
[0.28, 0.89) Gentle positive relative speed
(0.89, +\infty) Aggressive positive relative speed
Spacing [m𝑚mitalic_m] (-\infty,10.11) Close gap
[10.11,24.70) Normal gap
(24.70,+\infty) Long gap
Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Refer to caption
(d)
Refer to caption
(e)
Refer to caption
(f)
Refer to caption
(g)
Refer to caption
(h)
Refer to caption
(i)
Figure 4: Distributions of Driving Patterns for Driver 2, Driver 8 and Driver 10
Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Figure 5: Representative Driving Patterns for Driver 2, Driver 8 and Driver 10

5.2 Improved PIDL Model with LSTM

In order to overcome the challenges associated with accurate and explainable predictions, our study presents an improved PIDL model that integrates the IDM into LSTM networks. According to Zhu et al. [5], IDM demonstrates superior performance compared to other traditional car-following models, making it the ideal choice as the physical component of our model. Additionally, Mo et al. [7] found that LSTM-PIDL outperforms NN-based PIDL models, which motivates us to select LSTM as the neural network component of our model. Unlike previous data-driven car-following models, which directly output the acceleration of the FV based on network parameters, the proposed PIDL model in this paper leverages LSTM networks to output the IDM model parameters. These parameters are then used with the IDM model to calculate the acceleration of the following vehicle. This approach essentially constructs a dynamic time-varying IDM model, providing greater flexibility and adaptability. This task can be formulated as an optimization problem:

minθ,λsubscript𝜃𝜆\displaystyle\min_{\theta,\lambda}roman_min start_POSTSUBSCRIPT italic_θ , italic_λ end_POSTSUBSCRIPT i=1N(a(i)a^(i))2superscriptsubscript𝑖1𝑁superscriptsuperscript𝑎𝑖superscript^𝑎𝑖2\displaystyle\sum_{i=1}^{N}\left(a^{(i)}-\widehat{a}^{(i)}\right)^{2}∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ( italic_a start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT - over^ start_ARG italic_a end_ARG start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (7)
s.t.formulae-sequencest\displaystyle\mathrm{s.~{}t.}roman_s . roman_t . a(i)=fθ(𝐬(i)|θ),i=1,,N,formulae-sequencesuperscript𝑎𝑖subscript𝑓𝜃conditionalsuperscript𝐬𝑖𝜃𝑖1𝑁\displaystyle a^{(i)}=f_{\theta}\left({\mathbf{s}}^{(i)}|\theta\right),\quad i% =1,\ldots,N,italic_a start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT = italic_f start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT ( bold_s start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT | italic_θ ) , italic_i = 1 , … , italic_N ,
ψΨ,𝜓Ψ\displaystyle\psi\subseteq\Psi,italic_ψ ⊆ roman_Ψ ,

where ψ𝜓\psiitalic_ψ is the feasible domain of the parameters of the physics ΨΨ\Psiroman_Ψ, representing the physical range of the IDM parameter, 𝐬(i)superscript𝐬𝑖\mathbf{s}^{(i)}bold_s start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT is the i𝑖iitalic_i th observed state, fθsubscript𝑓𝜃f_{\theta}italic_f start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT represents the PIDL model, and a^^𝑎\widehat{a}over^ start_ARG italic_a end_ARG represents the predicted acceleration. The model architecture is shown as Fig. 6.

Refer to caption
Figure 6: LSTM-PIDL Model Architecture

5.3 MAML Updating

In MAML, the support set and query set are two essential components used in the meta-learning process. The support set refers to a small labeled dataset that is provided to the model during the meta-training phase. It contains examples from various tasks or domains that the model needs to learn from. For each task, the support set includes both input data and corresponding target labels. The query set, on the other hand, is another dataset used for evaluating the model’s performance after it has been trained on the support set. The query set usually consists of unseen examples from the same tasks or domains encountered in the support set. The MAML algorithm is employed to update the base learner, which is the improved PIDL model with IDM. This process involves two key stages: the inner-updating and the outer-updating.

  • Inner-updating: In the inner loop, model parameters are updated for each task using a small dataset (support set) from that task. The inner-updating for the PIDL model is performed using gradient descent and the task-specific loss function. The inner loop adaptation is performed as below:

    𝜽i𝜽α𝜽LTi(𝜽;Dtraini),superscriptsubscript𝜽𝑖𝜽𝛼subscript𝜽subscript𝐿subscript𝑇𝑖𝜽superscriptsubscript𝐷train𝑖\boldsymbol{\theta}_{i}^{\prime}\leftarrow\boldsymbol{\theta}-\alpha\nabla_{% \boldsymbol{\theta}}{L}_{{T}_{i}}(\boldsymbol{\theta};D_{\text{train}}^{i}),bold_italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ← bold_italic_θ - italic_α ∇ start_POSTSUBSCRIPT bold_italic_θ end_POSTSUBSCRIPT italic_L start_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( bold_italic_θ ; italic_D start_POSTSUBSCRIPT train end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ) , (8)

    where 𝜽𝜽\boldsymbol{\theta}bold_italic_θ represents initial model parameters, α𝛼\alphaitalic_α is the inner loop learning rate, and LTisubscript𝐿subscript𝑇𝑖{L}_{{T}_{i}}italic_L start_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT is the task-specific loss function for task Tisubscript𝑇𝑖{T}_{i}italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT.

  • Outer-updating: In the outer loop, initial model parameters are updated based on the performance of the adapted model on the query set for each task. The meta-objective is computed as the average loss across all tasks. The outer loop adaptation is performed as below:

    𝜽𝜽β𝜽Lmetai,𝜽𝜽𝛽subscript𝜽superscriptsubscript𝐿meta𝑖\boldsymbol{\theta}\leftarrow\boldsymbol{\theta}-\beta\nabla_{\boldsymbol{% \theta}}\sum{L}_{\text{meta}}^{i},bold_italic_θ ← bold_italic_θ - italic_β ∇ start_POSTSUBSCRIPT bold_italic_θ end_POSTSUBSCRIPT ∑ italic_L start_POSTSUBSCRIPT meta end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT , (9)

    where β𝛽\betaitalic_β is the outer loop learning rate, and Lmetaisuperscriptsubscript𝐿meta𝑖{L}_{\text{meta}}^{i}italic_L start_POSTSUBSCRIPT meta end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT is the meta loss for task Tisubscript𝑇𝑖{T}_{i}italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT computed from the adapted model parameters 𝜽isuperscriptsubscript𝜽𝑖\boldsymbol{\theta}_{i}^{\prime}bold_italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT.

5.4 Fine-tuning with MAML

By leveraging the MAML algorithm, we obtain a pre-trained MetaFollower model that captures driving styles from the training set of drivers. This pre-trained model can then be fine-tuned based on a limited amount of data from new drivers to obtain the final car-following model, named MetaFollower. The fine-tuning process is outlined in the following steps:

  1. 1.

    Initialize model parameters with the learned initialization from the MAML training: 𝜽𝜽superscript𝜽𝜽\boldsymbol{\theta}^{*}\leftarrow\boldsymbol{\theta}bold_italic_θ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ← bold_italic_θ.

  2. 2.

    For each new driver kK𝑘𝐾k\in{K}italic_k ∈ italic_K, sample a small dataset (support set) for training.

  3. 3.

    Perform the inner loop adaptation for each new driver using the support set and the learned initialization 𝜽superscript𝜽\boldsymbol{\theta}^{*}bold_italic_θ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT:

    𝜽k𝜽α𝜽LTk(𝜽;Dtraink).superscriptsubscript𝜽𝑘superscript𝜽𝛼subscriptsuperscript𝜽subscript𝐿subscript𝑇𝑘superscript𝜽superscriptsubscript𝐷train𝑘\boldsymbol{\theta}_{k}^{\prime}\leftarrow\boldsymbol{\theta}^{*}-\alpha\nabla% _{\boldsymbol{\theta}^{*}}{L}_{{T}_{k}}(\boldsymbol{\theta}^{*};D_{\text{train% }}^{k}).bold_italic_θ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ← bold_italic_θ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT - italic_α ∇ start_POSTSUBSCRIPT bold_italic_θ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT italic_L start_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( bold_italic_θ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ; italic_D start_POSTSUBSCRIPT train end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ) . (10)
  4. 4.

    Evaluate the adapted model’s performance on the query set for each new driver and assess the model’s ability to generalize to new driving styles with limited data.

The fine-tuning process enables the MetaFollower model to adapt quickly to the driving styles of the K𝐾Kitalic_K new drivers, effectively capturing their car-following behavior with minimal additional training data.

6 Data and Experiments

6.1 Shanghai Naturalistic Driving Study (SH-NDS)

This study utilized real-world car-following events from the Shanghai Naturalistic Driving Study (SH-NDS) to train and evaluate the proposed model. The SH-NDS, a collaborative effort by Tongji University, General Motors, and the Virginia Tech Transportation Institute, aimed to gain a deeper understanding of Chinese drivers’ vehicle usage, operation, and safety awareness. The data collection spanned from December 2012 to December 2015, during which five passenger vehicles equipped with the second Strategic Highway Research Program (SHRP 2) Driving Automation System (DAS) [59] were employed. The DAS system encompassed a multitude of components, including a forward radar for measuring distance and relative speed to vehicles ahead, an accelerometer for tracking longitudinal and lateral acceleration, a GPS sensor for precise location information, an interface box for collecting vehicle CAN Bus data, and four synchronized video cameras that captured crucial aspects such as the driver’s face, the view of the road ahead, the rear roadway, and the driver’s hand movements, as illustrated in Fig.7.

Refer to caption
Figure 7: SH-NDS Four Camera Views [37]

6.2 Car-Following Event Extraction

Specific criteria were applied to extract car-following data from the SH-NDS dataset. These criteria were based on previous studies conducted by Zhu et al. [37], Wang et al. [60], Zhao et al. [61]. The following criteria were used:

  • The identification number of the LV remained constant, indicating that the FV was consistently following the same vehicle.

  • The lateral distance between the LV and the LV was less than 2.5 meters, ensuring that they were driving in the same lane.

  • The duration of the car-following event was longer than 15 seconds, providing sufficient data for analysis.

  • Each individual driver’s dataset should have a minimum of 20 car-following events. This requirement ensured an adequate number of data samples for each driver, allowing for reliable statistical analysis and better capturing of individual driving behavior characteristics during the modeling.

Based on these criteria, 44 Shanghai drivers were selected for car-following model research. In our study, we regard the car following data of each individual driver as separate MAML tasks. We divide the data of 44 drivers into a training set and a test set. Specifically, the data of 33 drivers were used as the training data, while the remaining 11 drivers’ data were used as the test data. In the training dataset of the 33 drivers, each driver’s data was further divided into a support set (33.3% of the data) and a query set (66.7% of the data). For the 11 test drivers, each driver’s data was also divided into a support set (33.3% of the data) and a query set (66.7% of the data). In other words, the ratio of the support set to the query set is 1:3.

6.3 Baselines

Six types of baseline models are used for comparison to demonstrate the performance of our proposed model. All models use consistent input parameters, including spacing, FV’s speed, and relative speed. The primary objective of these models is to predict the acceleration of the FV based on these inputs.

  • IDM: The IDM model has been shown to outperform other traditional models according to [5]. We trained an IDM model using the genetic algorithm (GA) to minimize the Mean Squared Error (MSE) of spacing. By employing GA, we determined the optimal set of IDM parameters.

  • GHR: The GHR model is based on the assumption that the acceleration of the LV depends on the comparison between its own speed and that of the preceding vehicle. Similar to the IDM model, we employed GA to identify the optimal parameters for the GHR model [5].

  • LSTM without PIDL without meta without pre-train: This baseline model is trained directly on the training data of the test drivers without any pre-training or meta-learning. It uses the LSTM architecture to model the sequence data and predicts the acceleration of the FV based on the input parameters.

  • LSTM without PIDL without meta: In this model, pre-training is performed on the training drivers’ data using the LSTM architecture. The model is then fine-tuned on the support set of the test drivers to adapt it to the specific characteristics of the test drivers’ data. The model does not utilize meta-learning techniques and relies solely on the LSTM architecture for prediction.

  • LSTM without PIDL with meta: This model combines the LSTM architecture with meta-learning techniques. In this case, the model is trained on the training drivers’ data using meta-learning, which helps it adapt and generalize better to the new drivers’ data in the support set.

  • LSTM with PIDL without meta: This model combines the designed PIDL with the LSTM architecture. This allows the model to leverage both the temporal patterns in the data and the prior knowledge of the underlying physics. Unlike the previous model, it does not utilize meta-learning techniques.

6.4 Test Details

According to the general meta-learning data splitting method, as mentioned before, we divide the drivers into training drivers (numbered 1 to 33) and testing drivers (numbered 34 to 44). Each driver has its own support set and query set, and all models are evaluated on the query set of the testing drivers.

7 Results

7.1 Evaluation Metrics

To evaluate the performance of a car-following model, we choose two metrics [4] as the standard evaluation criteria: MSE of spacing and collision rate. The MSE of spacing measures the accuracy of the model in predicting the spacing between vehicles. A lower MSE score indicates a better fit of the model to the data and improved precision in predicting the spacing. For one car-following event, the MSE of spacing can be expressed as:

MSE=1Ni=1N(Sn1,n(t)Sn1,nobs(t))2MSE1𝑁superscriptsubscript𝑖1𝑁superscriptsubscript𝑆𝑛1𝑛𝑡superscriptsubscript𝑆𝑛1𝑛𝑜𝑏𝑠𝑡2\operatorname{MSE}={\frac{1}{N}\sum_{i=1}^{N}\left({{S_{n-1,n}}(t)-S_{n-1,n}^{% obs}(t)}\right)^{2}}roman_MSE = divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ( italic_S start_POSTSUBSCRIPT italic_n - 1 , italic_n end_POSTSUBSCRIPT ( italic_t ) - italic_S start_POSTSUBSCRIPT italic_n - 1 , italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_o italic_b italic_s end_POSTSUPERSCRIPT ( italic_t ) ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (11)

where N𝑁Nitalic_N is the total number of observations, and i𝑖iitalic_i is an observation index. S𝑆Sitalic_S and Sobssuperscript𝑆𝑜𝑏𝑠S^{obs}italic_S start_POSTSUPERSCRIPT italic_o italic_b italic_s end_POSTSUPERSCRIPT are the predicted and observed spacing between the FV and LV, respectively. Similarly, we define collision rate as the number of car-following events where the spacing between vehicles is less than zero divided by the total number of car-following events in the test dataset. The formula for collision rate is as follows:

CollisionRate=Number of events with spacing<0Total number of car-following events𝐶𝑜𝑙𝑙𝑖𝑠𝑖𝑜𝑛𝑅𝑎𝑡𝑒Number of events with spacing0Total number of car-following eventsCollision\ Rate=\frac{\text{Number of events with spacing}<0}{\text{Total % number of car-following events}}italic_C italic_o italic_l italic_l italic_i italic_s italic_i italic_o italic_n italic_R italic_a italic_t italic_e = divide start_ARG Number of events with spacing < 0 end_ARG start_ARG Total number of car-following events end_ARG (12)

7.2 MSE and Collision Rate in Testing Dataset

According to the analysis of Table 2, we can observe the performance of different models in terms of MSE of spacing and collision rate. Below is the analysis for each model:

  • GHR Model: It shows poor performance in spacing prediction but has a good collision avoidance rate.

  • IDM Model: It performs slightly better than the GHR model in spacing prediction accuracy.

  • LSTM without PIDL without meta without pre-train Model: It performs slightly better than GHR and IDM models in spacing prediction but struggles with collision rate.

  • LSTM without PIDL without meta Model: It shows improvements in both spacing prediction and collision rate compared to the previous model.

  • LSTM without PIDL with meta Model: It further improves spacing prediction and collision rate by incorporating meta-learning methods.

  • LSTM with PIDL without meta Model: It achieves significant enhancements in reducing collision rate by utilizing PIDL methods.

  • LSTM with PIDL with meta (MetaFollower) Model: It outperforms all other models, showing the best performance in terms of spacing prediction and collision rate.

Table 2: Model Performance: MSE of spacing and Collison rate
Model MSE of spacing Collison rate ‰
GHR 33.93 0
IDM 29.26 0
LSTM without PIDL without meta without pre-train 26.67 55.82 (34)
LSTM without PIDL without meta 16.86 24.63 (15)
LSTM without PIDL with meta 14.66 18.06 (11)
LSTM with PIDL without meta 14.06 6.57 (4)
LSTM with PIDL with meta (MetaFollower) 10.95 0

8 Summary and Conclusion

In this study, we propose a novel MetaFollower model for car-following behavior prediction, which, to the best of the authors’ knowledge, is the first car-following model that leverages the knowledge of meta-learning considering both driver and temporal heterogeneity. The evaluation was conducted on the naturalistic driving dataset using two key metrics: MSE of spacing and collision rate. We compared its performance with six baseline models. Results demonstrated that the MetaFollower model outperformed baseline models in terms of both accuracy and safety. It achieved a lower MSE of spacing and a significantly reduced collision rate, indicating improved precision in predicting spacing and an enhanced ability to avoid collisions. The superior performance of the MetaFollower model can be attributed to its unique architecture and the integration of meta-learning techniques. By leveraging insights gathered from multiple drivers under the MAML framework, the MetaFollower model synthesizes a comprehensive understanding of car-following behavior and captures nuances of various driving styles. In conclusion, our study introduces the MetaFollower model as an effective solution for car-following behavior prediction which also highlights its potential for enhancing driver assistance systems and autonomous vehicle technologies.

9 Future Work

Our research on the MetaFollower model for car-following behavior prediction opens up several avenues for future exploration and enhancement. Here are some potential directions for further investigation:

  • Data Augmentation: Augmenting the training dataset with a wider range of driving scenarios and conditions could enhance the model’s ability to generalize to diverse real-world situations. This could involve incorporating data from different geographic locations, weather conditions, and traffic patterns.

  • Domain Extension: While our study focuses on car-following behavior prediction, exploring the adaptability of the MetaFollower model to other driving tasks could be valuable. This could involve extending its capabilities to tasks such as merging, and trajectory forecasting.

  • Real-time Implementation: Investigating the feasibility of implementing the MetaFollower model in real-time systems, such as advanced driver assistance systems (ADAS) or autonomous vehicles, is crucial. This would involve optimizing the model’s architecture and computational efficiency to meet the real-time constraints of such applications.

  • Collaborative Learning: Investigating the potential for collaborative learning among multiple autonomous vehicles equipped with MetaFollower models could lead to enhanced cooperative behavior and further improve safety and efficiency in traffic.

10 Acknowledgements

This study was sponsored by Research on data-driven microscopic traffic simulation model for autonomous driving algorithm evaluation, Guangzhou. Basic and Applied Basic Research Project (SL2022A03J00083) and Guangzhou Municipal Science and Technology Project (2023A03J0011).

11 Author contributions statement

The authors confirm contribution to the paper as follows: study conception and design: Xianda Chen, Kehua Chen; data collection: Xuesong Wang, Meixin Zhu; analysis and interpretation of results: Xianda Chen, Kehua Chen; draft manuscript preparation: Xianda Chen, Kehua Chen. All authors reviewed the results and approved the final version of the manuscript.

References

  • Wang et al. [2016] Wang, X., M. Zhu, M. Chen, and P. Tremont, Drivers’ rear end collision avoidance behaviors under different levels of situational urgency. Transportation research part C: emerging technologies, Vol. 71, 2016, pp. 419–433.
  • Groelke et al. [2021] Groelke, B., C. Earnhardt, J. Borek, and C. Vermillion, A predictive command governor-based adaptive cruise controller with collision avoidance for non-connected vehicle following. IEEE Transactions on Intelligent Transportation Systems, Vol. 23, No. 8, 2021, pp. 12276–12286.
  • Wang et al. [2022] Wang, X., X. Zhang, F. Guo, Y. Gu, and X. Zhu, Effect of daily car-following behaviors on urban roadway rear-end crashes and near-crashes: A naturalistic driving study. Accident Analysis & Prevention, Vol. 164, 2022, p. 106502.
  • Chen et al. [2023] Chen, X., M. Zhu, K. Chen, P. Wang, H. Lu, H. Zhong, X. Han, and Y. Wang, FollowNet: A Comprehensive Benchmark for Car-Following Behavior Modeling, 2023.
  • Zhu et al. [2018a] Zhu, M., X. Wang, A. Tarko, et al., Modeling car-following behavior on urban expressways in Shanghai: A naturalistic driving study. Transportation research part C: emerging technologies, Vol. 93, 2018a, pp. 425–445.
  • Zhang et al. [2022] Zhang, Y., X. Chen, J. Wang, Z. Zheng, and K. Wu, A generative car-following model conditioned on driving styles. Transportation research part C: emerging technologies, Vol. 145, 2022, p. 103926.
  • Mo et al. [2021] Mo, Z., R. Shi, and X. Di, A physics-informed deep learning paradigm for car-following models. Transportation research part C: emerging technologies, Vol. 130, 2021, p. 103240.
  • Chen et al. [2024] Chen, X., X. Yuan, M. Zhu, X. Zheng, S. Shen, X. Wang, Y. Wang, and F.-Y. Wang, AggFollower: Aggressiveness Informed Car-Following Modeling. IEEE Transactions on Intelligent Vehicles, 2024.
  • Wang et al. [2018a] Wang, X., R. Jiang, L. Li, Y. Lin, X. Zheng, and F.-Y. Wang, Capturing car-following behaviors by deep learning. IEEE Transactions on Intelligent Transportation Systems, Vol. 19, No. 3, 2018a, pp. 910–920.
  • Ossen and Hoogendoorn [2011] Ossen, S. and S. P. Hoogendoorn, Heterogeneity in car-following behavior: Theory and empirics. Transportation research part C: emerging technologies, Vol. 19, No. 2, 2011, pp. 182–195.
  • Wang et al. [2018b] Wang, W., J. Xi, and D. Zhao, Driving style analysis using primitive driving patterns with Bayesian nonparametric approaches. IEEE Transactions on Intelligent Transportation Systems, Vol. 20, No. 8, 2018b, pp. 2986–2998.
  • Xu et al. [2020] Xu, D., Z. Ding, C. Tu, H. Zhao, M. Moze, F. Aioun, and F. Guillemard, Driver Identification through Stochastic Multi-State Car-Following Modeling. arXiv preprint arXiv:2005.11077, 2020.
  • Robbins and Chapman [2019] Robbins, C. and P. Chapman, How does drivers’ visual search change as a function of experience? A systematic review and meta-analysis. Accident Analysis & Prevention, Vol. 132, 2019, p. 105266.
  • Doroudgar et al. [2017] Doroudgar, S., H. M. Chuang, P. J. Perry, K. Thomas, K. Bohnert, and J. Canedo, Driving performance comparing older versus younger drivers. Traffic injury prevention, Vol. 18, No. 1, 2017, pp. 41–46.
  • Kim et al. [2013a] Kim, J.-K., G. F. Ulfarsson, S. Kim, and V. N. Shankar, Driver-injury severity in single-vehicle crashes in California: a mixed logit analysis of heterogeneity due to age and gender. Accident Analysis & Prevention, Vol. 50, 2013a, pp. 1073–1081.
  • Ravishankar and Mathew [2011] Ravishankar, K. and T. V. Mathew, Vehicle-type dependent car-following model for heterogeneous traffic conditions. Journal of transportation engineering, Vol. 137, No. 11, 2011, pp. 775–781.
  • Delitala and Tosin [2007] Delitala, M. and A. Tosin, Mathematical modeling of vehicular traffic: a discrete kinetic theory approach. Mathematical Models and Methods in Applied Sciences, Vol. 17, No. 06, 2007, pp. 901–932.
  • Chen et al. [2020] Chen, X., J. Sun, Z. Ma, J. Sun, and Z. Zheng, Investigating the long-and short-term driving characteristics and incorporating them into car-following models. Transportation research part C: emerging technologies, Vol. 117, 2020, p. 102698.
  • Chandler et al. [1958] Chandler, R. E., R. Herman, and E. W. Montroll, Traffic dynamics: studies in car following. Operations research, Vol. 6, No. 2, 1958, pp. 165–184.
  • Helly [1959] Helly, W., Simulation of bottlenecks in single-lane traffic flow. In: Proceedings of the Symposium on Theory of Traffic Flow, Research Laboratories, General Motors, New York., 1959.
  • Wiedemann [1974] Wiedemann, R., Simulation des StraBenverkehrsflusses In Schriftenreihe des tnstituts fir Verkehrswesen der Universitiit Karlsruhe, 1974.
  • Gipps [1981] Gipps, P. G., A behavioural car-following model for computer simulation. Transportation Research Part B: Methodological, Vol. 15, No. 2, 1981, pp. 105–111.
  • Bando et al. [1995] Bando, M., K. Hasebe, A. Nakayama, A. Shibata, and Y. Sugiyama, Dynamical model of traffic congestion and numerical simulation. Physical review E, Vol. 51, No. 2, 1995, p. 1035.
  • Treiber et al. [2000] Treiber, M., A. Hennecke, and D. Helbing, Congested traffic states in empirical observations and microscopic simulations. Physical review E, Vol. 62, No. 2, 2000, p. 1805.
  • Brackstone and McDonald [1999] Brackstone, M. and M. McDonald, Car-following: a historical review. Transportation Research Part F: Traffic Psychology and Behaviour, Vol. 2, No. 4, 1999, pp. 181–196.
  • Saifuzzaman and Zheng [2014] Saifuzzaman, M. and Z. Zheng, Incorporating human-factors in car-following models: a review of recent developments and research needs. Transportation research part C: emerging technologies, Vol. 48, 2014, pp. 379–403.
  • Hongfei et al. [2003] Hongfei, J., J. Zhicai, and N. Anning, Develop a car-following model using data collected by" five-wheel system". In Proceedings of the 2003 IEEE International Conference on Intelligent Transportation Systems, IEEE, 2003, Vol. 1, pp. 346–351.
  • Panwai and Dia [2007] Panwai, S. and H. Dia, Neural agent car-following models. IEEE Trans. Intell. Transp. Syst., Vol. 8, No. 1, 2007, pp. 60–70.
  • Chong et al. [2013] Chong, L., M. M. Abbas, A. M. Flintsch, and B. Higgs, A rule-based neural network approach to model driver naturalistic behavior in traffic. Transportation Research Part C: Emerging Technologies, Vol. 32, 2013, pp. 207–223.
  • Khodayari et al. [2012] Khodayari, A., A. Ghaffari, R. Kazemi, and R. Braunstingl, A modified car-following model based on a neural network model of the human driver effects. IEEE Transactions on Systems, Man, and Cybernetics-Part A: Systems and Humans, Vol. 42, No. 6, 2012, pp. 1440–1449.
  • Yang et al. [2018] Yang, D., L. Zhu, Y. Liu, D. Wu, and B. Ran, A novel car-following control model combining machine learning and kinematics models for automated vehicles. IEEE Transactions on Intelligent Transportation Systems, Vol. 20, No. 6, 2018, pp. 1991–2000.
  • Zhou et al. [2017] Zhou, M., X. Qu, and X. Li, A recurrent neural network based microscopic car following model to predict traffic oscillation. Transportation research part C: emerging technologies, Vol. 84, 2017, pp. 245–264.
  • Cho et al. [2014] Cho, K., B. Van Merriënboer, C. Gulcehre, D. Bahdanau, F. Bougares, H. Schwenk, and Y. Bengio, Learning phrase representations using RNN encoder-decoder for statistical machine translation. arXiv preprint arXiv:1406.1078, 2014.
  • Ma and Qu [2020] Ma, L. and S. Qu, A sequence to sequence learning based car-following model for multi-step predictions considering reaction delay. Transportation research part C: emerging technologies, Vol. 120, 2020, p. 102785.
  • Hochreiter and Schmidhuber [1997] Hochreiter, S. and J. Schmidhuber, Long short-term memory. Neural computation, Vol. 9, No. 8, 1997, pp. 1735–1780.
  • Zhu et al. [2022] Zhu, M., S. S. Du, X. Wang, Hao, Yang, Z. Pu, and Y. Wang, TransFollower: Long-Sequence Car-Following Trajectory Prediction through Transformer, 2022.
  • Zhu et al. [2018b] Zhu, M., X. Wang, and Y. Wang, Human-like autonomous car-following model with deep reinforcement learning. Transportation Research Part C: Emerging Technologies, Vol. 97, 2018b, pp. 348–368.
  • Chai and Wong [2015] Chai, C. and Y. D. Wong, Fuzzy cellular automata model for signalized intersections. Computer-Aided Civil and Infrastructure Engineering, Vol. 30, No. 12, 2015, pp. 951–964.
  • Zhu et al. [2020] Zhu, M., Y. Wang, Z. Pu, J. Hu, X. Wang, and R. Ke, Safe, efficient, and comfortable velocity control based on reinforcement learning for autonomous driving. Transportation Research Part C: Emerging Technologies, Vol. 117, 2020, p. 102662.
  • Gao et al. [2018] Gao, H., G. Shi, G. Xie, and B. Cheng, Car-following method based on inverse reinforcement learning for autonomous vehicle decision-making. International Journal of Advanced Robotic Systems, Vol. 15, No. 6, 2018, p. 1729881418817162.
  • Zhao et al. [2022] Zhao, Z., Z. Wang, K. Han, R. Gupta, P. Tiwari, G. Wu, and M. J. Barth, Personalized car following for autonomous driving with inverse reinforcement learning. In 2022 International Conference on Robotics and Automation (ICRA), IEEE, 2022, pp. 2891–2897.
  • Song et al. [2023] Song, D., B. Zhu, J. Zhao, J. Han, and Z. Chen, Personalized Car-Following Control Based on a Hybrid of Reinforcement Learning and Supervised Learning. IEEE Transactions on Intelligent Transportation Systems, 2023.
  • Mo and Di [2022] Mo, Z. and X. Di, Uncertainty quantification of car-following behaviors: physics-informed generative adversarial networks. In the 28th ACM SIGKDD in conjunction with the 11th International Workshop on Urban Computing (UrbComp2022), 2022.
  • Ding et al. [2022] Ding, Z., D. Xu, C. Tu, H. Zhao, M. Moze, F. Aioun, and F. Guillemard, Driver identification through heterogeneity modeling in car-following sequences. IEEE Transactions on Intelligent Transportation Systems, Vol. 23, No. 10, 2022, pp. 17143–17156.
  • Wen et al. [2022] Wen, X., Z. Cui, and S. Jian, Characterizing car-following behaviors of human drivers when following automated vehicles using the real-world dataset. Accident Analysis & Prevention, Vol. 172, 2022, p. 106689.
  • Huang et al. [2022] Huang, Y., X. Yan, X. Li, K. Duan, A. Rakotonirainy, and Z. Gao, Improving car-following model to capture unobserved driver heterogeneity and following distance features in fog condition. Transportmetrica A: Transport Science, 2022, pp. 1–24.
  • Kim et al. [2013b] Kim, I., T. Kim, and K. Sohn, Identifying driver heterogeneity in car-following based on a random coefficient model. Transportation research part C: emerging technologies, Vol. 36, 2013b, pp. 35–44.
  • Li and Malik [2016] Li, K. and J. Malik, Learning to optimize. arXiv preprint arXiv:1606.01885, 2016.
  • Santoro et al. [2016] Santoro, A., S. Bartunov, M. Botvinick, D. Wierstra, and T. Lillicrap, Meta-learning with memory-augmented neural networks. In International conference on machine learning, PMLR, 2016, pp. 1842–1850.
  • Finn et al. [2017] Finn, C., P. Abbeel, and S. Levine, Model-agnostic meta-learning for fast adaptation of deep networks. In International conference on machine learning, PMLR, 2017, pp. 1126–1135.
  • Lee and Choi [2018] Lee, Y. and S. Choi, Gradient-based meta-learning with learned layerwise metric and subspace. In International Conference on Machine Learning, PMLR, 2018, pp. 2927–2936.
  • Chen et al. [2021] Chen, Y., Z. Liu, H. Xu, T. Darrell, and X. Wang, Meta-baseline: Exploring simple meta-learning for few-shot learning. In Proceedings of the IEEE/CVF international conference on computer vision, 2021, pp. 9062–9071.
  • Lee et al. [2022] Lee, H.-y., S.-W. Li, and N. T. Vu, Meta learning for natural language processing: A survey. arXiv preprint arXiv:2205.01500, 2022.
  • Alet et al. [2018] Alet, F., T. Lozano-Pérez, and L. P. Kaelbling, Modular meta-learning. In Conference on robot learning, PMLR, 2018, pp. 856–868.
  • Schweighofer and Doya [2003] Schweighofer, N. and K. Doya, Meta-learning in reinforcement learning. Neural Networks, Vol. 16, No. 1, 2003, pp. 5–9.
  • Ye et al. [2021] Ye, F., P. Wang, C.-Y. Chan, and J. Zhang, Meta reinforcement learning-based lane change strategy for autonomous vehicles. In 2021 IEEE Intelligent Vehicles Symposium (IV), IEEE, 2021, pp. 223–230.
  • ** et al. [2022] **, Y., K. Chen, and Q. Yang, Selective cross-city transfer learning for traffic prediction via source city region re-weighting. In Proceedings of the 28th ACM SIGKDD Conference on Knowledge Discovery and Data Mining, 2022, pp. 731–741.
  • Bishop and Nasrabadi [2006] Bishop, C. M. and N. M. Nasrabadi, Pattern recognition and machine learning, Vol. 4. Springer, 2006.
  • Dingus et al. [2015] Dingus, T. A., J. M. Hankey, J. F. Antin, S. E. Lee, L. Eichelberger, K. E. Stulce, D. McGraw, M. Perez, and L. Stowe, Naturalistic driving study: Technical coordination and quality control. SHRP 2 Report S2-S06-RW-1, 2015.
  • Wang et al. [2017] Wang, X., R. Jiang, L. Li, Y. Lin, X. Zheng, and F.-Y. Wang, Capturing car-following behaviors by deep learning. IEEE Transactions on Intelligent Transportation Systems, Vol. 19, No. 3, 2017, pp. 910–920.
  • Zhao et al. [2017] Zhao, D., Y. Guo, and Y. J. Jia, Trafficnet: An open naturalistic driving scenario library. In 2017 IEEE 20th International Conference on Intelligent Transportation Systems (ITSC), IEEE, 2017, pp. 1–8.