Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Efficient Fusion and Task Guided Embedding for End-to-end Autonomous Driving

Yipin Guo    Yilin Lang    Qinyuan Ren {guoyipin, langyilin, renqinyuan}@zju.edu.cn College of Control Science and Engineering, Zhejiang University, Hangzhou 310027, P. R. China,
Abstract

To address the challenges of sensor fusion and safety risk prediction, contemporary closed-loop autonomous driving neural networks leveraging imitation learning typically require a substantial volume of parameters and computational resources to run neural networks. Given the constrained computational capacities of onboard vehicular computers, we introduce a compact yet potent solution named EfficientFuser. This approach employs EfficientViT for visual information extraction and integrates feature maps via cross attention. Subsequently, it utilizes a decoder-only transformer for the amalgamation of multiple features. For prediction purposes, learnable vectors are embedded as tokens to probe the association between the task and sensor features through attention. Evaluated on the CARLA simulation platform, EfficientFuser demonstrates remarkable efficiency, utilizing merely 37.6% of the parameters and 8.7% of the computations compared to the state-of-the-art lightweight method with only 0.4% lower driving score, and the safety score neared that of the leading safety-enhanced method, showcasing its efficacy and potential for practical deployment in autonomous driving systems.

keywords:
End-to-end Autonomous Driving, Hardware Efficient, Data Driving, Imitation Learning, Transformer
00footnotetext: Best Paper Award of the IEEE 13th Data-Driven Control and Learning Systems Conference. This work was supported by NSFC 62088101 Autonomous Intelligent Unmanned Systems.

1 Introduction

The success of deep neural networks (DNNs) has paved the way for data driven learning-based approaches to autonomous driving (AD), utilizing large-scale data and computation. This has made end-to-end autonomous driving (E2E AD) systems a reality. Defined as fully differentiable programs that directly map raw sensor data to planned actions or low-level control commands, E2E AD eliminates the need for intermediate modules. This not only simplifies development but also has the potential to improve performance.

E2E AD primarily advances along two main directions: reinforcement learning [1, 2] (RL) and imitation learning (IL). While RL robustifies against data distribution shifts, recent advancements in driving scene generation [3], coupled with the growing availability of data from electric vehicles, have made IL increasingly attractive.

Early E2E AD with IL [4] utilized convolutional neural networks (CNNs) to extract image features and directly imitate control actions. However, limited by data availability and computational power, these early systems struggled to achieve good performance. Subsequently, most research shifted towards predicting trajectories (i.e. waypoints). LBC [5] utilizes policy distillation where a teacher model, trained with ground-truth BEV semantic maps, predicts future waypoints. A student model, using only image data, learns from the teacher’s predictions. TransFuser [6, 7] uses two CNNs to extract image and LiDAR information, followed by transformers to fuse these information at each downsampling step. Similarly, InterFuser [8] extracts sensor features with CNNs, but leverages an encoder-decoder structure to additionally incorporate traffic rules and vehicle density information, aiming for a safe driving strategy. TCP [9] combines two kinds of prediction goals, introduces control actions and waypoints within a sequence of time into training, and uses a control method that mixes trajectory trackers and behavior prediction results.

Despite their promise, end-to-end approaches face a critical hurdle: the mismatch between DNNs’ hefty computational demands and AD’s need for real-time, low-latency operation. Current on-board hardware struggles to handle the complex architecture and massive parameters of DNNs, limiting the real-world feasibility of end-to-end systems. Recently, more and more works have focused on lightweight neural network (NN) designs. MobileNet [10, 11] utilize depthwise separable convolutions for computational efficiency without losing accuracy. MCUNet [12, 13] pushes the limits of miniaturization, fitting networks onto embedded platforms with only 256KB of memory, significantly expanding AI deployment possibilities in resource-constrained environments. Despite advancements in lightweight NN designs, there remains a scarcity of such networks specifically tailored for E2E AD systems.

To overcome the computational hurdles of E2E AD, we introduce EfficientFuser, a powerful and hardware-friendly model, which uses EfficientViT [14] for feature extraction from multi-view and employs a decoder-only transformer with task guided embedding for prediction. Cross attention seamlessly integrates multi-view information at different scales. Remarkably, EfficientFuser maintains strong capabilities and efficiency for requiring fewer parameters and computations, making it ideal for applications. The main contributions can be summarized as follows:

  • EfficientFuser fuses multiple camera views through cross attention, providing a richer understanding of the environment without straining much computation.

  • A decoder-only transformer is used for the prediction process. With learnable vectors as embedded tokens, the decoder finds the connection between tasks and sensor features through attention.

  • The predicted waypoints and control inputs are mixed dynamically, offering flexible adaptation to diverse driving scenarios and potentially safe behaviors.

Refer to caption
Figure 1: Architecture. Image feature extraction: Consider multi-view RGB images as inputs to EfficientFuser which uses several cross attention for the fusion of feature maps between different views. EfficientViT is used as the visual backbone, which makes it suitable for cross attention while maintaining a low computational resource footprint. Since the main-view and focus-view images are more important, larger backbones are used, and side-view images use smaller backbones. Decoder: A decoder-only transformer is used to make predictions. Visual features and sensor features are input as tokens. Additionally, two trainable vectors are set up as tokens, learning relevant information with other tokens at the early stage in the decoder layer.

2 EfficientFuser

In this work, an architecture for end-to-end driving is proposed, as shown in Fig. 1, with three components: (1) Cross Attention feature fusion. (2) Decoder-only transformer for prediction. (3) Dynamically mix for behavioral prediction results and waypoint tracker actions. The following sections detail the problem setting, input and output, and each component of the model.

2.1 Problem Setting

EfficientFuser’s decision-making process is anchored by an extensive input state, symbolized as x𝑥xitalic_x, which amalgamates multiple data sources: sensor signal i𝑖iitalic_i, capturing real-time environmental conditions via vehicle camera; vehicle speed v𝑣vitalic_v, indicating the current velocity; and high-level navigation information g𝑔gitalic_g, comprising discrete navigation commands and target coordinates from the global planner. To modulate the vehicle’s speed and direction, the system generates outputs for throttle[0,1]𝑡𝑟𝑜𝑡𝑡𝑙𝑒01throttle\in[0,1]italic_t italic_h italic_r italic_o italic_t italic_t italic_l italic_e ∈ [ 0 , 1 ], brake[0,1]𝑏𝑟𝑎𝑘𝑒01brake\in[0,1]italic_b italic_r italic_a italic_k italic_e ∈ [ 0 , 1 ] and steer[1,1]𝑠𝑡𝑒𝑒𝑟11steer\in[-1,1]italic_s italic_t italic_e italic_e italic_r ∈ [ - 1 , 1 ], controlling acceleration, deceleration, and steering respectively.

The goal of IL is to learn a policy π𝜋\piitalic_π that imitates the behavior of an expert πsuperscript𝜋\pi^{*}italic_π start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT. TCP believes that both trajectory and control actions contain important driving information, so both should be used as imitation targets. EfficientFuser follow the settings in TCP. The difference is that TCP first predicts the trajectory and then uses it to guide the prediction of control actions, while the information between the two can be fully exchanged through attention [15] in the decoder and predicted at the same time for EfficientFuser. Imitation target can be formulated as:

argminWED(x,[πt,πc])[(πt,πt)+(πc,πc)],subscript𝑊subscript𝐸similar-to𝐷𝑥superscriptsubscript𝜋𝑡superscriptsubscript𝜋𝑐delimited-[]subscript𝜋𝑡superscriptsubscript𝜋𝑡subscript𝜋𝑐superscriptsubscript𝜋𝑐\arg\min_{W}E_{D\sim(x,[\pi_{t}^{*},\pi_{c}^{*}])}[\mathcal{L}(\pi_{t},\pi_{t}% ^{*})+\mathcal{L}(\pi_{c},\pi_{c}^{*})],roman_arg roman_min start_POSTSUBSCRIPT italic_W end_POSTSUBSCRIPT italic_E start_POSTSUBSCRIPT italic_D ∼ ( italic_x , [ italic_π start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT , italic_π start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ] ) end_POSTSUBSCRIPT [ caligraphic_L ( italic_π start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT , italic_π start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) + caligraphic_L ( italic_π start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT , italic_π start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ) ] , (1)

where D(x,[πt,πc])similar-to𝐷𝑥superscriptsubscript𝜋𝑡superscriptsubscript𝜋𝑐D\sim(x,[\pi_{t}^{*},\pi_{c}^{*}])italic_D ∼ ( italic_x , [ italic_π start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT , italic_π start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ] ) is a dataset comprised of state-action pairs collected from the expert. πcsubscript𝜋𝑐\pi_{c}italic_π start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT denotes the policy of control branch and πtsubscript𝜋𝑡\pi_{t}italic_π start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT denotes the trajectory prediction. \mathcal{L}caligraphic_L is the loss measuring how close the action from the expert and the action from the model is.

Roach [16] is used as the expert, which is a relatively straightforward model that has been trained using RL with access to privileged information. This information encompasses various aspects of the driving environment such as roads, lanes, routes, vehicles, pedestrians, traffic lights, and stop signs, all of which are rendered into a 2D Bird’s Eye View (BEV) image. Compared with experts made by hand-crafted rules, it can provide latent features for control action prediction as intermediate supervision, making the training more stable.

2.2 Architecture Design

A structure that combines visual transformer (ViT) [17] and decoder-only transformer in large language model (LLM) [18, 19] is used. The transformer has since revolutionized various fields, including computer vision and even time-series modeling, which can be formulated as

out=Transformer(x)=MLP(Attention(x)).𝑜𝑢𝑡𝑇𝑟𝑎𝑛𝑠𝑓𝑜𝑟𝑚𝑒𝑟𝑥𝑀𝐿𝑃𝐴𝑡𝑡𝑒𝑛𝑡𝑖𝑜𝑛𝑥out=Transformer(x)=MLP(Attention(x)).italic_o italic_u italic_t = italic_T italic_r italic_a italic_n italic_s italic_f italic_o italic_r italic_m italic_e italic_r ( italic_x ) = italic_M italic_L italic_P ( italic_A italic_t italic_t italic_e italic_n italic_t italic_i italic_o italic_n ( italic_x ) ) . (2)

The core innovation of transformer lies in the attention mechanism, which allows the model to weigh the importance of different parts of the input data. The calculation of attention can be expressed as

out=Attention(Q,K,V)=softmax(QKTDk)V.𝑜𝑢𝑡𝐴𝑡𝑡𝑒𝑛𝑡𝑖𝑜𝑛𝑄𝐾𝑉𝑠𝑜𝑓𝑡𝑚𝑎𝑥𝑄superscript𝐾𝑇subscript𝐷𝑘𝑉out=Attention(Q,K,V)=softmax(\frac{QK^{T}}{\sqrt{D_{k}}})V.italic_o italic_u italic_t = italic_A italic_t italic_t italic_e italic_n italic_t italic_i italic_o italic_n ( italic_Q , italic_K , italic_V ) = italic_s italic_o italic_f italic_t italic_m italic_a italic_x ( divide start_ARG italic_Q italic_K start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG start_ARG square-root start_ARG italic_D start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG end_ARG ) italic_V . (3)

These mechanisms involve mapping the input to three vectors: Query (Q𝑄Qitalic_Q), Key (K𝐾Kitalic_K), and Value (V𝑉Vitalic_V), often through linear layers. Dksubscript𝐷𝑘D_{k}italic_D start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT is the dimension size of K𝐾Kitalic_K, which is scaled to prevent training instability. When dealing with a single input, these vectors all correspond to the same input, known as self-attention, allows the model to find inherent connections. When working with two inputs, the Q𝑄Qitalic_Q originates from one source, while the K𝐾Kitalic_K and V𝑉Vitalic_V come from the other. This cross-attention mechanism enables Q𝑄Qitalic_Q from different source to selectively focus on relevant features.

2.2.1 Image Backbone

EfficientViT [14] is used as the visual backbone. ViT pioneered the paradigm shift of applying the transformer architecture, originally designed for natural language processing, to the domain of computer vision. It firstly divides an input image into smaller patches (e.g., 16x16 pixels). These patches are then flattened and treated as a sequence of tokens, similar to words in a sentence. Then self-attention mechanism of the transformer operates on these image patch tokens. This allows ViT to learn complex relationships between different regions of the image.

While standard ViTs often require significant computational resources, EfficientViT with Cascaded Group Attention can maintain efficiency. This approach allows faster processing and reduced memory usage.

Unlike CNN-based fusion approaches Transfuser, which relies on multiple transformer layers for fusion after each downsampling, ViT can work with small image patches directly. For the Transfuser, CNNs still necessitate feature pooling before the transformer stage to reduce the computational burden, potentially followed by interpolation to restore resolution. This interpolation can compromise the integrity of the original feature map.

To retain the original image features, a method that leverages cross-attention between the two image backbones is employed, which enables knowledge exchange between both perspectives, guiding them to focus on relevant features in specific areas as Fig. 2.

Refer to caption
Figure 2: Fusion with Cross Attention.

2.2.2 Decoder Transformer

In the realm of sequence processing, transformers often rely on an encoder-decoder structure [15]. The encoder first processes the input sequence (image patches) and extracts essential information, culminating in a context vector that captures the entire input’s essence. The decoder then takes over, utilizing the context vector and query from additional information (speed, command, occupation map, etc.) to generate the output sequence element by element. This is how InterFuser works.

Inspired by the prevalent large language models like GPT [18], we employ a decoder-only transformer architecture for unified learning across all input data. Research has demonstrated that this decoder-only framework offers enhanced generalization capabilities [20], alongside a more streamlined structure, improved operational efficiency, and superior scalability.

Contrary to the typical transformer-based architecture that interacts token features and then feeds them into the prediction head, our approach introduces a unique embedding vector for predictions. This vector is initialized using random values drawn from a Gaussian distribution, with feature representations for the prediction task being learned throughout training. This technique not only enhances the scalability of the decoder component but also leverages the attention mechanism to identify advantageous relationships between tokens at an early stage. The schematic diagram is shown in Fig. 3.

Refer to caption
Figure 3: Decoder-only transformer with task guided learnable vector embedded.

2.2.3 Dynamic control

Contrary to TCP, which prioritizes the control volume predicted by the model and overlooks the control volume of the waypoint tracker, it’s argued that the preferences of the underlying controller should be dynamically adapted based on the driving scenario.

To accomplish this objective, a loss estimator is developed that utilizes the hidden features from the GRU within the waypoint head and the intermediate features from the Control head as its inputs, thereby modeling the training losses of both. The outcome of this process is leveraged as a measure of prediction confidence, which in turn is used to adjust the preference for the final control sequence. More specifically, this can be articulated as follows:

c~=Linear1(Xc)c,w~=Linear2(Xw)w,Prefer=1exp(kcc~)exp(kcc~)+exp(kww~),matrix~subscript𝑐𝐿𝑖𝑛𝑒𝑎subscript𝑟1subscript𝑋𝑐subscript𝑐~subscript𝑤𝐿𝑖𝑛𝑒𝑎subscript𝑟2subscript𝑋𝑤subscript𝑤𝑃𝑟𝑒𝑓𝑒𝑟1subscript𝑘𝑐~subscript𝑐subscript𝑘𝑐~subscript𝑐subscript𝑘𝑤~subscript𝑤\begin{matrix}\tilde{\mathcal{L}_{c}}=Linear_{1}(X_{c})\longrightarrow\mathcal% {L}_{c},\\ \tilde{\mathcal{L}_{w}}=Linear_{2}(X_{w})\longrightarrow\mathcal{L}_{w},\\ Prefer=1-\dfrac{\exp(k_{c}\tilde{\mathcal{L}_{c}})}{\exp(k_{c}\tilde{\mathcal{% L}_{c}})+\exp(k_{w}\tilde{\mathcal{L}_{w}})},\end{matrix}start_ARG start_ROW start_CELL over~ start_ARG caligraphic_L start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG = italic_L italic_i italic_n italic_e italic_a italic_r start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_X start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) ⟶ caligraphic_L start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT , end_CELL end_ROW start_ROW start_CELL over~ start_ARG caligraphic_L start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT end_ARG = italic_L italic_i italic_n italic_e italic_a italic_r start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_X start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT ) ⟶ caligraphic_L start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT , end_CELL end_ROW start_ROW start_CELL italic_P italic_r italic_e italic_f italic_e italic_r = 1 - divide start_ARG roman_exp ( italic_k start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT over~ start_ARG caligraphic_L start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG ) end_ARG start_ARG roman_exp ( italic_k start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT over~ start_ARG caligraphic_L start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG ) + roman_exp ( italic_k start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT over~ start_ARG caligraphic_L start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT end_ARG ) end_ARG , end_CELL end_ROW end_ARG (4)

where w~~subscript𝑤\tilde{\mathcal{L}_{w}}over~ start_ARG caligraphic_L start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT end_ARG and c~~subscript𝑐\tilde{\mathcal{L}_{c}}over~ start_ARG caligraphic_L start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG are the predicted waypoint loss and control loss, which need to be as close as possible to the real loss wsubscript𝑤\mathcal{L}_{w}caligraphic_L start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT and csubscript𝑐\mathcal{L}_{c}caligraphic_L start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. kc,kwsubscript𝑘𝑐subscript𝑘𝑤k_{c},k_{w}italic_k start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT , italic_k start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT are coefficients used to make up for the difference in the numerical range of the two losses, which is determined based on experience.

3 Experiments

Table 1: Driving performance. We report the mean deviation over 3 runs of each method. \downarrow indicates that the smaller the value, the better. \uparrow indicates that the bigger the value, the performance better.
\hhlineMethod Sensor Param(M)\downarrow Flops(G)\downarrow Town05 Short Town05 Long
Camera Lidar DS\uparrow RC\uparrow DS\uparrow RC\uparrow
CIL \checkmark ×\times× 12.7 2.4 7.5 13.4 3.7 7.2
LBC \checkmark ×\times× 13.0 3.16 31.0 55.0 7.1 32.1
AIM[6] \checkmark ×\times× 21.5 4.8 49.0 81.1 26.5 60.7
Late Fusion \checkmark \checkmark 32.6 7.1 51.6 81.1 26.5 60.7
Transfuser \checkmark \checkmark 66.1 11.4 54.5 78.4 33.1 56.3
TCP \checkmark ×\times× 25.8 17.1 80.9 96.1 45.1 71.5
InterFuser \checkmark \checkmark 82.2 46.1 86.9 92.7 33.6 56.7
EfficientFuser \checkmark ×\times× 9.7 0.51 75.8 85.2 40.6 56.7
EfficientFuser(wider view) \checkmark ×\times× 9.7 1.46 80.5 92.2 42.7 59.1
\hhline

3.1 Setup

Evaluation metrics. four metrics are used to evaluate the effectiveness of the methods: Driving Score (DS), Route Completion (RC), Number of parameters (Param), and Floating Point Operations (Flops). Among them, DS and RC are used to represent the driving effect, and Param and Flops are used to represent the efficiency of the neural network.

Driving Score (DS) is the main metric of the Carla leaderboard[21], serving as the product between the route completion and the infractions penalty. RC is the percentage of the route distance completed by an agent. Param refers to the total count of parameters that should be saved, which represents the size of the NN and usually determines the amount of storage space required to run the NN. Flops is a measure of the computational cost required to perform a forward pass (inference) through the network.

Dataset. The CARLA[22] simulator is used for training and testing, specifically CARLA 0.9.10. 7 towns are used for training and Town05 for evaluation. Town05 is selected for evaluation due to the large diversity in drivable regions compared to other CARLA towns, e.g. multi-lane and single-lane roads, highways and exits, bridges and underpasses. To further enhance the assessment of driving safety, scenarios involving the sudden appearance of pedestrians and erratic vehicle behavior have been integrated into the simulation environment. Two evaluation settings are considered: (1) Town05 Short: 32 short routes of 100-500m comprising 3 intersections each, (2) Town05 Long: 10 long routes of 1000-2000m comprising 10 intersections each. The weather condition is ClearNoon.

Training. EfficientViT-m1 and m0 are used as visual backbone and load ImageNet[23] pre-trained weights. Other parts are initialized with Gaussian random numbers. EfficientFuser is trained for 60 epochs with a learning rate of 0.0005, and then trained for 60 epochs with a learning rate of 0.0001. The batch size is 256. Adam is used with weightDecay=1e-7. The learning rate is reduced to half every 30 epochs. Four parts of loss are set, namely speed loss ssubscript𝑠\mathcal{L}_{s}caligraphic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, feature loss fsubscript𝑓\mathcal{L}_{f}caligraphic_L start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT, waypoint loss wsubscript𝑤\mathcal{L}_{w}caligraphic_L start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT, and control loss csubscript𝑐\mathcal{L}_{c}caligraphic_L start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. fsubscript𝑓\mathcal{L}_{f}caligraphic_L start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT and ssubscript𝑠\mathcal{L}_{s}caligraphic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT are the intermediate supervision to guide training, which is similar to TCP.

3.2 Results

Table 1 presents the comparative analysis of EfficientFuser alongside other notable studies within the public Carla Leaderboard framework. EfficientFuser’s performance is delineated in two distinct versions. The initial variant incorporates inputs from both the front view and the focus view (i.e., an enhanced frontal perspective), with an image resolution of 256x256. The ’Wider view’ iteration maintains identical camera orientation but expands the image width to 768 pixels, thereby capturing a broader spectrum of information.

EfficientFuser significantly reduces the size of closed-loop autonomous driving models derived from imitation learning to an unprecedented level. In comparison to the state-of-the-art lightweight approach, TCP, EfficientFuser’s model size is merely 37.6% as large, and its computational demand is just 8.5% of TCP’s. Despite its considerably smaller neural network (NN) size, EfficientFuser only experiences a marginal decrease of 0.4% in Driving Score (DS) in the Town05 Short scenario. When compared with CIL under similar parameters and computational loads, EfficientFuser substantially outperforms it, showing a remarkable 73-point advantage in DS.

As a cutting-edge solution for accessible closed-loop autonomous driving, InterFuser achieves a similar Route Completion (RC) to EfficientFuser but improves DS by 6.3 points. This improvement, however, comes at the expense of the number of parameters 8.4×\times× and computational requirements 31.6×\times×. Based on detection outcomes, these models often resort to a cautious approach, suggesting that the vehicle would proceed at a slow pace when it has been stationary for an extended period without any obstacles in its path. With its enhanced safety driving strategy, InterFuser would even recognize the red light at the next unreachable intersection far away and come to a halt — a behavior not aligned with typical human driving patterns, which EfficientFuser does not exhibit.

To highlight the security efficacy of EfficientFuser, the penalties incurred for infractions by various methods are reported on Town05 Short. EfficientFuser significantly surpasses the security benchmarks set by both TCP and Transfuser, and it only marginally falls short of Interfuser, which prioritizes security at its core.

Refer to caption
Figure 4: The penalties incurred for infractions.

EfficientFuser employs a novel approach by introducing a learnable embedding vector for predictions. This methodology diverges from traditional practices that rely on feeding sensor token information directly into the prediction head. By adopting this design, EfficientFuser begins to explore the correlation between the prediction target and the input data at an early stage, thereby generating task-specific representations. To illustrate the impact of this approach, two representative attention maps are visualized. The red lines divide the tokens from different information sources. Taking the abscissa as an example, from left to right they are prediction tokens, measurement tokens, side-view tokens and main-view tokens.

Refer to caption
(a) Layer0 Head3. Prediction tokens focus on the measurement tokens.
Refer to caption
(b) Layer4 Head2. Information from a specific patch in the main view is accentuated.
Figure 5: Attention Maps in Decoder.

3.3 Ablation Study

To thoroughly investigate the system architecture and evaluate the effectiveness of our proposed method, we have carried out a series of ablation studies. Initially, the impact of utilizing image backbones of varying sizes and the performance enhancement brought about by the cross-attention fuser were assessed. Subsequently, we explored the effects of altering the decoder layer depths and implementing learnable vector embeddings. Lastly, we compared the dynamic control adjustment with the static allocation approach employed in TCP.

The outcomes of these experiments were all gathered on Town05 Short, providing a comprehensive analysis of each component’s contribution to the overall performance.

3.3.1 Image Backbone

Fusion with cross attention is an efficient method, resulting in only 5.7% additional parameters and 7.2% computation in EfficientFuser.Despite its minimal computational requirement, Cross Attention plays a pivotal role in EfficientFuser. The absence of this multi-level image information interaction significantly impacts the system’s efficacy, leading to a notable reduction of 14.9 points in the driving score. This underlines the critical importance of Cross Attention in facilitating effective feature fusion and enhancing the model’s ability to make informed decisions.

Intuitively, one might assume that a larger image backbone would yield superior results, given its enhanced capabilities for feature extraction. Nevertheless, as demonstrated in Tab.2, an increase in backbone size actually detracts from the model’s performance. This counterintuitive outcome may be attributed to the complexity of the training tasks, which hinders the full training potential of larger backbones.

Table 2: Ablation on the image backbone
\hhlineBackbone Cross Attn Param(M) DS RC
EfficientViTm1𝐸𝑓𝑓𝑖𝑐𝑖𝑒𝑛𝑡𝑉𝑖subscript𝑇𝑚1EfficientViT_{m1}italic_E italic_f italic_f italic_i italic_c italic_i italic_e italic_n italic_t italic_V italic_i italic_T start_POSTSUBSCRIPT italic_m 1 end_POSTSUBSCRIPT No 9.2 60.9 76.3
EfficientViTm0𝐸𝑓𝑓𝑖𝑐𝑖𝑒𝑛𝑡𝑉𝑖subscript𝑇𝑚0EfficientViT_{m0}italic_E italic_f italic_f italic_i italic_c italic_i italic_e italic_n italic_t italic_V italic_i italic_T start_POSTSUBSCRIPT italic_m 0 end_POSTSUBSCRIPT
EfficientViTm1𝐸𝑓𝑓𝑖𝑐𝑖𝑒𝑛𝑡𝑉𝑖subscript𝑇𝑚1EfficientViT_{m1}italic_E italic_f italic_f italic_i italic_c italic_i italic_e italic_n italic_t italic_V italic_i italic_T start_POSTSUBSCRIPT italic_m 1 end_POSTSUBSCRIPT Yes 9.7 75.8 85.2
EfficientViTm0𝐸𝑓𝑓𝑖𝑐𝑖𝑒𝑛𝑡𝑉𝑖subscript𝑇𝑚0EfficientViT_{m0}italic_E italic_f italic_f italic_i italic_c italic_i italic_e italic_n italic_t italic_V italic_i italic_T start_POSTSUBSCRIPT italic_m 0 end_POSTSUBSCRIPT
EfficientViTm5𝐸𝑓𝑓𝑖𝑐𝑖𝑒𝑛𝑡𝑉𝑖subscript𝑇𝑚5EfficientViT_{m5}italic_E italic_f italic_f italic_i italic_c italic_i italic_e italic_n italic_t italic_V italic_i italic_T start_POSTSUBSCRIPT italic_m 5 end_POSTSUBSCRIPT Yes 37.5 65.3 82.0
EfficientViTm4𝐸𝑓𝑓𝑖𝑐𝑖𝑒𝑛𝑡𝑉𝑖subscript𝑇𝑚4EfficientViT_{m4}italic_E italic_f italic_f italic_i italic_c italic_i italic_e italic_n italic_t italic_V italic_i italic_T start_POSTSUBSCRIPT italic_m 4 end_POSTSUBSCRIPT
\hhline

3.3.2 Decoder Layer

In evaluating our decoder layer, we established two sets of comparative analyses. The initial set investigates the impact of varying the depth of the decoder layer. Following this, we examine our proposed method of employing learnable vectors as prediction tokens. For comparative purposes, we conducted an additional experiment wherein, rather than utilizing the learnable vector and integrating it into the token dimension, we averaged the sensor features and sent them into the prediction head to observe the effect on predictions. The outcomes of these investigations are presented in Tab. 3.

Table 3: Ablation on decoder layers
\hhlineDepth Learnable Param(M) Flops(G) DS RC
Vector
1 Yes 6.6 0.39 51.4 62.4
4 Yes 8.0 0.44 71.2 80.4
8 No 9.7 0.51 70.0 82.7
8 Yes 9.7 0.51 75.8 85.2
12 Yes 11.5 0.59 67.0 90.2
\hhline

Although it introduces a minor computational overhead, the learnable vector approach utilized by EfficientFuser significantly enhances the driving performance. Concerning the depth of the decoder layer, it is evident that increasing the depth does not necessarily improve performance. While the RC of the vehicle may increase, the DS substantially decreases, indicating that the vehicle starts to overlook safety-related information. This phenomenon can also be observed in the visual attention maps; beginning from the seventh layer, the focus shifts away from the prediction token towards reinforcing certain measurement token information.

3.3.3 Dynamic Control

EfficientFuser adopts the hybrid control utilized by TCP and incorporates a dynamic allocation strategy. To validate the efficacy of the dynamic adjustment method, the experiments about the TCP settings and dynamic setting are introduced. The outcomes in Tab. 4 affirm the effectiveness of the dynamic allocation component, enhancing the DS and RC.

Table 4: Ablation on the control strategy
\hhlineMethod DS RC
Same with TCP 72.4 81.0
With dynamic adjustment 75.8 85.2
\hhline

4 Conclusion

EfficientFuser markedly diminishes both the size and computational demands of neural networks by incorporating a compact and efficient vision transformer as its visual backbone, coupled with the use of cross-attention mechanisms for the integration of information. The adoption of a decoder-only transformer architecture alongside learnable prediction vectors ensures that the system maintains commendable performance despite its reduced scale. Furthermore, alterations to the fundamental controller not only boost the driving efficiency but also significantly augment the safety features of the system. As the smallest closed-loop autonomous driving NN available to date, EfficientFuser stands out for its exceptional performance. However, similar to TCP, EfficientFuser not only learns from the results but also from intermediate features, which makes it suffer from some performance degradation when sim-to-real.

References

  • [1] A. Kendall, J. Hawke, D. Janz, P. Mazur, D. Reda, J.-M. Allen, V.-D. Lam, A. Bewley, and A. Shah, “Learning to drive in a day,” in 2019 International Conference on Robotics and Automation (ICRA), pp. 8248–8254, 2019.
  • [2] M. Toromanoff, E. Wirbel, and F. Moutarde, “End-to-end model-free reinforcement learning for urban driving using implicit affordances,” in Proceedings of the IEEE/CVF Conference on Computer Vision and Pattern Recognition (CVPR), June 2020.
  • [3] H. Shao, L. Wang, R. Chen, S. L. Waslander, H. Li, and Y. Liu, “Reasonnet: End-to-end driving with temporal and global reasoning,” in Proceedings of the IEEE/CVF Conference on Computer Vision and Pattern Recognition, pp. 13723–13733, 2023.
  • [4] M. Bojarski, D. D. Testa, D. Dworakowski, B. Firner, B. Flepp, P. Goyal, L. D. Jackel, M. Monfort, U. Muller, J. Zhang, X. Zhang, J. Zhao, and K. Zieba, “End to end learning for self-driving cars,” 2016.
  • [5] D. Chen, B. Zhou, V. Koltun, and P. Krähenbühl, “Learning by cheating,” in Proceedings of the Conference on Robot Learning, vol. 100 of Proceedings of Machine Learning Research, pp. 66–75, PMLR, 30 Oct–01 Nov 2019.
  • [6] A. Prakash, K. Chitta, and A. Geiger, “Multi-modal fusion transformer for end-to-end autonomous driving,” in 2021 IEEE/CVF Conference on Computer Vision and Pattern Recognition (CVPR), pp. 7073–7083, 2021.
  • [7] K. Chitta, A. Prakash, B. Jaeger, Z. Yu, K. Renz, and A. Geiger, “Transfuser: Imitation with transformer-based sensor fusion for autonomous driving,” IEEE Transactions on Pattern Analysis and Machine Intelligence, vol. 45, no. 11, pp. 12878–12895, 2023.
  • [8] H. Shao, L. Wang, R. Chen, H. Li, and Y. Liu, “Safety-enhanced autonomous driving using interpretable sensor fusion transformer,” 2022.
  • [9] P. Wu, X. Jia, L. Chen, J. Yan, H. Li, and Y. Qiao, “Trajectory-guided control prediction for end-to-end autonomous driving: A simple yet strong baseline,” in Advances in Neural Information Processing Systems, vol. 35, pp. 6119–6132, Curran Associates, Inc., 2022.
  • [10] M. Sandler, A. Howard, M. Zhu, A. Zhmoginov, and L.-C. Chen, “Mobilenetv2: Inverted residuals and linear bottlenecks,” in 2018 IEEE/CVF Conference on Computer Vision and Pattern Recognition(CVPR), pp. 4510–4520, 2018.
  • [11] A. Howard, M. Sandler, B. Chen, W. Wang, L.-C. Chen, M. Tan, G. Chu, V. Vasudevan, Y. Zhu, R. Pang, H. Adam, and Q. Le, “Searching for mobilenetv3,” in 2019 IEEE/CVF International Conference on Computer Vision (ICCV), pp. 1314–1324, 2019.
  • [12] J. Lin, W.-M. Chen, Y. Lin, j. cohn, C. Gan, and S. Han, “Mcunet: Tiny deep learning on iot devices,” in Advances in Neural Information Processing Systems, vol. 33, pp. 11711–11722, Curran Associates, Inc., 2020.
  • [13] J. Lin, W.-M. Chen, H. Cai, C. Gan, and S. Han, “Memory-efficient patch-based inference for tiny deep learning,” in Advances in Neural Information Processing Systems, vol. 34, pp. 2346–2358, Curran Associates, Inc., 2021.
  • [14] X. Liu, H. Peng, N. Zheng, Y. Yang, H. Hu, and Y. Yuan, “Efficientvit: Memory efficient vision transformer with cascaded group attention,” in 2023 IEEE/CVF Conference on Computer Vision and Pattern Recognition (CVPR), pp. 14420–14430, 2023.
  • [15] A. Vaswani, N. Shazeer, N. Parmar, J. Uszkoreit, L. Jones, A. N. Gomez, L. u. Kaiser, and I. Polosukhin, “Attention is all you need,” in Advances in Neural Information Processing Systems, vol. 30, Curran Associates, Inc., 2017.
  • [16] K. Cho, B. van Merrienboer, Çaglar Gülçehre, D. Bahdanau, F. Bougares, H. Schwenk, and Y. Bengio, “Learning phrase representations using rnn encoder–decoder for statistical machine translation,” in Conference on Empirical Methods in Natural Language Processing, 2014.
  • [17] A. Dosovitskiy, L. Beyer, A. Kolesnikov, D. Weissenborn, X. Zhai, T. Unterthiner, M. Dehghani, M. Minderer, G. Heigold, S. Gelly, J. Uszkoreit, and N. Houlsby, “An image is worth 16x16 words: Transformers for image recognition at scale,” ICLR, 2021.
  • [18] T. B. Brown, B. Mann, N. Ryder, M. Subbiah, J. Kaplan, P. Dhariwal, A. Neelakantan, P. Shyam, G. Sastry, A. Askell, S. Agarwal, A. Herbert-Voss, G. Krueger, T. Henighan, R. Child, A. Ramesh, D. M. Ziegler, J. Wu, C. Winter, C. Hesse, M. Chen, E. Sigler, M. Litwin, S. Gray, B. Chess, J. Clark, C. Berner, S. McCandlish, A. Radford, I. Sutskever, and D. Amodei, “Language models are few-shot learners,” 2020.
  • [19] L. Dong, N. Yang, W. Wang, F. Wei, X. Liu, Y. Wang, J. Gao, M. Zhou, and H.-W. Hon, Unified language model pre-training for natural language understanding and generation. Red Hook, NY, USA: Curran Associates Inc., 2019.
  • [20] T. Wang, A. Roberts, D. Hesslow, T. L. Scao, H. W. Chung, I. Beltagy, J. Launay, and C. Raffel, “What language model architecture and pretraining objective works best for zero-shot generalization?,” in Proceedings of the 39th International Conference on Machine Learning, vol. 162 of Proceedings of Machine Learning Research, pp. 22964–22984, PMLR, 17–23 Jul 2022.
  • [21] A. Dosovitskiy, G. Ros, F. Codevilla, A. Lopez, and V. Koltun, “Evaluation and metrics,” 2017. https://leaderboard.carla.org/, Last accessed on 2024-2-28.
  • [22] A. Dosovitskiy, G. Ros, F. Codevilla, A. Lopez, and V. Koltun, “Carla: An open urban driving simulator,” in Proceedings of the 1st Annual Conference on Robot Learning, pp. 1–16, 2017.
  • [23] J. Deng, W. Dong, R. Socher, L.-J. Li, K. Li, and L. Fei-Fei, “Imagenet: A large-scale hierarchical image database,” in 2009 IEEE Conference on Computer Vision and Pattern Recognition, pp. 248–255, 2009.