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

Learning Augmentation Policies from
A Model Zoo for Time Series Forecasting

Haochen Yuan1  Xuelin Li2  Yunbo Wang1  Xiaokang Yang1
1MoE Key Lab of Artificial Intelligence, AI Institute, Shanghai Jiao Tong University
2University of Electronic Science and Technology of China
{yuanhaochen,yunbow}@sjtu.edu.cn
Corresponding author: Yunbo Wang (yunbow@sjtu.edu.cn).
Abstract

Time series forecasting models typically rely on a fixed-size training set and treat all data uniformly, which may not effectively capture the specific patterns present in more challenging training samples. To address this issue, we introduce AutoTSAug, a learnable data augmentation method based on reinforcement learning. Our approach begins with an empirical analysis to determine which parts of the training data should be augmented. Specifically, we identify the so-called marginal samples by considering the prediction diversity across a set of pretrained forecasting models. Next, we propose using variational masked autoencoders as the augmentation model and applying the REINFORCE algorithm to transform the marginal samples into new data. The goal of this generative model is not only to mimic the distribution of real data but also to reduce the variance of prediction errors across the model zoo. By augmenting the marginal samples with a learnable policy, AutoTSAug substantially improves forecasting performance, advancing the prior art in this field with minimal additional computational cost.

1 Introduction

The time series forecasting problem has a wide range of applications, and previous literature has extensively explored this task from various perspectives. Recent deep learning approaches, particularly the powerful Transformer model (Vaswani et al., 2017), along with its subsequent models such as AutoFormer (Wu et al., 2021), PatchTST (Nie et al., 2023a), and iTransFormer (Liu et al., 2024), have shown remarkable capabilities in time series forecasting tasks. However, similar to previous learning-based time series forecasting models, these powerful frameworks are typically limited by the availability of training data. Current augmentation methods for time series data primarily rely on fixed-form preprocessing and lack effective (Wen et al., 2021; Cheung & Yeung, 2020), learnable data augmentation techniques for high-quality datasets. Since time series forecasting methods are commonly used in online systems with non-stationary streaming data that presents evolving distributions, we believe that developing effective data augmentation techniques is essential for improving the prediction accuracy and generalization ability of existing models.

Refer to caption
Figure 1: The fundamental idea of our approach involves: (i) identifying marginal data from the entire training set by assessing the prediction diversity within a zoo of pretrained models, and (ii) learning an augmentation model to transform the marginal data through reinforcement learning.

Nevertheless, implementing fixed-form data augmentation techniques without clear guidance can introduce additional noise and lead to increased computational costs. Although learning-based data augmentation has significantly advanced the field of computer vision (Cubuk et al., 2019; Shorten & Khoshgoftaar, 2019), it remains under-explored in the context of time series forecasting. In this work, we propose a novel method named AutoTSAug (see Figure 1), which adaptively identifies time series data that can largely benefit from augmentation and automatically searches for optimal augmentation policies based on these samples using reinforcement learning.

The fundamental idea of our approach is to leverage a zoo of pretrained forecasting models, which are readily available in the current deep learning landscape, to establish the data filtering criterion as well as the reinforcement learning objectives. This is inspired by our empirical findings below. When we analyze the fitting errors of multiple models, we can divide the training samples into two main groups: (i) data exhibiting high diversity in prediction errors across the model zoo, and (ii) data with low diversity in prediction errors across the model zoo. An interesting observation is that models trained on the low-diversity subset typically outperform those trained on the high-diversity subset when evaluated on the same test set. We provide detailed evidence for this result in Section 3.

Why is this the case? We believe that when the prediction results from the model zoo are consistent, it may indicate two scenarios: either the training samples are too simple or they are too difficult for the forecasting task. Conversely, when the predictions on the model zoo are diverse, it suggests that the training process on these data (which we refer to as “marginal samples”) is less stable. Intuitively, the most practical way to improve the training performance is to ensure that it maintains a great learning effect on the easy samples while also improving stability on the marginal samples. Notably, we can find similar learning principles in another field of curriculum learning. Therefore, a straightforward idea is that augmenting marginal samples is more likely to improve model performance, as they offer higher “marginal benefits111Marginal marginal benefit is a concept in microeconomics that describes the additional total revenue generated by increasing product sales by one unit. Here, we use it as an analogy to describe the performance gains from augmenting data by one unit.”. Accordingly, the first contribution of our method is leveraging the model zoo as a diversity criterion to identify marginal samples.

Given the identified marginal samples, the next step is to augment the training set using them as anchor points. Revisiting our preliminary findings discussed above, a straightforward idea emerges: since we can achieve better performance by training the model on low-diversity data, why not transform the marginal samples by incorporating patterns that are more conducive to training while preserving the unique patterns of the marginal data? In other words, we aim for the augmented data to yield more consistent prediction errors on the model zoo. However, because the variance calculated from backtesting the augmented data across multiple forecasting models is not differentiable for the augmentation model (formulated as neural networks), we specifically employ the REINFORCE algorithm (Williams, 1992) to optimize the augmentation policy. This approach allows us to use the variance from the model zoo as the reward function while treating the input tokens of the augmentation model, which takes the form of a variational masked autoencoder (He et al., 2022), as the actions to be optimized.

The contributions of this paper are as follows:

  • We introduce a novel method for learning data augmentation policies for time series forecasting. Our first contribution involves identifying marginal samples by backtesting the training set across a model zoo and using them as anchor points for augmentation. The key intuition is that improving training stability on these marginal samples is crucial for enhancing overall performance.

  • The core of our method is to train a generative sequence model to predict suitable transformations of marginal samples through a REINFORCE algorithm. Unlike existing approaches that focus on generating realistic data points, our method automatically searches for optimal data augmentation policies aimed at minimizing the variance of prediction errors across the model zoo.

  • Our approach shows significant improvements over common heuristic data augmentation methods across a wide range of time series datasets, demonstrating that augmenting marginal samples is a more effective and time-efficient strategy for enhancing model performance.

2 Related Work

2.1 Time Series Forecasting Models

With the remarkable advancements in natural language processing, various sequence-to-sequence methods have emerged to address time series forecasting (Zhang & Yan, 2022; Wang et al., 2022; Cao et al., 2024; Torres et al., 2021). Recurrent neural networks (RNNs) (Che et al., 2018; Sagheer & Kotb, 2019) are widely utilized in this area. However, due to their sequential architecture, RNNs often struggle with long-term time series forecasting. The Transformer-based models (Li et al., 2019; Wu et al., 2021; Zhou et al., 2021; Liu et al., 2021; Zhou et al., 2022; Zeng et al., 2023) have also demonstrated superior performance in long-term time series forecasting. For instance, Autoformer (Wu et al., 2021) enhances the Transformer by employing a deep decomposition architecture that progressively separates the trend and seasonal components throughout the forecasting process. PatchTST (Nie et al., 2023b) vectorizes time series data into patches of a specified size, encoding the resulting sequence of vectors through a Transformer that outputs forecasts of the desired length using an appropriate prediction head. LSTF-Linear (Zeng et al., 2023) was developed to simplify complex time series forecasting problems (Yi et al., 2024) and outperforms many leading Transformers by utilizing a set of remarkably simple one-layer linear models. iTransformer (Liu et al., 2024) adopts components with inverted dimensions and a modified architecture, demonstrating superior performance on multivariate time series data. Considering both efficiency and accuracy, we select iTransformer as the primary base model for our work and use iTransformer, PatchTST, Autoformer, and the original Transformer to construct the so-called model zoo. It is worth noting that recent literature has introduced large foundation models specifically designed for time series forecasting (Das et al., 2024; Jin et al., 2024; Bian et al., 2024). While incorporating these models into the model zoo for our data augmentation method would be an intriguing avenue for exploration, it is not the focus of this work.

2.2 Time Series Augmentation Methods

Various methods for time series augmentation (Demirel & Holz, 2024; Schneider et al., 2024) have been proposed, particularly inspired by image augmentation (Shorten & Khoshgoftaar, 2019), demonstrating excellent performance in time series forecasting tasks. Specifically, Iglesias et al. (2023) proposed a taxonomy of augmentation techniques spanning from simple to advanced approaches. Simple transformation methods involve time, frequency, and magnitude domain augmentation techniques such as slicing (Cao et al., 2020), frequency warping (Cui et al., 2015), and jittering (Flores et al., 2021). Additionally, advanced generative models have been specifically designed to generate realistic time series data, including the GAN-based approaches (such as TimeGAN (Yoon et al., 2019) and Conditional Sig-Wasserstein GAN (Liao et al., 2020)), the VAE-based approaches (such as Conditional VAE (Sohn et al., 2015) and Smoothness-Including Sequential VAE (Li et al., 2020)), and Diffusion-based approaches (Huang et al., 2023). The generated samples can be used for further training of the forecasting models. The third category of data augmentation techniques for time series is the reinforcement learning-based methods (Ko & Ok, 2022; Fu et al., 2022), dynamically scheduling data augmentation operations during reinforcement learning. It is important to note that the first category of methods employs fixed data transformation functions, which may introduce significant noise. While the second and third categories are learning-based, they fail to analyze the characteristics that augmented data should possess or to identify the most effective parts of the data to serve as anchor points for augmentation.

3 Preliminary Findings

Following previous work (Wu et al., 2021; Nie et al., 2023a; Liu et al., 2024), we initially conducted explorations on several classical real-world multivariate time series datasets, including ETT (4444 subsets), Weather and Electricity. A more detailed description of the datasets can be found in the supplementary materials. To evaluate the impact of varying training data on model performance, we conduct the following experiments. We hypothesize that the model zoo can thoroughly analyze the patterns within each sample. To illustrate the intuition behind our proposed method, we pretrain a model zoo that includes iTransformer, Transformer, PatchTST, and LTSF-Linear, and then backtest them on the same training dataset. By calculating the mean squared error (MSE) for each model, we sort the variance of MSE across different models. We then divide the training data into two subsets based on variance: (A) the top 50%percent5050\%50 % of high-variance data and (B) the top 50%percent5050\%50 % of low-variance data, and compare the test results. We train the forecasting model separately using data from groups A and B. Given its superior average performance across the model zoo, we choose iTransformer as the preferred model for these experiments.

Refer to caption
Figure 2: The comparison of the performance of the state-of-the-art forecasting model, iTransformer, on the same test set after training with different datasets. We divide the original training set based on the variance of the prediction error across the pretrained model zoo.

The forecasting results are presented in Figure 2, where a lower MAE indicates more accurate predictions. Our experiments reveal that the choice of sub-training sets significantly impacts the training outcomes. Predictions made using group B as training data consistently outperform those using group A across all datasets, and in some instances, even surpass the performance achieved with the full training set. This novel finding suggests that marginal samples, which typically induce high variance in model zoo evaluations, are critical factors that can negatively affect model performance. As a result, we hypothesize that augmenting these specific samples is a more effective strategy.

Therefore, our data augmentation methods are exclusively focused on these marginal samples. By applying learnable data augmentation to enhance these samples, we aim to help the model better capture their patterns, ultimately leading to improved performance.

4 Method

4.1 Overall Training Pipeline

Inspired by the preliminary results, we can broadly categorize the training data into three groups: easy, hard, and marginal samples. We can partially identify these groups by analyzing the variance in prediction errors across a batch of pretrained forecasting models. The idea is to handle each group individually during the augmentation process. For easy samples, we do not perform augmentation, because repeated playback of data with high similarity and strong regularity does not promote the model to fit more complex samples. For marginal samples, due to the varying ability of different models to learn specific data patterns, only a subset of models can effectively capture these patterns and produce accurate results. In this case, predictions from the model zoo tend to show higher variance. For hard samples, most models consistently make biased predictions with large mean errors but relatively low variance. This can be caused by strong noises or outlier distributions within the dataset, making these samples less suitable for data augmentation.

By leveraging the model zoo as a criterion to identify marginal samples—–where model performances diverge—–we select these marginal samples as anchor points for data augmentation. As a result, our augmentation efforts will focus primarily on the samples with higher variance in the model zoo. Subsequently, the entire training pipeline consists of three stages:

  • Stage A: Train a probabilistic generative model to initialize the neural augmentor, using marginal data and minimizing the reconstruction loss. This model is implemented as a variational masked autoencoder, which takes partially masked time series data as input and generates complete data.

  • Stage B: Perform a REINFORCE algorithm (Williams, 1992) to enable the neural augmentor to generate data beyond merely replicating the original data distribution.

  • Stage C: Train the forecasting model using both the original data and the augmented data. Specifically, we augment the top 50%percent5050\%50 % samples with the highest variance over the model zoo, tripling the size of the original training set.

Refer to caption
(a) The V-MAE training phase
Refer to caption
(b) The REINFORCE training phase
Figure 3: An illustration of the training phases of the neural augmentor. During the V-MAE training phase (a), we optimize the entire probabilistic generative model to approximate the original distribution of the marginal data. In REINFORCE (b), we finetune the prior network in V-MAE, treating it as the policy network, while keeping the parameters of the other V-MAE modules fixed.

4.2 Variational Masked Autoencoder

By using a probabilistic generative model as the neural augmentor, denoted as s^1:L(i)G(s1:L(i),z(i))similar-tosuperscriptsubscript^𝑠:1𝐿𝑖𝐺superscriptsubscript𝑠:1𝐿𝑖superscript𝑧𝑖\hat{s}_{1:L}^{(i)}\sim{G}(s_{1:L}^{(i)},z^{(i)})over^ start_ARG italic_s end_ARG start_POSTSUBSCRIPT 1 : italic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ∼ italic_G ( italic_s start_POSTSUBSCRIPT 1 : italic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT , italic_z start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT ), we can generate an infinite amount of data by learning a transformation function based on the marginal data s1:L(i)superscriptsubscript𝑠:1𝐿𝑖s_{1:L}^{(i)}italic_s start_POSTSUBSCRIPT 1 : italic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_i ) end_POSTSUPERSCRIPT, where i𝑖iitalic_i is the data index and L𝐿Litalic_L is the length of the data sequence. The key is to learn an appropriate distribution of the latent variable z𝑧zitalic_z, balancing the diversity of the augmented data with its similarity to the original data. For simplicity, we omit the data index in the following notations. The initial learning step involves optimizing G𝐺Gitalic_G with a data reconstruction objective, minimizing the divergence between the masked data and the original data distribution.

We design a Variational Masked Autoencoder (V-MAE), as illustrated in Figure 3a, which takes masked time series data m1:Lsubscript𝑚:1𝐿m_{1:L}italic_m start_POSTSUBSCRIPT 1 : italic_L end_POSTSUBSCRIPT as input and outputs the complete corresponding data. The entire architecture contains four modules, including (i) the prior network, which learns the prior distribution of z𝑧zitalic_z based on the masked data, (ii) the posterior module, which learns the posterior distribution of z𝑧zitalic_z based on the original data, (iii) the data encoder, which extracts significant features from m1:Lsubscript𝑚:1𝐿m_{1:L}italic_m start_POSTSUBSCRIPT 1 : italic_L end_POSTSUBSCRIPT, (iv) the decoder, which generates data from the encoding features and the latent variables. These modules are parametrized by θ1:3subscript𝜃:13\theta_{1:3}italic_θ start_POSTSUBSCRIPT 1 : 3 end_POSTSUBSCRIPT and ϕitalic-ϕ\phiitalic_ϕ, respectively. We have

Prior: z~p(m1:L;θ1),similar-to~𝑧𝑝subscript𝑚:1𝐿subscript𝜃1\displaystyle\quad\tilde{z}\sim p(m_{1:L};\theta_{1}),over~ start_ARG italic_z end_ARG ∼ italic_p ( italic_m start_POSTSUBSCRIPT 1 : italic_L end_POSTSUBSCRIPT ; italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) , (1)
Posterior: zq(s1:L;θ2),similar-to𝑧𝑞subscript𝑠:1𝐿subscript𝜃2\displaystyle\quad z\sim q(s_{1:L};\theta_{2}),italic_z ∼ italic_q ( italic_s start_POSTSUBSCRIPT 1 : italic_L end_POSTSUBSCRIPT ; italic_θ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) ,
Encoding & Decoding: s^1:L=Dec(concat(Enc(m1:L;θ3),z);ϕ).subscript^𝑠:1𝐿DecconcatEncsubscript𝑚:1𝐿subscript𝜃3𝑧italic-ϕ\displaystyle\quad\hat{s}_{1:L}=\operatorname{Dec}\left(\operatorname{concat}% \left(\operatorname{Enc}(m_{1:L};\theta_{3}),z\right);\phi\right).over^ start_ARG italic_s end_ARG start_POSTSUBSCRIPT 1 : italic_L end_POSTSUBSCRIPT = roman_Dec ( roman_concat ( roman_Enc ( italic_m start_POSTSUBSCRIPT 1 : italic_L end_POSTSUBSCRIPT ; italic_θ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) , italic_z ) ; italic_ϕ ) .

We draw the latent variables, which control the diversity of the generated data, from parametrized Gaussian distributions. The mean and standard deviation of these distributions are produced by the prior and posterior modules. We minimize the Kullback–Leibler (KL) divergence between the prior and posterior distributions. After this training stage, we replace the posterior z𝑧zitalic_z with the prior z~~𝑧\tilde{z}over~ start_ARG italic_z end_ARG as input to the decoder. The overall objective function is

=𝔼sDss^1:Ls1:L22+β𝒟KL(q(zs1:L)p(z~m1:L)),\displaystyle\mathcal{L}=\mathbb{E}_{s\sim D_{s}}\left\|\hat{s}_{1:L}-s_{1:L}% \right\|_{2}^{2}+\beta\ \mathcal{D}_{KL}\left(q(z\mid s_{1:L})\parallel p(% \tilde{z}\mid m_{1:L})\right),caligraphic_L = blackboard_E start_POSTSUBSCRIPT italic_s ∼ italic_D start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_POSTSUBSCRIPT ∥ over^ start_ARG italic_s end_ARG start_POSTSUBSCRIPT 1 : italic_L end_POSTSUBSCRIPT - italic_s start_POSTSUBSCRIPT 1 : italic_L end_POSTSUBSCRIPT ∥ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_β caligraphic_D start_POSTSUBSCRIPT italic_K italic_L end_POSTSUBSCRIPT ( italic_q ( italic_z ∣ italic_s start_POSTSUBSCRIPT 1 : italic_L end_POSTSUBSCRIPT ) ∥ italic_p ( over~ start_ARG italic_z end_ARG ∣ italic_m start_POSTSUBSCRIPT 1 : italic_L end_POSTSUBSCRIPT ) ) , (2)

where Dssubscript𝐷𝑠D_{s}italic_D start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT is the set of the marginal data obtained by measuring the prediction diversity over the model zoo. In line with previous work, we use the 2subscript2\mathcal{L}_{2}caligraphic_L start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT loss to measure the reconstruction error. Notably, the posterior module is used exclusively to constrain the prior learner and is not utilized in subsequent training stages.

In general, this V-MAE design can be integrated into any encoder-decoder-based time series forecasting architecture. In this work, we specifically adopt the encoder and decoder (implemented as a linear projector) from iTransformer (Liu et al., 2024) as the feature extraction and decoding modules.

4.3 Consistency-Driven REINFORCE with Model Zoo

Our fundamental idea is that merely approximating the distribution of the original data is insufficient, as the aforementioned objective function of V-MAE does not provide clear guidance for transforming the marginal data. This approach can be viewed as broadly spreading the data distribution of the marginal samples in the hope that it will fortuitously cover important patterns in the test set, rather than learning a targeted shift for data augmentation.

An interesting finding from the preliminary experiments is that training the forecasting model on samples with higher prediction error variances across the model zoo leads to better performance. The model zoo, denoted by \mathcal{M}caligraphic_M, consists of K𝐾Kitalic_K forecasting models with pretrained parameters ω1:Ksubscript𝜔:1𝐾\omega_{1:K}italic_ω start_POSTSUBSCRIPT 1 : italic_K end_POSTSUBSCRIPT. This suggests a potential strategy for tuning the neural augmentor — augmenting the dataset with samples that exhibit lower prediction variance across the model zoo. However, a practical challenge is that the “model-zoo variance” is not differentiable and thus cannot be directly used to optimize the augmentor through gradient descent.

To tackle this problem, we propose to finetune the neural augmentor using a REINFORCE algorithm. In this framework, the latent space generated by the prior module in V-MAE serves as the action space, while the variance of the prediction error across the model zoo, evaluated on the augmented data s^1:Lsubscript^𝑠:1𝐿\hat{s}_{1:L}over^ start_ARG italic_s end_ARG start_POSTSUBSCRIPT 1 : italic_L end_POSTSUBSCRIPT, is incorporated as part of the reward function. As illustrated in Figure 3b, we first backtest the generated data s^1:Lsubscript^𝑠:1𝐿\hat{s}_{1:L}over^ start_ARG italic_s end_ARG start_POSTSUBSCRIPT 1 : italic_L end_POSTSUBSCRIPT using each forecasting model within the model zoo. We calculate the model-zoo variance as

Var(s^1:L,)=1K1K(s^L/2:L(ωk)s¯L/2:L)2,s.t.s¯L/2:L=1K1Ks^L/2:Ls^L/2:L(ωk)22,formulae-sequenceVarsubscript^𝑠:1𝐿1𝐾subscriptsuperscript𝐾1superscriptsuperscriptsubscript^𝑠:𝐿2𝐿subscript𝜔𝑘subscriptsuperscript¯𝑠:𝐿2𝐿2s.t.subscriptsuperscript¯𝑠:𝐿2𝐿1𝐾subscriptsuperscript𝐾1superscriptsubscriptdelimited-∥∥subscript^𝑠:𝐿2𝐿superscriptsubscript^𝑠:𝐿2𝐿subscript𝜔𝑘22\begin{split}\operatorname{Var}\big{(}\hat{s}_{1:L},\mathcal{M}\big{)}&=\frac{% 1}{K}\sum^{K}_{1}{\big{(}\hat{s}_{L/2:L}^{\prime}(\omega_{k})-\bar{s}^{\prime}% _{L/2:L}\big{)}^{2}},\\ \text{s.t.}\quad\bar{s}^{\prime}_{L/2:L}&=\frac{1}{K}\sum^{K}_{1}{\|\hat{s}_{L% /2:L}-\hat{s}_{L/2:L}^{\prime}(\omega_{k})\|_{2}^{2}},\\ \end{split}start_ROW start_CELL roman_Var ( over^ start_ARG italic_s end_ARG start_POSTSUBSCRIPT 1 : italic_L end_POSTSUBSCRIPT , caligraphic_M ) end_CELL start_CELL = divide start_ARG 1 end_ARG start_ARG italic_K end_ARG ∑ start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( over^ start_ARG italic_s end_ARG start_POSTSUBSCRIPT italic_L / 2 : italic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) - over¯ start_ARG italic_s end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_L / 2 : italic_L end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , end_CELL end_ROW start_ROW start_CELL s.t. over¯ start_ARG italic_s end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_L / 2 : italic_L end_POSTSUBSCRIPT end_CELL start_CELL = divide start_ARG 1 end_ARG start_ARG italic_K end_ARG ∑ start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ∥ over^ start_ARG italic_s end_ARG start_POSTSUBSCRIPT italic_L / 2 : italic_L end_POSTSUBSCRIPT - over^ start_ARG italic_s end_ARG start_POSTSUBSCRIPT italic_L / 2 : italic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) ∥ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , end_CELL end_ROW (3)

where s^L/2:L(ωk)superscriptsubscript^𝑠:𝐿2𝐿subscript𝜔𝑘\hat{s}_{L/2:L}^{\prime}(\omega_{k})over^ start_ARG italic_s end_ARG start_POSTSUBSCRIPT italic_L / 2 : italic_L end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_ω start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ) represents the prediction results from the k𝑘kitalic_k-th pretrained forecasting model. We then finetune the V-MAE’s prior module through REINFORCE, treating the prior learner as the policy network while keeping the other network parameters fixed. The learning objective is to maximize the following reward function:

r=11+ekf(s^1:L,),s.t.f(s^1:L,)=1s^1:Ls1:L22Var(s^1:L,).formulae-sequence𝑟11superscript𝑒𝑘𝑓subscript^𝑠:1𝐿s.t.𝑓subscript^𝑠:1𝐿1superscriptsubscriptnormsubscript^𝑠:1𝐿subscript𝑠:1𝐿22Varsubscript^𝑠:1𝐿r=\frac{1}{1+e^{-k\cdot f(\hat{s}_{1:L},\mathcal{M})}},\quad\text{s.t.}\quad f% (\hat{s}_{1:L},\mathcal{M})=\frac{1}{\|\hat{s}_{1:L}-s_{1:L}\|_{2}^{2}\cdot% \operatorname{Var}\big{(}\hat{s}_{1:L},\mathcal{M}\big{)}}.italic_r = divide start_ARG 1 end_ARG start_ARG 1 + italic_e start_POSTSUPERSCRIPT - italic_k ⋅ italic_f ( over^ start_ARG italic_s end_ARG start_POSTSUBSCRIPT 1 : italic_L end_POSTSUBSCRIPT , caligraphic_M ) end_POSTSUPERSCRIPT end_ARG , s.t. italic_f ( over^ start_ARG italic_s end_ARG start_POSTSUBSCRIPT 1 : italic_L end_POSTSUBSCRIPT , caligraphic_M ) = divide start_ARG 1 end_ARG start_ARG ∥ over^ start_ARG italic_s end_ARG start_POSTSUBSCRIPT 1 : italic_L end_POSTSUBSCRIPT - italic_s start_POSTSUBSCRIPT 1 : italic_L end_POSTSUBSCRIPT ∥ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⋅ roman_Var ( over^ start_ARG italic_s end_ARG start_POSTSUBSCRIPT 1 : italic_L end_POSTSUBSCRIPT , caligraphic_M ) end_ARG . (4)

Our goal is to generate new samples that not only resemble the original data but also effectively supplement the training set. This approach encourages that the transformations of the marginal data can also be marginal data. In the reward function, a scaled-sigmoid function is employed to minimize the likelihood of rewards clustering around 00 or 1111 controlled by hyperparameter k𝑘kitalic_k. This approach ensures that, despite potential order-of-magnitude differences in reconstruction error and backtest variance within the model zoo, the reward function can learn effectively from these variations. Based on this reward function, the optimization of the policy network can be formulated as follows via gradient ascent, where α𝛼\alphaitalic_α is the learning rate:

θ1θ1+αrθ1logp(z~m1:L;θ1).subscript𝜃1subscript𝜃1𝛼𝑟subscriptsubscript𝜃1𝑝conditional~𝑧subscript𝑚:1𝐿subscript𝜃1\theta_{1}\leftarrow\theta_{1}+\alpha\cdot r\cdot\nabla_{\theta_{1}}\log p% \left(\tilde{z}\mid m_{1:L};\theta_{1}\right).italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ← italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_α ⋅ italic_r ⋅ ∇ start_POSTSUBSCRIPT italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUBSCRIPT roman_log italic_p ( over~ start_ARG italic_z end_ARG ∣ italic_m start_POSTSUBSCRIPT 1 : italic_L end_POSTSUBSCRIPT ; italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) . (5)

To summarize, by using the REINFORCE algorithm to optimize the neural augmentor, we effectively leverage the model zoo’s ability to assess data properties, resulting in the generation of samples with patterns that time series forecasting models can more easily capture.

5 Experiments

5.1 Experimental Setup

Datasets.

Following previous work in model zoo (Liu et al., 2024; Nie et al., 2023a; Zeng et al., 2023; Vaswani et al., 2017), we thoroughly evaluate the proposed AutoTSAug on four real-world datasets that are publicly available, including ETT (4444 subsets), Traffic, Electricity, Weather. Please refer to supplementary materials for more details on these datasets.

Model zoo.

We use a set of pretrained forecasting models, referred to as the model zoo, to establish the diversity criterion for identifying marginal samples and to construct the reward function in the REINFORCE algorithm. The model zoo consists of four models: Transformer (Vaswani et al., 2017), PatchTST (Nie et al., 2023a), LTSF-Linear (Zeng et al., 2023), and iTransformer (Liu et al., 2024). Our experiments demonstrate that AutoTSAug is consistently effective across the model zoo, as it improves the performance of each model when it is used as the final forecasting model.

Augmentation baseline with Gaussian noises.

Inspired by prior literature (Iglesias et al., 2023), we employ a traditional data augmentor based on simple data transformations. It enhances the diversity of the training data by adjusting the controllable variances and means of the added Gaussian noise. As in previous work, this method applies consistent augmentation across all training data.

Table 1: The performance of iTransformer without vs. with data augmentation. We use the same random seed for training in all three scenarios. For AutoTSAug, since it employs a stochastic neural augmentor, we conduct the training process three times, each with a different set of augmentation data. We report the mean and standard deviation of these experiments.
Dataset Original GaussianAugmentor AutoTSAug
MAE MSE MAE MSE MAE MSE
ETTh1 0.405 0.387 0.407±plus-or-minus\pm±0.02 0.392±plus-or-minus\pm±0.01 0.397±plus-or-minus\pm±0.02 0.382±plus-or-minus\pm±0.01
ETTh2 0.350 0.301 0.352±plus-or-minus\pm±0.01 0.307±plus-or-minus\pm±0.01 0.347±plus-or-minus\pm±0.01 0.294±plus-or-minus\pm±0.01
ETTm1 0.377 0.341 0.374±plus-or-minus\pm±0.02 0.340±plus-or-minus\pm±0.02 0.363±plus-or-minus\pm±0.01 0.327±plus-or-minus\pm±0.01
ETTm2 0.272 0.186 0.272±plus-or-minus\pm±0.01 0.187±plus-or-minus\pm±0.00 0.262±plus-or-minus\pm±0.01 0.179±plus-or-minus\pm±0.01
Weather 0.219 0.178 0.227±plus-or-minus\pm±0.01 0.187±plus-or-minus\pm±0.01 0.207±plus-or-minus\pm±0.00 0.170±plus-or-minus\pm±0.00
Electricity 0.239 0.148 0.243±plus-or-minus\pm±0.01 0.150±plus-or-minus\pm±0.00 0.237±plus-or-minus\pm±0.01 0.147±plus-or-minus\pm±0.01
Traffic 0.269 0.392 0.269±plus-or-minus\pm±0.01 0.394±plus-or-minus\pm±0.02 0.268±plus-or-minus\pm±0.01 0.391±plus-or-minus\pm±0.01
Table 2: The impact of AutoTSAug applied to each model in the model zoo. Specifically, we report the average performance of the models trained with three different augmented datasets.
Train Data iTransfomer Transformer PatchTST LTSF-Linear AVG
MAE MSE MAE MSE MAE MSE MAE MSE MAE MSE
ETTh1 (Orig.) 0.405 0.387 0.750 0.884 0.408 0.394 0.396 0.383 0.490 0.512
+AutoTSAug 0.397 0.382 0.684 0.758 0.396 0.379 0.395 0.383 0.468 0.476
Promotion 1.98% 1.30% 8.90% 14.3% 2.94% 3.81% 0.25% 0% 4.49% 7.03%
ETTh2 (Orig.) 0.350 0.301 0.974 1.362 0.343 0.294 0.380 0.329 0.512 0.572
+AutoTSAug 0.347 0.294 0.849 1.134 0.343 0.293 0.378 0.326 0.479 0.512
Promotion 0.86% 2.33% 12.8% 16.7% 0% 0.34% 0.53% 0.91% 6.45% 10.5%
ETTm1 (Orig.) 0.377 0.341 0.477 0.461 0.360 0.321 0.374 0.346 0.397 0.367
+AutoTSAug 0.363 0.327 0.460 0.452 0.361 0.322 0.362 0.329 0.387 0.358
Promotion 3.71% 4.11% 3.56% 1.95% -0.28% -0.31% 3.21% 4.91% 2.52% 2.45%
ETTm2 (Orig.) 0.272 0.186 0.521 0.434 0.260 0.178 0.271 0.182 0.331 0.245
+AutoTSAug 0.262 0.179 0.501 0.411 0.259 0.177 0.264 0.180 0.322 0.237
Promotion 3.68% 2.76% 3.84% 5.30% 0.38% 0.56% 2.58% 1.10% 2.72% 3.27%
Weather (Orig.) 0.219 0.178 0.464 0.409 0.217 0.175 0.255 0.197 0.289 0.240
+AutoTSAug 0.207 0.170 0.368 0.310 0.216 0.175 0.235 0.194 0.257 0.212
Promotion 5.48% 4.49% 20.69% 24.21% 0.46% 0% 7.84% 1.52% 11.07% 11.67%
Elec. (Orig.) 0.239 0.148 0.355 0.261 0.249 0.163 0.278 0.195 0.280 0.192
+AutoTSAug 0.237 0.147 0.350 0.257 0.251 0.165 0.274 0.193 0.277 0.191
Promotion 0.84% 0.68% 1.41% 1.53% -0.80% -1.23% 1.44% 1.03% 1.07% 0.52%
Traffic (Orig.) 0.269 0.392 0.484 0.866 0.276 0.434 0.397 0.650 0.357 0.585
+AutoTSAug 0.268 0.391 0.466 0.837 0.276 0.433 0.387 0.647 0.349 0.577
Promotion 0.37% 0.26% 3.72% 3.35% 0% 0.23% 2.52% 0.46% 2.24% 1.37%
Table 3: The correspondence of the final results with how many marginal samples (with high prediction variance across the model zoo) are used as the augmentation anchor points in AutoTSAug.
Augmented Data ETTh1 ETTh2 ETTm1 ETTm2 Weather
MAE MSE MAE MSE MAE MSE MAE MSE MAE MSE
30% Top-Variance Data 0.400 0.383 0.346 0.295 0.369 0.343 0.265 0.184 0.207 0.170
50% Top-Variance Data 0.397 0.382 0.347 0.294 0.363 0.327 0.262 0.179 0.209 0.173
All Training Data 0.408 0.390 0.351 0.304 0.372 0.337 0.271 0.186 0.224 0.185

5.2 Forecasting Results

Main results.

We employ a fixed lookback length of 96969696 timestamps across all datasets and report the multivariate sequence prediction results with prediction lengths of 96969696 timestamps. Under this setting, we find that iTransformer outperforms other models on average. Consequently, we use iTransformer as the primary forecasting model and compare the results of training iTransformer without and with different augmentation methods. It is noteworthy that both the traditional Gaussian augmentor and AutoTSAug augment the original training set threefold. In most cases in Table 1, the inclusion of random noise data negatively impacts training. In contrast, data augmented by AutoTSAug consistently results in better testing performance across all datasets.

Consistent improvements upon other forecasting models.

In addition to using iTransformer as the default forecasting model in Stage C, we further investigate the potential benefits of AutoTSAug for improving other models from the model zoo. As illustrated in Table 2, AutoTSAug consistently improves the prediction accuracy across various forecasting models.

Few-shot experiments.

Effective data augmentation methods should be able to deal with the training difficulty due to insufficient training data. To explore this, we conduct few-shot experiments using 25%percent2525\%25 % of the original training data. As shown in Figure 4, limited training data significantly reduces the prediction accuracy of the model. However, after applying AutoTSAug, the performance of the model is significantly closer to the performance when we use the full dataset.

Refer to caption
Figure 4: Few-shot experiments: We compare the results on a partial training set (green) and the full training set (blue). After applying AutoTSAug, the model’s performance on the 25%percent2525\%25 % subset is significantly close to the performance on the full dataset.

5.3 Model Analyses

Shall we augment all training samples?

Inspired by preliminary findings, we only augment the training data with higher variance in the model zoo to enhance the model’s learning from marginal samples. Specifically, we augment the top 50%percent5050\%50 % periods with the highest variance. What would be the impact of augmenting more or fewer samples? To investigate this, we conduct the following experiments: augmenting the top 30%percent3030\%30 % high-variance samples, the top 50%percent5050\%50 % high-variance samples, and the entire training set. For consistency, the total amount of augmented data was kept constant at three times the original training set and applied to iTransformer. As shown in Table 3, augmenting the entire dataset directly was less effective than data-dependent augmentation, which is in line with our preliminary findings. Furthermore, augmenting different percentages of high-variance samples yields varying degrees of improvement, demonstrating that our data augmentation model can adaptively enhance marginal samples, thereby improving the performance of the forecasting model.

Ablation study on REINFORCE.

The design of REINFORCE enables our model to optimize the distribution of the latent variable z𝑧zitalic_z based on the backtest results across the model zoo, thereby generating augmented data that balances data diversity and similarity to the original data. Intuitively, this design is inspired by preliminary findings. However, V-MAE can also serve as a complete model before undergoing REINFORCE training. To investigate the significance of the REINFORCE method, we train the iTransformer model using augmented data generated by V-MAE without REINFORCE optimization. As shown in Table 4, REINFORCE improves the prediction accuracy on multiple datasets, demonstrating the significance of the module to the quality of the augmented data.

Table 4: Comparison of the impact of the REINFORCE algorithm on data augmentation. We similarly use augmented data which is three times the amount of the original data to train iTransformer. By comparing the test results, we assess the quality of the augmented data.
REINF. ETTh1 ETTh2 ETTm1 ETTm2 Weather Elec. Traffic
MAE MSE MAE MSE MAE MSE MAE MSE MAE MSE MAE MSE MAE MSE
×\times× 0.405 0.386 0.349 0.300 0.377 0.341 0.272 0.186 0.221 0.182 0.240 0.149 0.269 0.393
\checkmark 0.397 0.382 0.347 0.294 0.363 0.327 0.262 0.179 0.207 0.170 0.237 0.147 0.268 0.391

Computational cost.

The proposed data augmentation method involves training V-MAE and employing the REINFORCE algorithm, and it requires backtesting across multiple models within the model zoo. These processes introduce additional computational overhead compared to basic prediction models. As shown in Table 5, we report the training time for V-MAE (Stage A), REINFORCE (Stage B), and the forecasting model (Stage C). The increased training time (including Stage A and Stage B) is considered acceptable due to the significant performance gains. Besides, it is notably shorter than the time needed for training the forecasting models. We evaluate the computational cost using a single NVIDIA 3090 GPU.

Table 5: Computational cost for each training stage. The total training time required for dataset augmentation, including Stage A and Stage B, is notably shorter than the time needed for training the forecasting models, indicating our method’s reasonable efficiency.
Dataset Stage A: V-MAE Stage B: REINFORCE Stage C: Forecasting
iTransformer PatchTST
ETTh1 2min 4min 9min 14min
Elec. 1h 41min 2h 48min 6h 37min 10h 04min
Traffic 4h 42min 7h 42min 17h 43min 27h 13min

6 Conclusions and Limitations

In this paper, we proposed AutoTSAug, a novel data augmentation method driven by reinforcement learning. There are two contributions in methodology: First, our method automatically identifies the critical training data for augmentation, termed the marginal samples, based on prediction diversity from a set of pretrained forecasting models. Second, AutoTSAug exploits a variational masked autoencoder in conjunction with the REINFORCE algorithm to generate new data from these marginal samples. AutoTSAug significantly boosts forecasting performance while maintaining minimal computational overhead by leveraging a learnable policy to transform the marginal samples.

One unresolved issue in this study is the need for multiple pretrained models. The presence of models with poor performance could potentially impact the quality of the generated data. To address this, we incorporate recent competitive models in the model zoo. Future work will focus on enhancing the robustness of our approach to effectively adapt to an imperfect model zoo.

Acknowledgments

This work was supported by the National Natural Science Foundation of China (Grant No. 62250062, 62106144), the Shanghai Municipal Science and Technology Major Project (Grant No. 2021SHZDZX0102), and the Fundamental Research Funds for the Central Universities.

References

  • Bian et al. (2024) Yuxuan Bian, Xuan Ju, Jiangtong Li, Zhijian Xu, Dawei Cheng, and Qiang Xu. Multi-patch prediction: Adapting llms for time series representation learning. In ICML, 2024.
  • Cao et al. (2024) Defu Cao, Furong Jia, Sercan O Arik, Tomas Pfister, Yixiang Zheng, Wen Ye, and Yan Liu. Tempo: Prompt-based generative pre-trained transformer for time series forecasting. In ICLR, 2024.
  • Cao et al. (2020) Ping Cao, Xinyi Li, Kedong Mao, Fei Lu, Gangmin Ning, Luping Fang, and Qing Pan. A novel data augmentation method to enhance deep neural networks for detection of atrial fibrillation. Biomedical Signal Processing and Control, 56:101675, 2020.
  • Che et al. (2018) Zhengping Che, Sanjay Purushotham, Kyunghyun Cho, David Sontag, and Yan Liu. Recurrent neural networks for multivariate time series with missing values. Scientific Reports, 8(1):6085, 2018.
  • Cheung & Yeung (2020) Tsz-Him Cheung and Dit-Yan Yeung. Modals: Modality-agnostic automated data augmentation in the latent space. In ICLR, 2020.
  • Cubuk et al. (2019) Ekin D Cubuk, Barret Zoph, Dandelion Mane, Vijay Vasudevan, and Quoc V Le. Autoaugment: Learning augmentation strategies from data. In CVPR, pp.  113–123, 2019.
  • Cui et al. (2015) Xiaodong Cui, Vaibhava Goel, and Brian Kingsbury. Data augmentation for deep neural network acoustic modeling. IEEE/ACM Transactions on Audio, Speech, and Language Processing, 23(9):1469–1477, 2015.
  • Das et al. (2024) Abhimanyu Das, Weihao Kong, Rajat Sen, and Yichen Zhou. A decoder-only foundation model for time-series forecasting. In ICML, 2024.
  • Demirel & Holz (2024) Berken Utku Demirel and Christian Holz. Finding order in chaos: A novel data augmentation method for time series in contrastive learning. In NeurIPS, volume 36, 2024.
  • Flores et al. (2021) Anibal Flores, Hugo Tito-Chura, and Honorio Apaza-Alanoca. Data augmentation for short-term time series prediction with deep learning. In Intelligent Computing: Proceedings of the 2021 Computing Conference, pp.  492–506, 2021.
  • Fu et al. (2022) Yuwei Fu, Di Wu, and Benoit Boulet. Reinforcement learning based dynamic model combination for time series forecasting. In AAAI, volume 36, pp.  6639–6647, 2022.
  • He et al. (2022) Kaiming He, Xinlei Chen, Saining Xie, Yanghao Li, Piotr Dollár, and Ross Girshick. Masked autoencoders are scalable vision learners. In CVPR, pp.  16000–16009, 2022.
  • Huang et al. (2023) Hongbin Huang, Minghua Chen, and Xiao Qiao. Generative learning for financial time series with irregular and scale-invariant patterns. In ICLR, 2023.
  • Iglesias et al. (2023) Guillermo Iglesias, Edgar Talavera, Ángel González-Prieto, Alberto Mozo, and Sandra Gómez-Canaval. Data augmentation techniques in time series domain: a survey and taxonomy. Neural Computing and Applications, 35(14):10123–10145, 2023.
  • Jin et al. (2024) Ming Jin, Shiyu Wang, Lintao Ma, Zhixuan Chu, James Y Zhang, Xiaoming Shi, Pin-Yu Chen, Yuxuan Liang, Yuan-Fang Li, Shirui Pan, et al. Time-llm: Time series forecasting by reprogramming large language models. In ICLR, 2024.
  • Ko & Ok (2022) Byungchan Ko and Jungseul Ok. Efficient scheduling of data augmentation for deep reinforcement learning. In NeurIPS, volume 35, pp.  33289–33301, 2022.
  • Li et al. (2020) Longyuan Li, Junchi Yan, Haiyang Wang, and Yaohui Jin. Anomaly detection of time series with smoothness-inducing sequential variational auto-encoder. IEEE Transactions on Neural Networks and Learning Systems, 32(3):1177–1191, 2020.
  • Li et al. (2019) Shiyang Li, Xiaoyong Jin, Yao Xuan, Xiyou Zhou, Wenhu Chen, Yu-Xiang Wang, and Xifeng Yan. Enhancing the locality and breaking the memory bottleneck of transformer on time series forecasting. In NeurIPS, volume 32, 2019.
  • Liao et al. (2020) Shujian Liao, Hao Ni, Lukasz Szpruch, Magnus Wiese, Marc Sabate-Vidales, and Baoren Xiao. Conditional sig-wasserstein gans for time series generation. arXiv preprint arXiv:2006.05421, 2020.
  • Liu et al. (2021) Shizhan Liu, Hang Yu, Cong Liao, Jianguo Li, Weiyao Lin, Alex X Liu, and Schahram Dustdar. Pyraformer: Low-complexity pyramidal attention for long-range time series modeling and forecasting. In ICLR, 2021.
  • Liu et al. (2024) Yong Liu, Tengge Hu, Haoran Zhang, Haixu Wu, Shiyu Wang, Lintao Ma, and Mingsheng Long. itransformer: Inverted transformers are effective for time series forecasting. In ICLR, 2024.
  • Nie et al. (2023a) Yuqi Nie, Nam H. Nguyen, Phanwadee Sinthong, and Jayant Kalagnanam. A time series is worth 64 words: Long-term forecasting with transformers. In ICLR, 2023a.
  • Nie et al. (2023b) Yuqi Nie, Nam H Nguyen, Phanwadee Sinthong, and Jayant Kalagnanam. A time series is worth 64 words: Long-term forecasting with transformers. In ICLR, 2023b.
  • Sagheer & Kotb (2019) Alaa Sagheer and Mostafa Kotb. Time series forecasting of petroleum production using deep lstm recurrent networks. Neurocomputing, 323:203–213, 2019.
  • Schneider et al. (2024) Nora Schneider, Shirin Goshtasbpour, and Fernando Perez-Cruz. Anchor data augmentation. In NeurIPS, volume 36, 2024.
  • Shorten & Khoshgoftaar (2019) Connor Shorten and Taghi M Khoshgoftaar. A survey on image data augmentation for deep learning. Journal of Big Data, 6(1):1–48, 2019.
  • Sohn et al. (2015) Kihyuk Sohn, Honglak Lee, and Xinchen Yan. Learning structured output representation using deep conditional generative models. In NeurIPS, volume 28, 2015.
  • Torres et al. (2021) José F Torres, Dalil Hadjout, Abderrazak Sebaa, Francisco Martínez-Álvarez, and Alicia Troncoso. Deep learning for time series forecasting: a survey. Big Data, 9(1):3–21, 2021.
  • Vaswani et al. (2017) Ashish Vaswani, Noam Shazeer, Niki Parmar, Jakob Uszkoreit, Llion Jones, Aidan N Gomez, Łukasz Kaiser, and Illia Polosukhin. Attention is all you need. In NeurIPS, volume 30, 2017.
  • Wang et al. (2022) Huiqiang Wang, Jian Peng, Feihu Huang, Jince Wang, Junhui Chen, and Yifei Xiao. Micn: Multi-scale local and global context modeling for long-term series forecasting. In ICLR, 2022.
  • Wen et al. (2021) Qingsong Wen, Liang Sun, Fan Yang, Xiaomin Song, Jingkun Gao, Xue Wang, and Huan Xu. Time series data augmentation for deep learning: A survey. In IJCAI, 2021.
  • Williams (1992) Ronald J Williams. Simple statistical gradient-following algorithms for connectionist reinforcement learning. Machine Learning, 8:229–256, 1992.
  • Wu et al. (2021) Haixu Wu, Jiehui Xu, Jianmin Wang, and Mingsheng Long. Autoformer: Decomposition transformers with auto-correlation for long-term series forecasting. In NeurIPS, volume 34, pp.  22419–22430, 2021.
  • Yi et al. (2024) Kun Yi, Qi Zhang, Wei Fan, Shoujin Wang, Pengyang Wang, Hui He, Ning An, Defu Lian, Longbing Cao, and Zhendong Niu. Frequency-domain mlps are more effective learners in time series forecasting. In NeurIPS, volume 36, 2024.
  • Yoon et al. (2019) Jinsung Yoon, Daniel Jarrett, and Mihaela Van der Schaar. Time-series generative adversarial networks. In NeurIPS, volume 32, 2019.
  • Zeng et al. (2023) Ailing Zeng, Muxi Chen, Lei Zhang, and Qiang Xu. Are transformers effective for time series forecasting? In AAAI, volume 37, pp.  11121–11128, 2023.
  • Zhang & Yan (2022) Yunhao Zhang and Junchi Yan. Crossformer: Transformer utilizing cross-dimension dependency for multivariate time series forecasting. In ICLR, 2022.
  • Zhou et al. (2021) Haoyi Zhou, Shanghang Zhang, Jieqi Peng, Shuai Zhang, Jianxin Li, Hui Xiong, and Wancai Zhang. Informer: Beyond efficient transformer for long sequence time-series forecasting. In AAAI, volume 35, pp.  11106–11115, 2021.
  • Zhou et al. (2022) Tian Zhou, Ziqing Ma, Qingsong Wen, Xue Wang, Liang Sun, and Rong Jin. Fedformer: Frequency enhanced decomposed transformer for long-term series forecasting. In ICML, pp.  27268–27286, 2022.