Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
skip to main content
research-article
Open access

ArieL: Adversarial Graph Contrastive Learning

Published: 12 February 2024 Publication History

Abstract

Contrastive learning is an effective unsupervised method in graph representation learning. The key component of contrastive learning lies in the construction of positive and negative samples. Previous methods usually utilize the proximity of nodes in the graph as the principle. Recently, the data-augmentation-based contrastive learning method has advanced to show great power in the visual domain, and some works have extended this method from images to graphs. However, unlike the data augmentation on images, the data augmentation on graphs is far less intuitive and it is much harder to provide high-quality contrastive samples, which leaves much space for improvement. In this work, by introducing an adversarial graph view for data augmentation, we propose a simple but effective method, Adversarial Graph Contrastive Learning (ArieL), to extract informative contrastive samples within reasonable constraints. We develop a new technique called information regularization for stable training and use subgraph sampling for scalability. We generalize our method from node-level contrastive learning to the graph level by treating each graph instance as a super-node. ArieL consistently outperforms the current graph contrastive learning methods for both node-level and graph-level classification tasks on real-world datasets. We further demonstrate that ArieL is more robust in the face of adversarial attacks.

1 Introduction

Contrastive learning is a widely used technique in various graph representation learning tasks. In contrastive learning, the model tries to minimize the distances among positive pairs and maximize the distances among negative pairs in the embedding space [10, 17, 20, 21, 28, 51, 52, 59, 62, 65]. The definition of positive and negative pairs is the key component in contrastive learning. Earlier methods such as DeepWalk [37] and node2vec [9] define positive and negative pairs based on the co-occurrence of node pairs in the random walks. For knowledge graph embedding, it is a common practice to define positive and negative pairs based on translations [3, 15, 29, 40, 53, 55, 60].
Recently, the breakthroughs of contrastive learning in computer vision have inspired some works to apply similar ideas from visual representation learning to graph representation learning. To name a few, Deep Graph Infomax (DGI) [51] extends Deep InfoMax [12] and achieves significant improvements over previous random walk–based methods. Graphical Mutual Information (GMI) [36] uses the same framework as DGI but generalizes the concept of mutual information from vector space to graph domain. Contrastive multi-view graph representation learning (MVGRL) [10] further improves DGI by introducing graph diffusion into the contrastive learning framework. The more recent works often follow the data augmentation–based contrastive learning methods [5, 11], which treat the data-augmented samples from the same instance as positive pairs and different instances as negative pairs. Graph Contrastive Coding (GCC) [38] uses random walks with restart [48] to generate two subgraphs for each node as two data-augmented samples. Graph Contrastive learning with Adaptive augmentation (GCA) [65] introduces an adaptive data augmentation method that perturbs both the node features and edges according to their importance. It is trained in a similar way as the famous visual contrastive learning framework SimCLR [5]. Its preliminary work, which uses uniform random sampling rather than adaptive sampling, is referred to as GRACE [64] in this article. Robinson et al. [39] propose a way to select hard negative samples based on the distances in the embedding space, which they use to obtain high-quality graph embedding. There are also many works [62, 63] systemically studying the data augmentation on the graphs.
However, unlike the rotation and color jitter operations on images, the transformations on graphs, such as edge dropping and feature masking, are far less intuitive to human beings. The data augmentation on the graph could be either too similar to or totally different from the original graph. This, in turn, leads to a crucial question, that is, how to generate a new graph that is hard enough for the model to discriminate from the original one plus also maintain the desired properties.
Inspired by some recent works [13, 16, 24, 26, 47], we introduce adversarial training on graph contrastive learning and propose a new framework called Adversarial GRaph ContrastIvE Learning(ArieL). Through the adversarial attack on both topology and node features, we generate an adversarial sample from the original graph. On the one hand, since the perturbation is under the constraint, the adversarial sample still stays close enough to the original one. On the other hand, the adversarial attack makes sure that the adversarial sample is hard to discriminate from the other view by increasing the contrastive loss. In addition, we propose a new constraint called information regularization, which could stabilize the training of ArieL and prevent collapsing. We bridge the gap between node-level graph contrastive learning and graph-level contrastive learning by treating each graph instance as a super-node in node-level graph contrastive learning. Thus, we make ArieL a universal graph representation learning framework. We demonstrate that the proposed ArieL outperforms the existing graph contrastive learning frameworks in the node classification and graph classification tasks on both real-world graphs and adversarially attacked graphs.
In summary, we make the following contributions.
First, we introduce an adversarial view as a new form of data augmentation in graph contrastive learning, which makes the data augmentation more informative under mild perturbations.
Second, we propose a new technique called information regularization to stabilize the training of adversarial graph contrastive learning by regularizing the mutual information among positive pairs.
Third, we bridge the gap between node-level graph contrastive learning and graph-level contrastive learning and we unify their formulation under our framework.
Finally, we empirically demonstrate that ArieL can achieve better performance and higher robustness compared with previous graph contrastive learning methods.
The rest of the article is organized as follows. Section 2 gives the problem definition of graph representation learning and the preliminaries. Section 3 describes the proposed algorithm. The experimental results are presented in Section 4. After reviewing related work in Section 5, we conclude the article in Section 6.

2 Problem Definition

In this section, we will introduce all the notations used in this article and give a formal definition of our problem. In addition, we briefly introduce the preliminaries of our method.

2.1 Graph Representation Learning

For graph representation learning, let \(G=\lbrace \mathcal {V},\mathcal {E},\mathbf {X}\rbrace\) be an attributed graph, where \(\mathcal {V}=\lbrace v_1,v_2,\dots ,v_n\rbrace\) denotes the set of nodes, \(\mathcal {E}\subseteq \mathcal {V}\times \mathcal {V}\) denotes the set of edges, and \(\mathbf {X}\in \mathbb {R}^{n\times d}\) denotes the feature matrix. Each node \(v_i\) has a d-dimensional feature \(\mathbf {X}[i,:]\) , and all edges are assumed to be unweighted and undirected. We use a binary adjacency matrix \(\mathbf {A}\in \lbrace 0,1\rbrace ^{n\times n}\) to represent the information of nodes and edges, where \(\mathbf {A}[i,j]=1\) if and only if the node pair \((v_i,v_j) \in \mathcal {E}\) . In the following text, we will use \(G=\lbrace \mathbf {A},\mathbf {X}\rbrace\) to represent the graph.
The objective of the graph representation learning is to learn an encoder \(f: \mathbb {R}^{n\times n}\times \mathbb {R}^{n\times d}\rightarrow \mathbb {R}^{n\times d^{\prime }}\) , which maps the nodes in the graph into low-dimensional embeddings. Denote the node embedding matrix \(\mathbf {H}=f(\mathbf {A},\mathbf {X})\) , where \(\mathbf {H}[i,:]\in \mathbb {R}^{d^{\prime }}\) is the embedding for node \(v_i\) . This representation could be used for downstream tasks such as node classification. Based on the node-embedding matrix, we can further obtain the graph embedding through an order-invariant readout function \(R(\cdot)\) , which generates the graph representation as \(R(\mathbf {H})\in \mathbb {R}^{d^{\prime \prime }}\) .

2.2 InfoNCE Loss

InfoNCE loss [49] is the predominant workhorse of the contrastive learning loss, which maximizes the lower bound of the mutual information between two random variables. For each positive pair \((\mathbf {x},\mathbf {x}^+)\) associated with k negative samples of \(\mathbf {x}\) , denoted as \(\lbrace \mathbf {x}_1^-, \mathbf {x}_2^-, \cdots ,\mathbf {x}_k^-\rbrace\) , InfoNCE loss could be written as
\begin{align} L_k = -\log {\left(\frac{g(\mathbf {x},\mathbf {x}^+)}{g(\mathbf {x},\mathbf {x}^+)+\sum \nolimits _{i=1}^kg(\mathbf {x},\mathbf {x}_i^-)}\right)}. \end{align}
(1)
Here, \(g(\cdot)\) is the density ratio with the property that \(g(\mathbf {a},\mathbf {b}) \propto \frac{p(\mathbf {a}|\mathbf {b})}{p(\mathbf {a})}\) , where \(\propto\) stands for proportional to. It has been shown by the authors of [49] that \(-L_k\) actually serves as the lower bound of the mutual information \(I(\mathbf {x};\mathbf {x}^+)\) with
\begin{align} I(\mathbf {x};\mathbf {x}^+)\ge \log {(k)}-L_k. \end{align}
(2)

2.3 Graph Contrastive Learning

We build the proposed method upon the framework of SimCLR [5], which is also the basic framework that GCA [65] and GraphCL [62] are built on.

2.3.1 Node-Level Contrastive Learning.

Given a graph G, two views of the graph \(G_1=\lbrace \mathbf {A}_1,\mathbf {X}_1\rbrace\) and \(G_2=\lbrace \mathbf {A}_2,\mathbf {X}_2\rbrace\) are first generated. This step can be treated as the data augmentation on the original graph, and various augmentation methods can be used herein. We use random edge dropping and feature masking as GCA does. The node-embedding matrix for each graph can be computed as \(\mathbf {H}_1=f(\mathbf {A}_1,\mathbf {X}_1)\) and \(\mathbf {H}_2=f(\mathbf {A}_2,\mathbf {X}_2)\) . The corresponding node pairs in two graph views are the positive pairs and all other node pairs are negative. Define \(\theta (\mathbf {u},\mathbf {v})\) to be the similarity function between vectors \(\mathbf {u}\) and \(\mathbf {v}\) ; in practice, it is usually chosen as the cosine similarity on the projected embedding of each vector, using a two-layer neural network as the projection head. Denote \(\mathbf {u}_i=\mathbf {H}_1[i,:]\) and \(\mathbf {v}_i=\mathbf {H}_2[i,:]\) ; the contrastive loss is defined as
\begin{align} L_{\text{con}}(G_1,G_2) &= \frac{1}{2n}\sum _{i=1}^n(l(\mathbf {u}_i,\mathbf {v}_i)+l(\mathbf {v}_i,\mathbf {u}_i)), \end{align}
(3)
\begin{align} l(\mathbf {u}_i,\mathbf {v}_i) &= -\log {\frac{e^{\theta (\mathbf {u}_i,\mathbf {v}_i)/\tau }}{e^{\theta (\mathbf {u}_i,\mathbf {v}_i)/\tau } +\sum \nolimits _{j\ne i}e^{\theta (\mathbf {u}_i,\mathbf {v}_j)/\tau }+\sum \nolimits _{j\ne i}e^{\theta (\mathbf {u}_i,\mathbf {u}_j)/\tau }}}, \end{align}
(4)
where \(\tau\) is a temperature parameter. \(l(\mathbf {v}_i,\mathbf {u}_i)\) is symmetrically defined by exchanging the variables in \(l(\mathbf {u}_i,\mathbf {v}_i)\) . This loss is basically a variant of InfoNCE loss, which is symmetrically defined instead.

2.3.2 Graph-Level Contrastive Learning.

Graph-level contrastive learning is closer to contrastive learning in the visual domain. For a batch of graphs \(\mathcal {B}=\lbrace G_1,\cdots , G_b\rbrace\) , we obtain the augmentation of each graph as \(\mathcal {B}^+=\lbrace G_1^+,\cdots ,G_b^+\rbrace\) through node dropping, subgraph sampling, edge perturbation, and feature masking as in GraphCL [62]. The loss function is thus defined on these two batches of graphs as
\begin{equation} L_{\text{con}}(\mathcal {B},\mathcal {B}^+) = \mathbb {E}\left[-\log {\frac{e^{\theta (\mathbf {R}_i, \mathbf {R}_i^+)/\tau }}{e^{\theta (\mathbf {R}_i, \mathbf {R}_i^+)/\tau }+\sum \nolimits _{j\ne i}e^{\theta (\mathbf {R}_i, \mathbf {R}_j^+)/\tau }+\sum \nolimits _{j\ne i}e^{\theta (\mathbf {R}_i, \mathbf {R}_j)/\tau }}}\right], \end{equation}
(5)
where \(\mathbf {R}_i = R(\mathbf {H}_i)\) and \(\mathbf {R}_i^+ = R(\mathbf {H}_i^+)\) .
By abuse of notation, we also use \(L_{\text{con}}\) to denote the loss function for graph-level contrastive learning. The actual meaning of \(L_{\text{con}}\) is dependent on the input type, graph or set, in the following text.
Specifically, we notice that a set of graphs with \(G_i=\lbrace \mathbf {A}_i, \mathbf {X}_i\rbrace\) can be combined into one graph as
\begin{equation} G^*=\lbrace \text{block}\_\text{diag}(\mathbf {A}_1,\cdots ,\mathbf {A}_b), \text{Concat}(\mathbf {X}_1, \cdots ,\mathbf {X}_b)\rbrace . \end{equation}
(6)
Under this transformation, graph embedding of \(G_i\) can be treated as the embedding of a super-node in \(G^*\) . This observation helps us bridge the gap between node-level contrastive learning and graph-level contrastive learning; the only difference between them is the granularity of the instance in the contrastive learning loss. Therefore, we can build a universal framework for graph contrastive learning that can be used for both node-level and graph-level downstream tasks.

2.3.3 Graph Encoder.

In principle, our framework could be applied on any graph neural network (GNN) architecture for an encoder as long as it could be attacked. For simplicity, we employ a two-layer Graph Convolutional Network (GCN) [27] for node-level contrastive learning and a three-layer Graph Isomorphism Network (GIN) [58] for graph-level contrastive learning in this work. Define the symmetrically normalized adjacency matrix
\begin{align} \hat{\mathbf {A}} = \tilde{\mathbf {D}}^{-\frac{1}{2}}\tilde{\mathbf {A}}\tilde{\mathbf {D}}^{-\frac{1}{2}}, \end{align}
(7)
where \(\tilde{\mathbf {A}} = \mathbf {A} +\mathbf {I}_n\) is the adjacency matrix with self-connections added and \(\mathbf {I}_n\) is the identity matrix, and \(\tilde{\mathbf {D}}\) is the diagonal degree matrix of \(\tilde{\mathbf {A}}\) with \(\tilde{\mathbf {D}}[i,i] = \sum _j\tilde{\mathbf {A}}[i,j]\) . The two-layer GCN is given as
\begin{align} f(\mathbf {A},\mathbf {X}) = \sigma (\hat{\mathbf {A}}\sigma (\hat{\mathbf {A}}\mathbf {X}\mathbf {W}^{(1)})\mathbf {W}^{(2)}), \end{align}
(8)
where \(\mathbf {W}^{(1)}\) and \(\mathbf {W}^{(2)}\) are the weights of the first and second layer, respectively, and \(\sigma (\cdot)\) is the activation function.
The Graph Isomorphism operator could be defined as
\begin{align} \mathbf {X}^{\prime } = h((\mathbf {A}+\epsilon \mathbf {I})\mathbf {X}), \end{align}
(9)
where \(h(\cdot)\) is a neural network such as multi-layer perceptrons (MLPs) and \(\epsilon\) is a non-negative scalar. A three-layer GIN is the stack of three Graph Isomorphism operators. In this work, \(h(\cdot)\) is a two-layer MLP followed by an activation function and Batch Normalization [14], and \(\epsilon\) is set as 0 for all operators. Use \(\mathbf {X}^{(i)}\) to denote the node embeddings after the i-th operator; the final node embeddings are the concatenation of \(\mathbf {X}^{(i)}\) , \(\mathbf {H}=\text{Concat}(\mathbf {X}^{(i)}|i=1,2,3)\) , and the graph embedding is the concatenation of the node embeddings after mean pooling, \(R(\mathbf {H})=\text{Concat}(\text{Mean}(\mathbf {X}^{(i)})|i=1,2,3)\) .

2.4 Projected Gradient Descent Attack

A Projected Gradient Descent (PGD) attack [32] is an iterative attack method that projects perturbation onto the ball of interest at the end of each iteration. Assuming that the loss \(L(\cdot)\) is a function of the input matrix \(\mathbf {Z} \in \mathbb {R}^{n\times d}\) , at the t-th iteration, the perturbation matrix \(\mathbf {\Delta }_t\in \mathbb {R}^{n\times d}\) under an \(l_{\infty }\) -norm constraint could be written as
\begin{align} \mathbf {\Delta }_{t} = \Pi _{\Vert \mathbf {\Delta }\Vert _{\infty }\le \delta } (\mathbf {\Delta }_{t-1}+ \eta \cdot \text{sgn}(\nabla _{\mathbf {\Delta }}L(\mathbf {Z}+\mathbf {\Delta }_{t-1})), \end{align}
(10)
where \(\eta\) is the step size, \(\text{sgn}(\cdot)\) takes the sign of each element in the input matrix, and \(\Pi _{\Vert \mathbf {\Delta }\Vert _{\infty }\le \delta }\) projects the perturbation onto the \(\delta\) -ball in the \(l_{\infty }\) -norm.

3 Method

In this section, we will first investigate the vulnerability of the graph contrastive learning, then we will spend the remaining section discussing each part of ArieL in detail. Based on the connection we build upon the node-level contrastive learning and graph-level contrastive learning, we will illustrate our method from the perspective of node-level contrastive learning and extend it to the graph level.

3.1 Vulnerability of the Graph Contrastive Learning

Many GNNs are known to be vulnerable to adversarial attacks [2, 66]. Thus, we first investigate the vulnerability of the GNNs trained with the contrastive learning objective in Equation (3). We generate a sequence of 60 graphs by iteratively dropping edges and masking the features. Let \(G_0=G\) ; for the t-th iteration, we generate \(G_t\) from \(G_{t-1}\) by randomly dropping the edges in \(G_{t-1}\) and randomly masking the unmasked features, both with probability \(p=0.03\) . Since \(G_t\) is guaranteed to contain less information than \(G_{t-1}\) , \(G_t\) should be less similar to \(G_0\) than \(G_{t-1}\) on both the graph and node level. Denote the node embeddings of \(G_t\) as \(\mathbf {H}_t\) ; we measure the similarity \(\theta (\mathbf {H}_t[i,:], \mathbf {H}_0[i,:])\) and it is expected that the similarity decreases as the iteration goes on.
We generate the sequences on two datasets, Amazon-Computers and Amazon-Photo [43]. The results are shown in Figure 1. At the 30-th iteration, with \(0.97^{30}=40.10\%\) edges and features left, the average similarity of the positive samples is under 0.5 on Amazon-Photo. At the 60-th iteration, with \(0.97^{60}=16.08\%\) edges and features left, the average similarity drops under 0.2 on both Amazon-Computers and Amazon-Photo. Additionally, starting from the 30-th iteration, the cosine similarity has around 0.3 standard deviation for both datasets, which indicates that a lot of nodes are actually very sensitive to the external perturbations, even if we do not add any adversarial component but just mask out some information. These results demonstrate that the current graph contrastive learning framework is not trained over enough high-quality contrastive samples and is not robust to adversarial attacks.
Fig. 1.
Fig. 1. Average cosine similarity between the node embeddings of the original graph and the perturbed graph; results are on datasets Amazon-Computers and Amazon-Photo. The shaded area represents the standard deviation.
Given this observation, we are motivated to build an adversarial graph contrastive learning framework that could improve the performance and robustness of the previous graph contrastive learning methods. The overview of our framework is shown in Figure 2.
Fig. 2.
Fig. 2. Overview of the proposed ArieL framework. For each iteration, two augmented views are generated from the original graph by data augmentation (purple arrows). Then, an adversarial view is generated (red arrow) from the original graph by maximizing the contrastive loss against one of the augmented views. The similarities of the corresponding nodes (dashed lines) will be penalized by the information regularization if they exceed the estimated upper bound. The objective of ArieL is to minimize the contrastive loss (gray arrows) between the augmented views, the adversarial view, and the corresponding augmented view, and the information regularization. Best viewed in color.

3.2 Adversarial Training

Adversarial training uses the samples generated through the adversarial attack methods to improve the generalization ability and robustness of the original method during training. Although most existing attack frameworks are targeted at supervised learning, it is natural to generalize these methods to contrastive learning by replacing the classification loss with the contrastive loss. The goal of the adversarial attack on graph contrastive learning is to maximize the contrastive loss by adding a small perturbation on the contrastive samples, which can be formulated as
\begin{align} G_{\text{adv}} = \arg \max _{G^{\prime }} L_{\text{con}}(G_1, G^{\prime }), \end{align}
(11)
where \(G^{\prime }=\lbrace \mathbf {A}^{\prime }, \mathbf {X}^{\prime }\rbrace\) is generated from the original graph G, and the change is constrained by the budget \(\Delta _{\mathbf {A}}\) and \(\Delta _{\mathbf {X}}\) as
\begin{align} &\sum _{i,j}|\mathbf {A}^{\prime }[i,j]-\mathbf {A}[i,j]|\le \Delta _{\mathbf {A}}, \end{align}
(12)
\begin{align} &\sum _{i,j}|\mathbf {X}^{\prime }[i,j]-\mathbf {X}[i,j]|\le \Delta _{\mathbf {X}}. \end{align}
(13)
We treat adversarial attacks as one kind of data augmentation. Although we find it effective to make the adversarial attack on one or two augmented views as well, we follow the typical contrastive learning procedure as in SimCLR [5] to make the attack on the original graph in this work. In addition, it does not matter whether \(G_1\) , \(G_2\) , or G is chosen as the anchor for the adversary; each choice works in our framework and it can also be sampled as a third view. In our experiments, we use the PGD attack [32] as our attack method.
We generally follow the method proposed by Xu et al. [57] to make the PGD attack on the graph structure and apply the regular PGD attack method on the node features. Define the supplement of the adjacency matrix as \(\mathbf {\bar{A}}=\mathbf {1}_{n\times n}-\mathbf {I}_n-\mathbf {A}\) , where \(\mathbf {1}_{n\times n}\) is the ones matrix of size \(n\times n\) . The perturbed adjacency matrix can be written as
\begin{align} \mathbf {A}_{\text{adv}} &= \mathbf {A}+ \mathbf {C}\circ \mathbf {L}_{\mathbf {A}}, \end{align}
(14)
\begin{align} \mathbf {C} &= \bar{\mathbf {A}} - \mathbf {A}, \end{align}
(15)
where \(\circ\) is the element-wise product and \(\mathbf {L}_{\mathbf {A}}\in \lbrace 0,1\rbrace ^{n\times n}\) is a symmetric matrix with each element \(\mathbf {L}_{\mathbf {A}}[i,j]\) corresponding to the modification (e.g., add, delete, or no modification) of the edge between the node pair \((v_i,v_j)\) . The perturbation on \(\mathbf {X}\) follows the regular PGD attack procedure and the perturbed feature matrix can be written as
\begin{align} \mathbf {X}_{\text{adv}} = \mathbf {X} + \mathbf {L}_{\mathbf {X}}, \end{align}
(16)
where \(\mathbf {L}_{\mathbf {X}}\in \mathbb {R}^{n\times d}\) is the perturbation on the feature matrix.
For ease of optimization, \(\mathbf {L}_{\mathbf {A}}\) is relaxed to its convex hull \(\tilde{\mathbf {L}}_{\mathbf {A}}\in [0,1]^{n\times n}\) , which satisfies \(\mathcal {S}_{\mathbf {A}}=\lbrace \tilde{\mathbf {L}}_{\mathbf {A}}|\sum \nolimits _{i,j}\tilde{\mathbf {L}}_{\mathbf {A}}\le \Delta _{\mathbf {A}}, \tilde{\mathbf {L}}_{\mathbf {A}}\in [0,1]^{n\times n}\rbrace\) . The constraint on \(\mathbf {L}_{\mathbf {X}}\) can be written as \(\mathcal {S}_{\mathbf {X}}=\lbrace \mathbf {L}_{\mathbf {X}}|\Vert \mathbf {L}_{\mathbf {X}}\Vert _{\infty }\le \delta _{\mathbf {X}}, \mathbf {L}_{\mathbf {X}}\in \mathbb {R}^{n\times d}\rbrace\) , where we directly treat \(\delta _{\mathbf {X}}\) as the constraint on the feature perturbation. In each iteration, we make the updates
\begin{align} \tilde{\mathbf {L}}_{\mathbf {A}}^{(t)} &= \Pi _{\mathcal {S}_{\mathbf {A}}}\left[\tilde{\mathbf {L}}_{\mathbf {A}}^{(t-1)}+\alpha \cdot \mathbf {G}_{\mathbf {A}}^{(t)}\right], \end{align}
(17)
\begin{align} \mathbf {L}_{\mathbf {X}}^{(t)} & =\Pi _{\mathcal {S}_{\mathbf {X}}}\left[\mathbf {L}_{\mathbf {X}}^{(t-1)}+\beta \cdot \text{sgn}(\mathbf {G}_{\mathbf {X}}^{(t)})\right], \end{align}
(18)
where t denotes the current number of iterations, and
\begin{align} \mathbf {G}_{\mathbf {A}}^{(t)}&=\nabla _{\tilde{\mathbf {L}}_{\mathbf {A}}} L_{\text{con}}(G_1,G_{\text{adv}}^{(t-1)}), \end{align}
(19)
\begin{align} \mathbf {G}_{\mathbf {X}}^{(t)}&=\nabla _{\mathbf {L}_{\mathbf {X}}} L_{\text{con}}(G_1,G_{\text{adv}}^{(t-1)}), \end{align}
(20)
denote the gradients of the loss with respect to \(\tilde{\mathbf {L}}_{\mathbf {A}}\) at \(\tilde{\mathbf {L}}_{\mathbf {A}}^{(t-1)}\) and \(\mathbf {L}_{\mathbf {X}}\) at \(\mathbf {L}_{\mathbf {X}}^{(t-1)}\) , respectively. Here, \(G_{\text{adv}}^{(t-1)}\) is defined as \(\lbrace \mathbf {A}+ \mathbf {C}\circ \tilde{\mathbf {L}}_{\mathbf {A}}^{(t-1)}, \mathbf {X} + \mathbf {L}_{\mathbf {X}}^{(t-1)}\rbrace\) . The projection operation \(\Pi _{\mathcal {S}_{\mathbf {X}}}\) simply clips \(\mathbf {L}_{\mathbf {X}}\) into the range \([-\delta _{\mathbf {X}}, \delta _{\mathbf {X}}]\) element-wisely. The projection operation \(\Pi _{\mathcal {S}_{\mathbf {A}}}\) is calculated as
\begin{align} \Pi _{\mathcal {S}_{\mathbf {A}}}(\mathbf {Z}) = {\left\lbrace \begin{array}{ll} P_{[0,1]}[\mathbf {Z}-\mu \mathbf {1}_{n\times n}], & \text{if } \mu \gt 0, \text{and } \sum \nolimits _{i,j}{P_{[0,1]}[\mathbf {Z}-\mu \mathbf {1}_{n\times n}]}=\Delta _{\mathbf {A}},\\ \\ P_{[0,1]}[\mathbf {Z}], & \text{if } \sum \nolimits _{i,j}{P_{[0,1]}[\mathbf {Z}]}\le \Delta _{\mathbf {A}}, \end{array}\right.} \end{align}
(21)
where \(P_{[0,1]}[\mathbf {Z}]\) clips \(\mathbf {Z}\) into the range \([0,1]\) . We use the bisection method [4] to solve the equation \(\sum \nolimits _{i,j} P_{[0,1]}[\mathbf {Z}-\mu \mathbf {1}_{n\times n}]=\Delta _{\mathbf {A}}\) with respect to the dual variable \(\mu\) .
To finally obtain \(\mathbf {L}_{\mathbf {A}}\) from \(\tilde{\mathbf {L}}_{\mathbf {A}}\) , each element is independently sampled from a Bernoulli distribution as \(\mathbf {L}_{\mathbf {A}}[i,j]\sim \text{Bernoulli}(\tilde{\mathbf {L}}_{\mathbf {A}}[i,j])\) . To obtain a symmetric matrix, we only sample the upper triangular part (the elements on the diagonal are known to be 0 in our formulation) and obtain the lower triangular part through transposition.

3.3 Adversarial Graph Contrastive Learning

To assimilate the graph contrastive learning and adversarial training together, we treat the adversarial view \(G_{\text{adv}}\) obtained from Equation (11) as another view of the graph. We define the adversarial contrastive loss as the contrastive loss between \(G_1\) and \(G_{\text{adv}}\) . The adversarial contrastive loss is added to the original contrastive loss in Equation (3), which becomes
\begin{align} L(G_1,G_2, G_{\text{adv}}) = L_{\text{con}}(G_1,G_2)+\epsilon _1 L_{\text{con}}(G_1,G_{\text{adv}}), \end{align}
(22)
where \(\epsilon _1 \gt 0\) is the adversarial contrastive loss coefficient. We further adopt two additional subtleties on top of this basic framework: subgraph sampling and curriculum learning. For each iteration, a subgraph \(G_s\) with a fixed size is first sampled from the original graph G. Then, the data augmentation and adversarial attack are both conducted on this subgraph. The subgraph sampling could avoid the gradient derivation on the whole graph, which will lead to heavy computation on a large network. We also observe that subgraph sampling could increase the randomness of the sample and sometimes boost the performance. To avoid the imbalanced sample on the isolated nodes, we uniformly sample a random set of nodes and then construct the subgraph on top of them. For every T epochs, the adversarial contrastive loss coefficient is multiplied by a weight \(\gamma\) . When \(\gamma \gt 1\) , the portion of the adversarial contrastive loss is gradually increasing and the contrastive learning becomes harder as the training goes on.

3.4 Information Regularization

Adversarial training could effectively improve the model’s robustness to perturbations. Nonetheless, we find that these hard training samples could impose the additional risk of training collapsing, i.e., the model will be located at a bad parameter area at the early stage of the training, assigning higher probability to a highly perturbed sample than a mildly perturbed one. In our experiment, we find that this vanilla adversarial training method may fail to converge in some cases (e.g., Amazon-Photo dataset). To stabilize the training, we add one constraint termed information regularization, whose main goal is to regularize the instance similarity in the feature space.
Data processing inequality [6] states that for three random variables \(\mathbf {Z}_1\) , \(\mathbf {Z}_2\) , and \(\mathbf {Z}_3\in \mathbb {R}^{n\times d^{\prime }}\) , if they satisfy the Markov relation \(\mathbf {Z_1}\rightarrow \mathbf {Z}_2\rightarrow \mathbf {Z}_3\) , then the inequality \(I(\mathbf {Z}_1;\mathbf {Z}_3)\le I(\mathbf {Z}_1;\mathbf {Z}_2)\) holds. As proved by Zhu et al. [65], since the node embeddings of two views \(\mathbf {H}_1\) and \(\mathbf {H}_2\) are conditionally independent given the node embeddings of the original graph \(\mathbf {H}\) , they also satisfy the Markov relation with \(\mathbf {H}_1\rightarrow \mathbf {H}\rightarrow \mathbf {H}_2\) and vice versa. Therefore, we can derive the following properties over their mutual information:
\begin{align} I(\mathbf {H}_1; \mathbf {H}_2) &\le I(\mathbf {H};\mathbf {H}_1), \end{align}
(23)
\begin{align} I(\mathbf {H}_1; \mathbf {H}_2) &\le I(\mathbf {H};\mathbf {H}_2). \end{align}
(24)
In fact, this inequality holds on each node. A sketch of the proof is that the embedding of each node \(v_i\) is determined by all the nodes from its l-hop neighborhood if an l-layer GNN is used as the encoder, and this subgraph composed of its l-hop neighborhood also satisfies the Markov relation. Therefore, we can derive the more strict inequalities:
\begin{align} I(\mathbf {H}_1[i,:]; \mathbf {H}_2[i,:]) &\le I(\mathbf {H}[i,:];\mathbf {H}_1[i,:]), \end{align}
(25)
\begin{align} I(\mathbf {H}_1[i,:]; \mathbf {H}_2[i,:]) &\le I(\mathbf {H}[i,:];\mathbf {H}_2[i,:]). \end{align}
(26)
Since \(-L_{\text{con}}(G_1,G_2)\) is only a lower bound of the mutual information, directly applying the above constraints is har; we only consider the constraints on the density ratio. Using the Markov relation for each node, we give the following theorem.
Theorem 1.
For two graph views \(G_1\) and \(G_2\) independently transformed from the graph G, the density ratio of their node embeddings \(\mathbf {H}_1\) and \(\mathbf {H}_2\) should satisfy \(g(\mathbf {H}_2[i,:], \mathbf {H}_1[i,:])\le g(\mathbf {H}_2[i,:], \mathbf {H}[i,:]),\) and \(g(\mathbf {H}_1[i,:], \mathbf {H}_2[i,:])\le g(\mathbf {H}_1[i,:], \mathbf {H}[i,:])\) , where \(\mathbf {H}\) is the node embeddings of the original graph.
Proof.
Following the Markov relation of each node, we get that
\begin{equation} \begin{split} p(\mathbf {H}_2[i,:]|\mathbf {H}_1[i,:])&=p(\mathbf {H}_2[i,:]|\mathbf {H}[i,:])p(\mathbf {H}[i,:]|\mathbf {H}_1[i,:])\\ &\le p(\mathbf {H}_2[i,:]|\mathbf {H}[i,:]) \end{split} \end{equation}
(27)
and, consequently,
\begin{align} \frac{p(\mathbf {H}_2[i,:]|\mathbf {H}_1[i,:])}{p(\mathbf {H}_2[i,:])}\le \frac{p(\mathbf {H}_2[i,:]|\mathbf {H}[i,:])}{p(\mathbf {H}_2[i,:])}. \end{align}
(28)
Since \(g(\mathbf {a},\mathbf {b}) \propto \frac{p(\mathbf {a}|\mathbf {b})}{p(\mathbf {a})}\) , we get that \(g(\mathbf {H}_2[i,:], \mathbf {H}_1[i,:])\le g(\mathbf {H}_2[i,:], \mathbf {H}[i,:])\) . A similar proof applies to the other inequality. □
Note that \(g(\cdot ,\cdot)\) is symmetric for the two inputs. Thus, we get two upper bounds for \(g(\mathbf {H}_1[i,:], \mathbf {H}_2[i,:])\) . According to the previous definition, \(g(\mathbf {a},\mathbf {b})=e^{\theta (\mathbf {a},\mathbf {b})/\tau }\) , we can simply replace \(g(\cdot , \cdot)\) with \(\theta (\cdot ,\cdot)\) in the inequalities. Then, we combine these two upper bounds into one:
\begin{align} 2\cdot \theta (\mathbf {H}_1[i,:], \mathbf {H}_2[i,:]) \le \theta (\mathbf {H}_2[i,:], \mathbf {H}[i,:])+\theta (\mathbf {H}_1[i,:], \mathbf {H}[i,:]). \end{align}
(29)
This bound intuitively requires the similarity between \(\mathbf {H}_1[i,:]\) and \(\mathbf {H}_2[i,:]\) to be less than the similarity between \(\mathbf {H}[i,:]\) and \(\mathbf {H}_1[i,:]\) or \(\mathbf {H}_2[i,:]\) . Equipped with this upper bound, we define the following information regularization to penalize the higher probability of a less similar contrastive pair:
\begin{align} &d_i = 2\cdot \theta (\mathbf {H}_1[i,:], \mathbf {H}_2[i,:]) -(\theta (\mathbf {H}_2[i,:], \mathbf {H}[i,:])+\theta (\mathbf {H}_1[i,:], \mathbf {H}[i,:])), \end{align}
(30)
\begin{align} &L_{I}(G_1, G_2, G) = \frac{1}{n}\sum \limits _{i=1}^n{\max \lbrace d_i, 0\rbrace }. \end{align}
(31)
Specifically, information regularization could be defined over any three graphs that satisfy the Markov relation. However, for our framework, to save memory and time complexity, we avoid additional sampling and directly ground information regularization on the existing graphs. It is also fine to apply information regularization on G, \(G_1\) and \(G_{\text{adv}}\) or G, \(G_2\) and \(G_{\text{adv}}\) .
The final loss of ArieL can be written as
\begin{equation} L(G_1,G_2, G_{\text{adv}}) = L_{\text{con}}(G_1,G_2)+\epsilon _1 L_{\text{con}}(G_1,G_{\text{adv}}) + \epsilon _2 L_{I}(G_1, G_2, G), \end{equation}
(32)
where \(\epsilon _2 \gt 0\) controls the strength of the information regularization.
The entire algorithm of ArieL is summarized in Algorithm 1.

3.5 Extension to Graph-Level Contrastive Learning

For a batch of graphs \(\mathcal {B}\) and the batch of their augmentation views \(\mathcal {B}^+\) , we aim to generate a batch of adversarial views, which we denote as \(\mathcal {B}_{\text{adv}}\) . Denote the combined graph of each batch as \(G^*\) , \({G^+}^*\) , and \(G_{\text{adv}}^*\) . The objective of adversarial graph contrastive learning on the graph level can be formulated as
\begin{align} \mathcal {B}_{\text{adv}} = \arg &\max _{\mathcal {B}^{\prime }} L_{\text{con}}(\mathcal {B}^+, \mathcal {B}^{\prime }), \end{align}
(33)
\begin{align} \text{subject to } & \sum _{i,j}|{\mathbf {A}^{\prime }}^*[i,j]-{\mathbf {A}}^*[i,j]|\le \Delta _{\mathbf {A}}, \end{align}
(34)
\begin{align} & \sum _{i,j}|{\mathbf {X}^{\prime }}^*[i,j]-{\mathbf {X}}^*[i,j]|\le \Delta _{\mathbf {X}}. \end{align}
(35)
It is worth noting that the constraints we use here are applied on the batch rather than each graph, i.e., we only constrain the total perturbations over all graphs rather than the perturbations on each graph. This can greatly reduce the computational cost in solving the above-constrained maximization problem in that it reduces the number of constraints from twice the batch size to 2. However, it also introduces the additional risk that the perturbations could be severely imbalanced among the graphs in the batch, e.g., a graph is heavily perturbed while others are almost unchanged. In our experiment, we do not observe this problem but it could theoretically happen. A good practice is to start from this simple form and then gradually add constraints to the vulnerable graphs in the batch if the imbalanced perturbations are observed.
During the attack stage, the perturbation matrix \(\mathbf {L}_{\mathbf {A}}\) and its convex hull \(\tilde{\mathbf {L}}_{\mathbf {A}}\) are further subject to the constraints that they should be block diagonal matrices with 0 at position \((i,j)\) if node i and node j are the nodes from two graphs in the batch. This could be easily implemented by using a block diagonal mask to zero out the gradients during the forward propagation:
\begin{align} \mathbf {L}_{\mathbf {A}}&=\text{block}\_\text{diag}(\mathbf {1}_{n_i\times n_i}|i=1\cdots ,b)\circ \mathbf {L}_{\mathbf {A}}, \end{align}
(36)
\begin{align} \tilde{\mathbf {L}}_{\mathbf {A}}&=\text{block}\_\text{diag}(\mathbf {1}_{n_i\times n_i}|i=1\cdots ,b)\circ \tilde{\mathbf {L}}_{\mathbf {A}}, \end{align}
(37)
where \(n_i\) is the number of nodes in the graph \(G_i\) in the batch. With this processing, the projection operation on the adjacency matrix remains the same as in Equation (21). In the case that we need to apply the constraints for each graph in the batch, we just need to apply the projection operation defined in Equation (21) on the adjacency matrix of each graph using the bisection method to solve \(\mu\) for each graph separately. The projection operation on the feature perturbation matrix is not affected on the graph level, which still clips \(\mathbf {L}_{\mathbf {X}}\) into the range \([-\delta _{\mathbf {X}},\delta _{\mathbf {X}}]\) element-wisely.
Information regularization also applies to graph-level contrastive learning, in which we only need to replace the node embedding with the graph embedding in Equation (30). Hence, we can derive the bound atop different views of the same graph in \(\mathcal {B}\) , \(\mathcal {B}^+\) , and \(\mathcal {B}_{\text{adv}}\) :
\begin{align} &d_i = 2\cdot \theta (R(\mathbf {H}^+_i), R(\mathbf {H}_{\text{adv},i})) -(\theta (R(\mathbf {H}^+_i), R(\mathbf {H}_i))+\theta (R(\mathbf {H}_{\text{adv,i}}), R(\mathbf {H}_i))), \end{align}
(38)
\begin{align} &L_{I}(\mathcal {B}, \mathcal {B}^+, \mathcal {B}_{\text{adv}}) = \frac{1}{b}\sum \limits _{i=1}^b{\max \lbrace d_i, 0\rbrace }. \end{align}
(39)
The final loss of ArieL for the graph-level contrastive learning could be written as
\begin{equation} L(\mathcal {B},\mathcal {B}^+, \mathcal {B}_{\text{adv}}) = L_{\text{con}}(\mathcal {B},\mathcal {B}^+)+\epsilon _1 L_{\text{con}}(\mathcal {B}^+,\mathcal {B}_{\text{adv}}) + \epsilon _2 L_{I}(\mathcal {B}, \mathcal {B}^+, \mathcal {B}_{\text{adv}}). \end{equation}
(40)
The graph-level adversarial contrastive learning could also follow the steps outlined in Algorithm 1 for training by simply replacing the input graph with the input batch in loss functions.

4 Experiments

In this section, we conduct empirical evaluations that are designed to answer the following three questions:
RQ1.
How effective is the proposed ArieL in comparison with previous graph contrastive learning methods on the node classification and graph classification tasks?
RQ2.
To what extent does ArieL gain robustness over the attacked graph?
RQ3.
How does each part of ArieL contribute to its performance?
We evaluate our method with the node classification task and graph classification task on real-world graphs and further evaluate the robustness of it with the node classification task on the attacked graphs. The node/graph embeddings are first learned by the proposed ArieL algorithm; then, the embeddings are fixed to perform the classification with a simple classifier trained over it. All our experiments are conducted on the NVIDIA Tesla V100S GPU with 32 G memory.

4.1 Experimental Setup

4.1.1 Datasets.

For node-level contrastive learning, we use eight datasets for the evaluation, including Cora, CiteSeer, Amazon-Computers, Amazon-Photo, Coauthor-CS, Coauthor-Physics, Facebook, and LastFM Asia. Cora and CiteSeer [61] are citation networks, in which nodes represent documents and edges correspond to citations. Amazon-Computers and Amazon-Photo [43] are extracted from the Amazon co-purchase graph. In these graphs, nodes are the goods and they are connected by an edge if they are frequently bought together. Coauthor-CS and Coauthor-Physics [43] are the coauthorship graphs, in which each node is an author and the edge indicates the coauthorship on a paper. Facebook [41] is a page-page graph of verified Facebook pages in which edges correspond to the likes of each other. LastFM Asia [42] is a social network of Asian users; each node represents a user and they are connected via friendship.
For graph-level contrastive learning, we evaluate ArieL on four datasets from the benchmark TUDataset [33], including the biochemical molecules graphs NCI1, PROTEINS, DD, and MUTAG.
Summaries of the dataset1 statistics are presented in Table 1 and Table 2.
Table 1.
DatasetNodesEdgesFeaturesClasses
Cora2,7085,4291,4337
CiteSeer3,3274,7323,7036
Amazon-Computers13,752245,86176710
Amazon-Photo7,650119,0817458
Coauthor-CS18,33381,8946,80515
Coauthor-Physics34,493247,9628,4155
Facebook22,470342,0041284
LastFM Asia7,62455,61212818
Table 1. Node-Level Contrastive Learning Dataset Statistics — Number of Nodes, Edges, Node Feature Dimensions, and Classes
Table 2.
DatasetGraphsNodesDegreeFeaturesClasses
NCI1411029.871.08372
PROTEINS111339.061.8632
DD1178284.322.52892
MUTAG18817.931.1072
Table 2. Graph-Level Contrastive Learning Dataset Statistics — Number of Graphs, Average Number of Nodes and Degree, and Number of Node Feature Dimensions and Classes

4.1.2 Baselines.

We consider seven graph contrastive learning methods for node-level contrastive learning, including DeepWalk [37], DGI [51], Robust DGI (RDGI) [56], GMI [36], MVGRL [10], GRACE [64], and GCA [65]. Since DeepWalk only generates the embeddings for the graph topology, we concatenate the node features to the generated embeddings for evaluation so that the final embeddings can incorporate both topology and attribute information. We also compare our method with two supervised methods: GCN [27] and Graph Attention Network (GAT) [50].
For graph-level contrastive learning, we compare ArieL with the state-of-the-art graph kernel methods, including graphlet kernel (GL), Weisfeiler-Lehman sub-tree kernel (WL) and deep graph kernel (DGK), and recent unsupervised graph representation learning methods, including node2vec [9], sub2vec [1], graph2vec [34], InfoGraph [44], and GraphCL [62].

4.1.3 Evaluation Protocol.

For each dataset, we randomly select \(10\%\) nodes/graphs for training, \(10\%\) nodes/graphs for validation, and the remaining for testing. For contrastive learning methods, a logistic regression classifier is trained to do the node classification over the node embeddings, whereas a support vector machine is trained to do the graph classification over the graph embeddings. Accuracy is used as the evaluation metric.
For node-level contrastive learning, we search each method over 6 different random seeds, including 5 random seeds from our own and the best random seed of GCA on each dataset. For each seed, we evaluate the method on 20 random training-validation-testing dataset splits and report the mean and the standard deviation of the accuracy on the best seed. Specifically, for the supervised learning methods, we abandon the existing splits, for example, on Cora and CiteSeer. Instead, we do a random split before the training and report the results over 20 splits.
For graph-level contrastive learning, we keep the evaluation protocol the same as the setting in [44] and [62], where the experiments are conducted on 5 random seeds, each corresponding to a 10-fold evaluation.
Besides testing on the original, clean graphs, we also evaluate our method on the attacked graphs for node-level contrastive learning. We use Metattack [67] to perform the poisoning attack. Since Metattack is targeted at graph structure only and computationally inefficient on large graphs, we first randomly sample a subgraph of 5,000 nodes. If the number of nodes in the original graph is greater than 5,000, then we randomly mask out \(20\%\) of the node features and use Metattack to perturb \(20\%\) of the edges to generate the final attacked graph. For ArieL, we use the hyperparameters of the best models we obtain on the clean graphs for evaluation. For GCA, we report the performance in our main results for its three variants, GCA-DE, GCA-PR, and GCA-EV, which correspond to the adoption of degree, eigenvector, and PageRank [25, 35] centrality measures, and use the best variant on each dataset for the evaluation on the attacked graphs.

4.2 Hyperparameters

For node-level contrastive learning, we use the same parameters and design choices for ArieL’s network architecture, optimizer, and training scheme as in GRACE and GCA on each dataset. However, we find that GCA does not behave well on Cora, with a significant performance drop. Thus, we re-search the parameters for GCA on Cora separately and use a different temperature for it. Other contrastive learning–specific parameters are kept the same over GRACE, GCA, and ArieL. On graph-level contrastive learning, we keep ArieL’s hyperparameters the same as the ones used by GraphCL except for its own parameters.
All GNN-based baselines on node-contrastive learning use a two-layer GCN as the encoder. For each method, we compare its default hyperparameters and the ones used by ArieL and use the hyperparameters leading to better performance. Other algorithm-specific hyperparameters all respect the default setting in its official implementation. For graph-level contrastive learning, ArieL uses a three-layer GIN as the encoder, and we take the results for each baseline from its original paper under the same experimental setting.
Other hyperparameters of ArieL are summarized as follows:
Adversarial contrastive loss coefficient \(\epsilon _1\) and information regularization strength \(\epsilon _2\) . We search them over \(\lbrace 0.5, 1, 1.5, 2\rbrace\) and use the one with the best performance on the validation set of each dataset. Specifically, we first fix \(\epsilon _2\) as 0 and decide the optimal value for all other parameters. Then, we search \(\epsilon _2\) on top of the model with other hyperparameters fixed.
Number of attack steps and perturbation constraints. These parameters are fixed on all datasets. For node-level contrastive learning, we set the number of attack steps to 5, edge perturbation constraint \(\Delta _{\mathbf {A}}=0.1\sum _{i,j}\mathbf {A}[i,j],\) and feature perturbation constraint \(\delta _{\mathbf {X}}=0.5\) . For graph-level contrastive learning, we set the number of attack steps to 5, edge perturbation constraint \(\Delta _{\mathbf {A}}=0.05\sum _{i,j}\mathbf {A}[i,j],\) and feature perturbation constraint \(\delta _{\mathbf {X}}=0.04\) .
Curriculum learning weight \(\gamma\) and change period T. In our experiments, we simply fix \(\gamma =1.1\) and \(T=20\) for node-level contrastive learning and \(\gamma =1\) for graph-level contrastive learning.
Graph perturbation rate \(\alpha\) and feature perturbation rate \(\beta\) . We search both over \(\lbrace 0.001, 0.01, 0.1\rbrace\) and take the best one on the validation set of each dataset.
Subgraph size. On node-level contrastive learning, we keep the subgraph size at 500 for ArieL on all datasets except Facebook and LastFM Asia, where we use a subgraph size of 3,000. We do not do the subgraph sampling on graph-level contrastive learning. Instead, we control the batch size b, where we fix \(b=32\) for DD and \(b=128\) for the other three datasets.

4.3 Main Results

The comparison results of node classification on all eight datasets are summarized in Table 3. Our method ArieL outperforms baselines over all datasets except on Cora and Facebook, with results only \(0.11\%\) and \(0.40\%\) lower in accuracy than MVGRL. It can be seen that the previous state-of-the-art method GCA does not bear significant improvements over previous methods. In contrast, ArieL can achieve consistent improvements over GRACE and GCA on all datasets, especially on Amazon-Computers, with almost \(3\%\) gain.
Table 3.
MethodCoraCiteSeerAmazon-ComputersAmazon-PhotoCoauthor-CSCoauthor-PhysicsFacebookLastFM Asia
GCN \(84.14\pm 0.68\) \(69.02\pm 0.94\) \(88.03\pm 1.41\) \(92.65\pm 0.71\) \(92.77\pm 0.19\) \(95.76\pm 0.11\) 89.98 \(\pm\) 0.2683.96 \(\pm\) 0.47
GAT \(83.18\pm 1.17\) \(69.48\pm 1.04\) \(85.52\pm 2.05\) \(91.35\pm 1.70\) \(90.47\pm 0.35\) \(94.82\pm 0.21\) 89.97 \(\pm\) 0.3983.04 \(\pm\) 0.39
DeepWalk \(79.82\pm 0.85\) \(67.14\pm 0.81\) \(86.23\pm 0.37\) \(90.45\pm 0.45\) \(85.02\pm 0.44\) \(94.57\pm 0.20\) 86.67 \(\pm\) 0.2283.93 \(\pm\) 0.61
DGI \(84.24\pm 0.75\) \(69.12\pm 1.29\) \(88.78\pm 0.43\) \(92.57\pm 0.23\) \(92.26\pm 0.12\) \(95.38\pm 0.07\) 89.80 \(\pm\) 0.2782.88 \(\pm\) 0.52
RDGI \(81.84\pm 1.07\) \(65.92\pm 1.26\) \(88.07\pm 0.28\) \(92.17\pm 0.27\) OOMOOMOOM \(77.34\pm 0.69\)
GMI \(82.43\pm 0.90\) \(70.14\pm 1.00\) \(83.57\pm 0.40\) \(88.04\pm 0.59\) OOMOOMOOM74.71 \(\pm\) 0.70
MVGRL84.39 \(\pm\) 0.77 \(69.85\pm 1.54\) \(89.02\pm 0.21\) \(92.92\pm 0.25\) \(92.22\pm 0.22\) \(95.49\pm 0.17\) 90.60 \(\pm\) 0.2883.83 \(\pm\) 0.85
GRACE \(83.40\pm 1.08\) \(69.47\pm 1.36\) \(87.77\pm 0.34\) \(92.62\pm 0.25\) \(93.06\pm 0.08\) \(95.64\pm 0.08\) 88.95 \(\pm\) 0.3179.52 \(\pm\) 0.64
GCA-DE \(82.57\pm 0.87\) \(72.11\pm 0.98\) \(88.10\pm 0.33\) \(92.87\pm 0.27\) \(93.08\pm 0.18\) \(95.62\pm 0.13\) 89.73 \(\pm\) 0.3782.42 \(\pm\) 0.46
GCA-PR \(82.54\pm 0.87\) \(72.16\pm 0.73\) \(88.18\pm 0.39\) \(92.85\pm 0.34\) \(93.09\pm 0.15\) \(95.58\pm 0.12\) 89.68 \(\pm\) 0.3682.44 \(\pm\) 0.51
GCA-EV \(81.80\pm 0.92\) \(67.07\pm 0.79\) \(87.95\pm 0.43\) \(92.63\pm 0.33\) \(93.06\pm 0.14\) \(95.64\pm 0.08\) 89.68 \(\pm\) 0.3882.35 \(\pm\) 0.46
ARIEL84.28 \(\pm\) 0.9672.74 \(\pm\) 1.1091.13 \(\pm\) 0.3494.01 \(\pm\) 0.2393.83 \(\pm\) 0.1495.98 \(\pm\) 0.0590.20 \(\pm\) 0.2384.04 \(\pm\) 0.44
Table 3. Node Classification Accuracy in Percentage on Eight Real-World Datasets
We bold the results with the best mean accuracy. The methods above the line are the supervised ones, and the ones below the line are unsupervised. OOM stands for Out-of-Memory on our 32-G GPUs.
In addition, we find MVGRL a solid baseline whose performance is close to or even better than GCA on these datasets. It achieves the highest score on Cora and Facebook, and the second-highest on Amazon-Computers and Amazon-Photo. However, it does not behave well on CiteSeer, where GCA can effectively increase the score of GRACE. To sum up, previous modifications over the grounded frameworks are mostly based on specific knowledge, for example, MVGRL introduces the diffusion matrix to DGI and GCA defines the importance on the edges and features, and they cannot consistently take effect on all datasets. However, ArieL uses the adversarial attack to automatically construct high-quality contrastive samples and achieves more stable performance improvements.
In comparison with the supervised methods, ArieL also achieves a clear advantage over all of them. Although it would be premature to conclude that ArieL is more powerful than these supervised methods since they are usually tested under the specific training–testing split, these results do demonstrate that ArieL can indeed generate highly expressive node embeddings for the node classification task, which can achieve comparable performance to the supervised methods.
The graph classification results are summarized in Table 4. Compared with our basic framework GraphCL, which uses naïve augmentation methods, ArieL achieves even stronger performance on all datasets. GraphCL does not show a clear advantage against previous baselines such as InfoGraph and it does not behave well on datasets with small graph sizes (e.g., NCI1 and MUTAG). However, ArieL can take the lead on three of the datasets and greatly reduce the performance gap on NCI1 with the graph kernel methods. It can be clearly seen that ArieL behaves better than GraphCL on NCI1 and MUTAG, with at least \(1\%\) improvement in accuracy. In comparison with another graph contrastive learning method, InfoGraph, we can also see that ArieL takes an overall lead on all datasets, even on MUTAG, where InfoGraph shows a dominant advantage against other baselines.
Table 4.
MethodNCI1PROTEINSDDMUTAG
GL81.66 \(\pm\) 2.11
WL80.11 \(\pm\) 0.5072.92 \(\pm\) 0.5680.72 \(\pm\) 3.00
DGK80.31 \(\pm\) 0.4673.30 \(\pm\) 0.8287.44 \(\pm\) 2.72
node2vec54.89 \(\pm\) 1.6157.49 \(\pm\) 3.5772.63 \(\pm\) 10.20
sub2vec52.84 \(\pm\) 1.4753.03 \(\pm\) 5.5561.05 \(\pm\) 15.80
graph2vec73.22 \(\pm\) 1.8173.30 \(\pm\) 2.05-83.15 \(\pm\) 9.25
InfoGraph76.20 \(\pm\) 1.0674.44 \(\pm\) 0.3172.85 \(\pm\) 1.7889.01 \(\pm\) 1.13
GraphCL77.87 \(\pm\) 0.4174.39 \(\pm\) 0.4578.62 \(\pm\) 0.4086.80 \(\pm\) 1.34
ARIEL78.91 \(\pm\) 0.3675.22 \(\pm\) 0.2679.15 \(\pm\) 0.5389.25 \(\pm\) 1.18
Table 4. Graph Classification Accuracy in Percentage on Four Real-World Datasets
We bold the results with the best mean accuracy. The methods above the double line belong to the graph kernel methods, and the ones below the double line are unsupervised representation learning methods. The compared numbers are from the original paper under the same experimental setting.
The above empirical results on the node classification and graph classification tasks clearly demonstrate the advantage of ArieL on real-world graphs, which indicates the better augmentation strategy of ArieL.

4.4 Results Under Attack

The results on attacked graphs are summarized in Table 5. Specifically, we evaluate all these methods on the attacked subgraph of Amazon-Computers, Amazon-Photo, Coauthor-CS, Coauthor-Physics, Facebook, and LastFM Asia. Thus, their results are not directly comparable to the results in Table 3. To compare with the previous results, we look at the datasets in which ArieL takes the lead and then find the performance of the second-best method on each dataset for both the original graph and the attacked one. If ArieL outperforms the second-best method by a much larger margin on the attacked graph compared with that on the original graph, we claim that ArieL is significantly robust on that dataset.
Table 5.
MethodCoraCiteSeerAmazon-ComputersAmazon-PhotoCoauthor-CSCoauthor-PhysicsFacebookLastFM Asia
GCN \(80.03\pm 0.91\) \(62.98\pm 1.20\) \(84.10\pm 1.05\) \(91.72\pm 0.94\) \(80.32\pm 0.59\) \(87.47\pm 0.38\) 70.07 \(\pm\) 0.7473.22 \(\pm\) 0.85
GAT \(79.49\pm 1.29\) \(63.30\pm 1.11\) \(81.60\pm 1.59\) \(90.66\pm 1.62\) \(77.75\pm 0.80\) \(86.65\pm 0.41\) 72.02 \(\pm\) 0.7873.21 \(\pm\) 0.64
DeepWalk \(74.12\pm 1.02\) \(63.20\pm 0.80\) \(79.08\pm 0.67\) \(88.06\pm 0.41\) \(49.30\pm 1.23\) \(79.26\pm 1.38\) 59.07 \(\pm\) 1.0167.61 \(\pm\) 0.80
DGI \(80.84\pm 0.82\) \(64.25\pm 0.96\) \(83.36\pm 0.55\) \(91.27\pm 0.29\) \(78.73\pm 0.50\) \(85.88\pm 0.37\) 70.52 \(\pm\) 0.9371.80 \(\pm\) 0.59
RDGI \(77.29\pm 1.01\) \(59.94\pm 1.29\) \(82.35\pm 0.59\) \(90.63\pm 0.41\) \(83.09\pm 0.64\) \(83.58\pm 0.75\) \(67.85\pm 1.19\) \(63.59\pm 0.91\)
GMI79.17 \(\pm\) 0.76 \(65.37\pm 1.03\) \(77.42\pm 0.59\) \(89.44\pm 0.47\) \(80.92\pm 0.64\) \(87.72\pm 0.45\) 68.93 \(\pm\) 0.8358.89 \(\pm\) 0.95
MVGRL80.90 \(\pm\) 0.75 \(64.81\pm 1.53\) \(83.76\pm 0.69\) \(91.76\pm 0.44\) \(79.49 \pm 0.75\) \(86.98\pm 0.61\) 71.76 \(\pm\) 0.6973.42 \(\pm\) 1.11
GRACE \(78.55\pm 0.81\) \(63.17\pm 1.81\) \(84.74\pm 1.13\) \(91.26\pm 0.37\) \(80.61\pm 0.63\) \(85.71\pm 0.38\) 71.97 \(\pm\) 0.9869.39 \(\pm\) 0.63
GCA \(76.79\pm 0.97\) \(64.89\pm 1.33\) \(85.05\pm 0.51\) \(91.71\pm 0.34\) \(82.72\pm 0.58\) \(89.00\pm 0.31\) 69.54 \(\pm\) 0.8271.83 \(\pm\) 1.03
ARIEL80.33 \(\pm\) 1.2569.13 \(\pm\) 0.9488.61 \(\pm\) 0.4692.99 \(\pm\) 0.2184.43 \(\pm\) 0.5989.09 \(\pm\) 0.3171.15 \(\pm\) 1.1973.94 \(\pm\) 0.78
Table 5. Node Classification Accuracy in Percentage on the Graphs Under Metattack, where Subgraphs of Amazon-Computers, Amazon-Photo, Coauthor-CS and Coauthor-Physics, Facebook and LastFM Asia are used for Attack and Their Results are Not Directly Comparable to Those in Table 3
We bold the results with the best mean accuracy. GCA is evaluated on its best variant on each clean graph.
Under this principle, we can see that ArieL is significantly robust on CiteSeer, with the margin to the second-best method increasing from \(0.58\%\) to \(3.96\%\) , Amazon-Computers, with the margin increasing from \(2.11\%\) to \(3.56\%\) , and Coauthor-CS, with the margin increasing from \(0.74\%\) to \(1.71\%\) . On Coauthor-Physics, ArieL and GCA both show clear robustness against the remaining methods. Although some baselines are robust on specific datasets, for example, MVGRL on Cora, GMI on CiteSeer, GCN on Facebook, and GCA on Coauthor-CS and Coauthor-Physics, they fail to achieve consistent robustness over all datasets. Although GCA indeed makes GRACE more robust for most datasets, it is still vulnerable on Cora, CiteSeer, and Amazon-Computers, with more than \(3\%\) lower than ArieL in the final accuracy.
We can also see that ArieL still shows high robustness on the datasets in which it cannot take the lead. On Cora and Facebook, ArieL is only less than \(1\%\) lower in accuracy than the best method and it is still better than most baselines. It does not show a sudden performance drop on any dataset, such as MVGRL on CiteSeer and GCA on Facebook.
Basically, MVGRL and GCA can improve the robustness of their respective grounded frameworks over specific datasets. However, we find this kind of improvement to be relatively minor. Instead, ArieL has more significant improvements and greatly increases robustness. It is worth noting that though RDGI is also developed to improve the robustness of graph representation learning, it does not show a clear advantage against DGI in our evaluation. This is mainly because the original RDGI considers the attack at test time and what we evaluate is the robustness against the attack at training time, which is more common for the graph learning tasks [2, 66, 67]. Based on the comparative results, we claim that ArieL is more robust than previous graph contrastive learning methods in the face of an adversarial attack.

4.5 Ablation Study

For this section, we first set \(\epsilon _2\) as 0 and investigate the role of adversarial contrastive loss. The adversarial contrastive loss coefficient \(\epsilon _1\) controls the portion of the adversarial contrastive loss in the final loss. When \(\epsilon _1=0\) , the final loss reduces to the regular contrastive loss in Equation (3). To explore the effect of the adversarial contrastive loss, we fix other parameters in our best models on Cora and CiteSeer and gradually increase \(\epsilon _1\) from 0 to 2. The changes in the final performance are shown in Figure 3.
Fig. 3.
Fig. 3. Effect of adversarial contrastive loss coefficient \(\epsilon _1\) on Cora and CiteSeer. The dashed line represents the performance of GRACE with subgraph sampling.
The dashed line represents the performance of GRACE with subgraph sampling, i.e., \(\epsilon _1=0\) . Although there exist some variations, ArieL is always above the baseline under a positive \(\epsilon _1\) with around \(2\%\) improvement. The subgraph sampling trick may sometimes help the model; for example, it improves GRACE without subgraph sampling by \(1\%\) on CiteSeer. However, it could be detrimental as well, such as on Cora. This is understandable since subgraph sampling can simultaneously enrich data augmentation and lessen the number of negative samples, both critical to contrastive learning. At the same time, for the adversarial contrastive loss, it has a stable and significant improvement on GRACE with subgraph sampling, which demonstrates that the performance improvement of ArieL mainly stems from the adversarial loss rather than the subgraph sampling.
Next, we fix all other parameters and check the behavior of \(\epsilon _2\) . Information regularization is mainly designed to stabilize the training of ArieL. We find that ArieL would experience collapsing at the early training stage and that information regularization could mitigate this issue. We choose the best run on Amazon-Photo, where the collapsing frequently occurs, and similar to before, we gradually increase \(\epsilon _2\) from 0 to 2. The results are shown in Figure 4 (left). As can be seen, without using information regularization, ArieL could collapse without learning anything, whereas setting \(\epsilon _2\) greater than 0 can effectively avoid this situation. To further illustrate this, we draw the training curve of the regular contrastive loss in Figure 4 (right), for the best ArieL model on Amazon-Photo and the same model by simply removing the information regularization. Without information regularization, the model could get stuck in a bad parameter area and fail to converge, whereas information regularization can resolve this issue.
Fig. 4.
Fig. 4. Effect of information regularization on Amazon-Photo. The left figure shows the model performance under different \(\epsilon _2\) and the right figure plots the training curve of ArieL under \(\epsilon _2=0\) and \(\epsilon _2=1.0\) .

4.6 Training Analysis

Here, we compare the training of ArieL on node-level contrastive learning to other methods on our NVIDIA Tesla V100S GPU with 32 G memory.
Adversarial attacks on graphs tend to be highly computationally expensive since the attack requires the gradient calculation over the entire adjacency matrix, which is of size \(O(n^2)\) . For ArieL, we resolve this bottleneck with subgraph sampling on large graphs and empirically show that the adversarial training on the subgraph still yields significant improvements without increasing the number of training iterations. In our experiments, we find GMI to be the most memory inefficient, which cannot be trained on Coauthor-CS, Coauthor-Physics, and Facebook. For DGI, MVGRL, GRACE, and GCA, their training also amounts to 30G GPU memory on Coauthor-Physics whereas the training of ArieL requires no more than 8G GPU memory. In terms of the training time, DGI and MVGRL are the fastest to converge. However, it takes MVGRL a long time to compute the diffusion matrix on large graphs. ArieL is slower than GRACE and GCA on Cora and CiteSeer. However, it is faster on large graphs such as Coauthor-CS and Coauthor-Physics, with the training time for each iteration invariant to the graph size due to the subgraph sampling. On the largest graph, Coauthor-Physics, each iteration takes GRACE 0.875 second and GCA 1.264 seconds, while it only takes ArieL 0.082 second. This demonstrates that ArieL has even better scalability than GRACE and GCA.
Subgraph sampling, under some mild assumptions, could be an efficient way to reduce the computational cost for any node-level contrastive learning algorithm. In addition to this general trick, we want to point out that the attack is in fact not always needed on the whole graph to generate the adversarial view. Another solution to avoid explosive memory is to select some anchor nodes and only perturb the edges among these anchor nodes and their features. Since the scalability issue has been resolved by subgraph sampling on all datasets appearing in this work, we will not further discuss the details of this method and empirically prove its effectiveness. We leave this for future work.

5 Related Work

In this section, we review the related work in the following three categories: graph contrastive learning, adversarial attack on graphs, and adversarial contrastive learning.

5.1 Graph Contrastive Learning

Contrastive learning is known for its simplicity and strong expressivity. Traditional methods ground the contrastive samples on the node proximity in the graph, such as DeepWalk [37] and node2vec [9], which use random walks to generate the node sequences and approximate the co-occurrence probability of node pairs. However, these methods can only learn the embeddings for the graph structures regardless of the node features.
GNNs [27, 50] can easily capture the local proximity and node features [19, 22, 51, 62, 63]. To further improve the performance, the Information Maximization (InfoMax) principle [30] has been introduced. DGI [51] is adapted from Deep InfoMax [12] to maximize the mutual information between the local and global features. It generates the negative samples with a corrupted graph and contrasts the node embeddings with the original graph embedding and the corrupted one. Based on a similar idea, GMI [36] generalizes the concept of mutual information to the graph domain by separately defining the mutual information on the features and edges. Graph Community Infomax (GCI) [45] instead tries to maximize the mutual information between the community representation and the node representation for those positive pairs. Another follow-up work of DGI, MVGRL [10], maximizes the mutual information between the first-order neighbors and graph diffusion. On the graph level, InfoGraph [44] makes use of a similar idea to maximize the mutual information between the global representation and patch representation from the same graph. HDI [18] introduces high-order mutual information to consider both intrinsic and extrinsic training signals. However, mutual information-based methods generate the corrupted graphs by simply randomly shuffling the node features. Recent methods exploit the graph topology and features to generate better-augmented graphs. GCC [38] adopts a random walk–based strategy to generate different views of the context graph for a node, but it ignores the augmentation on the feature level. GCA [64], instead, considers the data augmentation from both the topology and feature level, and introduces the adaptive augmentation by considering the importance of each edge and feature dimension. To investigate the power of different data augmentations in graph domains, GraphCL [62] systematically studies the different combinations of graph augmentation strategies and applies them to different graph learning settings. Unlike the above methods, which construct the data augmentation samples based on domain knowledge, ArieL uses an adversarial attack to construct the view that maximizes the contrastive loss, which is more informative with broader applicability.

5.2 Adversarial Attack on Graphs

Deep learning methods are known to be vulnerable to adversarial attacks; this is also the case in the graph domain. As shown by Bojchevski and Günnemann [2], both random walk–based methods and GNN-based methods could be attacked by flipping a small portion of edges. Xu et al. [57] propose a PGD attack and min-max attack on the graph structure from the optimization perspective. NETTACK [66] is the first to attack GNNs using both structure attack and feature attack, causing a significant performance drop of GNNs on the benchmarks. After that, Metattack [67] formulates the poisoning attack of GNNs as a meta-learning problem and achieves remarkable performance by only perturbing the graph structure. Node Injection Poisoning Attacks [46] use a hierarchical reinforcement learning approach to sequentially manipulate the labels and links of the injected nodes. Recently, InfoMax [31] formulated the adversarial attack on GNNs as an influence maximization problem.

5.3 Adversarial Contrastive Learning

The concept of adversarial contrastive learning is first proposed on visual domains [13, 16, 26]. All of these works propose a similar idea to use the adversarial sample as a form of data augmentation in contrastive learning. This can bring better downstream task performance and higher robustness. ACL [26] studies the different paradigms of adversarial contrastive learning by replacing one or two of the augmentation views with the adversarial view generated by PGD attack [32]. CLAE [13] and RoCL [16] use FGSM [8] to generate an additional adversarial view atop the two standard augmentation views. RDGI [56] and AD-GCL [47] are the most relevant work to ours in graph domains. RDGI quantifies the robustness of node representation as the decrease in mutual information between the graph and its embedding under adversarial attacks. It learns a robust node representation by simultaneously minimizing the standard contrastive learning loss and improving the robustness. Nonetheless, its objective sacrifices the expressiveness of the node representation for robustness while ArieL can improve both of them. AD-GCL formulates adversarial graph contrastive learning in a min-max form and uses a parameterized network for edge dropping. However, AD-GCL is designed for the graph classification task only and does not explore the robustness of graph contrastive learning. Finally, all previous adversarial contrastive learning methods do not take scalability into consideration, with visual models and AD-GCL dealing with independent instances and RDGI only working on small graphs, but ArieL can work for both interconnected instances (node embedding) and independent instances (graph embedding) on a large scale.
Some recent theoretical analyses further reveal the vulnerability of contrastive learning. Jing et al. [23] show that dimensional collapse could happen if the variation of the data augmentation exceeds the variation of the data itself in contrastive learning. Wang et al. [54] prove that contrastive learning could cluster the instances from the same class only when the support of different intra-class samples overlaps under data augmentation. The representations learned by contrastive learning may fail in downstream tasks when either under-overlapping or excessive overlapping happens. From these perspectives, searching for adversarial contrastive samples in a safe area is more likely to generate useful representations for downstream tasks.

6 Conclusion

In this article, we propose a universal framework for graph contrastive learning by introducing an adversarial view, scaling it through subgraph sampling, and stabilizing it through information regularization. It consistently outperforms the state-of-the-art graph contrastive learning methods in the node classification and graph classification tasks and exhibits a higher degree of robustness to the adversarial attack. Our framework is not limited to the graph contrastive learning frameworks we build on in this article, and it can be naturally extended to other graph contrastive learning methods as well. In the future, we plan to further investigate (1) the adversarial attack on graph contrastive learning and (2) the integration of graph contrastive learning and supervised methods.

Acknowledgements

The content of the information in this document does not necessarily reflect the position or the policy of the Government or Amazon, and no official endorsement should be inferred. The U.S. Government is authorized to reproduce and distribute reprints for Government purposes notwithstanding any copyright notation hereon.

Footnote

1
All of the datasets are from PyTorch Geometric 2.0.4: https://pytorch-geometric.readthedocs.io/en/latest/modules/datasets.html

References

[1]
Bijaya Adhikari, Yao Zhang, Naren Ramakrishnan, and B. Aditya Prakash. 2018. Sub2Vec: Feature learning for subgraphs. In PAKDD.
[2]
Aleksandar Bojchevski and Stephan Günnemann. 2019. Adversarial Attacks on Node Embeddings via Graph Poisoning. arxiv:1809.01093 [cs.LG]
[3]
Antoine Bordes, Nicolas Usunier, Alberto Garcia-Duran, Jason Weston, and Oksana Yakhnenko. 2013. Translating embeddings for modeling multi-relational data. In Advances in Neural Information Processing Systems, C. J. C. Burges, L. Bottou, M. Welling, Z. Ghahramani, and K. Q. Weinberger (Eds.), Vol. 26. Curran Associates, Inc., 2787–2795. https://proceedings.neurips.cc/paper/2013/file/1cecc7a77928ca8133fa24680a88d2f9-Paper.pdf
[4]
Stephen Boyd and Lieven Vandenberghe. 2004. Convex Optimization. Cambridge University Press.
[5]
Ting Chen, Simon Kornblith, Mohammad Norouzi, and Geoffrey Hinton. 2020. A Simple Framework for Contrastive Learning of Visual Representations. arxiv:2002.05709 [cs.LG]
[6]
Thomas M. Cover and Joy A. Thomas. 2006. Elements of Information Theory, 2nd Edition (Wiley Series in Telecommunications and Signal Processing). Wiley-Interscience.
[7]
Shengyu Feng, Baoyu Jing, Yada Zhu, and Hanghang Tong. 2022. Adversarial Graph Contrastive Learning with Information Regularization.
[8]
Ian J. Goodfellow, Jonathon Shlens, and Christian Szegedy. 2015. Explaining and Harnessing Adversarial Examples. arxiv:1412.6572 [stat.ML]
[9]
Aditya Grover and Jure Leskovec. 2016. node2vec: Scalable Feature Learning for Networks. arxiv:1607.00653 [cs.SI]
[10]
Kaveh Hassani and Amir Hosein Khasahmadi. 2020. Contrastive Multi-View Representation Learning on Graphs. arxiv:2006.05582 [cs.LG]
[11]
Kaiming He, Haoqi Fan, Yuxin Wu, Saining Xie, and Ross Girshick. 2020. Momentum Contrast for Unsupervised Visual Representation Learning. arxiv:1911.05722 [cs.CV]
[12]
R Devon Hjelm, Alex Fedorov, Samuel Lavoie-Marchildon, Karan Grewal, Phil Bachman, Adam Trischler, and Yoshua Bengio. 2019. Learning Deep Representations by Mutual Information Estimation and Maximization. arxiv:1808.06670 [stat.ML]
[13]
Chih-Hui Ho and Nuno Vasconcelos. 2020. Contrastive Learning with Adversarial Examples. arxiv:2010.12050 [cs.CV]
[14]
Sergey Ioffe and Christian Szegedy. 2015. Batch Normalization: Accelerating Deep Network Training by Reducing Internal Covariate Shift.
[15]
Guoliang Ji, Shizhu He, Liheng Xu, Kang Liu, and Jun Zhao. 2015. Knowledge graph embedding via dynamic mapping matrix. In Proceedings of the 53rd Annual Meeting of the Association for Computational Linguistics and the 7th International Joint Conference on Natural Language Processing (Volume 1: Long Papers). Association for Computational Linguistics, Beijing, China, 687–696.
[16]
Ziyu Jiang, Tianlong Chen, Ting Chen, and Zhangyang Wang. 2020. Robust Pre-Training by Adversarial Contrastive Learning. arxiv:2010.13337 [cs.CV]
[17]
Baoyu Jing, Shengyu Feng, Yuejia Xiang, Xi Chen, Yu Chen, and Hanghang Tong. 2021. X-GOAL: Multiplex heterogeneous graph prototypical contrastive learning. (2021).
[18]
Baoyu Jing, Chanyoung Park, and Hanghang Tong. 2021. HDMI: High-order deep multiplex infomax. In Proceedings of the Web Conference 2021. 2414–2424.
[19]
Baoyu Jing, Hanghang Tong, and Yada Zhu. 2021. Network of tensor time series. In Proceedings of the Web Conference 2021. 2425–2437.
[20]
Baoyu Jing, Yuchen Yan, Kaize Ding, Chanyoung Park, Yada Zhu, Huan Liu, and Hanghang Tong. 2023. STERLING: Synergistic representation learning on bipartite graphs. arXiv preprint arXiv:2302.05428 (2023).
[21]
Baoyu Jing, Yuchen Yan, Yada Zhu, and Hanghang Tong. 2022. COIN: Co-cluster infomax for bipartite graphs. In NeurIPS 2022 Workshop: New Frontiers in Graph Learning.
[22]
Baoyu Jing, Zeyu You, Tao Yang, Wei Fan, and Hanghang Tong. 2021. Multiplex graph neural network for extractive text summarization. In Proceedings of the 2021 Conference on Empirical Methods in Natural Language Processing. 133–139.
[23]
Li Jing, Pascal Vincent, Yann LeCun, and Yuandong Tian. 2021. Understanding dimensional collapse in contrastive self-supervised learning. (2021).
[24]
Nikola Jovanović, Zhao Meng, Lukas Faber, and Roger Wattenhofer. 2021. Towards Robust Graph Contrastive Learning. arxiv:2102.13085 [cs.LG]
[25]
Jian Kang, Meijia Wang, Nan Cao, Yinglong Xia, Wei Fan, and Hanghang Tong. 2018. Aurora: Auditing PageRank on large graphs. In 2018 IEEE International Conference on Big Data (Big Data). IEEE, 713–722.
[26]
Minseon Kim, Jihoon Tack, and Sung Ju Hwang. 2020. Adversarial Self-Supervised Contrastive Learning. arxiv:2006.07589 [cs.LG]
[27]
Thomas N. Kipf and Max Welling. 2017. Semi-Supervised Classification with Graph Convolutional Networks. arxiv:1609.02907 [cs.LG]
[28]
Bolian Li, Baoyu Jing, and Hanghang Tong. 2022. Graph communal contrastive learning. In Proceedings of the ACM Web Conference 2022. 1203–1213.
[29]
Yankai Lin, Zhiyuan Liu, Maosong Sun, Yang Liu, and Xuan Zhu. 2015. Learning entity and relation embeddings for knowledge graph completion. In Proceedings of the 29th AAAI Conference on Artificial Intelligence (Austin, Texas) (AAAI’15). AAAI Press, 2181–2187.
[30]
Ralph Linsker. 1988. Self-organization in a perceptual network. Computer 21, 3 (1988), 105–117.
[31]
Jiaqi Ma, Junwei Deng, and Qiaozhu Mei. 2021. Adversarial Attack on Graph Neural Networks as an Influence Maximization Problem. arxiv:2106.10785 [cs.LG]
[32]
Aleksander Madry, Aleksandar Makelov, Ludwig Schmidt, Dimitris Tsipras, and Adrian Vladu. 2019. Towards Deep Learning Models Resistant to Adversarial Attacks. arxiv:1706.06083 [stat.ML]
[33]
Christopher Morris, Nils M. Kriege, Franka Bause, Kristian Kersting, Petra Mutzel, and Marion Neumann. 2020. TUDataset: A collection of benchmark datasets for learning with graphs.
[34]
Annamalai Narayanan, Mahinthan Chandramohan, Rajasekar Venkatesan, Lihui Chen, Yang Liu, and Shantanu Jaiswal. 2017. graph2vec: Learning Distributed Representations of Graphs.
[35]
Lawrence Page, Sergey Brin, Rajeev Motwani, and Terry Winograd. 1999. The PageRank Citation Ranking: Bringing Order to the Web.Technical Report 1999-66. Stanford InfoLab. http://ilpubs.stanford.edu:8090/422/Previous number = SIDL-WP-1999-0120.
[36]
Zhen Peng, Wenbing Huang, Minnan Luo, Qinghua Zheng, Yu Rong, Tingyang Xu, and Junzhou Huang. 2020. Graph Representation Learning via Graphical Mutual Information Maximization. arxiv:2002.01169 [cs.LG]
[37]
Bryan Perozzi, Rami Al-Rfou, and Steven Skiena. 2014. DeepWalk: Online learning of social representations. In Proceedings of the 20th ACM SIGKDD International Conference on Knowledge Discovery and Data Mining (New York, New York, USA) (KDD ’14). ACM, New York, NY, USA, 701–710.
[38]
Jiezhong Qiu, Qibin Chen, Yuxiao Dong, Jing Zhang, Hongxia Yang, Ming Ding, Kuansan Wang, and Jie Tang. 2020. GCC: Graph contrastive coding for graph neural network pre-training. In Proceedings of the 26th ACM SIGKDD International Conference on Knowledge Discovery & Data Mining. 1150–1160.
[39]
Joshua David Robinson, Ching-Yao Chuang, Suvrit Sra, and Stefanie Jegelka. 2021. Contrastive learning with hard negative samples. In International Conference on Learning Representations. https://openreview.net/forum?id=CR1XOQ0UTh-
[40]
Andrea Rossi, Denilson Barbosa, Donatella Firmani, Antonio Matinata, and Paolo Merialdo. 2021. Knowledge graph embedding for link prediction. ACM Transactions on Knowledge Discovery from Data 15, 2 (Apr2021), 1–49.
[41]
Benedek Rozemberczki, Carl Allen, and Rik Sarkar. 2019. Multi-scale Attributed Node Embedding.
[42]
Benedek Rozemberczki and Rik Sarkar. 2020. Characteristic functions on graphs: Birds of a feather, from statistical descriptors to parametric models. (2020).
[43]
Oleksandr Shchur, Maximilian Mumme, Aleksandar Bojchevski, and Stephan Günnemann. 2019. Pitfalls of Graph Neural Network Evaluation. arxiv:1811.05868 [cs.LG]
[44]
Fan-Yun Sun, Jordan Hoffmann, Vikas Verma, and Jian Tang. 2019. InfoGraph: Unsupervised and Semi-supervised Graph-Level Representation Learning via Mutual Information Maximization.
[45]
Heli Sun, Yang Li, Bing Lv, Wujie Yan, Liang He, Shaojie Qiao, and Jianbin Huang. 2021. Graph community infomax. ACM Trans. Knowl. Discov. Data 16, 3, Article 53 (nov2021), 21 pages.
[46]
Yiwei Sun, Suhang Wang, Xianfeng Tang, Tsung-Yu Hsieh, and Vasant Honavar. 2020. Adversarial attacks on graph neural networks via node injections: A hierarchical reinforcement learning approach. In Proceedings of The Web Conference 2020 (Taipei, Taiwan) (WWW ’20). Association for Computing Machinery, New York, NY, USA, 673–683.
[47]
Susheel Suresh, Pan Li, Cong Hao, and Jennifer Neville. 2021. Adversarial Graph Augmentation to Improve Graph Contrastive Learning. arxiv:2106.05819 [cs.LG]
[48]
Hanghang Tong, Christos Faloutsos, and Jia-Yu Pan. 2006. Fast random walk with restart and its applications. In 6th International Conference on Data Mining (ICDM’06). IEEE, 613–622.
[49]
Aaron van den Oord, Yazhe Li, and Oriol Vinyals. 2019. Representation Learning with Contrastive Predictive Coding. arxiv:1807.03748 [cs.LG]
[50]
Petar Veličković, Guillem Cucurull, Arantxa Casanova, Adriana Romero, Pietro Liò, and Yoshua Bengio. 2018. Graph Attention Networks. arxiv:1710.10903 [stat.ML]
[51]
Petar Veličković, William Fedus, William L. Hamilton, Pietro Liò, Yoshua Bengio, and R. Devon Hjelm. 2018. Deep Graph Infomax. arxiv:1809.10341 [stat.ML]
[52]
Haohui Wang, Baoyu Jing, Kaize Ding, Yada Zhu, and Dawei Zhou. 2023. Characterizing long-tail categories on graphs. arXiv preprint arXiv:2305.09938 (2023).
[53]
Ruijie Wang, Yuchen Yan, Jialu Wang, Yuting Jia, Ye Zhang, Weinan Zhang, and Xinbing Wang. 2018. AceKG: A large-scale knowledge graph for academic data mining. In Proceedings of the 27th ACM International Conference on Information and Knowledge Management. 1487–1490.
[54]
Yifei Wang, Qi Zhang, Yisen Wang, Jiansheng Yang, and Zhouchen Lin. 2022. Chaos is a ladder: A new theoretical understanding of contrastive learning via augmentation overlap. In International Conference on Learning Representations. https://openreview.net/forum?id=ECvgmYVyeUz
[55]
Zhen Wang, Jianwen Zhang, Jianlin Feng, and Zheng Chen. 2014. Knowledge graph embedding by translating on hyperplanes. In Proceedings of the 28th AAAI Conference on Artificial Intelligence (Québec City, Québec, Canada) (AAAI’14). AAAI Press, 1112–1119.
[56]
Jiarong Xu, Junru Chen, Yang Yang, Yizhou Sun, Chunping Wang, and Jiangang Lu. 2020. Unsupervised adversarially robust representation learning on graphs. In AAAI Conference on Artificial Intelligence. https://api.semanticscholar.org/CorpusID:227305618
[57]
Kaidi Xu, Hongge Chen, Sijia Liu, Pin-Yu Chen, Tsui-Wei Weng, Mingyi Hong, and Xue Lin. 2019. Topology Attack and Defense for Graph Neural Networks: An Optimization Perspective. arxiv:1906.04214 [cs.LG]
[58]
Keyulu Xu, Weihua Hu, Jure Leskovec, and Stefanie Jegelka. 2018. How Powerful Are Graph Neural Networks?
[59]
Yuchen Yan, Baoyu Jing, Lihui Liu, Ruijie Wang, Jinning Li, Tarek Abdelzaher, and Hanghang Tong. 2023. Reconciling competing sampling strategies of network embedding. In 37th Conference on Neural Information Processing Systems.
[60]
Yuchen Yan, Lihui Liu, Yikun Ban, Baoyu Jing, and Hanghang Tong. 2021. Dynamic knowledge graph alignment. In Proceedings of the AAAI Conference on Artificial Intelligence, Vol. 35. 4564–4572.
[61]
Zhilin Yang, William W. Cohen, and Ruslan Salakhutdinov. 2016. Revisiting Semi-Supervised Learning with Graph Embeddings. arxiv:1603.08861 [cs.LG]
[62]
Yuning You, Tianlong Chen, Yongduo Sui, Ting Chen, Zhangyang Wang, and Yang Shen. 2020. Graph Contrastive Learning with Augmentations. arxiv:2010.13902 [cs.LG]
[63]
Tong Zhao, Yozen Liu, Leonardo Neves, Oliver Woodford, Meng Jiang, and Neil Shah. 2020. Data Augmentation for Graph Neural Networks. arxiv:2006.06830 [cs.LG]
[64]
Yanqiao Zhu, Yichen Xu, Feng Yu, Qiang Liu, Shu Wu, and Liang Wang. 2020. Deep Graph Contrastive Representation Learning. arxiv:2006.04131 [cs.LG]
[65]
Yanqiao Zhu, Yichen Xu, Feng Yu, Qiang Liu, Shu Wu, and Liang Wang. 2021. Graph contrastive learning with adaptive augmentation. Proceedings of the Web Conference 2021 (Apr2021).
[66]
Daniel Zügner, Amir Akbarnejad, and Stephan Günnemann. 2018. Adversarial attacks on neural networks for graph data. Proceedings of the 24th ACM SIGKDD International Conference on Knowledge Discovery and Data Mining (Jul2018).
[67]
Daniel Zügner and Stephan Günnemann. 2019. Adversarial attacks on graph neural networks via meta learning. In International Conference on Learning Representations. https://openreview.net/forum?id=Bylnx209YX

Recommendations

Comments

Information & Contributors

Information

Published In

cover image ACM Transactions on Knowledge Discovery from Data
ACM Transactions on Knowledge Discovery from Data  Volume 18, Issue 4
May 2024
707 pages
EISSN:1556-472X
DOI:10.1145/3613622
Issue’s Table of Contents
This work is licensed under a Creative Commons Attribution International 4.0 License.

Publisher

Association for Computing Machinery

New York, NY, United States

Publication History

Published: 12 February 2024
Online AM: 22 December 2023
Accepted: 13 December 2023
Revised: 29 October 2023
Received: 27 August 2022
Published in TKDD Volume 18, Issue 4

Check for updates

Author Tags

  1. Graph representation learning
  2. contrastive learning
  3. adversarial training
  4. mutual information

Qualifiers

  • Research-article

Funding Sources

  • National Science Foundation
  • National Science Foundation Program on Fairness in AI in collaboration with Amazon
  • DARPA
  • NIFA
  • DHS
  • ARO
  • C3.ai Digital Transformation Institute, MIT-IBM Watson AI Lab, and IBM-Illinois Discovery Accelerator Institute

Contributors

Other Metrics

Bibliometrics & Citations

Bibliometrics

Article Metrics

  • 0
    Total Citations
  • 934
    Total Downloads
  • Downloads (Last 12 months)934
  • Downloads (Last 6 weeks)143
Reflects downloads up to 11 Sep 2024

Other Metrics

Citations

View Options

View options

PDF

View or Download as a PDF file.

PDF

eReader

View online with eReader.

eReader

Get Access

Login options

Full Access

Media

Figures

Other

Tables

Share

Share

Share this Publication link

Share on social media