b

DiscoverSearch
About
My stuff
Learning Adversarially Robust Representations via Worst-Case Mutual Information Maximization
2020·arXiv
Abstract
Abstract

Training machine learning models that are robust against adversarial inputs poses seemingly insurmountable challenges. To better understand adversarial robustness, we consider the underlying problem of learning robust representations. We develop a notion of representation vulnerability that captures the maximum change of mutual information between the input and output distributions, under the worst-case input perturbation. Then, we prove a theorem that establishes a lower bound on the minimum adversarial risk that can be achieved for any downstream classifier based on its representation vulnerability. We propose an unsupervised learning method for obtaining intrinsically robust representations by maximizing the worst-case mutual information between the input and output distributions. Experiments on downstream classification tasks support the robustness of the representations found using unsupervised learning with our training principle.

Machine learning has made remarkable breakthroughs in many fields, including computer vision (He et al., 2016) and natural language processing (Devlin et al., 2019), especially when evaluated on classification accuracy on a given dataset. However, adversarial vulnerability (Szegedy et al., 2014; Engstrom et al., 2017), remains a serious problem that impedes the deployment of the state-of-the-art machine learning models in safety-critical applications, such as autonomous driving (Eykholt et al., 2018) and face recognition (Sharif et al., 2016). Despite extensive efforts to improve model robustness, state-of-the-art adversarially robust training methods (M ˛adry et al., 2018; Zhang et al., 2019) still fail to produce robust models, even for simple classification tasks on CIFAR-10 (Krizhevsky et al., 2009).

In addition to many ineffective empirical attempts for achieving model robustness, recent studies have identified intrinsic difficulties for learning in the presence of adversarial examples. For instance, a line of works (Gilmer et al., 2018; Fawzi et al., 2018; Mahloujifar et al., 2019; Shafahi et al., 2018) proved that adversarial vulnerability is inevitable if the underlying input distribution is concentrated. Schmidt et al. (2018) showed that for certain learning problems, adversarially robust generalization requires more sample complexity compared with standard one, whereas Bubeck et al. (2019) constructed a specific task on which adversarially robust learning is computationally intractable.

Motivated by the apparent empirical and theoretical diffi-culties of robust learning with adversarial examples, we focus on the underlying problem of learning adversarially robust representations (Garg et al., 2018; Pensia et al., 2020). Given an input space  X ⊆ Rd and a feature space  Z ⊆ Rn,any function  g : X → Zis called a representation with respect to (X, Z). Adversarially robust representations denote the set of functions from X to Z that are less sensitive to adversarial perturbations with respect to some metric  ∆defined on X. Note that one can always get an overall clas-sification model by learning a downstream classifier given a representation, thus learning representations that are robust can be viewed as an intermediate step for the ultimate goal of finding adversarially robust models. In this sense, learning adversarially robust representations may help us better understand adversarial examples, and perhaps more importantly, bypass some of the aforementioned intrinsic difficulties for achieving model robustness.

In this paper, we give a general definition for robust representations based on mutual information, then study its implications on model robustness for a downstream clas-sification task. Finally, we propose empirical methods for estimating and inducing representation robustness.

Contributions. Motivated by the empirical success of standard representation learning using the mutual information maximization principle (Bell & Sejnowski, 1995; Hjelm et al., 2018), we first give a formal definition on representa- tion vulnerability as the maximum change of mutual information between the representation’s input and output against adversarial input perturbations bound in an  ∞-Wasserstein ball (Section 3). Under a Gaussian mixture model, we established theoretical connections between the robustness of a given representation and the adversarial gap of the best classifier that can be based on it (Section 3.1). In addition, based on the standard mutual information and the representation vulnerability, we proved a fundamental lower bound on the minimum adversarial risk that can be achieved for any downstream classifiers built upon a representation with given representation vulnerability (Section 3.2).

To further study the implication of robust representations, we first propose a heuristic algorithm to empirically estimate the vulnerability of a given representation (Section 4), and then by adding a regularization term on representation vulnerability in the objective of mutual information maximization principle, provide an unsupervised way for training meaningful and robust representations (Section 5). We observe a direct correlation between model and representation robustness in experiments on benchmark image datasets MINST and CIFAR-10 (Section 6.1). Experiments on downstream classification tasks and saliency maps further show the effectiveness of our proposed training method in obtaining more robust representations (Section 6.2).

Related Work. With similar motivations, several different definitions of robust features have been proposed in literature. The pioneering work of Garg et al. (2018) considered a feature to be robust if it is insensitive to input perturbations in terms of the output values. However, their definition of feature robustness is not invariant to scale changes. Based on the linear correlation between feature outputs and true labels, Ilyas et al. (2019) proposed a definition of robust features to understand adversarial examples, whereas Eykholt et al. (2019) proposed to study robust features whose outputs will not change with respect to small input perturbations. However, these two definitions either require the additional label information or restrict the feature space to be discrete, thus are not general. The most closely related work to ours is Pensia et al. (2020), which considered Fisher information of the output distribution as the indicator of feature robustness and proposed a robust information bottleneck method for extracting robust features. Compared with Pensia et al. (2020), our definition is defined for the worst-case input distribution perturbation, whereas Fisher information can only capture feature’s sensitivity near the input distribution. In addition, our proposed training method for robust representations is better in the sense that it is unsupervised.

Notation. We use small boldface letters such as x to denote vectors and capital letters such as X to denote random variables. Let  (X, ∆)be a metric space, where  ∆ : X ×X → Ris some distance metric. Let P(X) denote the set of all probability measures on  X and δxbe the dirac measure at  x ∈ X.Let  B(x, ϵ, ∆) = {x′ ∈ X : ∆(x′, x) ≤ ϵ}be the ball around x with radius  ϵ. When ∆is free of context, we simply write  B(x, ϵ) = B(x, ϵ, ∆). Denote by  sgn(·)the sign function such that  sgn(x) = 1 if x ≥ 0; sgn(x) = −1 other-wise. Given  f : X → Y and g : Y → Z, define g◦f as theircomposition such that for any  x ∈ X, (g◦f)(x) = g(f(x)).We use [m] to denote {1, 2, . . . , m} and |A| to denote the cardinality of a finite set  A. For any x ∈ Rd, the ℓp-norm ofx is defined as  ∥x∥p = (�i∈[d] xpi )1/pfor any  p ≥ 1. For any  θ ∈ Rd and positive definite matrix  Σ ∈ Rd×d, denoteby  N(θ, Σ)the d-dimensional Gaussian distribution with mean vector  θand covariance matrix  Σ.

This section introduces the main ideas we build upon: mutual information, Wasserstein distance and adversarial risk.

Mutual information. Mutual information is an entropybased measure of the mutual dependence between variables:

Definition 2.1. Let (X, Z) be a pair of random variables with values over the space  X × Z. Themutual information of (X, Z) is defined as:

image

where  pXZis the joint probability density function of (X, Z), and  pX, pZare the marginal probability density functions of X and Z, respectively.

Intuitively, I(X; Z) tells us how well one can predict Z from X (and X from Z, since it is symmetrical). By definition, I(X; Z) = 0 if X and Z are independent; when X and Z are identical, I(X; X) equals to the entropy H(X).

Wasserstein distance. Wasserstein distance is a distance function defined between two probability distributions on a given metric space:

Definition 2.2. Let  (X, ∆)be a metric space with bounded support. Given two probability measures  µ and ν on (X, ∆),the p-th Wasserstein distance, for any  p ≥ 1, is defined as:

image

where  Γ(µ, ν)is the collection of all probability measures on  X × X with µ and νbeing the marginals of the first and second factor, respectively. The p-th Wasserstein ball with respect to  µand radius  ϵ ≥ 0is defined as:

image

The  ∞-Wasserstein distance is defined as the limit of p-th Wasserstein distance,  W∞(µ, ν) = limp→∞ Wp(µ, ν).

Adversarial risk. Adversarial risk captures the vulnerability of a given classification model to input perturbations:

Definition 2.3. Let  (X, ∆)be the input metric space and Y be the set of labels. Let  µXYbe the underlying distribution of the input and label pairs. For any classifier  f : X → Y, the adversarial risk of f with respect to  ϵ ≥ 0is defined as:

image

Adversarial risk with  ϵ = 0is equivalent to standard risk, namely  AdvRisk0(f) = Risk(f) = Pr(x,y)∼µ[f(x) ̸= y]. For any classifier  f : X → Y, we define the adversarial gap of f with respect to  ϵ as:

image

In this section, we first propose a definition of representation vulnerability, and then prove several theorems that bound achievable model robustness based on representation vulnerability. Let  X ⊆ Rdbe the input space and  Z ⊆ Rnbe some feature space. In this work, we define a representation to be a function g that maps any input x in X to some vector g(x) ∈ Z. A classifier,  f = h ◦ g, maps an input to a label in a label space Y, and is a composition of a downstream classifier,  h : Z → Y, with a representation,  g : X → Z. As is done in previous works (Garg et al., 2018; Ilyas et al., 2019), we define a feature as a function from X to R, so can think of a representation as an array of features.

Inspired by the empirical success of standard representation learning using the mutual information maximization principle (Hjelm et al., 2018), we propose the following definition of representation vulnerability, which captures the robustness of a given representation against input distribution perturbations in terms of mutual information between its input and output.

Definition 3.1. Let  (X, µX, ∆)be a metric probability space of inputs and Z be some feature space. Given a representation  g : X → Zand  ϵ ≥ 0, the representation vulnerability of g with respect to perturbations bounded in an  ∞-Wasserstein ball with radius  ϵis defined as:

image

where X and  X′denote random variables that follow  µXand  µX′, respectively.

Representation vulnerability is always non-negative, and higher values indicate that the representation is less robust to adversarial input distribution perturbations. More formally, given parameters  ϵ ≥ 0and  τ ≥ 0, a representation g is called  (ϵ, τ)-robust if RVϵ(g) ≤ τ.

Notably, using the  ∞-Wasserstein distance does not restrict the choice of the metric function  ∆of the input space. This metric  ∆corresponds to the perturbation metric for defining adversarial examples. Thus, based on our definition of representation vulnerability, our following theoretical results and empirical methods work with any adversarial perturbation, including any  ℓp-norm based attack.

Compared with existing definitions of robust features (Garg et al., 2018; Ilyas et al., 2019; Eykholt et al., 2019), our definition is more general and enjoys several desirable properties. As it does not impose any constraint on the feature space, it is invariant to scale change1 and it does not require the knowledge of the labels. The most similar definition to ours is from Pensia et al. (2020), who propose to use statistical Fisher information as the evaluation criteria for feature robustness. However, Fisher information can only capture the average sensitivity of the log conditional density to small changes on the input distribution (when  ϵ → 0), whereas our definition is defined with respect to the worst-case input distribution perturbations in an  ∞-Wasserstein ball, which is more aligned with the adversarial setting. As will be shown next, our representation robustness notion has a clear connection with the potential model robustness of any classifier that can be built upon a representation.

3.1. Gaussian Mixture

We first study the implications of representation vulnerability under a simple Gaussian mixture model. We consider X ⊆ Rd as the input space and  Y = {−1, 1}as the space of binary labels. Assume  µXYis the underlying joint probability distribution defined over  X × Y, where all the examples (x, y) ∼ µXYare generated according to

image

where  θ∗ ∈ Rd and Σ∗ ∈ Rd×d are given parameters. The following theorem, proven in Appendix A.1, connects the vulnerability of a given representation with the adversarial gap of the best classifier based on the representation.

Theorem 3.2. Let  (X, ∥ · ∥p)be the input metric space and Y = {−1, 1}be the label space. Assume the underlying data are generated according to (3.1). Consider the feature space  Z = {−1, 1}and the set of representations,

image

Let  H = {h : Z → Y}be the set of non-trivial downstream

classifiers.2 Given  ϵ ≥ 0, for any g ∈ Gbin, we have

image

where  f ∗ = argminh∈H AdvRiskϵ(h ◦ g)is the optimal classifier based on  g, H2(θ) = −θ log θ−(1−θ) log(1−θ)is the binary entropy function and  H′2 denotes its derivative.

For this theoretical model for a simple case, Theorem 3.2 reveals the strong connection between representation vulnerability and the adversarial gap achieved by the optimal downstream classifier based on the representation. Note that the binary entropy function  H2(θ)is monotonically increasing over (0, 1/2), thus the first inequality suggests that low representation vulnerability guarantees a small adversarial gap if we train the downstream classifier properly. On the other hand, the second inequality implies that adversarial robustness cannot be achieved for any downstream classifier, if the vulnerability of the representation it uses is too high. As discussed in Section 6.1, the connection between representation vulnerability and adversarial gap is also found to hold empirically for image classification benchmarks.

3.2. General Case

This section presents our main theoretical results regarding robust representations. First, we present the following lemma, proven in Appendix A.2, that characterizes the connection between adversarial risk and input distribution perturbations bounded in an  ∞-Wasserstein ball.

Lemma 3.3. Let  (X, ∆)be the input metric space and Y be the set of labels. Assume all the examples are generated from a joint probability distribution  (X, Y ) ∼ µXY. Let  µXbe the marginal distribution of X. Then, for any classifier  f : X → Y and ϵ > 0, we have

image

where  X′ denotes the random variable that follows  µX′.

The next theorem, proven in Appendix A.3, gives a lower bound for the adversarial risk for any downstream classi-fier, using the worst-case mutual information between the representation’s input and output distributions.

Theorem 3.4. Let  (X, ∆)be the input metric space, Y be the set of labels and  µXYbe the underlying joint probability distribution. Assume the marginal distribution of labels  µYis a uniform distribution over Y. Consider the feature space Z and the set of downstream classifiers  H = {h : Z → Y}.Given  ϵ ≥ 0, for any g : X → Z, we have

image

where X is the random variable that follows the marginal distribution of inputs  µX and Z = g(X).

Theorem 3.4 suggests that adversarial robustness cannot be achieved if the available representation is highly vulnerable or the standard mutual information between X and g(X) is low. Note that  I(X; g(X)) − RVϵ(g) = inf{I(X′; g(X′)) :X′ ∼ µX′ ∈ BW∞(µX, ϵ)}, which corresponds to the worst-case mutual information between input and output of g. Therefore, if we assume robust classification as the downstream task for representation learning, then the representation having high worst-case mutual information is a necessary condition for achieving adversarial robustness for the overall classifier.

In addition, we remark that Theorem 3.4 can be extended to general p-th Wasserstein distances, if the downstream classifiers are evaluated based on robustness under distributional shift3, instead of adversarial risk. To be more specific, if using  Wpmetric to define representation vulnerability, we can then establish an upper bound on the maximum distributional robustness with respect to the considered  Wpmetric for any downstream classifier based on similar proof techniques of Theorem 3.4.

This section presents an empirical method for estimating the vulnerability of a given representation using i.i.d. samples. Recall from Definition 3.1, for any  g : X → Z, the representation vulnerability of g with respect to the input metric probability space  (X, µX, ∆) and ϵ ≥ 0is defined as:

image

(4.1) To measure representation vulnerability, we need to compute both terms  J1and  J2. However, the main challenge is that we do not have the knowledge of the underlying probability distribution  µXfor real-world problem tasks. Instead, we only have access to a finite set of data points sampled from the distribution. Therefore, it is natural to consider samplebased estimator for  J1 and J2for practical use.

The first term  J1is essentially the mutual information between X and Z = g(X). A variety of methods have been proposed for estimating mutual information (Moon et al., 1995; Darbellay & Vajda, 1999; Suzuki et al., 2008; Kan- dasamy et al., 2015; Moon et al., 2017). The most effective estimator is the mutual information neural estimator (MINE) (Belghazi et al., 2018), based on the dual representation of KL-divergence (Donsker & Varadhan, 1983):

image

where  Tθ : X × Z → Ris the function parameterized by a deep neural network with parameters  θ ∈ Θ, and �µ(m)XZ, �µ(m)X and �µ(m)Zdenote the empirical distributions4 of random variables (X, Z), X and Z respectively, based on m samples. In addition, Belghazi et al. (2018) empirically demonstrates the superiority of the proposed estimator in terms of estimation accuracy and efficiency, and prove that it is strongly consistent: for all  ε > 0, there exists  M ∈ Zsuch that for any  m ≥ M, |�Im(X; Z) − I(X; Z)| ≤ εalmost surely. Given the established effectiveness of this method, we implement MINE to estimate I(X; g(X)) as the first step.

Compared with  J1, the second term  J2is much more diffi-cult to estimate, as it involves finding the worst-case perturbations on  µXin a  ∞-Wasserstein ball in terms of mutual information. As with the estimation of  J1, we only have a finite set of instances sampled from  µX. On the other hand, due to the non-linearity and the lack of duality theory with respect to the  ∞-Wasserstein distance (Champion et al., 2008), it is inherently difficult to directly solve an ∞-Wasserstein constrained optimization problem, even if we work with the empirical distribution of  µX. To deal with the first challenge, we replace  µXwith its empirical measure  �µ(m)Xbased on i.i.d. samples. Then, to avoid the need to search through the whole  ∞-Wasserstein ball, we restrict the search space of  µX′to be the following set of empirical distributions:

image

where  S = {xi : i ∈ [m]}denotes the given set of m data points sampled from  µX. Note that the considered set A(S, ϵ) ⊆ BW∞(�µ(m)X , ϵ), since each perturbed point  x′i isat most  ϵ-away from  xi. Finally, making use of the dual formulation of KL-divergence that is used in MINE, we propose the following empirical optimization problem for estimating  J2:

image

where we simply set the empirical distribution  �µ(m)X′to be the same as  µX′. In addition, we propose a heuristic alternating minimization algorithm to solve (4.3) (see Appendix B for the pseudocode and a complexity analysis of the proposed algorithm). More specifically, our algorithm alternatively performs gradient ascent on  θfor the inner maximization problem of estimating �Im(X′; g(X′))given µ′X, and searches for the set of worst-case perturbations on {x′i : i ∈ [m]} given θbased on projected gradient descent.

In this section, we present our method for learning adversarially robust representations. First, we introduce the mutual information maximization principle for representation learning (Linsker, 1989; Bell & Sejnowski, 1995). Mathematically, given an input probability distribution  µXand a set of representations  G = {g : X → Z}, the maximization principle proposes to solve this optimization problem:

image

Although this principle has been shown to be successful for learning good representations under the standard setting (Hjelm et al., 2018), it becomes ineffective when considering adversarial perturbations (see Table 1 for an illustration). Motivated by the theoretical connections between feature sensitivity and adversarial risk for downstream robust classification shown in Section 3, we stimulate robust representations by adding a regularization term based on representation vulnerability:

image

where  β ≥ 0is the trade-off parameter between I(X; g(X)) and  RVϵ(g). When β = 0, (5.2)is same as the objective for learning standard representations (5.1). Increasing the value of  βwill produce representations with lower vulnerability, but may undesirably affect the standard mutual information I(g(X); X) if  βis too large. In particular, we set  β = 1in the following discussions, which allows us to simplify (5.2) to obtain the following optimization problem:

image

The proposed training principle (5.3) aims to maximize the mutual information between the representation’s input and output under the worst-case input distribution perturbation bounded in a  ∞-Wasserstein ball. We remark that optimization problem (5.3) aligns well with the results of Theorem 3.4, which shows the importance of the learned feature representation achieving high worst-case mutual information for a downstream robust classification task.

As with estimating the feature sensitivity in Section 4, we do not have access to the underlying  µX. However, the inner minimization problem is exactly the same as estimating the worst-case mutual information  J2 in (4.1), thus we can simply adapt the proposed empirical estimator (4.3) to solve (5.3). To be more specific, we reparameterize g using a neural network with parameter  ψ ∈ Ψand use the following min-max optimization problem:

image

Based on the proposed algorithm for the inner minimization problem, (5.4) can be efficiently solved using a standard optimizer, such as stochastic gradient descent.

This section reports on experiments to study the implications of robust representations on benchmark image datasets. Instead of focusing directly improving model robustness, our experiments focus on understanding the proposed defi-nition of robust representations as well as its implications. Based on the proposed estimator in Section 4, Section 6.1 summarizes experiments to empirically test the relationship between representation vulnerability and model robustness, by extracting internal representations from the state-of-the-art pre-trained standard and robust classification models. In addition, we empirically evaluate the general lower bound on adversarial risk presented in Theorem 3.4. In Section 6.2, we evaluate the proposed training principle for learning robust representations on image datasets, and test its performance with comparisons to the state-of-the-art standard representation learning method in a downstream robust clas-sification framework. We also visualize saliency maps as an intuitive criteria for evaluating representation robustness.

We conduct experiments on MNIST (LeCun & Cortes, 2010), Fashion-MNIST (Xiao et al., 2017), SVHN (Net- zer et al., 2011), and CIFAR-10 (Krizhevsky et al., 2009), considering typical  ℓ∞-norm bounded adversarial perturbations for each dataset (ϵ = 0.3for MNIST, 0.1 for FashionMNIST, 4/255 for SVHN, and 8/255 for CIFAR-10). We use the PGD attack (M ˛adry et al., 2018) for both generating adversarial distributions in the estimation of worst-case mutual information and evaluating model robustness. To implement our proposed estimator (4.3), we adopt the encode-and-dot-product model architecture in Hjelm et al. (2018) and adjust it to adapt to different forms of representations. We leverage implementations from Engstrom et al. (2019a) and Hjelm et al. (2018) in our implementation5. Implementation details are provided in Appendix D.1.

6.1. Representation Robustness

To evaluate our proposed definition on representation vulnerability and its implications for downstream classification

image

Figure 1. Correlations between the representation vulnerability and the CIFAR-10 model’s natural-adversarial accuracy gap. Filled points indicate robust models (trained with  ϵ = 8/255), half-filled are models adversarially trained with  ϵ = 2/255, and unfilled points are standard models.

models, we conduct experiments on image benchmarks using various classifiers, including VGG (Simonyan & Zis- serman, 2015), ResNet (He et al., 2016), DenseNet (Huang et al., 2017) and the simple convolutional neural network in Hjelm et al. (2018) denoted as Baseline-H.

Correlation with model robustness. Theorem 3.2 establishes a direct correlation between our representation vulnerability definition and achievable model robustness for the synthetic Gaussian-mixture case, but we are not able to theoretically establish that correlation for arbitrary distributions. Here, we empirically test this correlation on image benchmark datasets. Figure 1 summarizes the results of these experiments for CIFAR-10, where we set the logit layer as the considered representation space. The adversarial gap decreases with decreasing representation vulnerability in an approximately consistent relationship. Models with low logit layer representation vulnerability tend to have low natural-adversarial accuracy gap, which is consistent with the intuition behind our definition and with the theoretical result on the synthetic Gaussian-mixture case. This suggests the correlation between representation vulnerability and model robustness may hold for general case.

Adversarial risk lower bound. Theorem 3.4 provides a lower bound on the adversarial risk that can be achieved by any downstream classifier as a function of representation vulnerability. To evaluate the tightness of this bound, we estimate the normal-case and worst-case mutual information I(X; g(X)) of layer representation g for different models, and empirically evaluate the adversarial risk of the models. Figure 2 shows the results, where we again set the logit layer as the feature space for a more direct comparison. The lower bound of adversarial risk is calculated according to Theorem 3.4 and is converted to the upper bound of ad-

image

Figure 2. Normal and worst case mutual information for logit-layer representations. Each pair of points shows the result of a specific model—the left point indicates the worst case mutual information and the right for the normal mutual information. Filled points are robust models; hollow points are standard models.

versarial accuracy for reference. In particular, for standard models, both the estimated worst-case mutual information and the adversarial accuracy are close to zero, whereas the computed upper bounds on adversarial accuracy are around 30%. We empirically observed around 50% adversarial accuracy for robust models, whereas the bounds computed using the estimated worst-case mutual information and Theorem 3.4 are about 75%. This shows that Theorem 3.4 gives a reasonably tight bound for a model’s adversarial accuracy with respect to the logit-layer representation robustness.

Figure 2 also indicates that even the robust models produced by adversarial training have representations that are not suf-ficiently robust to enable robust downstream classifications. For example, robust DenseNet121 in our evaluations has the highest logit layer worst-case mutual information of 1.08, yet the corresponding adversarial accuracy is upper bounded by 77.0% which is unsatisfactory for CIFAR-10. Such information theoretic limitation also justifies our training principle of worst-case mutual information maximization, since on the other hand the adversarial accuracy upper bound calculated by normal-case mutual information does not constitute a limitation for most robust models in our experiments (as in Figure 1, most robust models achieve adversarial accuracy close to 100%).

Internal feature robustness. We further investigate the implications of our proposed definition from the level of individual features. Specifically for neural networks, we consider the function from the input to each individual neuron within a layer as a feature. The motivations for considering feature robustness comes from the fact that mutual information in terms of the whole representation is controlled by the sum of all the features’ mutual information (see Appendix C for a rigorous argument) and robust features are potentially easier to train (Garg et al., 2018). As an illustration, we

image

Figure 3. Distribution of mutual information I(X; g(X)) and feature vulnerability in the second convolutional layer of Baseline-H. The upper plots are for standard models, and the lower plots are for robust models. The total number of neurons is 128.

evaluate the robustness of all the convolutional kernels in the second layer of the Baseline-H model. Each neuron evaluated here is a composite convolutional kernel (all kernels in the first layer connected to a second layer kernel) with image input size  10×10. Figure 3shows the results that are averaged over two independently trained models for each type. This result reveals the apparent difference in feature robustness between a standard model and the adversarially-trained robust model, even in lower layers. Although in this case the result does not prohibit a robust downstream model for lower layers neurons, for neurons in higher layers the difference becomes more distinct and the vulnerability of neurons can thus be the bottleneck of achieving high model robustness. The different feature robustness according to our definition also coincide with the saliency maps of features (see Figure 5 in Appendix D.2), where the saliency maps of robust features are apparently more interpretable compared to those of standard features.

6.2. Learning Robust Representations

Our worst-case mutual information maximization training principle provides an unsupervised way to learn adversarially robust representations. Since there are no established ways to measure the robustness of a representation, empirically testing the robustness of representations learned by our training principle poses a dilemma. To avoid circular reasoning, we evaluate the learned representations by running a series of downstream adversarial classification tasks and comparing the performance of the best models we are able to find for each representation. In addition, recent work shows that the interpretability of saliency map has certain connections with robustness (Etmann et al., 2019; Ilyas et al., 2019), thus we study the saliency map as an alternative criteria for evaluating robust representations.

The unsupervised representation learning approach based on mutual information maximization principle in Hjelm

image

Table 1. Comparisons of different representation learning methods on CIFAR-10 in downstream classification settings. E.S. denotes early stopping under the criterion of the best adversarial accuracy. We present mean accuracy and the standard deviation over 4 repeated trials.

et al. (2018) achieves the state-of-the-art results in many downstream tasks, including standard classification. We further adopt their encoder architecture in our implementation, and extend their evaluation settings to adversarially robust classification. Specifically, we truncate the front part of Baseline-H with a 64-dimensional latent layer output as the representation g and train it by the worst-case mutual information maximization principle using only unlabeled data (removing the labels from the normal training data). We test two architectures (two-layer multilayer perceptron and linear classifier) for implementing the downstream clas-sifier h and train it using labeled data after the encoder g has been trained using unlabeled data. Appendix D.1 provides additional details on the experimental setup.

Downstream classification tasks. Comparison results on CIFAR-10 are demonstrated in Table 1 (see Appendix D.2 for a similar results for MNIST, Fashion-MNIST, and SVHN). The fully-supervised models are trained for reference, from which we can see the simple model architecture we use achieves a decent natural accuracy of 86.3%; the adversarially-trained robust model reduces accuracy to around 70% with adversarial accuracy of 40.5%. The baseline, with g and h both trained normally, resembles the setting in Hjelm et al. (2018) and achieves a natural accuracy of 58.8%. For representations learned using worst-case mutual information maximization, the composition with standard two-layer multilayer perceptron (MLP) h achieves a non-trivial (compared to the 0.2% for the standard representation) adversarial accuracy of 14.1%. When h is further trained using adversarial training, the robust accuracy increases to 31.5% which is comparable to the result of the robust fully-supervised model. As an ablation, the robust h based on standard g achieves an adversarial accuracy of 15.1%, yet the natural accuracy severely drops below 30%, indicating that a robust classifier cannot be found using the vulnerable representation. The case where h is a simple linear classifier shows similar results. These comparisons show that the representation learned using worst-case mutual information maximization can make the downstream classification more robust over the baseline and approaches the robustness of fully-supervised adversarial training. This provides evidence that our training principle produces adversarially robust representations.

Another interesting implication given by results in Table 1 is that robustly learned representations may also have better natural accuracy (62.5%) over the standard representation (58.8%) in downstream classification tasks on CIFAR-10. This matches our experiments in Figure 2 where logit layer representations in robust models conveys more normal-case mutual information (up to 1.75) than those in standard models (up to 1.25). However, this is not the case on MNIST dataset as in Table 3. We conjecture that this is because the information conveyed by robust representations has better generalizations, and the generalization is more of a problem on CIFAR-10 than on MNIST (Schmidt et al., 2018).

Saliency maps. A saliency map is commonly defined as the gradient of a model’s loss with respect to the model’s input (Etmann et al., 2019). For a classification model, it intuitively illustrates what the model looks for in changing its classification decision for a given sample. Recent work (Etmann et al., 2019; Ilyas et al., 2019) indicates, at least in some synthetic settings, that the more alignment the saliency map has with the input image, the more adversarially robust the model is. As an additional test of representation robustness, we calculate the saliency maps of standard and robust representations g by the mutual information maximization loss with respect to the input. Figure 4 shows that the saliency maps of the robust representation appear to be much less noisy and more interpretable in terms of the alignment with original images. Intuitively, this shows that robust representations capture relatively higher level visual concepts instead of pixel-level statistical clues (Engstrom et al., 2019b). The more interpretable saliency maps of representation learned by our training principle further support its effectiveness in learning adversarially robust representation.

image

Figure 4. Visualization of saliency maps of different models on CIFAR-10: (a) original images (b) representations learned using Hjelm et al. (2018) (c) representations learned using our method.

We proposed a novel definition of representation robustness based on the worst-case mutual information, and showed both theoretical and empirical connections between our defi-nition and model robustness for a downstream classification task. In addition, by developing estimation and training methods for representation robustness, we demonstrated the connection and the usefulness of the proposed method on benchmark datasets. Our results are not enough to produce strongly robust models, but they provide a new approach for understanding and measuring achievable adversarial robustness at the level of representations.

This work was partially funded by an award from the National Science Foundation SaTC program (Center for Trustworthy Machine Learning, #1804603) and additional support from Amazon, Baidu, and Intel.

Belghazi, M. I., Baratin, A., Rajeshwar, S., Ozair, S., Ben- gio, Y., Hjelm, R. D., and Courville, A. C. Mutual information neural estimation. In International Conference on Learning Representations, 2018.

Bell, A. J. and Sejnowski, T. J. An informationmaximization approach to blind separation and blind deconvolution. Neural computation, 7(6):1129–1159, 1995.

Bubeck, S., Lee, Y. T., Price, E., and Razenshteyn, I. Ad- versarial examples from computational constraints. In International Conference on Machine Learning, 2019.

Champion, T., De Pascale, L., and Juutinen, P. The  ∞-wasserstein distance: Local solutions and existence of optimal transport maps. SIAM Journal on Mathematical Analysis, 40(1):1–20, 2008.

Cover, T. M. and Thomas, J. A. Elements of Information Theory. John Wiley & Sons, 2012.

Darbellay, G. A. and Vajda, I. Estimation of the information by an adaptive partitioning of the observation space. IEEE Transactions on Information Theory, 45(4):1315–1321, 1999.

Devlin, J., Chang, M.-W., Lee, K., and Toutanova, K. BERT: Pre-training of deep bidirectional transformers for language understanding. In North American Chapter of the Association for Computational Linguistics: Human Language Technologies, 2019.

Donsker, M. D. and Varadhan, S. S. Asymptotic evaluation of certain markov process expectations for large time, IV. Communications on Pure and Applied Mathematics, 36 (2):183–212, 1983.

Engstrom, L., Tran, B., Tsipras, D., Schmidt, L., and M ˛adry, A. A rotation and a translation suffice: Fooling CNNs with simple transformations. arXiv:1712.02779, 2017.

Engstrom, L., Ilyas, A., Santurkar, S., and Tsipras, D. Ro- bustness (python library), 2019a. URL https://github.com/ MadryLab/robustness.

Engstrom, L., Ilyas, A., Santurkar, S., Tsipras, D., Tran, B., and M ˛adry, A. Adversarial robustness as a prior for learned representations. arXiv:1906.00945, 2019b.

Etmann, C., Lunz, S., Maass, P., and Schoenlieb, C. On the connection between adversarial robustness and saliency map interpretability. In International Conference on Machine Learning, 2019.

Eykholt, K., Evtimov, I., Fernandes, E., Li, B., Rahmati, A., Xiao, C., Prakash, A., Kohno, T., and Song, D. Robust physical-world attacks on deep learning visual classifi-cation. In IEEE Conference on Computer Vision and Pattern Recognition, 2018.

Eykholt, K., Gupta, S., Prakash, A., and Zheng, H. Ro- bust classification using robust feature augmentation. arXiv:1905.10904, 2019.

Fawzi, A., Fawzi, H., and Fawzi, O. Adversarial vulnerabil- ity for any classifier. In Advances in Neural Information Processing Systems, 2018.

Garg, S., Sharan, V., Zhang, B., and Valiant, G. A spectral view of adversarially robust features. In Advances in Neural Information Processing Systems, 2018.

Gilmer, J., Metz, L., Faghri, F., Schoenholz, S. S., Raghu, M., Wattenberg, M., and Goodfellow, I. Adversarial spheres. arXiv:1801.02774, 2018.

He, K., Zhang, X., Ren, S., and Sun, J. Deep residual learning for image recognition. In IEEE Conference on Computer Vision and Pattern Recognition, 2016.

Hjelm, R. D., Fedorov, A., Lavoie-Marchildon, S., Grewal, K., Bachman, P., Trischler, A., and Bengio, Y. Learning deep representations by mutual information estimation and maximization. In International Conference on Learning Representations, 2018.

Huang, G., Liu, Z., Van Der Maaten, L., and Weinberger, K. Q. Densely connected convolutional networks. In IEEE Conference on Computer Vision and Pattern Recognition, 2017.

Ilyas, A., Santurkar, S., Tsipras, D., Engstrom, L., Tran, B., and M ˛adry, A. Adversarial examples are not bugs, they are features. In Advances in Neural Information Processing Systems, 2019.

Kandasamy, K., Krishnamurthy, A., Poczos, B., Wasser- man, L., et al. Nonparametric von mises estimators for entropies, divergences and mutual informations. In Advances in Neural Information Processing Systems, 2015.

Kolouri, S., Park, S. R., Thorpe, M., Slepcev, D., and Rohde, G. K. Optimal mass transport: Signal processing and machine-learning applications. IEEE Signal Processing Magazine, 34(4):43–59, 2017.

Krizhevsky, A., Hinton, G., et al. Learning multiple layers of features from tiny images. 2009.

LeCun, Y. and Cortes, C. MNIST handwritten digit database. 2010. URL http://yann.lecun.com/exdb/mnist/.

Linsker, R. How to generate ordered maps by maximizing the mutual information between input and output signals. Neural Computation, 1(3):402–411, 1989.

Mahloujifar, S., Diochnos, D. I., and Mahmoody, M. The curse of concentration in robust learning: Evasion and poisoning attacks from concentration of measure. In AAAI Conference on Artificial Intelligence, 2019.

M ˛adry, A., Makelov, A., Schmidt, L., Tsipras, D., and Vladu, A. Towards deep learning models resistant to adversarial attacks. In International Conference on Learning Representations, 2018.

Moon, K. R., Sricharan, K., and Hero, A. O. Ensemble estimation of mutual information. In IEEE International Symposium on Information Theory. IEEE, 2017.

Moon, Y.-I., Rajagopalan, B., and Lall, U. Estimation of mutual information using kernel density estimators. Physical Review E, 52(3):2318, 1995.

Netzer, Y., Wang, T., Coates, A., Bissacco, A., Wu, B., and Ng, A. Y. Reading digits in natural images with unsupervised feature learning. In NeurIPS Workshop on Deep Learning and Unsupervised Feature Learning, 2011.

Pensia, A., Jog, V., and Loh, P.-L. Extracting robust and ac- curate features via a robust information bottleneck. IEEE Journal on Selected Areas in Information Theory, 2020.

Scarlett, J. and Cevher, V. An introductory guide to Fano’s inequality with applications in statistical estimation. arXiv:1901.00555, 2019.

Schmidt, L., Santurkar, S., Tsipras, D., Talwar, K., and M ˛adry, A. Adversarially robust generalization requires more data. In Advances in Neural Information Processing Systems, 2018.

Shafahi, A., Huang, W. R., Studer, C., Feizi, S., and Gold- stein, T. Are adversarial examples inevitable? In International Conference on Learning Representations, 2018.

Sharif, M., Bhagavatula, S., Bauer, L., and Reiter, M. K. Accessorize to a crime: Real and stealthy attacks on state-of-the-art face recognition. In ACM Conference on Computer and Communications Security, 2016.

Simonyan, K. and Zisserman, A. Very deep convolutional networks for large-scale image recognition. In International Conference on Learning Representations, 2015.

Sinha, A., Namkoong, H., and Duchi, J. Certifying some dis- tributional robustness with principled adversarial training. In International Conference on Learning Representations, 2018.

Suzuki, T., Sugiyama, M., Sese, J., and Kanamori, T. Ap- proximating mutual information by maximum likelihood density ratio estimation. In New challenges for feature selection in data mining and knowledge discovery, 2008.

Szegedy, C., Zaremba, W., Sutskever, I., Bruna, J., Erhan, D., Goodfellow, I., and Fergus, R. Intriguing properties of neural networks. In International Conference on Learning Representations, 2014.

Xiao, H., Rasul, K., and Vollgraf, R. Fashion-MNIST: a novel image dataset for benchmarking machine learning algorithms. arXiv:1708.07747, 2017.

Zhang, H., Yu, Y., Jiao, J., Xing, E., El Ghaoui, L., and Jordan, M. I. Theoretically principled trade-off between robustness and accuracy. In International Conference on Learning Representations, 2019.

In this section, we provide the proofs of our theoretical results in Section 3.

For ease of presentation, we introduce the following notations and an alternative definition of the  ∞-Wasserstein distance. Let  (X, µ) and (Y, ν)be two probability spaces. We say that  T : X → Ytransports  µ ∈ P(X) to ν ∈ P(Y), and we call T a transport map, if  ν(B) = µ(T −1(B)), for all ν-measurable sets B. In addition, for any measurable map  T : X → Y, wedefine the pushforward of  µ through T as (T#(µ)) given by

image

Alternative definition of  ∞-Wasserstein distance. From the perspective of transportation theory, given two probability measures  µand  νon  (X, ∆), any joint probability distribution  γ ∈ Γ(µ, ν)corresponds to a specific transport map T : X → Xthat moves  µto  ν. Then, the p-th Wasserstein distance can viewed as finding the optimal transport map to move from  µ to νthat minimizes some cost functional depending on p (Kolouri et al., 2017). For the case where  p = ∞, ifwe let T be the transport map induced by a given  γ ∈ Γ(µ, ν), then the cost functional can be informally understood as the maximum of all the transport distances  ∆(T(x), x). More rigorously, the  ∞-Wasserstein distance can be alternatively defined as

image

A more detailed discussion of  ∞-Wasserstein distance can be found in Champion et al. (2008).

A.1. Proof of Theorem 3.2

Theorem 3.2 (restated here) connects the vulnerability of a given representation with the minimum adversarial gap of any classifier based on that representation.

Theorem 3.2. Let  (X, ∥ · ∥p)be the input metric space and  Y = {−1, 1}be the label space. Assume the underlying data are generated according to (3.1). Consider the feature space  Z = {−1, 1}and the set of representations,

image

Let  H = {h : Z → Y}be the set of non-trivial downstream classifiers.6 Given  ϵ ≥ 0, for any g ∈ Gbin, we have

image

where  f ∗ = argminh∈H AdvRiskϵ(h ◦ g)is the optimal classifier based on  g, H2(θ) = −θ log θ − (1 − θ) log(1 − θ)is the binary entropy function and  H′2 denotes its derivative.

Proof. Let  µXYbe the underlying joint probability distribution of the examples according to (3.1) and  µXbe corresponding the marginal distribution of X. To begin with, we compute the explicit formulation for the defined representation vulnerability in Definition 3.1. Note that  I(U; V ) = H(U) − H(U|V ) = H(V ) − H(V |U). Thus, for any  gw ∈ G we have

image

where the first equality holds because  H(gw(U) | U) = 0for any random variable U, and the second equality is due to the fact that the distribution of X is symmetric with respect to  w⊤x = 0. Note that the binary entropy function H2(θ) = −θ log θ−(1−θ) log(1−θ)is monotonically increasing with respect to  θ in [0, 1/2)and monotonically decreasing in (1/2, 1]. Therefore, the optimal value of  RVϵ(gw)is achieved when  µX′either minimizes  Prx′∼µX′ (w⊤x′ ≥ 0)or maximizes  Prx′∼µX′ (w⊤x′ ≥ 0).

According to the Hölder’s inequality, we have  |⟨a, b⟩| ≤ ∥a∥p · ∥b∥qfor any  a, b ∈ Rd, where 1/p + 1/q = 1. By the alternative definition of  ∞-Wasserstein distance, for any  µX′that satisfies  W∞(µX′, µX) ≤ ϵ, it induces a transport map T : X → Xsuch that  µ′X = T#(µX)and  ∥∆(T(X), X)∥p ≤ ϵholds almost surely with respect to the randomness of X and T. Thus, we have

image

which implies

We remark that the equality can be achieved when the the transport map T is constructed by perturbing the i-th element of any sampled  x ∼ µXby  ϵ · (wqi / �i wiq)1/p, for any i = 1, 2, . . . , d. In addition, according to the assumed Gaussian Mixture model (3.1), we have

image

Similarly, we have

image

Therefore, we derive the explicit formulation for  RVϵ(gw)

image

where  H2(·)is denotes binary entropy function.

Next, given  gw ∈ Gbin, we are going to compute the adversarial gap of  f ◦ gw for each h ∈ H. To begin with, we consider the first case  h1(z) = z for any z ∈ Z. According to the definition of adversarial risk, we have

image

where the equality is due to the fact that  B(0, ϵ)is symmetric with respect to 0, and the last equality holds because of the Hölder’s inequality: for any  a, b ∈ Rn, it holds that  a⊤b ≥ −∥a∥p · ∥b∥qand the equality is achieved when (ai/∥a∥p)p = (bi/∥b∥q)q for any i ∈ {1, 2, . . . , d}.

Similarly, the standard risk can be computed as:

image

Thus, we derive the gap between standard and adversarial risk with respect to  h1 ◦ gw:

image

For the other case where  h2(z) = −z for any z ∈ Z, note that  h1 ◦ gw = h2 ◦ g−w for any gw ∈ Gbin. Thus, a similar proof technique can be applied to compute the adversarial risk,

image

and the adversarial gap,

Note that  f ∗ : X → Yis the optimal classifier based on  gwthat minimizes the adversarial risk of  h ◦ gw for any h ∈ H. Bycomparing the adversarial risk of  h1 ◦ gwand  h2 ◦ gw, we have  f ∗ = h1 ◦ gw, if  w⊤θ∗ ≥ 0; f ∗ = h2 ◦ gw, otherwise. Thus, we derive the adversarial gap with respect to  f ∗ as follows

image

Based on (A.2), we further obtain the following inequality

image

which completes the proof.

A.2. Proof of Lemma 3.3

Lemma 3.3, restated below, connects adversarial risk and input distribution perturbations bounded in an  ∞-Wasserstein ball.

Lemma 3.3. Let  (X, ∆)be the input metric space and Y be the set of labels. Assume all the examples are generated from a joint probability distribution  (X, Y ) ∼ µXY . Let µXbe the marginal distribution of X. Then, for any classifier  f : X → Yand  ϵ ≥ 0, we have

image

where  X′ denotes the random variable that follows  µX′.

Proof. Our proof proves the equality by proving  ≤inequalities in both directions. First, we prove

image

For any classifier  f : X → Y, according to Definition 2.3, we have

image

Since f is a given deterministic function, the optimal perturbation scheme that achieves  AdvRiskϵ(f)essentially defines a transport map  T : X → X. More specifically, let  Cy(f) = {x ∈ X : f(x) ̸= y}. Then, for any sampled pair  (x, y) ∼ µXY ,we can construct T such that

image

Let (X, Y ) be the random variable that follows  µXY. By construction, it can be easily verified that  T#(µX) ∈ BW∞(µX, ϵ)and  AdvRiskϵ(f) = Pr�f(T(X)) ̸= Y�. Therefore, we have proven (A.3).

It remains to prove the other direction of the inequality:

According to the alternative definition of  ∞-Wasserstein distance, the optimal solution  µ∗X′that achieves the supremum of the right hand side of (A.4) can be captured by a transport map  T ∗ : X → X such that µ∗X′ = T ∗#(µX) and ∆(T ∗(X), X) ≤ ϵholds almost surely with respect to the randomness of  X and T ∗. Thus, we have

image

Therefore, we have proven the second direction and completed the proof.

A.3. Proof of Theorem 3.4

Theorem 3.4, restated below, gives a lower bound for the adversarial risk for any downstream classifier in terms of the worst-case mutual information between the representation’s input and output distributions.

Theorem 3.4. Let  (X, ∆)be the input metric space, Y be the set of labels and  µXYbe the underlying joint probability distribution. Assume the marginal distribution of labels  µYis a uniform distribution over Y. Consider the feature space Z and the set of downstream classifiers  H = {h : Z → Y}. Given ϵ ≥ 0, for any g : X → Z, we have

image

where X is the random variable that follows the marginal distribution of inputs  µX and Z = g(X).

Before starting the proof, we state two useful lemmas on Markov chains. A Markov chain is defined to be a collection of random variables  {Xt}t∈Zwith the property that given the present, the future is conditionally independent of the past. Namely,

image

Lemma A.1 (Fano’s Inequality). Let X be a random variable uniformly distributed over a finite set of outcomes X. For any estimator �X such that X → Y → �Xforms a Markov chain, we have

image

Lemma A.2 (Data-Processing Inequality). For any Markov chain  X → Y → Z, we have

image

Chapter 2 in Cover & Thomas (2012) provides proofs of Lemmas A.1 and A.2.

Proof of Theorem 3.4. For any classifier  h : Z → Y, according to Lemma 3.3, we have

image

Let  µX′ ∈ BW∞(µX, ϵ)be a probability measure over  (X, ∆). According to the alternative definition of  ∞-Wasserstein distance using optimal transport,  µX′corresponds to a transport map  T : X → X such that µX′ = T#(µX). Thus, for any given  µX′ ∈ BW∞(µX, ϵ) and h ∈ H, we have the Markov chain

image

where X, Y are random variables for input and label distributions respectively. The first Markov chain  Y → Xcan be understood as a generative model for generating inputs according to the conditional probability distribution  µX|Y. Therefore, applying Lemmas A.1 and A.2, we obtain the inequality,

image

Taking the supremum over the distribution of  X′ in BW∞(µX, ϵ)and infimum over  h ∈ Hon both sides of (A.6) yields

image

where the first equality is due to (A.5) and the inequality holds because of (A.6). Thus, we completed the proof.

This section presents the pseudocode of our heuristic algorithm for solving the empirical estimation problem (4.3). More specifically, given a training sample set  Strain, our algorithm alternatively optimizes for the worst-case input perturbations using projected gradient descent (Algorithm 1) and conducts gradient ascent for the network parameters  θ(training phase in Algorithm 2). Based on the best parameter  θoptselected from the training phase, our algorithm then estimates the worst-case mutual information with respect to the given representation g using a testing sample set  Stest(testing phase in Algorithm 2). Since we only have assess to a finite set of data sampled from  µX, we use an additional testing phase in Algorithm 2 to minimize the overfitting effect of the training samples on the optimal network parameter  θoptfor mutual information neural estimation (MINE).

Moreover, we adopt the negative sampling scheme (Hjelm et al., 2018) to estimate the expectation term with respect to  �µ(m)X ⊗ �µ(m)Zin mutual information neural estimation for better performance. Here, the pairing scheme defines a correspondence from each input to a set of inputs for a given sample set. To be more specific, given a set of samples {xi}i∈[B], a pairing scheme with negative sampling size  N ≤ Bcorresponds to a set of vectors  {πi}i∈[B]such that each  πiis a randomly selected subset from  {1, 2, . . . , B} with size N, and πijdenotes the j-th element of  πi. Compared with the algorithm in Hjelm et al. (2018) for estimating standard mutual information, Algorithm 2 requires additional  B · S steps offorward and backward propagations with respect to the input for finding the worst-case input perturbations in each iteration.

The following tensorization inequality (Scarlett & Cevher, 2019) characterizes the connection between the mutual information of individual neuron features and that of the whole representation. According to Theorem 3.4, such connection suggests the necessity of enough worst-case mutual information for each individual neuron.

image

Lemma C.1 (Tensorization of Mutual Information). Let  Z = (Z1, ..., Zn)be a product distributions over random variables. If  Z1, ..., Znare mutually independent conditioned on X, then

image

Suppose neurons within a single layer have no interconnection, then each neuron’s output is mutually independent conditioned on the model input. If a perturbation imposed on the input distribution makes the perturbed mutual information  I(X′; Z′i)relatively low for each neuron, then the perturbed mutual information with respect to the entire layer  I(X′; Z′)will also be low, which further implies a low adversarial accuracy for any downstream classifier based on Theorem 3.4.

D.1. Implementation Details

Here, we provide additional implementation details of our experiments presented in Section 6.

Model architectures. For all experiments, we follow Hjelm et al. (2018) in implementing the MINE estimator. We adopt the encode-and-dot-product model architecture in Hjelm et al. (2018) which maps x and z respectively to two high-dimensional vectors and then takes the dot-product to calculate the output. The basic modules used in our experiments are listed in Table 2. A slight difference in training the feature (encoder) is that Hjelm et al. (2018) shares parameters between parts of the mutual information estimator and the encoder, while we separate the two parts completely to be consistent with our mutual information estimation experiments.

image

Hyperparameters. We use simple hyperparameter settings to control their effect on our various ablation experiments. We use  l∞constrained perturbations and PGD attack (M ˛adry et al., 2018) on all datasets. For CIFAR-10, we set the radius ϵ = 8/255and use 7 attack steps with step size 0.01. For MNIST, we set the radius  ϵ = 0.3and use 10 attack steps with step size 0.1. For Fashion-MNIST, we set the radius  ϵ = 0.1 and use 10attack steps with step size 0.02. For SVHN, we set the radius  ϵ = 4/255and use 10 attack steps with step size 0.005. The batch size is set as 128 for both datasets, and our results are consistent with different batch sizes between 128 to 512 (we did not test other sizes). A total training epochs of 200 is set for VGG, ResNet, and DenseNet, with an initial learning rate of 0.1 which decays by a factor of 10 every 50 epochs. For the Baseline-H model and the similar mutual information estimator, we set the training epoch to 300 and use a fixed learning rate of 0.0001 as in Hjelm et al. (2018).

D.2. Additional Results

Saliency maps of internal features. In section 6.1, we evaluated the internal feature vulnerability of all the convolutional kernels in the second layer of Baseline-H. Here, we further visualize the saliency maps of the those internal features to evaluate the underlying correlations. As shown in Figure 5, features in robust model have less noisy saliency maps, which is consistent with the observations of lower representation vulnerability shown in Figure 3.

Saliency maps of learned representations. More comparison results of saliency maps are given in Figure 6. The saliency maps of representations learned using our unsupervised training method shows comparable interpretability results to the models learned using fully-supervised adversarial training. Saliency maps computed by different losses also show consistent interpretability results. This indicates that our training principle indeed produces adversarially robust representations.

image

MLP h Linear h Representation (g) Classifier (h) Natural Adversarial Natural Adversarial

image

Table 3. Comparisons of different representation learning methods for downstream classification on MNIST. E.S. denotes early stopping under the criterion of the best adversarial accuracy. We present the mean accuracy and the standard deviation over 4 repeated trials.

MLP h Linear h Representation (g) Classifier (h) Natural Adversarial Natural Adversarial

image

Table 4. Comparisons of different representation learning methods for downstream classification on Fashion-MNIST. E.S. denotes early stopping under the criterion of the best adversarial accuracy. We present the mean accuracy and the standard deviation over 4 repeated trials.

MLP h Linear h Representation (g) Classifier (h) Natural Adversarial Natural Adversarial

image

Table 5. Comparisons of different representation learning methods for downstream classification on SVHN. E.S. denotes early stopping under the criterion of the best adversarial accuracy. We present the mean accuracy and the standard deviation over 4 repeated trials.

image

Figure 5. Saliency maps of four arbitrarily selected features in the second convolutional layer of Baseline-H. The feature saliency map is computed by the gradient of a kernel’s averaged activation over a input image. Each row presents saliency maps of a specific convolutional kernel.

image

Figure 6. Saliency maps of different models on CIFAR-10: (a) original images (b) fully-supervised standard model (c) fully-supervised robust model (d) representations learned using Hjelm et al. (2018) with cross-entropy loss (e) representations learned using Hjelm et al. (2018) with mutual information maximization loss (f) representations learned using our method with cross-entropy loss (g) representations learned using our method with mutual information maximization loss.


Designed for Accessibility and to further Open Science