1 Introduction

Prediction and estimation problems play a major role in modern mathematical statistics. The aim is to approximate, in one way or another, an unknown random variable Y by a function from a given class F, defined on a probability space \((\Omega ,\mu )\). The given data is a random sample \((X_i,Y_i)_{i=1}^N\), distributed according to the N-product of the joint distribution of \(\mu \) and Y, endowed on the product space \((\Omega \times \mathbb {R})^N\).

The notion of approximation may change from problem to problem. It is reflected by different choices of loss functions, which put a price tag on predicting f(X) instead of Y. Although it is not the most general form possible, we also assume throughout this article that if \(\ell \) is the loss function, the cost of predicting f(X) instead of Y is \(\ell (f(X)-Y)\):

Definition 1.1

A loss is a real-valued function that is even, increasing in \(\mathbb {R}_+\) and convex, and vanishes at 0. We will assume that it is sufficiently smooth—for example, that it has a second derivative, except, perhaps at \(\pm x_0\) for some fixed \(x_0\)—although, as will be clear from what follows, this assumption can be relaxed further.

Once the loss is selected, one can define the best element in the class, namely, a function in F that minimizes the average loss, or risk, \(\mathbb {E}\ell (f(X)-Y)\) (with the obvious underlying assumption that the minimizer exists). We denote that minimizer by \(f^*\).

Next, one may choose a procedure that uses the data \((X_i,Y_i)_{i=1}^N\) to produce a (random) function \(\hat{f} \in F\). The effectiveness of \(\hat{f}\) may be measured in several ways, and the two we focus on here lead to the notions of prediction and estimation.

Problem 1.2

Given a procedure \(\hat{f}\), find the ‘smallest’ functions \({\mathcal {E}}_p\) and \({\mathcal {E}}_{e}\) possible for which the following holds. If \(F \subset L_2(\mu )\) is a class of functions and Y is the unknown target, then with probability at least \(1-\delta \) over samples \((X_i,Y_i)_{i=1}^N\),

$$\begin{aligned} \mathbb {E}\left( \ell (\hat{f}(X)-Y)\big |(X_i,Y_i)_{i=1}^N\right) \le \inf _{f \in F} \mathbb {E}\ell (f(X)-Y) + {\mathcal {E}}_p. \end{aligned}$$

Alternatively, with probability at least \(1-\delta \),

$$\begin{aligned} \left\| \hat{f}-f^*\right\| _{L_2}^2 = \mathbb {E}\left( (\hat{f}-f^*)^2(X) \big | (X_i,Y_i)_{i=1}^N \right) \le {\mathcal {E}}_e. \end{aligned}$$

The functions \({\mathcal {E}}_p\) and \({\mathcal {E}}_e\) may depend on the structure of F, the sample size N, the confidence level \(\delta \), some ‘global’ properties of Y (e.g., its \(L_q\) norm), etc.

The prediction error \({\mathcal {E}}_p\) measures the ‘predictive capabilities’ of \(\hat{f}\), specifically whether \(\hat{f}\) is likely to be almost as effective as the best possible in the class—the latter being \(f^*\). The estimation error \({\mathcal {E}}_e\) measures the distance between \(\hat{f}\) and \(f^*\), with respect to the underlying \(L_2(\mu )\) metric.

Literature devoted to the study of prediction and estimation is extensive and goes well beyond what can be reasonably surveyed here. We refer the reader to the manuscripts [2, 4, 6, 9, 15, 24, 25] as possible starting points for information on the history of Problem 1.2, as well as for more recent progress.

The procedure we focus on here is empirical risk minimization (ERM), in which \({\hat{f}}\) is selected to be a function in F that minimizes the empirical risk

$$\begin{aligned} P_N \ell _f \equiv \frac{1}{N} \sum _{i=1}^N \ell (f(X_i)-Y_i); \end{aligned}$$

here, and throughout the article, \(P_N\) denotes the empirical mean associated with the random sample \((X_i,Y_i)_{i=1}^N\).

Since it is impossible to obtain nontrivial information on the performance of any procedure, including ERM, without imposing some assumptions on the class F, the target Y and the loss \(\ell \), one has to select a framework that, on the one hand, is general enough to include natural problems that one would like to study, but on the other, still allows one to derive significant results on prediction and estimation. Unfortunately, some of the assumptions that are commonly used in literature are highly restrictive, though seemingly benign. And among the more harmful assumptions are that the loss is a Lipschitz function and that functions in F and Y are uniformly bounded.

The origin of these assumptions is technical: they are an outcome of the ‘classical’ method of analysis used to tackle Problem 1.2. The method itself is based on tools from empirical processes theory, most notably, on contraction and concentration arguments that are simply false without imposing the right assumptions on the class, the target and the loss. However, these assumptions leave a large number of natural problems out of reach.

To explain why concentration and contraction arguments are so appealing, let us outline the standard method of analyzing data driven procedures like ERM.

The basic underlying assumption behind such procedures, and in particular, behind ERM, is that sampling mimics reality. Since one’s goal is to identify the function \(f^*\) which minimizes in F the functional \(f \rightarrow \mathbb {E}\ell (f(X)-Y)\), a natural course of action is to compare empirical means of the loss functional to the actual means in the hope that an empirical minimizer will be close to the true minimizer.

To that end, one may consider the excess loss functional associated with \(f \in F\)

$$\begin{aligned} {\mathcal {L}}_f(X,Y) = \ell (f(X)-Y)-\ell (f^*(X)-Y), \end{aligned}$$

observe that for every \(f \in F\), \(\mathbb {E}{\mathcal {L}}_f \ge 0\) and that if \(f^*\) is unique, equality is achieved only by \(f^*\). Also, since \({\mathcal {L}}_{f^*}=0\), it is evident that the empirical minimizer \(\hat{f}\) satisfies that \(P_N {\mathcal {L}}_{\hat{f}} \le 0\); thus, for every sample, the empirical minimizer belongs to the random set

$$\begin{aligned} \left\{ f \in F : P_N {\mathcal {L}}_f \le 0\right\} {.} \end{aligned}$$
(1.1)

The key point in the analysis of ERM is that the random set of potential minimizers consists of functions for which sampling behaves in an a-typical manner: \(P_N {\mathcal {L}}_f \le 0\) while \(\mathbb {E}{\mathcal {L}}_f > 0\). The hope is that one may identify the set by exploiting the discrepancy between the ‘empirical’ and ‘actual’ behaviour of means. For example, a solution to the prediction problem follows if the set (1.1) consists only of functions with ‘predictive capabilities’ that are close to the optimal in F, while the estimation problem may be resolved if (1.1) consists only of functions that are close to \(f^*\) with respect to the \(L_2\) distance.

What makes the nature of the set \(\{f : P_N {\mathcal {L}}_f \le 0\}\) rather elusive is not only the fact that it is random, but also that one has no real knowledge of the functions \(\ell _f=\ell (f(X)-Y)\) and \({\mathcal {L}}_f=\ell _f -\ell _{f^*}\): the two have unknown components—the target Y and the true minimizer \(f^*\).

Concentration and contraction help in identifying the set (1.1). By applying concentration results to a well-chosen subset of excess loss functions \(\{{\mathcal {L}}_f : f \in F^\prime \}\), one may show that there is a large subset \(F^\prime \subset F\), on which \(P_N {\mathcal {L}}_f\) cannot be too far from \(\mathbb {E}{\mathcal {L}}_f\) (or, for more sophisticated results, that the ratios \(P_N {\mathcal {L}}_f/\mathbb {E}{\mathcal {L}}_f\) cannot be too far from 1). Since \(\mathbb {E}{\mathcal {L}}_f > 0\) if \(f \not = f^*\), this forces \(f \rightarrow P_N {\mathcal {L}}_f\) to be positive on \(F^\prime \) and thus \(\hat{f} \in F \backslash F^\prime \).

Naturally, concentration results come at a cost, and for estimates such as

$$\begin{aligned} \sup _{f \in F^\prime } \left| P_N{\mathcal {L}}_f - \mathbb {E}{\mathcal {L}}_f\right|< \varepsilon \ \ {\mathrm{or}} \ \ \sup _{f \in F^\prime } \left| \frac{P_N {\mathcal {L}}_f}{\mathbb {E}{\mathcal {L}}_f}-1\right| < \varepsilon \end{aligned}$$
(1.2)

to hold with high probability requires strong assumptions on the random variables involved—for example, that functions in F and Y are uniformly bounded (see the books [3, 11] for more details on concentration of measure phenomena).

Finally, and what is possibly the most costly step in the classical method of analysis, is contraction. The contraction argument is based on the fact that class members and the target are uniformly bounded functions and that the loss is Lipschitz on the ranges of the functions \(f(X)-Y\). It implies that the supremum of the empirical process indexed by the (unknown!) excess loss class may be controlled in terms of the supremum of an empirical process indexed by functions of the form \(f-f^*\) (see, for example, [1, 9, 18] for more details).

One result that is based on the classical method and that utilizes the full strength of the two assumptions—that class members and the target are uniformly bounded and that the loss is Lipschitz—is Theorem 1.3 below, proved originally in [1]. It serves as a preliminary benchmark for our discussion.

Assume that F is a class of functions that are bounded by 1 and let Y be the target random variable that is also bounded by 1. Let \(\ell \) be a Lipschitz function with constant \(\Vert \ell \Vert _{\mathrm{lip}}\) on \([-2,2]\), which is an interval containing all the ranges of \(f(X)-Y\). Assume further that \(f^*\) exists and is unique and that for every \(f \in F\), \(\Vert f-f^*\Vert _{L_2}^2 \le B \mathbb {E}{\mathcal {L}}_f\), which is the significant part of the so-called Bernstein condition (see, e.g., [13, 16, 17]).

A standard example in which all these conditions hold is when F is a closed, convex class consisting of functions into \([-1,1]\), Y also maps into \([-1,1]\), and \(\ell (t)=t^2\). In that case it is straightforward to show that \(B=1\) and \(\Vert \ell \Vert _{\mathrm{lip}}=4\).

Let \(D_{f^*}\) be the \(L_2(\mu )\) ball of radius 1, centred at \(f^*\). Thus, \(\{f \in F : \Vert f-f^*\Vert _{L_2} \le r\} = F \cap rD_{f^*}\). For every \(r>0\), let

$$\begin{aligned} k_N(r)=\sup _{f \in F \cap rD_{f^*}} \left| \frac{1}{\sqrt{N}}\sum _{i=1}^N\varepsilon _i (f-f^*)(X_i)\right| , \end{aligned}$$
(1.3)

and

$$\begin{aligned} \bar{k}_N(r)=\mathbb {E}\sup _{f \in F \cap rD_{f^*}} \left| \frac{1}{\sqrt{N}}\sum _{i=1}^N\varepsilon _i (f-f^*)(X_i)\right| . \end{aligned}$$
(1.4)

Here \((\varepsilon _i)_{i=1}^N\) are independent, symmetric, \(\{-1,1\}\)-valued random variables that are independent of \((X_i)_{i=1}^N\) and the expectation is taken with respect to both \((X_i)_{i=1}^N\) and \((\varepsilon _i)_{i=1}^N\). Finally, set

$$\begin{aligned} k_N^*(\gamma ,\delta )=\inf \left\{ r>0: Pr \left( k_N(r/\left\| \ell \right\| _{\mathrm{lip}})\le \gamma r^2\sqrt{N}\right) \ge 1-\delta \right\} \end{aligned}$$

and

$$\begin{aligned} \bar{k}_N^*(\gamma )=\inf \left\{ r>0: \bar{k}_N(r/\left\| \ell \right\| _{\mathrm{lip}})\le \gamma r^2\sqrt{N}\right\} . \end{aligned}$$

Theorem 1.3

There exist absolute constants \(c_1\) and \(c_2\) for which the following holds. If F, Y and \(\ell \) are as above, then for every \(0<\delta <1\), with probability at least \(1-\delta \)

$$\begin{aligned} \mathbb {E}{\mathcal {L}}_{\hat{f}} \le c_1\max \left\{ \left( k_N^*\left( c_2\left( B\left\| \ell \right\| _{\mathrm{lip}}\right) ^{-1},\delta \right) \right) ^2,\frac{\left\| \ell \right\| _{\mathrm{lip}}^2B}{N}\right\} , \end{aligned}$$
(1.5)

and

$$\begin{aligned} \mathbb {E}{\mathcal {L}}_{\hat{f}} \le c_1\max \left\{ \left( \bar{k}_N^*\left( c_2(B\left\| \ell \right\| _{\mathrm{lip}})^{-1}\right) \right) ^2,(\left\| \ell \right\| _{\mathrm{lip}}^2B)\frac{\log (1/\delta )}{N}\right\} . \end{aligned}$$
(1.6)

Recalling that \(\Vert f-f^*\Vert _{L_2}^2 \le B \mathbb {E}{\mathcal {L}}_f\) for every \(f \in F\), analogous results hold for the estimation problem.

The proof of Theorem 1.3 and results of a similar nature may be found in [1, 9, 18].

Although Theorem 1.3 is one of the main benchmarks of the performance of ERM, it truly requires rather restrictive assumptions, as do all the results that are based on the concentration-contraction mechanism. For example, Theorem 1.3 cannot be used to tackle one of the most fundamental problems in Statistics—linear regression in \(\mathbb {R}^n\) relative to the squared loss and with independent additive gaussian noise:

Example 1.4

Let \(\ell (x)=x^2\). Given \(T \subset \mathbb {R}^n\), set \(F_T =\left\{ \bigl < t,\cdot \bigr > : t \in T\right\} \) to be the class of linear functionals on \(\mathbb {R}^n\) associated with T. Let \(\mu \) be a probability measure on \(\mathbb {R}^n\) and set X to be a random vector distributed according to \(\mu \). Let W be a standard gaussian variable that is independent of X and the target is \(Y=\bigl < t_0,\cdot \bigr >+W\) for some fixed but unknown \(t_0 \in T\).

Observe that

  • Y is not bounded (because of the gaussian noise).

  • Unless \(\mu \) is supported in a bounded set in \(\mathbb {R}^n\), functions in \(F_T\) are not bounded.

  • The loss \(\ell (x)=x^2\) satisfies a Lipschitz condition in \([-a,a]\) with a constant 2a. Unless \(\mu \) has a bounded support and Y is bounded, \(\ell \) does not satisfy a Lipschitz condition on an interval containing the ranges of the functions \(f(X)-Y\).

Each one of these observations is enough to place linear regression with independent additive gaussian noise outside the scope of Theorem 1.3, and what is equally alarming is that the same holds even if \(\mu \) is the standard gaussian measure on \(\mathbb {R}^n\), regardless of T, the choice of noise or even its existence.

An additional downside of Theorem 1.3 is that even in situations that do fall within its scope, resulting bounds are often less than satisfactory (see, for example, the discussion in [19]).

The suboptimal behaviour of Theorem 1.3 and, in fact, of the entire concentration-contraction mechanism, happens to be endemic: it is caused by the nature of the complexity parameter used to govern the rates \({\mathcal {E}}_p\) and \({\mathcal {E}}_e\). Indeed, when considering likely sources of error in prediction and estimation problems, two generic reasons come to mind:

  • \((X_1,..,X_N)\) is merely a sample and two functions in F can coincide on that sample but still be far from one another in \(L_2(\mu )\). This leads to the notion of the version space: a random subset of F, defined by

    $$\begin{aligned} \left\{ f \in F: f(X_i)=f^*(X_i) \ {\mathrm{for }}\,{\mathrm{every }}\ 1 \le i \le N\right\} \end{aligned}$$

    and which measures the way a random sample can be used to distinguish between class members. Clearly, the \(L_2(\mu )\) diameter of the version space is an intrinsic property of the class F and has nothing to do with the noiseFootnote 1 \(\xi =f^*(X)-Y\). Standard arguments show (see, e.g. [10]) that even in noise-free problems, when \(Y=f_0(X)\) for some \(f_0 \in F\), it is impossible to construct a learning procedure whose error rate consistently outperforms the \(L_2\) diameter of the version space.

  • Measurements are noisy: one does not observe \(f^*(X_i)\) but rather \(Y_i\). Known results as well as common sense indicate that the ‘closer’ Y is to F, the better the behaviour of \({\mathcal {E}}_p\) and \({\mathcal {E}}_e\) should be. Thus, it stands to reason that \({\mathcal {E}}_p\) and \({\mathcal {E}}_e\) should depend on the ‘noise level’ of the problem, i.e., on a natural distance between the target and the class.

With that in mind, it is reasonable to conjecture that \({\mathcal {E}}_p\) and \({\mathcal {E}}_e\) exhibit two regimes, captured by two different complexity parameters. Firstly, a ‘low noise’ regime, in which the ‘noise’ \(\xi =f^*(X)-Y\) is sufficiently close to zero in the right sense, and the behaviour of ERM is similar to its behaviour in the noise-free problem—essentially the \(L_2\) diameter of the version space. Secondly, a ‘high noise’ regime, in which mistakes occur because of the way the loss affects the interaction between class members and the noise.

Theorem 1.3 yields only one regime, and that regime is governed by a single complexity parameter. This parameter does not depend on the noise \(\xi =f^*(X)-Y\), except via a trivial \(L_\infty \) bound; rather, it depends on the correlation of the set \(\{(f(X_i))_{i=1}^N : f \in F\}\) (the so-called random coordinate projection of F) with a generic random noise model, represented by a random point in \(\{-1,1\}^N\). Obviously, the generic noise may have nothing to do with the actual noise one faces.

The main goal of this article is to address Problem 1.2 by showing that \({\mathcal {E}}_p\) and \({\mathcal {E}}_e\) indeed have two regimes. Each one of those regimes is captured by a different parameter: firstly, an ‘intrinsic parameter’ that governs low-noise problems (when Y is sufficiently close to F) and depends only on the class and not on the target or on the loss; secondly, an external parameter that captures the interaction of the class with the noise and with the loss, and dominates in high-noise situations, when Y is far from F.

Moreover, a solution to Problem 1.2 has to hold without the restrictive assumptions of the concentration-contraction mechanism, namely:

  • The class F need not be bounded in \(L_\infty \), but rather satisfies significantly weaker tail conditions.

  • The target Y need not be bounded (in fact, \(Y \in L_2\) suffices in most cases).

  • The loss function \(\ell \) need not be Lipschitz on an interval containing the ranges of \(f(X)-Y\).

The problem of ERM’s estimation error relative to the squared loss was studied in [19], leading to a satisfactory solution in that case. The aim here is to explore more general loss functions, not merely for the sake of generality, but because of a real side-effect of the squared loss: the combination of its rapid growth with heavy-tailed sampling inevitably leads to outliers. Roughly put, outliers are sample points that are misleading because they capture some a-typical behaviour of the sampled object. Outliers occur more frequently when dealing with heavy-tailed functions, and have a significant impact on ERM when the loss grows quickly. It is highly desirable to find a way of removing the ill-effects of outliers, and thanks to the general theory we develop here we are able to do just that: we show that if one chooses a loss that is calibrated to fit the noise level and the intrinsic structure of the underlying class, the ill-effects of outliers caused by the ‘noise’ are removed.

1.1 Basic definitions and some notation

Throughout the article, absolute constants are denoted by \(c_1,c_2,\ldots \); their values may change from line to line. We write \(A \lesssim B\) if there is an absolute constant \(c_1\) for which \(A \le c_1 B\), and \(A \sim B\) if \(c_1 A \le B \le c_2 A\) for absolute constants \(c_1\) and \(c_2\). \(A \lesssim _r B\) or \(A \sim _r B\) means that the constants depend on some parameter r. \(\kappa _0\), \(\kappa _1\),..., denote constants whose values remain unchanged.

Given a probability measure \(\mu \), set \(D=B(L_2(\mu ))\) to be the unit ball of \(L_2(\mu )\), let rD be the \(L_2(\mu )\) ball of radius r and put \(rD_f\) to be the \(L_2(\mu )\) ball centred at f and of radius r; \(S(L_2)\) denotes the unit sphere in \(L_2(\mu )\). From this point onward we do not specify the \(L_2\) space to which the functions in question belong, as that will be clear from the context.

Given \(1 \le p < \infty \), let \(B_p^n = \{x \in \mathbb {R}^n : \sum _{i=1}^n |x_i|^p \le 1 \}\) be the unit ball in the space \(\ell _p^n = (\mathbb {R}^n, \Vert \ \Vert _p)\), with the obvious modification when \(p=\infty \); set \(S^{n-1}\) to be the Euclidean unit sphere in \(\mathbb {R}^n\).

Let \(\{G_f : f \in F\}\) be the canonical gaussian process indexed by F with a covariance structure endowed by \(L_2(\mu )\) (see, e.g., [7] for a detailed survey on gaussian processes) and set

$$\begin{aligned} \mathbb {E}\left\| G\right\| _F = \sup \left\{ \mathbb {E}\sup _{h \in H} G_h : \ H \subset F, \ H \ {\mathrm{is}} \ {\mathrm{finite}} \right\} . \end{aligned}$$

Put \(d_F(L_2)=\sup _{f \in F} \Vert f\Vert _{L_2}\) and let

$$\begin{aligned} k_F=\left( \frac{\mathbb {E}\left\| G\right\| _F}{d_F(L_2)}\right) ^2, \end{aligned}$$

which is an extension of the celebrated Dvoretzky–Milman dimension of a convex body in \(\mathbb {R}^n\). We refer the reader to [22, 23] for more details on the Dvoretzky–Milman dimension and its role in asymptotic geometric analysis.

For \(\alpha \ge 1\), \(L_{\psi _\alpha }\) is the Orlicz space of all measurable functions for which the \(\psi _\alpha \) norm, defined by

$$\begin{aligned} \left\| f\right\| _{\psi _\alpha }=\inf \left\{ c>0 : \mathbb {E}\exp \left( \left| f/c\right| ^\alpha \right) \le 2 \right\} , \end{aligned}$$

is finite. Some basic facts on Orlicz spaces may be found, for example, in [25].

The most important Orlicz norm in our context is the \(\psi _2\) norm, which calibrates the subgaussian tail behaviour of a function. A class is L-subgaussian if the \(\psi _2\) and \(L_2\) norms are L-equivalent on F, that is, if for every \(f,h \in F \cup \{0\}\), \(\Vert f-h\Vert _{\psi _2} \le L\Vert f-h\Vert _{L_2}\). In particular, such norm equivalence implies that there are absolute constants \(c_1\) and \(c_2\) such that for every \(p \ge 2\), \(\Vert f-h\Vert _{L_p} \le c_1L\sqrt{p}\Vert f-h\Vert _{L_2}\) and for \(u \ge 1\), \(Pr(|f-h| > c_2uL\Vert f-h\Vert _{L_2}) \le 2\exp (-u^2/2)\).

A class of functions H is star-shaped around 0 if for every \(h \in H\) and every \(\lambda \in [0,1]\), \(\lambda h \in H\). In other words, if \(h \in H\) then H contains the entire interval connecting h to 0.

It is straightforward to verify that if F is convex and \(f \in F\) then \(H_f=F-f=\{h-f : h \in F\}\) is star-shaped around 0.

A class that is star-shaped around zero has some regularity. The star-shape property implies that if \(r < \rho \), then \(H \cap r S(L_2)\) contains a ‘scaled-down’ version of \(H \cap \rho S(L_2)\). Indeed, if \(h \in H \cap \rho S(L_2)\) and since \(r/\rho \in [0,1]\), it follows that \((r/\rho )h \in H \cap r S(L_2)\). In particular, normalized ‘layers’ of a star-shaped class become richer the closer the layer is to zero.

Finally, if A is a finite set, we denote by |A| its cardinality.

2 Beyond the squared loss

As noted previously, our goal is to extend the results established in [19] from the squared loss \(\ell (t)=t^2\) to a general smooth convex loss. For reasons we clarify later, the most interesting choices of loss functions satisfy some strong convexity property—either globally or in a neighbourhood of 0.

Definition 2.1

A function \(\ell \) is strongly convex with a constant \(c_0>0\) in the interval I if

$$\begin{aligned} \ell (y) \ge \ell (x)+\ell ^\prime (x)(y-x)+\frac{c_0}{2}(y-x)^2 \end{aligned}$$

for every \(x,y \in I\).

Note that if \( \inf _{x \in \mathbb {R}} \ell ^{\prime \prime }(x) \ge c>0\) then \(\ell \) is strongly convex in \(\mathbb {R}\) with a constant c, and one such example is the squared loss \(\ell (t)=t^2\).

If one wishes the loss to be convex, as we do, its growth from any point must be at least linear. Therefore, it seems natural to consider loss functions that are strongly convex in an interval around zero, thus mimicking the local behaviour of the squared loss, while away from zero exhibit a linear, or almost linear growth, hopefully limiting the negative effect of outliers.

Typical examples of such losses are the Huber loss with parameter \(\gamma \), defined by

$$\begin{aligned} \ell _\gamma (t)={\left\{ \begin{array}{ll} \frac{1}{2}t^2 &{} {\mathrm{if}} \ |t| \le \gamma \\ \gamma |t|-\frac{\gamma ^2}{2} &{} {\mathrm{if}} \ |t| > \gamma , \end{array}\right. } \end{aligned}$$
(2.1)

and a version of the log-lossFootnote 2

$$\begin{aligned} \ell (t)=-\log \left( \frac{4\exp (t)}{(1+\exp (t))^2}\right) , \end{aligned}$$
(2.2)

which is strongly convex in any bounded interval, but with a constant that decays exponentially to zero with the interval’s length (because \(\ell ^{\prime \prime }(t)= 2\exp (t)/(\exp (t)+1)^2\)).

As noted previously, the analysis of ERM is usually based on exclusion: showing that a large (random) part of the class cannot contain the empirical minimizer because the empirical risk functional is positive on functions that belong to that part. Strong convexity properties of the loss come in handy when exploring the decomposition of \({\mathcal {L}}_f(X,Y)\) via a Taylor expansion—a decomposition that leads naturally to an exclusion argument:

If \(\ell \) has a second derivative then for every (XY) there is a mid-point Z for which

$$\begin{aligned} {\mathcal {L}}_f (X,Y)&= \ell (f(X)-Y)-\ell \left( f^*(X)-Y\right) \\&= \ell ^\prime (\xi )(f-f^*)(X)+\frac{1}{2} \ell ^{\prime \prime }(Z)(f-f^*)^2(X)=(1)+(2). \end{aligned}$$

One may exclude \(H \subset F\) by showing that the empirical mean of the quadratic term (2) is positive on H, while the empirical mean of the multiplier component (1) cannot be very negative there. For such functions, \(P_N {\mathcal {L}}_f >0\), implying that \(\hat{f} \in F \backslash H\).

If \(\ell \) does not have a second derivative everywhere, one may modify this decomposition by noting that for every \(x_1\) and \(x_2\),

$$\begin{aligned} \ell (x_2)-\ell (x_1)=\int _{x_1}^{x_2} \ell ^\prime (w)dw=\ell ^\prime (x_1)(x_2-x_1)+\int _{x_1}^{x_2} \left( \ell ^\prime (w)-\ell ^\prime (x_1)\right) dw. \end{aligned}$$

Therefore, when applied to (XY) and a fixed \(f \in F\), and setting \(\xi =f^*(X)-Y\), the quadratic component in the decomposition is

$$\begin{aligned} \int _\xi ^{\xi +(f-f^*)(X)} \left( \ell ^\prime (w)-\ell ^\prime (\xi )\right) dw. \end{aligned}$$

In particular, it is straightforward to show that if \(\ell \) is twice differentiable in \(\mathbb {R}\), except, perhaps, at \(\pm x_0\), then for every XY one has

$$\begin{aligned} {\mathcal {L}}_f (X,Y) \ge \ell ^\prime (\xi )(f-f^*)(X)+\frac{1}{16} \ell ^{\prime \prime }(Z)(f-f^*)^2(X) \end{aligned}$$

for a well-chosen midpoint Z.

Definition 2.2

For every \(f,f^* \in F\) and (XY), set

$$\begin{aligned} {\mathcal {M}}_{f-f^*}(X,Y) = \ell ^\prime (\xi )(f-f^*)(X), \end{aligned}$$

and put

$$\begin{aligned} {\mathcal {Q}}_{f-f^*}(X,Y)=\int _\xi ^{\xi +(f-f^*)(X)} \left( \ell ^\prime (w)-\ell ^\prime (x)\right) dw, \end{aligned}$$

representing the multiplier and quadratic components of the excess loss \({\mathcal {L}}_f(X,Y)\).

Observe that a nontrivial, uniform, lower bound on \(\ell ^{\prime \prime }\) significantly simplifies the question of lower bounding the empirical means of the quadratic component. In fact, in such a case the problem reverts to the study of the quadratic component of the squared loss, explored in [19]. Our focus is on situations in which no such lower bound exists.

2.1 The exclusion argument in prediction and estimation

A structural assumption that is needed throughout this exposition is the following:

Assumption 2.1

Assume that for every \(f \in F\),

$$\begin{aligned} \mathbb {E}\ell ^\prime (\xi ) (f-f^*)(X) \ge 0. \end{aligned}$$

Assumption 2.1 is not really restrictive:

  • If \(\xi (X,Y)=f^*(X)-Y\) is independent of X (e.g., when \(Y=f_0(X)-W\) for some unknown \(f_0 \in F\) and an independent, mean-zero random variable W), then \(\mathbb {E}\ell ^\prime (\xi ) (f-f^*)(X)=0\), because \(\ell ^\prime \) is odd.

  • If F is a convex class of functions and \(\ell \) satisfies minimal integrability conditions, then \(\mathbb {E}\ell ^\prime (\xi ) (f-f^*)(X) \ge 0\) for every \(f \in F\). Indeed, if there is some \(f_1 \in F\) for which \(\mathbb {E}\ell ^\prime (\xi )(f_1-f^*)(X) < 0\), then by considering \(f_\lambda = \lambda f_1 + (1-\lambda )f^* \in F\) for \(\lambda \) close to 0,

    $$\begin{aligned} \mathbb {E}\ell \left( f_\lambda (X)-Y\right) < \mathbb {E}\ell \left( f^*(X)-Y\right) , \end{aligned}$$

    which is impossible.

Under Assumption 2.1, if \(\ell \) is twice differentiable then given a sample \((X_i,Y_i)_{i=1}^N\) and \(f \in F\), there are mid-points \(Z_i\) that belong to the interval whose end points are \(f^*(X_i)-Y_i\) and \(f(X_i)-Y_i=(f-f^*)(X_i)+\xi _i\), such that

$$\begin{aligned} P_N {\mathcal {L}}_f =&\frac{1}{N} \sum _{i=1}^N \ell (f(X_i)-Y_i) - \ell \left( f^*(X_i)-Y_i\right) \nonumber \\ \ge&\frac{1}{N} \sum _{i=1}^N \ell ^\prime (\xi _i) (f-f^*)(X_i) + \frac{1}{2N} \sum _{i=1}^N \ell ^{\prime \prime } (Z_i) (f-f^*)^2(X_i) \nonumber \\ \ge&-\left| \frac{1}{N} \sum _{i=1}^N \ell ^\prime (\xi _i) (f-f^*)(X_i)-\mathbb {E}\ell ^\prime (\xi ) (f-f^*)\right| + \mathbb {E}\ell ^\prime (\xi ) (f-f^*) \nonumber \\&+ \frac{1}{2N} \sum _{i=1}^N \ell ^{\prime \prime } (Z_i) (f-f^*)^2(X_i). \end{aligned}$$
(2.3)

Assume that on a high-probability event \({\mathcal {A}}\), for every \(f \in F\),

$$\begin{aligned} \left| \frac{1}{N}\sum _{i=1}^N \ell ^\prime (\xi _i)(f-f^*)(X_i) - \mathbb {E}\ell ^\prime (\xi )(f-f^*)(X)\right| \le \frac{\theta }{4}\max \left\{ \left\| f-f^*\right\| _{L_2}^2,r_M^2\right\} , \end{aligned}$$

for well chosen values \(r_M\) and \(\theta \). Assume further that on a high probability event \({\mathcal {B}}\), for every \(f \in F\) with \(\Vert f-f^*\Vert _{L_2} \ge r_Q\),

$$\begin{aligned} \frac{1}{N}\sum _{i=1}^N \ell ^{\prime \prime }(Z_i)(f-f^*)^2(X_i) \ge \theta \left\| f-f^*\right\| _{L_2}^2. \end{aligned}$$

Theorem 2.3

If F satisfies Assumption 2.1, then on the event \({\mathcal {A}} \cap {\mathcal {B}}\), \(\Vert \hat{f}-f^*\Vert _{L_2} \le \max \{r_M,r_Q\}\).

Proof

By Assumption 2.1 and (2.3),

$$\begin{aligned} P_N{\mathcal {L}}_f \ge&\frac{1}{2N}\sum _{i=1}^N \ell ^{\prime \prime }(Z_i)(f-f^*)^2(X_i)\\&- \left| \frac{1}{N}\sum _{i=1}^N \ell ^\prime (\xi _i)(f-f^*)(X_i) - \mathbb {E}\ell ^\prime (\xi )(f-f^*)(X)\right| . \end{aligned}$$

Hence, on the event \({\mathcal {A}} \cap {\mathcal {B}}\), if \(\Vert f-f^*\Vert _{L_2} \ge \max \{r_M,r_Q\}\) then \(P_N{\mathcal {L}}_f \ge (\theta /4)\Vert f-f^*\Vert _{L_2}^2 >0\), and f cannot be an empirical minimizer. \(\square \)

Theorem 2.3 implies that to bound the performance of ERM in the estimation problem it suffices to identify \(r_M\) and \(r_Q\) for which the event \({\mathcal {A}} \cap {\mathcal {B}}\) is sufficiently large.

Turning to the prediction problem, there is an additional assumption that is needed, namely, that \(\mathbb {E}{\mathcal {Q}}_{f-f^*}\) does not increase too quickly when f is close to \(f^*\).

Assumption 2.2

Assume that there is a constant \(\beta \) for which, for every \(f \in F\) with \(\Vert f-f^*\Vert _{L_2} \le \max \{r_M,r_Q\}\), one has

$$\begin{aligned} \mathbb {E}{\mathcal {Q}}_{f-f^*} \le \beta \left\| f-f^*\right\| _{L_2}^2. \end{aligned}$$

Clearly, if \(\ell ^\prime \) is a Lipschitz function, one may take \(\beta =\Vert \ell ^\prime \Vert _{\mathrm{lip}}\); hence, if \(\ell ^{\prime \prime }\) exists everywhere and is a bounded function, \(\beta \le \Vert \ell ^{\prime \prime }\Vert _{L_\infty }\). Moreover, even when \(\ell ^{\prime \prime }\) is not bounded, such a \(\beta \) exists if the functions \(f-f^*\) have well behaved tails relative to the growth of \(\ell ^{\prime \prime }\). Since the analysis required in these cases is rather obvious, we will not explore this issue further.

Theorem 2.4

Assume that the loss \(\ell \) is twice differentiable and satisfies Assumption 2.1 and Assumption 2.2. Using the notation introduced above, on the event \({\mathcal {A}} \cap {\mathcal {B}}\) one has

$$\begin{aligned} \mathbb {E}{\mathcal {L}}_{\hat{f}} \le 2(\theta +\beta ) \max \left\{ r_M^2,r_Q^2\right\} . \end{aligned}$$

Proof

Fix a sample in \({\mathcal {A}} \cap {\mathcal {B}}\). By Theorem 2.3, \(\Vert \hat{f}-f^*\Vert _{L_2} \le \max \{r_M,r_Q\}\). Thus, it suffices to show that if \(\Vert f-f^*\Vert _{L_2} \le \max \{r_M,r_Q\}\) and \(\mathbb {E}{\mathcal {L}}_f \ge 2(\theta +\beta ) \max \{r_M^2,r_Q^2\}\), then \(P_N{\mathcal {L}}_f >0\); in particular, such a function cannot be an empirical minimizer.

Note that for every \(f \in F\), \({\mathcal {L}}_f ={\mathcal {M}}_{f-f^*} + {\mathcal {Q}}_{f-f^*}\) and thus either \(\mathbb {E}{\mathcal {L}}_f \le 2\mathbb {E}{\mathcal {M}}_{f-f^*}=2\mathbb {E}\ell ^\prime (\xi )(f-f^*)(X)\) or \(\mathbb {E}{\mathcal {L}}_f \le 2\mathbb {E}{\mathcal {Q}}_{f-f^*}\).

However, if f satisfies the above, only the first option is possible; indeed, if \(\mathbb {E}{\mathcal {Q}}_{f-f^*}\) is dominant, then by Assumption 2.2,

$$\begin{aligned} \mathbb {E}{\mathcal {L}}_f \le 2\mathbb {E}{\mathcal {Q}}_{f-f^*} \le 2\beta \left\| f-f^*\right\| _{L_2}^2 \le 2\beta \max \left\{ r_M^2,r_Q^2\right\} , \end{aligned}$$

which is impossible by the choice of f. Therefore, it suffices to treat the case in which \(\mathbb {E}{\mathcal {L}}_f \le 2\mathbb {E}\ell ^\prime (\xi )(f-f^*)(X)\).

Fix such an \(f \in F\). Since \(\ell \) is convex, \(P_N {\mathcal {L}}_f \ge \frac{1}{N}\sum _{i=1}^N \xi _i(f-f^*)(X_i)\), and on \({\mathcal {A}} \cap {\mathcal {B}}\),

$$\begin{aligned} P_N {\mathcal {L}}_f \ge&\, \mathbb {E}\ell ^\prime (\xi )(f-f^*)(X) - \left| \frac{1}{N} \sum _{i=1}^N \ell ^\prime (\xi )(f-f^*)(X_i) - \mathbb {E}\ell ^\prime (\xi )(f-f^*)(X)\right| \\ \ge&\,\frac{1}{2}\mathbb {E}{\mathcal {L}}_f - \frac{\theta }{4} \max \left\{ r_M^2,\left\| f-f^*\right\| _{L_2}^2\right\} \\ \ge&\, (\theta +\beta ) \max \left\{ r_M^2,r_Q^2\right\} - \frac{\theta }{4}\max \left\{ r_M^2,r_Q^2\right\} >0. \end{aligned}$$

\(\square \)

Remark 2.5

When \(\ell \) is twice differentiable except perhaps at \(\pm x_0\), then Theorem 2.3 and Theorem 2.4 are still true though with modified constants.

In the following sections we develop the necessary machinery leading to a uniform lower estimate on the quadratic term \(f \rightarrow P_N {\mathcal {Q}}_{f-f^*}\) and to an upper estimate on the multiplier term \(f \rightarrow P_N{\mathcal {M}}_{f-f^*}\). Combining the two, we identify the values \(r_Q\) and \(r_M\), as well as the right choice of \(\theta \).

3 Preliminary estimates

Let \((Z_i)_{i=1}^N\) be independent copies of a random variable Z and set \((Z_i^*)_{i=1}^N\) to be the monotone non-increasing rearrangement of \((|Z_i|)_{i=1}^N\). Below we obtain upper and lower estimates on various functions of \((Z_i^*)_{i=1}^N\). All the estimates we present are straightforward applications of either a concentration inequality for \(\{0,1\}\)-valued random variables (selectors) with mean \(\delta \), or, alternatively, a rather crude binomial estimate. Indeed, given a property \({\mathcal {P}}\) set \(\delta _i = \mathbbm {1}_{\{Z_i \in {\mathcal {P}}\}}\)—the characteristic function of the event that \(Z_i\) satisfies property \({\mathcal {P}}\). Let \(\delta = Pr(Z \in {\mathcal {P}})\) and note that \(|\{i : Z_i \in {\mathcal {P}}\}| = \sum _{i=1}^N \delta _i\). By Bernstein’s inequality (see, for example, [3, 25]),

$$\begin{aligned} Pr\left( \left| \frac{1}{N}\sum _{i=1}^N \delta _i - \delta \right| \le t \right) \ge 1-2\exp \left( -cN \min \left\{ t^2/\delta ,t\right\} \right) \end{aligned}$$

for a suitable absolute constant c. Hence, taking \(t = u \delta \),

$$\begin{aligned} N\delta (1-u) \le \left| \left\{ i : Z_i \in {\mathcal {P}}\right\} \right| \le N\delta (1+u) \end{aligned}$$
(3.1)

with probability at least \(1-2\exp (-cN\delta \min \{u^2,u\})\).

The binomial estimate we employ is equally simple: it is based on the fact that

$$\begin{aligned} Pr\left( \left| \left\{ i : Z_i \in {\mathcal {P}}\right\} \right| \ge k\right) \le \left( {\begin{array}{c}N\\ k\end{array}}\right) Pr^k(Z \in {\mathcal {P}}) \le \left( \frac{eN}{k} \cdot Pr(Z \in {\mathcal {P}})\right) ^k. \end{aligned}$$

3.1 Tail-based upper estimates

Assume that one has information on \(\Vert Z\Vert _{L_q}\) for some \(q > 2\) and set \(L = \Vert Z\Vert _{L_q}/\Vert Z\Vert _{L_2}\). Applying Chebyshev’s inequality,

$$\begin{aligned} Pr\left( \left| Z\right| \ge w \left\| Z\right\| _{L_2}\right) \le \frac{\mathbb {E}\left| Z\right| ^q}{\left\| Z\right\| _{L_2}^q w^q} = \frac{L^q}{w^q}. \end{aligned}$$

Hence, if \({\mathcal {P}}=\{|Z| < w\Vert Z\Vert _{L_2}\}\) it follows that \(Pr(Z \in {\mathcal {P}}) \ge 1-(L/w)^q\), which can be made arbitrarily close to 1 by selecting w that is large enough. This observation implies that with high probability, an arbitrary large proportion of \(\{|Z_1|,\ldots ,|Z_N|\}\) are not very large.

Lemma 3.1

There exists absolute constants \(c_1\) and \(c_2\) for which the following holds. Let \(Z \in L_2\). For every \(0<\varepsilon <1\), with probability at least \(1-2\exp (-c_1\varepsilon N)\) there exists a subset \(I \subset \{1,\ldots ,N\}\), \(|I| \ge (1-\varepsilon )N\), and for every \(i \in I\),

$$\begin{aligned} \left| Z_i\right| \le c_2\varepsilon ^{-1/2}\left\| Z\right\| _{L_2}. \end{aligned}$$

Proof

Fix \(\varepsilon \) as above and note that \(Pr(|Z| \ge 2\Vert Z\Vert _{L_2}/\sqrt{\varepsilon }) \le \varepsilon /4\). Hence, by a binomial estimate,

$$\begin{aligned}&Pr\left( \left| \left\{ i: \left| Z_i\right| \ge 2\left\| Z\right\| _{L_2}/\sqrt{\varepsilon }\right\} \right| \ge N\varepsilon \right) \le \left( {\begin{array}{c}N\\ \varepsilon N\end{array}}\right) Pr^{\varepsilon N} \left( \left| Z\right| \ge 2\left\| Z\right\| _{L_2}/\sqrt{\varepsilon }\right) \\&\quad \le \left( \frac{e}{\varepsilon }\right) ^{N \varepsilon } \cdot \left( \frac{\varepsilon }{4}\right) ^{N\varepsilon } \le \exp \left( -c N \varepsilon \right) , \end{aligned}$$

for a suitable absolute constant c. \(\square \)

Given a vector \(a=(a_i)_{i=1}^N\), the \(L_q\) norm of a, when considered as a function on \(\Omega =\{1,\ldots ,N\}\) endowed with the uniform probability measure, is

$$\begin{aligned} \left\| a\right\| _{L_q^N}=\left( \frac{1}{N}\sum _{i=1}^N \left| a_i\right| ^q\right) ^{1/q}. \end{aligned}$$

The weak-\(L_q\) norm of the vector a is

$$\begin{aligned} \left\| a\right\| _{L_{q,\infty }^N}=\inf \left\{ c>0: d_a(t) \le (c/t)^q \ {\mathrm{for}} \ {\mathrm{every }}\ t>0\right\} , \end{aligned}$$

where \(d_a(t)=N^{-1}|\{i: |a_i|>t\}|\).

The next observation is that sampling preserves the \(L_q\) structure of Z, in the sense that if \(Z \in L_q\), then with high probability, \(\Vert (Z_i)_{i=1}^N\Vert _{L_{q,\infty }^N} \lesssim \Vert Z\Vert _{L_q}\).

Lemma 3.2

Let \(1 \le q \le r\). If \(Z \in L_r\), \(u \ge 2\) and \(1 \le k \le N/2\), then

$$\begin{aligned} Z_k^* \le u(N/k)^{1/q}\left\| Z\right\| _{L_r} \end{aligned}$$

with probability at least \(1-u^{-kr}\left( \frac{eN}{k}\right) ^{-k((r/q)-1)}\).

In particular, with probability at least \(1-2u^{-r}N^{-((r/q)-1)}\),

$$\begin{aligned} \left\| (Z_i)\right\| _{L_{q,\infty }^N} \le u\left\| Z\right\| _{L_r}. \end{aligned}$$

Proof

Let \(\eta =(r/q)-1\), fix \(1 \le k \le N/2\) and set \(v>0\) to be named later. The binomial estimate implies that

$$\begin{aligned}&Pr\left( Z_k^* \ge v(eN/k)^{(1+\eta )/r} \left\| Z\right\| _{L_r}\right) \le \left( {\begin{array}{c}N\\ k\end{array}}\right) Pr^k \left( \left| Z\right| \ge v(eN/k)^{(1+\eta )/r}\left\| Z\right\| _{L_r}\right) \\&\quad \le \left( \frac{eN}{k}\right) ^k \left( \frac{k}{eN}\right) ^{(1+\eta )k} \cdot v^{-kr} =\left( \frac{e N}{k}\right) ^{-\eta k} v^{-k r}. \end{aligned}$$

In particular, for \(v=u (eN/k)^{1/q - (1+\eta )/r}\),

$$\begin{aligned} Z_k^* \le u(N/k)^{1/q} \left\| Z\right\| _{L_r} \end{aligned}$$

with probability at least

$$\begin{aligned} 1-u^{-kr}\left( \frac{eN}{k}\right) ^{k\left( (r/q)-1\right) }. \end{aligned}$$

The second part of the claim follows by summing up the probabilities for \(k \le N/2\), using that \(Z_k^* \le Z_{N/2}^*\) for \(k \ge N/2\) and that \((u^{-kr})_{k=1}^{N/2}\) is a geometric progression. \(\square \)

Remark 3.3

Note that similar statements to Lemma 3.2 are true if one simply assumes that \(Pr(|Z| \ge t)<\varepsilon \), even without moment assumptions. Of course, under such an assumption one has no information whatsoever on the largest \(\varepsilon N\) coordinates of \((|Z_1|,\ldots ,|Z_N|)\), but rather, only on a certain proportion that is slightly smaller than \((1-\varepsilon )N\) of the smallest coordinates.

Also, observe that \(\Vert (Z_i^*)_{i \ge j}\Vert _{L_{q,\infty }^N} \lesssim \Vert Z\Vert _{L_q}\) with a probability estimate that improves exponentially in j.

3.2 Lower estimates using a small-ball property

A similar line of reasoning to the one used above is true for lower estimates, and is based on a small-ball condition:

Definition 3.4

A random variable Z satisfies a small-ball condition with constants \(\kappa >0\) and \(0<\varepsilon <1\) if

$$\begin{aligned} Pr\left( \left| Z\right| \ge \kappa \left\| Z\right\| _{L_2}\right) \ge \varepsilon . \end{aligned}$$

A class of functions F satisfies a small-ball property with constants \(\kappa \) and \(0<\varepsilon <1\) if for every \(f \in F\),

$$\begin{aligned} Pr\left( \left| f\right| \ge \kappa \left\| f\right\| _{L_2}\right) \ge \varepsilon . \end{aligned}$$

Remark 3.5

Because the applications considered below require that many of the \(|Z_i|\)’s are at least of the order of \(\Vert Z\Vert _{L_2}\), the \(L_2\) norm is used as a point of reference in the definition of the small-ball condition—though the notion of ‘small-ball’ can be modified to fit other norms, as well as in situations in which Z need not even be integrable.

This small-ball condition was introduced in the context of estimation problems in [19], and is the most important feature of our presentation. It is a rather weak assumption that is almost universally satisfied and eliminates the need to guarantee two-sided concentration, which is a far-more restrictive phenomenon. Unlike concentration, a small-ball condition captures that behaviour of a random variable close to zero: by considering \(h=Z/\Vert Z\Vert _{L_2}\), it quantifies the weight that h assigns to a neighbourhood of zero. In particular, if \(Z \in L_2\) and \(Z \not \equiv 0\), then it satisfies a small-ball condition for some constants \(\varepsilon \) and \(\kappa \). It is also important to emphasize that a small-ball condition has nothing to do with the tail behaviour of Z (other than the obvious assumption that Z is square-integrable). For example, if Z is mean-zero, variance 1, random variable that has a density bounded by M, it satisfies a small-ball condition with constants that depend only on M—though it is possible that Z does not have any moment beyond the second one.

The proofs we present below are based on the assumption that the class satisfies a small-ball property, which is simply a small-ball condition with fixed constants that holds for every function in the classFootnote 3.

Naturally, for such an approach to be of any use, one must show that there are enough natural situations in which a small-ball property holds, and as indications let us describe several cases involving classes of linear functionals on \(\mathbb {R}^n\).

Let \(\mu \) be a probability measure on \(\mathbb {R}^n\) and let X be distributed according to \(\mu \). If there are constants \(\kappa \) and \(\varepsilon \) such that for every \(t \in \mathbb {R}^n\),

$$\begin{aligned} Pr\left( \left| \bigl< t,X \bigr>\right| \ge \kappa \left\| \bigl < t,X \bigr >\right\| _{L_2} \right) \ge \varepsilon \end{aligned}$$
(3.2)

then it immediately follows that any class of functions \(\{\bigl < t,\cdot \bigr > : t \in T\}\) satisfies the small-ball property with constants \(\kappa \) and \(\varepsilon \).

The simplest situation in which (3.2) holds for every \(t \in \mathbb {R}^n\) is when there is some \(q>2\) for which the \(L_q\) and \(L_2\) norms of linear functionals are equivalent; that is,

$$\begin{aligned} \sup _{t \in S^{n-1}} \frac{\left\| \bigl< t,X \bigr>\right\| _{L_q}}{\left\| \bigl < t,X \bigr >\right\| _{L_2}} \le L. \end{aligned}$$
(3.3)

A straightforward application of the Paley-Zygmund inequality (see, e.g. [5]) shows that (3.2) holds for constants \(\kappa \) and \(\varepsilon \) that depend only on q and L, and in particular, are independent of the dimension n of the underlying space. As an extreme example, if \(0<\alpha \le 2\) and X is a \(\Psi _\alpha \) random vector in \(\mathbb {R}^n\) (e.g., if X is subgaussian or log-concave) then (3.2) holds—though this is an obvious overkill: a small-ball condition like in (3.2) is true under a minimal norm equivalence like (3.3) and does not require the high moment equivalence \(\Vert \bigl< X,t \bigr>\Vert _{L_p} \le cp^{1/\alpha }\Vert \bigl < X,t \bigr >\Vert _{L_2}\) that is ensured by a \(\Psi _\alpha -L_2\) norm equivalence.

To indicate yet again the clear difference between a small-ball condition and tail estimates, let us construct random vectors that need not have any moment beyond the second one and still satisfy (3.2). One way of generating such random vectors is based on the following observation:

Lemma 3.6

There are absolute constants \(c_0\) and \(c_1\) for which the following holds.

  1. (1)

    Let z be a random variable that satisfies a small-ball condition with constants \(\kappa \) and \(\varepsilon \). Let \(z_1,\ldots ,z_n\) to be independent copies of z and put \(Z=(z_i)_{i=1}^n\). Then for every \(t \in \mathbb {R}^n\), \(\bigl < t,Z \bigr>\) satisfies a small-ball condition with constants \(c_0 \sqrt{\varepsilon } \kappa \) and \(c_1 \varepsilon \).

  2. (2)

    Let \(Z=(z_1,\ldots ,z_n)\) be an unconditional random vector (that is, Z has the same distribution as \((\varepsilon _i z_i)_{i=1}^n\) for any choice of signs \((\varepsilon _i)_{i=1}^n\)). Assume that for every \(1 \le i \le n\), \(\mathbb {E}z_i^2 =1\) and \(|z_i| \le M\) almost surely. Then, for every \(t \in \mathbb {R}^n\), \(\bigl < t,Z \bigr>\) satisfies a small-ball condition with constants \(c_2\) and \(c_3\) depending only on M.

The proof of Lemma 3.6 is rather standard and we only sketch it for the sake of completeness. The idea is that for every \(t \in S^{n-1}\), the random variable \(\bigl < t,Z \bigr >=\sum _{i=1}^n t_iz_i\) has the same distribution as \(\sum _{i=1}^n \varepsilon _i |t_i z_i|\), where \((\varepsilon _i)_{i=1}^n\) are independent random signs that are also independent of \((z_i)_{i=1}^n\). Note that conditioned on \((z_i)_{i=1}^n\), \(\sum _{i=1}^n \varepsilon _i |t_i z_i|\) satisfies a small-ball condition with absolute constants; indeed, this is an immediate outcome of the Paley-Zygmund inequality and the fact that there is an absolute constant L such that for every fixed \((a_i)_{i=1}^n \in \mathbb {R}^n\), \(\Vert \sum _{i=1}^n \varepsilon _i a_i\Vert _{L_2} \le L \Vert \sum _{i=1}^n \varepsilon _i a_i\Vert _{L_1}\). Therefore, there are constants \(c_1(L)\) and \(c_2(L)\) such that

$$\begin{aligned} Pr_\varepsilon \left( \big |\sum _{i=1}^n \varepsilon _i \left| t_i z_i\right| \big |> c_1\left( \sum _{i=1}^n t_i^2 z_i^2\right) ^{1/2}\right) \ge c_2 \end{aligned}$$

and all that remains is to show that with nontrivial probability, \(\sum _{i=1}^n t_i^2 z_i^2 \ge c \sum _{i=1}^n t_i^2 \mathbb {E}z_i^2\).

In (2), let \(w=\sum _{i=1}^n t_i^2 z_i^2\). By the \(L_\infty \) bound on the \(z_i\)’s and since \(\mathbb {E}z_i^2=1\),

$$\begin{aligned} \mathbb {E}w^2 = \mathbb {E}\left( \sum _{i=1}^n t_i^2 z_i^2\right) ^2 \le M^4 \left( \sum _{i=1}^n t_i^2\right) ^2 = M^4 \left( \mathbb {E}\left( \sum _{i=1}^n t_i^2 z_i^2\right) \right) ^2 = M^4 \left( \mathbb {E}\left| w\right| \right) ^2. \end{aligned}$$

Hence \(\Vert w\Vert _{L_2} \le M^2 \Vert w\Vert _{L_1}\) and the required estimate follows from the Paley-Zygmund inequality.

In (1), by the small-ball property of the \(z_i\)’s,

$$\begin{aligned} \sum _{i=1}^n t_i^2 z_i^2 \ge \kappa ^2 \sum _{i=1}^n t_i^2 (\mathbb {E}z_i^2) \cdot \mathbbm {1}_{\left\{ z_i^2 \ge \kappa \mathbb {E}z_i^2\right\} } = \kappa ^2 \sum _{i=1}^n t_i^2 (\mathbb {E}z_i^2) \cdot \delta _i, \end{aligned}$$

where \((\delta _i)_{i=1}^n\) independent selectors with mean at least \(\varepsilon \). Again, one concludes by invoking the Paley-Zygmund inequality for \(\sum _{i=1}^n t_i^2 (\mathbb {E}z_i^2) \cdot \delta _i\). \(\square \)

Lemma 3.7

There exists an absolute constant c for which the following holds. Assume that Z satisfies a small-ball condition with constants \(\kappa \) and \(\varepsilon \) and let \((Z_i)_{i=1}^N\) be independent copies of Z. Then, with probability at least \(1-2\exp (-cN\varepsilon )\), there is a subset I of \(\{1,\ldots ,N\}\) of cardinality at least \((3/4)\varepsilon N\) and for every \(i \in I\), \(|Z_i| \ge \kappa \Vert Z\Vert _{L_2}\).

The proof, which we omit, is an immediate application of Bernstein’s inequality for the i.i.d. selectors \(\delta _i = \mathbbm {1}_{\{|Z_i| \ge \kappa \Vert Z\Vert _{L_2}\}}\) for the choice of \(u=1/4\) in (3.1). Naturally, at a price of a weaker probability estimate the constant 3 / 4 in Lemma 3.7 can be made arbitrarily close to 1.

Combining the upper estimate from Lemma 3.1 and lower one from Lemma 3.7 yields the following corollary:

Corollary 3.8

There exist absolute constants \(c_1\) and \(c_2\) for which the following holds. Assume that \(Z \in L_2\) and that it satisfies a small-ball condition with constants \(\kappa \) and \(\varepsilon \). Then, with probability at least \(1-2\exp (-c_1\varepsilon N)\), there is \(J \subset \{1,\ldots ,N\}\), \(|J| \ge \varepsilon N/2\) and for every \(j \in J\),

$$\begin{aligned} \kappa \left\| Z\right\| _{L_2} \le \left| Z_j\right| \le c_2 \left\| Z\right\| _{L_2}/\sqrt{\varepsilon }. \end{aligned}$$

Corollary 3.8 allows one to control the behaviour of \((Z_i)_{i=1}^N\) on a subset of \(\{1,\ldots ,N\}\) of cardinality \(\sim \varepsilon N\), and with exponentially high probability. Moreover, by modifying \(c_1\) and \(c_2\), the cardinality of J can be made arbitrarily close to \(\varepsilon N\).

Remark 3.9

Note that by the union bound, a version of Corollary 3.8 holds uniformly for a collection of \(\exp (c_1N\varepsilon /2)\) random variables with probability at least \(1-2\exp (-c_1N\varepsilon /2)\).

3.3 Uniform estimates on a class

Next, we use the estimates obtained above and study the structure of a typical coordinate projection of a class H, \(P_\sigma H = \{ (h(X_i))_{i=1}^N : h \in H\}\). We show that with high probability, for every function in H of sufficiently large \(L_2\) norm, most of the coordinates of \(P_\sigma h\) are of the order of \(\Vert h\Vert _{L_2}\). Such a result is an extension of Corollary 3.8—from a single function to a class of functions—and the class we focus on in what follows is \(H_{f^*}=\{f-f^* : f \in F\}\).

To quantify what is meant by “sufficiently large \(L_2\) norm”, recall that for \(H \subset L_2(\mu )\), \(\{G_h : h \in H\}\) is the canonical gaussian process indexed by H with a covariance structure endowed by \(L_2(\mu )\).

Definition 3.10

Given a class of functions \(H \subset L_2(\mu )\), a sample size N and positive constants \(\zeta _1\) and \(\zeta _2\) set

$$\begin{aligned} r_{1,Q}(H,N,\zeta _1)=\inf \left\{ r>0 : \mathbb {E}\left\| G\right\| _{H \cap rD} \le \zeta _1 r \sqrt{N} \right\} , \end{aligned}$$
(3.4)

and put

$$\begin{aligned} r_{2,Q}(H,N,\zeta _2)=\inf \left\{ r>0 : \mathbb {E}\sup _{H \cap r D} \left| \frac{1}{\sqrt{N}} \sum _{i=1}^N \varepsilon _i h(X_i) \right| \le \zeta _2 r \sqrt{N}\right\} , \end{aligned}$$
(3.5)

where D is, as always, the unit ball of \(L_2(\mu )\).

When the class H and sample size N are obvious from the context, we denote the fixed points by \(r_{1,Q}(\zeta _1)\) and \(r_{2,Q}(\zeta _2)\) respectively.

Finally, set

$$\begin{aligned} r_Q(\zeta _1,\zeta _2)=r_Q(H,N,\zeta _1,\zeta _2) = \max \left\{ r_{1,Q}(\zeta _1),r_{2,Q}(\zeta _2)\right\} . \end{aligned}$$

If H is star-shaped around 0, it is straightforward to show that when \(r > r_{1,Q}(\zeta _1)\), one has \(\mathbb {E}\Vert G\Vert _{H \cap r D} \le \zeta _1 r \sqrt{N}\), while if \(r<r_{1,Q}(\zeta _1)\), \(\mathbb {E}\Vert G\Vert _{H \cap r D} \ge \zeta _1 r \sqrt{N}\). A similar observation is true for \(r_{Q,2}(\zeta _2)\).

Theorem 3.11

There exist absolute constants \(c_0,c_1,c_2,c_3,c_4\) and \(c_5\) for which the following holds. Let H be a class of functions that is star-shaped around 0 and that satisfies a small-ball property with constants \(\kappa _0\) and \(\varepsilon \). If \(\zeta _1=c_1\kappa _0\varepsilon ^{3/2}\), \(\zeta _2=c_2 \kappa _0\varepsilon \) and \(r>r_Q(\zeta _1,\zeta _2)\), there is \(V_r \subset H \cap rS(L_2)\) and an event \(\Omega ^\prime \) of probability at least \(1-2\exp (-c_0\varepsilon ^2 N)\), such that:

  1. (1)

    \(|V_r| \le \exp (c_3 \varepsilon N)\) and \(c_3 \le 1/1000\).

  2. (2)

    On the event \(\Omega ^\prime \), for every \(v \in V_r\) there is a subset \(I_v \subset \{1,\ldots ,N\}\), \(|I_v| \ge \varepsilon N/2\) and for every \(i \in I_v\),

    $$\begin{aligned} \kappa _0 r \le \left| v(X_i)\right| \le c_4r/\sqrt{\varepsilon }. \end{aligned}$$
  3. (3)

    On the event \(\Omega ^\prime \), for every \(h \in H \cap r S(L_2)\) there is some \(v \in V_r\) and a subset \(J_h \subset I_v\), consisting of at least 3 / 4 of the coordinates of \(I_v\) (and in particular, \(|J_h| \ge \varepsilon N/4\)), and for every \(j \in J_h\),

    $$\begin{aligned} (\kappa _0/2)\left\| h\right\| _{L_2} \le \left| h(X_j)\right| \le c_5 \left( \kappa _0+1/\sqrt{\varepsilon }\right) \left\| h\right\| _{L_2} \end{aligned}$$

    and

    $$\begin{aligned} {\mathrm{sgn}}(h(X_j))={\mathrm{sgn}}(v(X_j)). \end{aligned}$$

The idea of the proof is to find an appropriate net in \(H \cap r S(L_2)\) (the set \(V_r\)), and show that each point in the net has many ‘well-behaved’ coordinates in the sense of (2). Also, if \(\pi h\) denotes the best approximation of \(h \in H \cap r S(L_2)\) in \(V_r\) with respect to the \(L_2\) norm, then

$$\begin{aligned} \sup _{h \in H \cap r S(L_2)} \frac{1}{N} \sum _{i=1}^N \left| h-\pi h\right| (X_i) \end{aligned}$$

is not very big, implying that \(|(h-\pi h)(X_i)|\) cannot have too many large coordinates. Since \(h(X_i) = (\pi h)(X_i) + (h-\pi h)(X_i)\), the first term is dominant on a proportional number of coordinates, leading to (3).

Proof

Recall that by Corollary 3.8, if \(Z \in L_2\) satisfies the small-ball condition with constants \(\kappa _0\) and \(\varepsilon \) then with probability at least \(1-2\exp (-c_1\varepsilon N)\), there is \(I \subset \{1,\ldots ,N\}\), \(|I| \ge \varepsilon N/2\) and for every \(i \in I\),

$$\begin{aligned} \kappa _0 \left\| Z\right\| _{L_2} \le \left| Z_i\right| \le c_2 \left\| Z\right\| _{L_2}/\sqrt{\varepsilon }. \end{aligned}$$

Fix \(\zeta _1\) and \(\zeta _2\) to be named later, let \(r>r_Q(\zeta _1,\zeta _2)\) and set \(c_1^\prime =\min \{c_1,1/500\}\). Let \(V_r \subset H \cap r S(L_2)\) be a maximal separated set whose cardinality is at most \(\exp (c_1^\prime \varepsilon N/2)\) and denote its mesh by \(\eta \). Therefore, by Corollary 3.8 and the union bound, it follows that with probability at least \(1-2\exp (-c_1 \varepsilon N/2)\), for every \(v \in V_r\) there is a subset \(I_v\) as above, i.e., \(|I_v| \ge \varepsilon N/2\) and for every \(i \in I_v\),

$$\begin{aligned} \kappa _0 r = \kappa _0 \left\| v\right\| _{L_2} \le \left| v(X_i)\right| \le c_2 \left\| v\right\| _{L_2}/\sqrt{\varepsilon } = c_2 r/\sqrt{\varepsilon }. \end{aligned}$$

Moreover, by Sudakov’s inequality (see, e.g. [7, 12, 23]) and because \(r \ge r_{Q,1}(\zeta _1)\), it follows that

$$\begin{aligned} \eta \le c_3 \frac{\mathbb {E}\left\| G\right\| _{H \cap r S(L_2)}}{\sqrt{c_1^\prime N \varepsilon /2}} \le \left( c_4 \zeta _1/\sqrt{\varepsilon }\right) r \end{aligned}$$

for \(c_4=\sqrt{2}c_3/\sqrt{c_1^\prime }\).

For every \(h \in H \cap r S(L_2)\), let \(\pi h \in V_r\) such that \(\Vert h-\pi h\Vert _{L_2} \le \eta \), set \(u_h=\mathbbm {1}_{\{|h-\pi h| > \kappa _0r/2\}}\) and put

$$\begin{aligned} U_r=\left\{ u_h : h \in H \cap r S(L_2) \right\} . \end{aligned}$$

Let \(\phi (t)=t/(\kappa _0 r/2)\) and note that pointwise, for every \(u_h \in U_r\), \(u_h(X) \le \phi (|h-\pi h|(X))\).

Applying the Giné–Zinn symmetrization theorem (see, e.g., [25]) and recalling that \(r >r_{Q,2}(\zeta _2)\), one has

$$\begin{aligned}&\mathbb {E}\sup _{u_h \in U_r} \frac{1}{N} \sum _{i=1}^N u_h(X_i) \le \mathbb {E}\sup _{h \in H \cap r S(L_2)} \frac{1}{N}\sum _{i=1}^N \phi \left( \left| h-\pi h\right| (X_i)\right) \\&\quad \le \mathbb {E}\sup _{h \in H \cap rS(L_2)} \left| \frac{1}{N}\sum _{i=1}^N \phi \left( \left| h-\pi h\right| (X_i)\right) - \mathbb {E}\phi \left( \left| h-\pi h\right| (X_i)\right) \right| \\&\qquad +\,\sup _{h \in H \cap r S(L_2)} \mathbb {E}\phi \left( \left| h-\pi h\right| \right) \\&\quad \le \frac{4}{\kappa _0 r} \cdot \left( \mathbb {E}\sup _{h \in H \cap rS(L_2)} \left| \frac{1}{N} \sum _{i=1}^N \varepsilon _i (h-\pi h)(X_i) \right| +\sup _{h \in H \cap r S(L_2)} \left\| h-\pi h\right\| _{L_2} \right) \\&\quad \le \frac{4}{\kappa _0 r} \cdot \left( 2\zeta _2 r + \eta \right) \le \frac{\varepsilon }{32}, \end{aligned}$$

provided that \(\zeta _1 \sim \kappa _0 \varepsilon ^{3/2} \) and \(\zeta _2 \sim \kappa _0 \varepsilon \).

Let \(\psi (X_1,\ldots ,X_N)=\sup _{u \in U_r} \frac{1}{N} \sum _{i=1}^N u(X_i)\). By the bounded differences inequality (see, for example, [3]), with probability at least \(1-\exp (-c_5t^2)\),

$$\begin{aligned} \psi (X_1,\ldots ,X_N) \le \mathbb {E}\psi + \frac{t}{\sqrt{N}}. \end{aligned}$$

Thus, for \(t=\varepsilon \sqrt{N}/32\), with probability at least \(1-\exp (-c_6 \varepsilon ^2 N)\), \(\psi (X_1,\ldots ,X_N) \le \varepsilon /16\), implying that for every \(h \in H \cap r S(L_2)\),

$$\begin{aligned} \left| \left\{ i : \left| h-\pi h\right| (X_i) \le (\kappa _0/2)r\right\} \right| \ge \left( 1-\frac{\varepsilon }{16}\right) N. \end{aligned}$$

Recall that \(\pi h \in V_r\) and that \(|I_{\pi h}| \ge \varepsilon N/2\). Let

$$\begin{aligned} J_h=\left\{ j: \left| h-\pi h\right| (X_j) \le (\kappa _0/2)r\right\} \cap I_{\pi h} \end{aligned}$$

and thus \(J_h\) consists of at least 3 / 4 of the coordinates in \(I_{\pi h}\) and \(|J_h| \ge \varepsilon N/4\). Moreover, for every \(j \in J_h\),

$$\begin{aligned} \left| h(X_j)\right| \ge \left| \pi h (X_j)\right| - \left| (h-\pi h)(X_j)\right| \ge \kappa _0 r - (\kappa _0/2)r =(\kappa _0/2)r, \end{aligned}$$
(3.6)

which also shows that \({\mathrm{sgn}}(h(X_j))={\mathrm{sgn}}(\pi h(X_j))\).

The upper estimate follows from a similar argument, using that \(|h(X_j)| \le |\pi h (X_j)| + |(h-\pi h)(X_j)|\). \(\square \)

Remark 3.12

Observe that by the star-shape property of H, if \(\rho _1> \rho _2\), then

$$\begin{aligned} \frac{1}{\rho _1}\left( H \cap \rho _1 S(L_2)\right) \subset \frac{1}{\rho _2}\left( H \cap \rho _2 S(L_2)\right) . \end{aligned}$$

Therefore, certain features of \(H \cap \rho _2 S(L_2)\) are automatically transferred to \(H \cap \rho _1 S(L_2)\), and in particular, a version of Theorem 3.11 holds uniformly for every level that is ‘larger’ than \(2r_Q(\zeta _1,\zeta _2)\). Indeed, assume that one has chosen \(\rho _2 =2r_Q\) in Theorem 3.11 and fix \(h \in H \cap \rho _1 S(L_2)\). By applying Theorem 3.11 to \(h^\prime =(\rho _2/\rho _1)h \in H \cap \rho _2 S(L_2)\) it follows that on the event \(\Omega ^\prime \) there is a subset J of \(\{1,\ldots ,N\}\) of cardinality at least \(\varepsilon N/4\) on which

$$\begin{aligned} \left| h(X_j)\right| \ge (\kappa _0/2)\rho _1 \ \ {\mathrm{and}} \ \ {\mathrm{sgn}}(h(X_j))={\mathrm{sgn}}\left( \pi h^\prime \right) (X_j). \end{aligned}$$

Next, let \(F \subset L_2\) be a convex set, fix \(f^* \in F\) and put \(H_{f^*}=\{f-f^* : f \in F\}\). Since \(H_{f^*}\) is clearly star-shaped around 0 and \(H_{f^*} \subset F-F\) one has:

Corollary 3.13

If F is a convex class of functions, \(F-F\) satisfies a small-ball property with constants \(\kappa _0\) and \(\varepsilon \), and \(r=2r_Q(F-F,N,\zeta _1,\zeta _2)\), then with probability at least \(1-2\exp (-c_0\varepsilon ^2 N)\), the following holds. For every \(f_1,f_2 \in F\) that satisfy \(\Vert f_1-f_2\Vert _{L_2} \ge r\), there is a subset \(J_{f_1,f_2} \subset \{1,\ldots ,N\}\) of cardinality at least \(\varepsilon N/4\) and for every \(j \in J_{f_1,f_2}\),

$$\begin{aligned} \left| (f_1-f_2)(X_j)\right| \ge (\kappa _0/2)\left\| f_1-f_2\right\| _{L_2}. \end{aligned}$$

In particular, on the same event,

$$\begin{aligned} \inf _{\left\{ f \in F : \left\| f-f^*\right\| _{L_2} \ge 2r_Q\right\} } \frac{1}{N}\sum _{i=1}^N \left( \frac{{f-f^*}}{\left\| f-f^*\right\| _{L_2}}\right) ^2(X_i) \ge \frac{\varepsilon \kappa _0^2}{16}. \end{aligned}$$

4 The quadratic component of the loss

Following the path of the exclusion argument, the aim is to show that the quadratic component of the loss is sufficiently positive. And, although the results are formulated below in full generality, there are three examples that one should keep in mind: First, when \(\ell \) is strongly convex; second, when \(\ell \) is a general loss function and \(Y=f_0(X)+W\) for some \(f_0 \in F\) and a symmetric random variable W that is independent of X; finally, a situation that is, in some sense, a mixture of the two: a loss function that is guaranteed to be strongly convex only in a neighbourhood of 0 (for example, the loss functions (2.1) and (2.2)), and without assuming that the noise \(Y-f^*(X)\) is independent of X.

Throughout this section we assume that F is a convex class of functions and that \(F-F=\{f-h : f,h \in F\}\) satisfies a small-ball property with constants \(\kappa _0\) and \(\varepsilon \). Set \(r_Q=r_Q(F-F,N,\zeta _1,\zeta _2)\) with the choice of \(\zeta _1\) and \(\zeta _2\) as in Theorem 3.11—namely, \(\zeta _1 \sim \kappa _0\varepsilon ^{3/2}\) and \(\zeta _2 \sim \kappa _0\varepsilon \). Also assume that \(\ell ^{\prime \prime }\) exists everywhere except perhaps at \(\pm x_0\) and set for every \(0<t_1<t_2\)

$$\begin{aligned} \rho (t_1,t_2)=\inf \left\{ \ell ^{\prime \prime }(x) : x \in [t_1,t_2], \ x \not = \pm x_0\right\} . \end{aligned}$$
(4.1)

The following lower bound on the quadratic component in the strongly convex case is an immediate application of Corollary 3.13 and the fact that \({\mathcal {Q}}_{f-f^*}(X,Y) \gtrsim \ell ^{\prime \prime }(Z)(f-f^*)^2(X)\) for an appropriate mid-point Z. Its proof is omitted.

Theorem 4.1

There exists an absolute constant \(c_1\) for which the following holds. If \(\inf _{x \in \mathbb {R}\backslash \{\pm x_0\}} \ell ^{\prime \prime }(x) \ge 2c_0\), then with probability at least \(1-2\exp (-c_1 N \varepsilon ^2)\), for every \(f \in F\) with \(\Vert f-f^*\Vert _{L_2} \ge 2r_Q\),

$$\begin{aligned} P_N {\mathcal {Q}}_{f-f^*} \ge \frac{c_0\varepsilon \kappa _0^2}{16} \left\| f-f^*\right\| _{L_2}^2. \end{aligned}$$

Theorem 4.1 generalizes a similar result from [19] for the squared loss.

Turning to the more difficult (and interesting) problem of a loss that need not be strongly convex, we begin with the case of independent noise.

Assumption 4.1

Assume that \(Y=f_0(X)+W\) for \(f_0 \in F\) and a symmetric random variable \(W \in L_2\) that is independent of X and for which

$$\begin{aligned} Pr\left( \left| W\right| \le \kappa _1\left\| W\right\| _{L_2}\right) \le \varepsilon /1000. \end{aligned}$$
(4.2)

Clearly, \(f^*=f_0\); moreover, (4.2) is a rather minimal assumption, as a small-ball condition for a single function at one level holds when the function is absolutely continuous, by selecting the right value \(\kappa _1\).

Given a sample \((X_i,Y_i)_{i=1}^N\) let \(W_i=Y_i-f^*(X_i)\) and set \(Z_i\) to be the mid-points in the lower bound on the quadratic component of \(\ell \)—again using the fact that for the losses in question, \({\mathcal {Q}}_{f-f^*}(X_i,Y_i) \gtrsim \ell ^{\prime \prime }(Z_i)(f-f^*)^2(X_i)\).

Theorem 4.2

There exist absolute constants \(c_1,c_2,c_3\) and \(c_4\) for which the following holds. Let F and W be as above. With probability at least \(1-2\exp (-c_1\varepsilon ^2N)\), for every \(f \in F\) that satisfies \(\Vert f-f^*\Vert _{L_2} \ge 2r_Q\) one has

$$\begin{aligned} P_N {\mathcal {Q}}_{f-f^*} \ge c_2 \varepsilon \kappa _0^2 \rho (t_1,t_2) \left\| f-f^*\right\| _{L_2}^2. \end{aligned}$$

where

$$\begin{aligned} t_1=\kappa _1 \left\| W\right\| _{L_2} \ \ {\mathrm{and}} \ \ t_2= c_3\varepsilon ^{-1/2}\left\| W\right\| _{L_2} + c_4\left( \kappa _0+\varepsilon ^{-1/2}\right) \left\| f-f^*\right\| _{L_2}. \end{aligned}$$

The proof of Theorem 4.2 is based on several observations leading to accurate information on the ‘location’ of the midpoints \(Z_i\) in the lower bound on \({\mathcal {Q}}_{f-f^*}(X_i,Y_i)\). For every (XY), the corresponding mid-point belongs to interval whose end-points are \((f-f^*)(X)-W\) and \(-W\). If \(I_f\) is the set of coordinates on which \(|(f-f^*)(X_i)|\) is of the order of \(\Vert f-f^*\Vert _{L_2}\), and since X and W are independent and W is symmetric, then on roughly half of these coordinates the signs of \((f-f^*)(X_i)\) coincide with the signs of \(-W_i\). For those coordinates,

$$\begin{aligned} \left| Z_i\right| \in \left[ \left| W_i\right| ,\left| W_i\right| +\left| \left( f-f^*\right) (X_i)\right| \right] . \end{aligned}$$

Moreover, by excluding a further, sufficiently small proportion of the coordinates in \(I_f\) it follows that \(|W_i| \sim \Vert W\Vert _{L_2}\), as long as W is not highly concentrated around zero.

The difficulty is in making this argument uniform, in the sense that it should hold for every \(f \in F\) rather than for a specific choice of f. The first step towards such a uniform result is the following lemma.

Lemma 4.3

Let \(1 \le k \le m/40\) and set \({\mathcal {S}} \subset \{-1,0,1\}^m\) of cardinality at most \(\exp (k)\). For every \(s=(s(i))_{i=1}^m \in {\mathcal {S}}\) set \(I_s=\{i : s(i) \not = 0\}\) and assume that \(|I_s| \ge 40 k\). If \((\varepsilon _i)_{i=1}^m\) are independent, symmetric \(\{-1,1\}\)-valued random variables then with probability at least \(1-2\exp (-k)\), for every \(s \in {\mathcal {S}}\),

$$\begin{aligned} \left| \left\{ i \in I_s : {\mathrm{sgn}}(s(i))=\varepsilon _i\right\} \right| \ge \left| I_s\right| /3. \end{aligned}$$

Proof

For every fixed \(s \in {\mathcal {S}}\), the event \(|\{i: (s(i))_{i \in I_s}=\varepsilon _i\}| \ge \ell \) has the same distribution as \(|\{i \in I_s : \varepsilon _i =1\}| \ge \ell \). If \((\delta _i)_{i \in I_s}\) are selectors of mean 1 / 2 then

$$\begin{aligned} Pr\left( \left| \left\{ i \in I_s: s(i)=\varepsilon _i\right\} \right| \ge \ell \right) =Pr \left( \sum _{i=1}^{\left| I_s\right| } \delta _i \ge \ell \right) =(*). \end{aligned}$$

Applying Bernstein’s inequality for \(\ell =|I_s|/3\),

$$\begin{aligned} (*) \ge 1-Pr\left( \left| \frac{1}{|I_s|}\sum _{i=1}^{\left| I_s\right| } \delta _i -\frac{1}{2}\right| \ge \frac{1}{6}\right) \ge 1-2\exp \left( -\left| I_s\right| /20\right) , \end{aligned}$$

and by the union bound,

$$\begin{aligned}&Pr\left( {\mathrm{for}} \ {\mathrm{every }} \ s \in {\mathcal {S}}, \ \left| \left\{ i \in I_s : s(i)=\varepsilon _i\right\} \right| \ge \left| I_s\right| /3 \right) \ge 1-2\left| {\mathcal {S}}\right| \exp \left( -|I_s|/20\right) \\&\quad \ge 1-2\exp (-k). \end{aligned}$$

\(\square \)

Let r be as in Theorem 3.11 for the class \(H_{f^*}=F-f^*\) and set \(\Omega ^\prime \) to be the event on which the assertion of Theorem 3.11 holds. Using the notation of that theorem, consider \(r=2r_Q\) and the set \(V_r\). For every \(v \in V_r\) and a sample \((X_1,\ldots ,X_N) \in \Omega ^\prime \), let \(I_v = \{ i : \kappa _0 r \le |v(X_i)| \le c_1r/\sqrt{\varepsilon } \}\),

$$\begin{aligned} s_v=\bigl ({\mathrm{sgn}}(v(X_i)) \cdot \mathbbm {1}_{I_v}(X_i)\bigr )_{i=1}^N \ \ {\mathrm{and}} \ \ {\mathcal {S}}=\left\{ s_v : v \in V_r\right\} \subset \left\{ -1,0,1\right\} ^N. \end{aligned}$$

By Theorem 3.11, \(Pr(\Omega ^\prime ) \ge 1-2\exp (-c_2\varepsilon ^2 N)\) and on \(\Omega ^\prime \),

$$\begin{aligned} \left| {\mathcal {S}}\right| \le \exp \left( \varepsilon N/1000\right) \ \ {\mathrm{and}} \ \ \min _{v \in V} \left| I_{v}\right| \ge \varepsilon N/2. \end{aligned}$$

Lemma 4.4

Conditioned on \(\Omega ^\prime \), with probability at least \(1-2\exp (-c_0 \varepsilon N)\) with respect to the uniform measure on \(\{-1,1\}^N\), the following holds. For every \(h \in H_{f^*}\) with \(\Vert h\Vert _{L_2} \ge r\), there is subset \({\mathcal {I}}_h \subset \{1,\ldots ,N\}\) of cardinality at least \(\varepsilon N/24\), and for every \(i \in {\mathcal {I}}_h\),

$$\begin{aligned} (\kappa _0/2) \left\| h\right\| _{L_2} \le \left| h(X_i)\right| \le c_1 \left( \kappa _0+1/\sqrt{\varepsilon }\right) \left\| h\right\| _{L_2} \ \ {\mathrm{and}} \ \ {\mathrm{sgn}}(h(X_i)) = \varepsilon _i. \end{aligned}$$

Proof

Fix \(h \in H\) with \(\Vert h\Vert _{L_2} = r\) and let \(\pi h= v \in V_r\) be as in Theorem 3.11. Recall that there is a subset \(J_h \subset I_v\) consisting of at least 3 / 4 of the coordinates of \(I_v\), such that for every \(j \in J_h\),

$$\begin{aligned} (\kappa _0/2)r \le \left| h(X_j)\right| \le c_1 \left( \kappa _0+1/\sqrt{\varepsilon }\right) r \ \ \ {\mathrm{and}} \ \ \ {\mathrm{sgn}}(h(X_j))={\mathrm{sgn}}(v(X_j)). \end{aligned}$$

Applying Lemma 4.3 to the set \({\mathcal {S}}=\{s_v : v \in V_r\}\) for \(k=\varepsilon N /1000\), and noting that for every \(s_v \in {\mathcal {S}}\), \(|\{i: s_v(i) \not = 0\}| \ge \varepsilon N/2 \ge 40 k\), it follows that with probability at least \(1-2\exp (-c_2 \varepsilon N)\) (relative to the uniform measure on \(\{-1,1\}^N\)), for every \(v \in V_r\), \(s_v(i)=\varepsilon _i\) on at least 1 / 3 of the coordinates that belong to \(I_v\).

Since the set \(J_h\) contains at least 3 / 4 of the coordinates of \(I_v\) and \(v(X_i)=\varepsilon _i\) on at least a 1 / 3 of the coordinates of \(I_v\) it follows that on the coordinates that belong to the intersection of these two sets (at least 1 / 12 of the coordinates in \(I_v\)), both conditions hold, as claimed.

Finally, the claim is positive homogeneous and because \(H_{f^*}\) is star-shaped around 0, it holds on the same event when \(\Vert h\Vert _{L_2} \ge r\). \(\square \)

Corollary 4.5

There exist absolute constants \(c_0\) and \(c_1\) for which the following holds. Let F and W be as above. With probability at least \(1-2\exp (-c_0\varepsilon ^2 N)\) with respect to the product measure \((X \otimes W)^N\), for every \(f \in F\) with \(\Vert f-f^*\Vert _{L_2} \ge 2r_Q\) there is a (random) subset \({\mathcal {J}}_f \subset \{1,\ldots ,N\}\) of cardinality at least \(\varepsilon N/100\), and for every \(j \in {\mathcal {J}}_f\),

  1. 1.

    \((\kappa _0/2)\Vert f-f^*\Vert _{L_2} \le |(f-f^*)(X_j)| \le c_1 (\kappa _0+1/\sqrt{\varepsilon })\Vert f-f^*\Vert _{L_2}\),

  2. 2.

    \({\mathrm{sgn}}((f-f^*)(X_i)) = {\mathrm{sgn}}(-W)\), and

  3. 3.

    \(\kappa _1 \Vert W\Vert _{L_2} \le |W_j| \le c_2\Vert W\Vert _{L_2}/\sqrt{\varepsilon }\).

Proof

Recall that W is symmetric and therefore it has the same distribution as \(\eta |W|\) for a symmetric \(\{-1,1\}\)-valued random variable \(\eta \) that is independent of |W| and of X.

Considering \((\eta _i |W_i|)_{i=1}^N\), a direct application of Lemma 4.4 shows that with probability at least \(1-2\exp (-c_0\varepsilon ^2 N)\), if \(\Vert f-f^*\Vert _{L_2} \ge 2r_Q\), there is a subset \({\mathcal {I}}_f \subset \{1,\ldots ,N\}\) of cardinality at least \(\varepsilon N/24\), and for every \(i \in {\mathcal {I}}_f\),

$$\begin{aligned} (\kappa _0/2)\left\| f-f^*\right\| _{L_2} \le \left| (f-f^*)(X_j)\right| \le c_1 \left( \kappa _0+1/\sqrt{\varepsilon }\right) \left\| f-f^*\right\| _{L_2} \end{aligned}$$

and

$$\begin{aligned} {\mathrm{sgn}}\left( (f-f^*)(X_i)\right) = {\mathrm{sgn}}(-\eta _i). \end{aligned}$$

The final component is that for many of the coordinates in \({\mathcal {I}}_f\), \(|W_i| \sim \Vert W\Vert _{L_2}\). Indeed, by excluding the largest and smallest \(\varepsilon N/200\) coordinates of \((|W_i|)_{i \in {\mathcal {I}}_f}\), one obtains a subset \({\mathcal {J}}_f \subset {\mathcal {I}}_f\) of cardinality at least \(\varepsilon N/100\), and for every \(j \in {\mathcal {J}}_f\),

$$\begin{aligned} W^*_{N(1-\varepsilon /200)} \le \left| W_j\right| \le W^*_{\varepsilon N/200}, \end{aligned}$$

where \((W_i^*)_{i=1}^N\) is the non-increasing rearrangement of \((|W_i|)_{i=1}^N\).

Observe that by Lemma 3.1 applied to \(\varepsilon ^\prime =\varepsilon /200\), with probability at least \(1-2\exp (-c_2 N\varepsilon )\),

$$\begin{aligned} W^*_{\varepsilon N/200} \le c_3\varepsilon ^{-1/2}\left\| W\right\| _{L_2}. \end{aligned}$$

Moreover, recalling that \(Pr(|W| \le \kappa _1 \Vert W\Vert _{L_2}) \le \varepsilon /1000\), a straightforward application of a binomial estimate shows that with probability at least \(1-2\exp (-c_4 N \varepsilon )\), there are at most \(\varepsilon N/200\) \(W_i\)’s that satisfy \(|W_i| < \kappa _1 \Vert W\Vert _{L_2}\). Therefore, on that event,

$$\begin{aligned} W^*_{(1-\varepsilon /200)N} \ge \kappa _1 \left\| W\right\| _{L_2}, \end{aligned}$$

completing the proof. \(\square \)

Proof of Theorem 4.2

The convexity of \(\ell \) implies that \({\mathcal {Q}}_{f-f^*}\) is nonnegative. Consider the event from Corollary 4.5, and given \(f \in F\) for which \(\Vert f-f^*\Vert _{L_2} \ge 2r_Q\) let \({\mathcal {J}}_f \subset \{1,\ldots ,N\}\) be the set of coordinates as above: that is, for every \(j \in {\mathcal {J}}_f\),

$$\begin{aligned} (\kappa _0/2) \left\| f-f^*\right\| _{L_2} \le \left| (f-f^*)\right| (X_j) \le c_1\left( \kappa _0+ 1\sqrt{\varepsilon }\right) \left\| f-f^*\right\| _{L_2} \end{aligned}$$

and if \(j \in {\mathcal {J}}_f\), \(-W_j\) and \((f-f^*)(X_j)\) share the same sign—say positive. Thus, the mid-point \(Z_j\) belongs to the interval whose end-points are \(t_1=\kappa _1 \Vert W\Vert _{L_2}\) and \(t_2=c_1(\kappa _0+ 1\sqrt{\varepsilon })\Vert f-f^*\Vert _{L_2}+c_2\Vert W\Vert _{L_2}/\sqrt{\varepsilon }\), implying that

$$\begin{aligned} P_N {\mathcal {Q}}_{f-f^*} \ge \frac{1}{N}\sum _{j \in {\mathcal {J}}_f} \ell ^{\prime \prime }(Z_i) (f-f^*)^2(X_i) \ge c_3 \varepsilon \rho (t_1,t_2) \kappa _0^2 \left\| f-f^*\right\| _{L_2}^2. \end{aligned}$$

\(\square \)

Next, consider the general noise scenario, in which \(\xi =f^*(X)-Y\) need not be independent of X, nor does it necessarily satisfy a small-ball condition.

It is straightforward to verify that the only place in the proof above where the assumption that \(\xi \) and X are independent has been used, was to find a large subset of \(\{1,\ldots ,N\}\) on which \((f-f^*)(X_i)\) and \(\xi _i\) share the same sign. Also, the assumption that the noise satisfies a small-ball condition is only used to show that many of the \(|\xi _i|\)’s are sufficiently large—at least of the order of \(\Vert \xi \Vert _{L_2}\). Both components are not needed if one wishes to show that \(|Z_i| \le c(\kappa _0,\varepsilon )(\Vert \xi \Vert _{L_2}+\Vert f-f^*\Vert _{L_2})\) for a proportional number of coordinates.

Indeed, with high probability, if \(\Vert f-f^*\Vert _{L_2} \ge 2r_Q\), there is a subset of \(\{1,\ldots ,N\}\) of cardinality at least \(\varepsilon N/100\) on which

$$\begin{aligned} \left| Z_i\right| \lesssim \left( \kappa _0 + 1/\sqrt{\varepsilon }\right) \cdot \left( \left\| f-f^*\right\| _{L_2}+\left\| \xi \right\| _{L_2}\right) . \end{aligned}$$

Formally, one has:

Theorem 4.6

There exist absolute constants \(c_0,c_1\) and \(c_2\) for which the following holds. Let F be as above, set \(Y \in L_2\) and put \(\xi =f^*(X)-Y\). Then, with probability at least \(1-2\exp (-c_0\varepsilon ^2N)\), for every \(f \in F\) with \(\Vert f-f^*\Vert _{L_2} \ge 2r_Q\),

$$\begin{aligned} P_N {\mathcal {Q}}_{f-f^*} \ge c_1\varepsilon \kappa _0^2 \rho (0,t) \left\| f-f^*\right\| _{L_2}^2, \end{aligned}$$

for \(t=c_2(\kappa _0+\varepsilon ^{-1/2})\cdot (\Vert f-f^*\Vert _{L_2}+\Vert \xi \Vert _{L_2})\).

Remark 4.7

  1. (1)

    The assumption in Theorem 4.2 that the noise is independent of X allows one to obtain a positive lower bound on \(t_1\) of the order of the variance \(\Vert W\Vert _{L_2}\). This is significant when the loss function is not strongly convex in a large enough neighbourhood of zero (for example, when \(\ell (t)=t^p\) for \(p>2\)).

  2. (2)

    When F is bounded in \(L_2\), one may use the trivial bound \(\Vert f-f^*\Vert _{L_2} \le {\mathrm{diam}}(F,L_2)\) and replace \(t_2\) by \(c(\varepsilon ,\kappa _0)(\Vert \xi \Vert _{L_2} + {\mathrm{diam}}(F,L_2))\). This is of little importance when \(\ell ^{\prime \prime }\) decreases slowly, but matters a great deal when, for example, it has a compact support. Consider, for example, the Huber loss with parameter \(\gamma \). If \(\gamma \sim \Vert \xi \Vert _{L_2} + {\mathrm{diam}}(F,L_2)\) then \(\rho (0,t_2)=1\), but as stated, for a smaller value of \(\gamma \), \(\rho (0,t)=0\) leading to a useless estimate on the quadratic component.

    It turns out that one may improve Theorem 4.6 dramatically by ruling-out functions in F for which \(\Vert f-f^*\Vert _{L_2}\) is significantly larger than \(\Vert \xi \Vert _{L_2}\) as potential empirical minimizers, implying that one may set \(t=c(\kappa _0,\varepsilon )\Vert \xi \Vert _{L_2}\) rather than \(t=c(\kappa _0,\varepsilon )(\Vert \xi \Vert _{L_2}+{\mathrm{diam}}(F,L_2))\). We present this additional exclusion argument in Section 5.1.

5 Error estimates and oracle inequalities

Let us explore the multiplier component of the process, defined by \(f \rightarrow \frac{1}{N}\sum _{i=1}^N \ell ^\prime (\xi _i)(f-f^*)(X_i)\).

The following complexity term may be used to control the multiplier process, and is similar to the one used in [19].

Definition 5.1

Given a loss function \(\ell \), let \(\phi ^\ell _N(r)\) be the random function

$$\begin{aligned} \phi _N^\ell (r)=\frac{1}{\sqrt{N}} \sup _{\left\{ f \in F: \left\| f-f^*\right\| _{L_2} \le r\right\} } \left| \sum _{i=1}^N \varepsilon _i \ell ^\prime (\xi _i)(f-f^*)(X_i)\right| . \end{aligned}$$

Note that the randomness of \(\phi ^\ell _N(r)\) is in \((\xi _i)_{i=1}^N\), \((X_i)_{i=1}^N\) and the independent random signs \((\varepsilon _i)_{i=1}^N\).

Set

$$\begin{aligned} r_{M}^\prime (\kappa ,\delta )=\inf \left\{ r > 0 : Pr \left( \phi _N^\ell (r) \le r^2 \kappa \sqrt{N} \right) \ge 1-\delta \right\} , \end{aligned}$$

recall that \(H_{f^*}=F-f^*\), put

$$\begin{aligned} r_0(\kappa ) = \inf \left\{ r : \sup _{h \in H_{f^*} \cap rD} \left\| \ell ^\prime (\xi )h(X)\right\| _{L_2} \le \sqrt{N} \kappa r^2/4 \right\} \end{aligned}$$

and let

$$\begin{aligned} r_M(\kappa ,\delta )=r_M^\prime (\kappa ,\delta ) + r_0(\kappa ). \end{aligned}$$

The function \(\phi _N^\ell (r)\) and the definition of \(r_M\) arise naturally from a symmetrization argument. Indeed, it is well known that

$$\begin{aligned}&Pr\left( \sup _{h \in H_{f^*} \cap r D} \left| \frac{1}{N}\sum _{i=1}^N \ell ^\prime (\xi _i)h(X_i) - \mathbb {E}\ell ^\prime (\xi )h(X)\right|> x \right) \nonumber \\&\quad \le 2 Pr \left( \sup _{h \in H_{f^*} \cap r D} \left| \frac{1}{N}\sum _{i=1}^N \varepsilon _i \ell ^\prime (\xi _i)h(X_i)\right| > \frac{x}{4}\right) , \end{aligned}$$
(5.1)

provided that \(x \ge 4N^{-1/2}\sup _{h \in H_{f^*} \cap rD} \Vert \ell ^\prime (\xi )h(X)\Vert _{L_2}\) (see, e.g. [8])

Lemma 5.2

If F is a convex class of functions and \(r =2 r_M(\kappa /4,\delta /2)\) then with probability at least \(1-\delta \), for every \(f \in F\) such that \(\Vert f-f^*\Vert _{L_2} \ge r\), one has

$$\begin{aligned} \left| \frac{1}{N}\sum _{i=1}^N \ell ^\prime (\xi _i)(f-f^*)(X_i) - \mathbb {E}\ell ^\prime (\xi )(f-f^*)(X) \right| \le \kappa \max \left\{ \left\| f-f^*\right\| _{L_2}^2,r^2\right\} . \end{aligned}$$

Proof

The class F is convex and therefore \(H_{f^*}=F-f^*\) is star-shaped around 0. Also, \(r \ge r_0\), implying that \(4N^{-1/2}\sup _{h \in H_{f^*} \cap rD} \Vert \ell ^\prime (\xi )h(X)\Vert _{L_2} \le \kappa r^2\). By (5.1) for \(x=\kappa r^2\), one has

$$\begin{aligned}&Pr\left( \sup _{h \in H_{f^*} \cap r D} \left| \frac{1}{N}\sum _{i=1}^N \ell ^\prime (\xi _i)h(X_i) - \mathbb {E}\ell ^\prime (\xi )h(X)\right|> \kappa r^2 \right) \\&\quad \le 2 Pr \left( \phi _N^\ell (r)> \frac{ \kappa r^2}{4}\right) \le \delta , \end{aligned}$$

because \(r>r_M^\prime (\kappa /4,\delta /2)\).

Using once again that \(H_{f^*}\) is star-shaped around 0, it follows that if \(\Vert f-f^*\Vert _{L_2} \ge r\) then \(r(f-f^*)/\Vert f-f^*\Vert _{L_2} \in H_{f^*} \cap r S(L_2)\); thus,

$$\begin{aligned} \left| \frac{1}{N}\sum _{i=1}^N \ell ^\prime (\xi _i)(f-f^*)(X_i) - \mathbb {E}\ell ^\prime (\xi )(f-f^*)(X) \right| \le \kappa \left\| f-f^*\right\| _{L_2}^2. \end{aligned}$$

\(\square \)

The function \(\phi _N^\ell (r)\) is a natural geometric parameter: it is the ‘weighted width’ of the coordinate projection of \(H_{f^*} \cap rD\) in a random direction selected according to a symmetrized noise vector, which is a point-wise product of a vector that belongs to the combinatorial cube \(\{-1,1\}^N\) and the ‘noise multipliers’ \((\ell ^{\prime }(\xi _i))_{i=1}^N\) for \(\xi _i=f^*(X_i)-Y_i\).

The more standard counterparts of \(\phi _N^\ell (r)\), appearing in Theorem 1.3 and in similar results of that flavour, are the random function

$$\begin{aligned} \sup _{f \in F \cap rD_{f^*}} \left| \frac{1}{N}\sum _{i=1}^N \varepsilon _i (f-f^*)(X_i)\right| ; \end{aligned}$$

the conditional expectation—the so-called Rademacher average

$$\begin{aligned} \mathbb {E}_\varepsilon \left( \sup _{f \in F \cap rD_{f^*}} \left| \frac{1}{N}\sum _{i=1}^N \varepsilon _i (f-f^*)(X_i)\right| \ \big | (X_i)_{i=1}^N\right) ; \end{aligned}$$

and its expectation with respect to both \((\varepsilon _i)_{i=1}^N \otimes (X_i)_{i=1}^N\). Those represent the width or average width of

$$\begin{aligned} P_\sigma (H_{f^*} \cap r D) = \left\{ \left( (f-f^*)(X_i)\right) _{i=1}^N : f \in F, \ \left\| f-f^*\right\| _{L_2} \le r \right\} \end{aligned}$$

relative to a generic noise model, given by the random vector \((\varepsilon _i)_{i=1}^N\) weighted by the constant \(\Vert \ell \Vert _{\mathrm{lip}}\). A contraction argument shows that the typical width relative to a ‘generic noise’ vector dominates the typical width relative to the natural noise vector \((\varepsilon _i \ell ^\prime (\xi _i))_{i=1}^N\), implying that \(\phi _N^\ell \) is inherently superior to the generic complexity term.

Remark 5.3

It is straightforward to verify that when F consists of heavy-tailed random variables or if Y is a heavy-tailed random variable then the random sets

$$\begin{aligned} \left\{ \left( \ell ^\prime (\xi _i) (f-f^*)(X_i) \right) _{i=1}^N \ : \ f \in F, \ \left\| f-f^*\right\| _{L_2} \le r \right\} \end{aligned}$$

are weakly bounded. The unfortunate byproduct is that \(\phi ^\ell _N(r)\) may exhibit rather poor concentration around its conditional mean or its true mean. This is why \(r_M\) is defined using \(\phi _N^\ell \) rather than by the conditional mean or by its true mean: unlike bounded problems or subgaussian ones, there might be a substantial gap between the fixed point defined using \(\phi _N^\ell \) and the one defined using those means.

Combining the bounds on the quadratic and multiplier terms with Theorems 2.3 and 2.4, one has the following:

Theorem 5.4

For every \(\kappa _0\) and \(0<\varepsilon <1\) there exist constants \(c_0\), \(c_1\), \(c_2\) and \(c_3\) that depend only on \(\kappa _0\) and \(\varepsilon \), and an absolute constant \(c_4\) for which the following holds. Let F be a convex class of functions and assume that \(F-F\) satisfies a small-ball property with constants \(\kappa _0\) and \(\varepsilon \). Set \(t_1=0\), \(t_2=c_0(\kappa _0,\varepsilon )(\Vert \xi \Vert _{L_2}+{\mathrm{diam}}(F,L_2))\), \(\zeta _1=c_1(\kappa _0,\varepsilon )\), \(\zeta _2=c_2(\kappa _0,\varepsilon )\) and put \(\theta =c_3(\kappa _0,\varepsilon ) \rho (t_1,t_2)\). Then,

  • With probability at least \(1-\delta -2\exp (-c_4N\varepsilon ^2)\),

    $$\begin{aligned} \left\| \hat{f}-f^*\right\| _{L_2} \le 2\max \left\{ r_Q(\zeta _1,\zeta _2),r_M(\theta /16,\delta /2)\right\} . \end{aligned}$$
  • If \(\ell \) satisfies Assumption 2.2 with a constant \(\beta \), then with the same probability estimate,

    $$\begin{aligned} \mathbb {E}{\mathcal {L}}_{\hat{f}} \le 2(\theta +\beta )\max \left\{ r_Q(\zeta _1,\zeta _2), r_M(\theta /16,\delta /2)\right\} . \end{aligned}$$

Proof

By Theorem 4.6, there is an absolute constant \(c_0\) and an event of probability at least \(1-2\exp (-c_0\varepsilon ^2N)\), on which, if \(\Vert f-f^*\Vert _{L_2} \ge 2r_Q\) then

$$\begin{aligned} P_N {\mathcal {Q}}_{f-f^*} \ge \theta \left\| f-f^*\right\| _{L_2}^2. \end{aligned}$$

And, by Lemma 5.2, on an event of probability at least \(1-\delta \), if \(\Vert f-f^*\Vert _{L_2} \ge 2r_M(\theta /16,\delta /2)\), then

$$\begin{aligned} \left| P_N {\mathcal {M}}_{f-f^*} \right| \le (\theta /4) \left\| f-f^*\right\| _{L_2}^2. \end{aligned}$$

Using the notation of Theorems 2.3 and 2.4, the first event is \({\mathcal {B}}\), the second in \({\mathcal {A}}\) and the claim follow. \(\square \)

Theorem 5.4 is close to the result we would like to establish, with one significant step still missing: \(t_2\) is not of the order of \(\Vert \xi \Vert _{L_2}\). This is of little significance in the strongly convex case, but requires an additional argument when dealing with a general convex loss. That final step is presented next.

5.1 The main results

Let us begin by showing how to improve the choice of \(t_2\) from \(c(\kappa _0,\varepsilon ) (\Vert \xi \Vert _{L_2}+{\mathrm{diam}}(F,L_2))\) to the potentially much smaller \(2c(\kappa _0,\varepsilon )\Vert \xi \Vert _{L_2}\). To that end, we show that with high probability, the empirical minimizer does not belong to the set

$$\begin{aligned} \left\{ f \in F : \left\| f-f^*\right\| _{L_2} \ge \max \left\{ \left\| \xi \right\| _{L_2},2r_Q\right\} \right\} . \end{aligned}$$

Once that is established, the study of ERM in F may be reduced to the set \(F \cap \max \{\Vert \xi \Vert _{L_2},2r_Q\} D_{f^*}\), and in which case, Theorem 5.4 may be used directly, as the diameter of the class in question is \(\sim \max \{\Vert \xi \Vert _{L_2},r_Q\}\).

Recall that

$$\begin{aligned} {\mathcal {Q}}_{f-f^*}= \int _\xi ^{\xi +(f-f^*)(X)} \left( \ell ^\prime (w)-\ell ^\prime (\xi )\right) dw. \end{aligned}$$
(5.2)

Using Theorem 4.6, there are absolute constants \(c_0\), \(c_1\) and \(c_2\) for which, with probability at least \(1-2\exp (-c_0\varepsilon ^2N)\), if \(\Vert f-f^*\Vert _{L_2} \ge 2r_Q\), then

$$\begin{aligned} P_N {\mathcal {Q}}_{f-f^*} \ge c_1\varepsilon \kappa _0^2 \rho (0,t) \left\| f-f^*\right\| _{L_2}^2, \end{aligned}$$
(5.3)

where \(t=c_2(\kappa _0+\varepsilon ^{-1/2})\cdot (\Vert f-f^*\Vert _{L_2}+\Vert \xi \Vert _{L_2})\).

Let \(\theta =c_1\varepsilon \kappa _0^2 \rho (0,t)\) for \(t=2c_2(\kappa _0+\varepsilon ^{-1/2})\max \{\Vert \xi \Vert _{L_2},r_Q\}\) and assume further that

$$\begin{aligned} r_M(\theta /16,\delta /2) \le \max \left\{ \left\| \xi \right\| _{L_2},2r_Q\right\} . \end{aligned}$$

Theorem 5.5

On an event of probability at least \(1-\delta -2\exp (-c_0N\varepsilon ^2)\),

$$\begin{aligned} \left\| \hat{f}-f^*\right\| _{L_2} \le \max \left\{ \left\| \xi \right\| _{L_2},2r_Q\right\} . \end{aligned}$$

The proof of Theorem 5.5 is based on several observations.

Note that if \(h \in F\) and \(\Vert h-f^*\Vert _{L_2} > R\), there is some \(\lambda > 1\) and \(f \in F\) for which \(\Vert f-f^*\Vert _{L_2} = R\) and \(\lambda (f-f^*) = (h-f^*)\). Indeed, set \(\lambda = \Vert h-f^*\Vert _{L_2}/R>1\) and put \(f=h/\lambda + (1-1/\lambda )f^*\); by convexity, \(f \in F\). Hence, for every \(R>0\),

$$\begin{aligned}&\left\{ h-f^* : h \in F, \ \ \left\| h-f^*\right\| _{L_2} \ge R \right\} \nonumber \\&\quad \subset \left\{ \lambda (f-f^*) : \lambda \ge 1, \ f \in F, \ \left\| f-f^*\right\| _{L_2} = R \right\} . \end{aligned}$$
(5.4)

Lemma 5.6

On the event where (5.3) holds, if \(\Vert f-f^*\Vert _{L_2} = \max \{\Vert \xi \Vert _{L_2},2r_Q\}\) and \(\lambda \ge 1\) then

$$\begin{aligned} P_N {\mathcal {Q}}_{\lambda (f-f^*)} \ge \lfloor \lambda \rfloor \theta \max \left\{ \left\| \xi \right\| _{L_2}^2,4r_Q^2\right\} . \end{aligned}$$

Proof

Fix \(a,x \in \mathbb {R}\) and observe that for every \(\lambda \ge 1\),

$$\begin{aligned} \int _a^{a+\lambda x} \left( \ell ^\prime (w)-\ell ^\prime (a) \right) dw \ge \lfloor \lambda \rfloor \int _a^{a+x} \left( \ell ^\prime (w)-\ell ^\prime (a) \right) dw. \end{aligned}$$
(5.5)

To see this, let \(x>0\) and write

$$\begin{aligned}&\int _a^{a+\lambda x} \left( \ell ^\prime (w)-\ell ^\prime (a) \right) dw = \sum _{j=0}^{\lfloor \lambda \rfloor -1} \int _{a + jx}^{a+(j+1)x} \left( \ell ^\prime (w)-\ell ^\prime (a) \right) dw \\&\quad + \int _{a + \lfloor \lambda \rfloor }^{a+\lambda x} \left( \ell ^\prime (w)-\ell ^\prime (a) \right) dw. \end{aligned}$$

Since \(\ell ^\prime (w)-\ell ^\prime (a)\) is an increasing function in w, the first term in the sum is the smallest and

$$\begin{aligned} \int _a^{a+\lambda x} \left( \ell ^\prime (w)-\ell ^\prime (a) \right) dw \ge \lfloor \lambda \rfloor \int _a^{a+x} \left( \ell ^\prime (w)-\ell ^\prime (a) \right) dw. \end{aligned}$$

The case where \(x<0\) is equally simple.

When (5.5) is applied to (5.2), it follows that pointwise,

$$\begin{aligned} {\mathcal {Q}}_{\lambda (f-f^*)}=\int _\xi ^{\xi +\lambda (f-f^*)(X)} \left( \ell ^\prime (w)-\ell ^\prime (\xi )\right) dw \ge \lfloor \lambda \rfloor {\mathcal {Q}}_{f-f^*}, \end{aligned}$$

and by the lower bound on \(P_N{\mathcal {Q}}_{f-f^*}\) the claim follows. \(\square \)

Proof of Theorem 5.5

Recall that we assume that \(r_M(\theta /16,\delta /2) \le \max \{\Vert \xi \Vert _{L_2},2r_Q\}\); hence, with probability at least \(1-\delta \), if \(\Vert f-f^*\Vert _{L_2} \le \max \{\Vert \xi \Vert _{L_2},2r_Q\}\) then

$$\begin{aligned} \left| P_N {\mathcal {M}}_{f-f^*} \right| = \left| \frac{1}{N}\sum _{i=1}^N \ell ^\prime (\xi _i)(f-f^*)(X_i)\right| \le (\theta /4) \max \left\{ \left\| \xi \right\| _{L_2}^2,4r_Q^2\right\} . \end{aligned}$$

Also, \(P_N {\mathcal {M}}_{f-f^*}\) is linear in \(f-f^*\) and it follows that for every \(\lambda \ge 1\),

$$\begin{aligned} \left| P_N {\mathcal {M}}_{\lambda (f-f^*)}\right| =&\left| \frac{1}{N}\sum _{i=1}^N \ell ^\prime (\xi _i)\lambda (f-f^*)(X_i)\right| = \lambda \left| P_N {\mathcal {M}}_{f-f^*} \right| \\ \le&\, \lambda (\theta /4) \max \left\{ \left\| \xi \right\| _{L_2}^2,4r_Q^2\right\} . \end{aligned}$$

Combining this with the lower bound on \(P_N {\mathcal {Q}}_{f-f^*}\) shows that with probability at least \(1-\delta -2\exp (-c_0\varepsilon ^2N)\), if \(\Vert f-f^*\Vert _{L_2} = \max \{\Vert \xi \Vert _{L_2},2r_Q\}\) and \(\lambda \ge 1\) then

$$\begin{aligned} P_N {\mathcal {Q}}_{\lambda (f-f^*)} - \left| P_N {\mathcal {M}}_{\lambda (f-f^*)}\right| \ge \lambda (\theta /2) \max \left\{ \left\| \xi \right\| _{L_2}^2,4r_Q^2\right\} >0. \end{aligned}$$

Thus, by (5.4), on that event the empirical minimizer belongs to the set

$$\begin{aligned} F \cap \max \left\{ \left\| \xi \right\| _{L_2},2r_Q\right\} D_{f^*}. \end{aligned}$$

\(\square \)

Now we are finally ready to formulate and prove the main results of the article.

Theorem 5.7

For every \(\kappa _0\) and \(0<\varepsilon <1\) there exist constants \(c_0\), \(c_1\), \(c_2\) and \(c_3\) that depend only on \(\kappa _0\) and \(\varepsilon \), and an absolute constant \(c_4\) for which the following holds.

Let F be a convex class of functions and assume that \(F-F\) satisfies the small-ball property with constants \(\kappa _0\) and \(\varepsilon \). Set \(t_1=0\) and \(t_2=c_0(\varepsilon ,\kappa _0)\Vert \xi \Vert _{L_2}\), \(\zeta _1=c_1(\varepsilon ,\kappa _0)\) and \(\zeta _2=c_2(\varepsilon ,\kappa _0)\). Put \(\theta =c_3(\varepsilon ,\kappa _0) \rho (t_1,t_2)\).

If \(r_M(\theta /16,\delta /2) \le \max \{\Vert \xi \Vert _{L_2},2r_Q(\zeta _1,\zeta _2)\}\), then with probability at least \(1-\delta -2\exp (-c_4N\varepsilon ^2)\),

  • \(\Vert \hat{f}-f^*\Vert _{L_2} \le 2\max \{r_Q(\zeta _1,\zeta _2),r_M(\theta /16,\delta /2)\}\).

  • If \(\ell \) satisfies Assumption 2.2 with a constant \(\beta \) then with the same probability estimate,

    $$\begin{aligned} \mathbb {E}{\mathcal {L}}_{\hat{f}} \le 2(\theta +\beta )\max \left\{ r_Q(\zeta _1,\zeta _2),r_M(\theta /16,\delta /2)\right\} . \end{aligned}$$
  • If \(\xi \) is independent of X and satisfies a small-ball condition with constants \(\kappa _1\) and \(\varepsilon \), one may take \(t_1=c_5 \kappa _1 \Vert \xi \Vert _{L_2}\) for a constant \(c_5=c_5(\varepsilon )\), and the two assertions formulated above hold as well.

Proof

By the preliminary exclusion argument of Theorem 5.5, with probability at least \(1-\delta -2\exp (-c_0N\varepsilon ^2)\), \(\Vert \hat{f}-f^*\Vert _{L_2} \le \max \{\Vert \xi \Vert _{L_2},2r_Q\}\). If \(\Vert \xi \Vert _{L_2} \le 2r_Q\) then Theorem 5.5 suffices to prove the assertion of Theorem 5.7. Otherwise, the assertion follows from Theorem 5.4, applied to the class \(F \cap \Vert \xi \Vert _{L_2}D_{f^*}\). \(\square \)

The second main result deals with the case in which \(\ell \) is strongly convex in a neighbourhood of zero.

Theorem 5.8

For every \(\kappa _0\) and \(0<\varepsilon <1\) there exist constants \(c_0\), \(c_1\), \(c_2\) and \(c_3\) that depend only on \(\kappa _0\) and \(\varepsilon \), and an absolute constant \(c_4\) for which the following holds.

Assume that \(\ell \) is strongly convex in the interval \([-\gamma ,\gamma ]\) with a constant \(\kappa _2\) and that \(\Vert \xi \Vert _{L_2} \le c_0\gamma \).

Assume further that F is a convex class of functions and that \(F-F\) satisfies the small-ball property with constants \(\kappa _0\) and \(\varepsilon \). Set \(\zeta _1=c_1(\kappa _0,\varepsilon )\), \(\zeta _2=c_2(\kappa _0,\varepsilon )\) and \(\theta =c_3(\kappa _0,\varepsilon ) \kappa _2\).

If \(r_M(\theta /16,\delta /2) \le \gamma \), then with probability at least \(1-\delta -2\exp (-c_4N\varepsilon ^2)\),

  • \(\Vert \hat{f}-f^*\Vert _{L_2} \le 2\max \{r_Q(\zeta _1,\zeta _2),r_M(\theta /16,\delta /2)\}\).

  • If \(\ell \) satisfies Assumption 2.2 with a constant \(\beta \) then with the same probability estimate,

    $$\begin{aligned} \mathbb {E}{\mathcal {L}}_{\hat{f}} \le 2(\theta +\beta )\max \left\{ r_Q(\zeta _1,\zeta _2),r_M(\theta /16,\delta /2)\right\} . \end{aligned}$$

The proof of Theorem 5.8 is almost identical to that of Theorem 5.7, with one difference: when \(\gamma > 2r_Q\) then instead of considering the preliminary exclusion argument of Theorem 5.5 at the level \(\sim \max \{\Vert \xi \Vert _{L_2},2r_Q\}\), one performs preliminary exclusion at the level \(\gamma \), and with an identical proof. The rest of the argument remains unchanged and we omit its details.

At this point, let us return to the rather detailed ‘wish list’ that was outlined in the introduction regarding the parameters governing prediction and estimation problems and see where we stand.

Theorems 5.7 and 5.8 lead to bounds on \({\mathcal {E}}_p\) and \({\mathcal {E}}_e\) without assuming that the class consists of uniformly bounded or subgaussian functions, nor that the target is even in \(L_p\) for some \(p>2\). And, under a minor smoothness assumption, \(\ell \) need not be a Lipschitz function. Thus, all the restrictions of the classical method and of subgaussian learning have been bypassed successfully.

As for the complexity parameters involved, \(r_Q\) is indeed an intrinsic parameter of the class F and has nothing to do with the choice of the loss or with the target. It does measure (with the very high probability of \(1-2\exp (-c\varepsilon ^2N)\)), the \(L_2\) diameter of the version space of F associated with \(f^*\), and thus corresponds to the solution of the noise-free problem.

The noise and loss influence the problem in two places. In the quadratic component, the loss has to be calibrated to fit the noise level: the loss must be strongly convex in the interval \([0,c_1(\kappa _0,\varepsilon )\Vert \xi \Vert _{L_2}]\), or, when the noise is independent, it suffices that the loss is strongly convex in the smaller interval \([c_2(\kappa _1,\varepsilon )\Vert \xi \Vert _{L_2},c_1(\kappa _0,\varepsilon )\Vert \xi \Vert _{L_2}]\). The strong convexity constant in those intervals also determines the level \(\theta \) appearing in the definition of the multiplier component.

Although the noise and loss affect the quadratic component, their main impact is seen in the multiplier component, and thus in the external complexity parameter \(r_M\). Indeed, while the quadratic component and the level \(\sim \theta \) of a given class will be exactly the same for any loss that has the same strong convexity constant in the interval \([0,c_1(\kappa _0,\varepsilon )\Vert \xi \Vert _{L_2}]\), the difference between losses is ‘coded’ in \(r_M\). And, as expected, the interaction between the class, the noise and the loss is captured by a single parameter: the correlation (or width) of a random coordinate projection of the localized class with the random vector \((\varepsilon _i \ell ^\prime (\xi _i))_{i=1}^N\).

Therefore, the ‘wish list’ is satisfied in full: without any boundedness assumptions and for a rather general loss function, prediction and estimation problems exhibit the expected two-regime behaviour: a ‘low-noise’ regime captured by an intrinsic parameter and a ‘high-noise’ regime by an external one. The exact nature of the loss and noise determines the external parameter only through the multiplier vector \((\varepsilon _i \ell ^\prime (\xi _i))_{i=1}^N\), and this random vector also determines where the phase transition between the high-noise regime, in which the external parameter \(r_M\) is dominant, and the low-noise one, in which the intrinsic parameter \(r_Q\) is dominant, takes place.

The one remaining issue still left open is to show that selecting the loss with some care may be used to negate the effects of noise related outliers.

6 Loss functions and the removal of outliers

Damaging outliers appear when sample points are far from where one would expect them to be, and the loss assigns a large value to those points. This combination means that outliers actually have a true impact on the empirical mean \(P_N {\mathcal {L}}_f\) and therefore on the identity of the empirical minimizer.

The reason why outliers are not an issue in problems that feature a strong concentration phenomenon is clear: no matter what the loss is (as long as it does not grow incredibly quickly) only an insignificant fraction of the sample points fall outside the ‘right area’, and thus their impact is negligible.

The situation is different when either the class consists of heavy-tailed functions or when the noise \(Y-f^*(X)\) is heavy-tailed. In such cases, a more substantial fraction of a typical sample falls in a potentially misleading location, and if the effect is amplified by a fast-growing loss, outliers become a problem that has to be contended with. This problem may be partly resolved by ensuring that the loss is not very big outside the ‘expected area’ of \([-c\Vert \xi \Vert _{L_2},c\Vert \xi \Vert _{L_2}]\), which already hints towards the ‘right choice’ of a loss.

As noted previously, as long as \(\ell \) is strongly convex in \([0,c_1(\kappa _0,\varepsilon )\Vert \xi \Vert _{L_2}]\) (or in the smaller interval when the noise is independent of X) the effects of the loss and the noise are coded in the vector \((\varepsilon _i \ell ^\prime (\xi _i))_{i=1}^N\). If \(\ell \) is the squared loss, this vector is likely to have relatively many large coordinates when \(\xi \) is heavy-tailed. However, losses that grow almost linearly in \([a,\infty )\) lead to bounded multipliers because \(|\ell ^\prime (\xi )| \lesssim |\ell ^\prime (a)|\). This results in a better-behaved multiplier component and a smaller fixed point \(r_M\).

As examples, we focus on the three losses mentioned earlier: the squared loss, the log loss (2.2) and the Huber loss (2.1).

  • The squared loss is the canonical example of a strongly convex loss with a bounded second derivative. However, it is susceptible to the problem of noise related outliers because it grows rapidly: \(\ell ^\prime (\xi ) \sim \xi \).

  • The log-loss (2.2) exhibits a strongly convex behaviour in any bounded interval, but with a constant that decreases exponentially quickly to zero with the length of the interval. At the same time, its growth becomes close to linear for large values (i.e., \(\ell ^\prime \) tends to a constant). Therefore, on the one hand the lower estimate on the quadratic component becomes trivial as \(\Vert \xi \Vert _{L_2}\) increases, but the multiplier component displays a better behaviour than the squared loss.

  • The Huber loss with parameter \(\gamma \) is strongly convex in \((-\gamma ,\gamma )\) and grows linearly outside that interval. Hence, for \(\gamma \sim \Vert \xi \Vert _{L_2}\) one may hope to get the best of both worlds—a quadratic component that behaves like its counterpart for the squared loss, and a multiplier component that behaves as if \(\ell \) were linear.

We show that the three losses respond to outliers in very different ways. Our focus is on situations when F is subgaussian but \(\xi \) is potentially heavy-tailed, and the goal is to see if despite the heavy-tailed \(\xi \), ERM performs in the same way as if \(\xi \) were subgaussian. To simplify the presentation and to give each loss a level playing field, we consider targets of the form \(Y=f_0(X)+W\) for \(f_0 \in F\) and W that is symmetric an independent of X. The reason for this choice of targets is that for any ‘legal’ loss function, \(f^*=f_0\). Indeed, for any \(f \in F\),

$$\begin{aligned} \mathbb {E}\ell (f(X)-Y)-\mathbb {E}\ell (f_0(X)-Y) \ge \mathbb {E}\ell ^\prime (W)(f-f_0)(X)=0, \end{aligned}$$

because \(\ell ^\prime \) is odd and W is symmetric and independent of X. Hence, if the minimizer is also unique (as we assume), that minimizer must be \(f_0\). In particular, for all the three losses \({\mathrm{argmin}}_{f \in F} \mathbb {E}\ell (f(X)-Y)=f_0\).

We first present general estimates on the performance of ERM in terms the parameters \(r_Q\) and \(r_M\) for a general target Y, and then control the parameters for an arbitrary convex, L-subgaussian class and a heavy-tailed target. As an example we consider case of \(F=\{\bigl < t,\cdot \bigr > : t \in \mathbb {R}^n\}\), which is the class of all linear functionals on \(\mathbb {R}^n\).

In what follows \(F \subset L_2\) is a closed, convex class of functions and \(F-F\) satisfies a small-ball property with constants \(\kappa _0\) and \(\varepsilon \). The target one wishes to estimate is \(Y \in L_q\) for some \(q > 2\) and for the sake of simplicity we assume at times that \(q=4\), though that is not really an essential assumption.

6.1 Some facts on multiplier processes

The following is an upper estimate on multiplier and empirical processes indexed by a class that is L-subgaussian.

Theorem 6.1

[20] There exists an absolute constant \(c_0\) and for every \(L>1\) there are constants \(c_1\) and \(c_2\) that depend only on L and for which the following holds.

Assume that \(\Lambda \in L_q\) for some \(q>2\) and that F is L-subgaussian. Recall that \(k_F=(\mathbb {E}\Vert G\Vert _F/d_F(L_2))^2\) and let \(N \ge k_F\). If \((X_i,\Lambda _i)_{i=1}^N\) are N independent copies of \((X,\Lambda )\) then

  • for \(u>c_0\), with probability at least \(1-2\exp (-c_1u^2 k_F)\),

    $$\begin{aligned} \sup _{f \in F} \left| \frac{1}{N}\sum _{i=1}^N f(X_i) - \mathbb {E}f \right| \le c_2u \frac{\mathbb {E}\left\| G\right\| _F}{\sqrt{N}}. \end{aligned}$$
  • for \(u,\beta >c_0\), with probability at least \(1-2\beta ^{-q}N^{-((q/2)-1)}-2\exp (-c_1u^2 k_F)\),

    $$\begin{aligned} \sup _{f \in F} \left| \frac{1}{N}\sum _{i=1}^N \Lambda _i f(X_i) - \mathbb {E}\Lambda f \right| \le c_2\beta u \left\| \Lambda \right\| _{L_q} \frac{\mathbb {E}\left\| G\right\| _F}{\sqrt{N}}. \end{aligned}$$

6.2 The squared loss

If \(\ell (t)=t^2\) then for every \(t_1,t_2\), \(\rho (t_1,t_2) = 2\). Therefore, by Theorem 5.8 for an arbitrarily large \(\gamma \), it follows that with probability at least \(1-\delta -2\exp (-c_1 \varepsilon ^2 N)\),

$$\begin{aligned} \left\| \hat{f}-f^*\right\| _{L_2} \le \max \left\{ r_Q(\zeta _1,\zeta _2),r_M(c_2/4,\delta /2)\right\} , \end{aligned}$$

for constants \(\zeta _1\), \(\zeta _2\) and \(c_2\) that depend only on \(\kappa _0\) and \(\varepsilon \).

Clearly, \(\Vert \ell ^{\prime \prime }\Vert _{L_\infty } \le 2\), implying that with the same probability estimate,

$$\begin{aligned} \mathbb {E}{\mathcal {L}}_{\hat{f}} \le 2(c_2+1) \max \left\{ r_Q^2(\zeta _1,\zeta _2),r_M^2(c_2/4,\delta /2)\right\} . \end{aligned}$$

When F is, in addition, an L-subgaussian class, one may identify the parameters \(r_M\) and \(r_Q\). Recall that \(\Vert f\Vert _{\psi _2} \sim \sup _{p \ge 2} \Vert f\Vert _{L_p}/\sqrt{p}\) (which, as noted previously, suffices to ensure that the small-ball property holds for \(F-F\) with constants \(\kappa _0\) and \(\varepsilon \) that depend only on L).

Recall that

$$\begin{aligned} r_{2,Q}(\zeta _2)=\inf \left\{ r>0: \mathbb {E}\sup _{f \in F \cap rD_{f^*}} \left| \frac{1}{\sqrt{N}}\sum _{i=1}^N \varepsilon _i (f-f^*)(X_i) \right| \le \zeta _2 r \sqrt{N} \right\} ; \end{aligned}$$

setting \(F_r=\{ f-f^* : f \in F \cap rD_{f^*} \}\), it follows from Theorem 6.1 that

$$\begin{aligned} \mathbb {E}\sup _{f \in F \cap rD_{f^*}} \left| \frac{1}{\sqrt{N}}\sum _{i=1}^N \varepsilon _i (f-f^*)(X_i) \right| \le c_3 \mathbb {E}\left\| G\right\| _{F_r}. \end{aligned}$$

Therefore,

$$\begin{aligned} r_Q(\zeta _1,\zeta _2) \le \inf \left\{ r>0: \mathbb {E}\left\| G\right\| _{F_r} \le c_4 \min \left\{ \zeta _1,\zeta _2\right\} r \sqrt{N} \right\} . \end{aligned}$$

Turning to \(r_M\), one has to identify

$$\begin{aligned} r_0=\inf \left\{ r>0: \sup _{f \in F \cap r D_{f^*}} \left\| \xi (f-f^*)(X)\right\| _{L_2} \le \sqrt{N} r^2 (c_2/16)\right\} \end{aligned}$$

and the ‘lowest’ level r for which

$$\begin{aligned} Pr\left( \frac{1}{\sqrt{N}} \sup _{f \in F \cap r D_{f^*}} \left| \sum _{i=1}^N \varepsilon _i \xi _i (f-f^*)(X_i) \right| \le r^2 (c_2/4) \sqrt{N} \right) \ge 1-\delta . \end{aligned}$$

Applying the \(L_4-L_2\) norm equivalence, \(\Vert f-f^*\Vert _{L_4} \le 2L\Vert f-f^*\Vert _{L_2}\) and

$$\begin{aligned} \left\| \xi (f-f^*)(X)\right\| _{L_2} \le 2L\left\| \xi \right\| _{L_4} r \lesssim _L \sqrt{N} r^2, \end{aligned}$$

provided that \(r \gtrsim \Vert \xi \Vert _{L_4}/\sqrt{N}\). Moreover, by Theorem 6.1 for \(q=4\), it follows that with probability at least \(1-2/(\beta ^{4}N)-2\exp (-c_3(L)u^2 k_{F_r})\),

$$\begin{aligned} \sup _{f \in F_r} \left| \frac{1}{\sqrt{N}}\sum _{i=1}^N \varepsilon _i \xi _i (f-f^*)(X_i)\right| \le c_4(L)\beta u \left\| \xi \right\| _{L_4} \mathbb {E}\left\| G\right\| _{F_r}. \end{aligned}$$

Fix \(0<\delta <1\). If \(k_{F_r} \ge \log (2/\delta )\) one may take \(u=c_5(L)\) and if the reverse inequality is satisfied, one may set \(u \sim _L (k_{F_r}^{-1}\log (2/\delta ))^{1/2}\), leading to a probability estimate of \(1-\delta \). Therefore, if

$$\begin{aligned} u(r,\delta )=c_6(L) \left( 1+k_{F_r}^{-1}\log (2/\delta )\right) ^{1/2} \end{aligned}$$

and

$$\begin{aligned} \beta \sim \max \left\{ \frac{1}{(\delta N)^{1/4}},1\right\} , \end{aligned}$$

then with probability at least \(1-\delta \),

$$\begin{aligned} \sup _{f \in F_r} \left| \frac{1}{\sqrt{N}}\sum _{i=1}^N \varepsilon _i \xi _i (f-f^*)(X_i)\right| \le c_7(L)\beta u \left\| \xi \right\| _{L_4} \mathbb {E}\left\| G\right\| _{F_r}, \end{aligned}$$

and

$$\begin{aligned} r_M(c_2/4,\delta /2) \le \frac{\left\| \xi \right\| _{L_4}}{\sqrt{N}} + \inf \left\{ r>0: \mathbb {E}\left\| G\right\| _{F_r} \le c_8(L) \sqrt{N} \left\| \xi \right\| _{L_4}^{-1}(\beta u)^{-1}r^2\right\} . \end{aligned}$$
(6.1)

Thus, \(r_M(c_2/4,\delta /2)\) dominates \(r_Q(\zeta _1,\zeta _2)\) as long as \(\Vert \xi \Vert _{L_4}\) is not very small.

As a point of reference, consider the case in which the Dvoretzky–Milman dimension \(k_{F_r}\) is at least of the order of \(\log (2/\delta )\), and thus the infimum in (6.1) is attained for a value r for which \(u(r,\delta )=c(L)\). Therefore,

$$\begin{aligned} r_M(c_2/4,\delta /2) \le \frac{\left\| \xi \right\| _{L_4}}{\sqrt{N}} + \inf \left\{ r>0: \mathbb {E}\left\| G\right\| _{F_r} \le c_9(L) \sqrt{N} \left\| \xi \right\| _{L_4}^{-1}\beta ^{-1}r^2\right\} . \end{aligned}$$
(6.2)

The difference between (6.2) and the analogous estimate in the purely subgaussian case (see [10]) is the factor \(\beta ^{-1}\). Since \(\beta \sim \max \{1/(\delta N)^{1/4},1\}\), it causes a slower rate when the desired confidence level is high. Indeed, if \(\delta \ll 1/N\), then \(\beta \gg 1\) leading to a larger value of \(r_M\) than in a purely subgaussian problem, and to a weaker accuracy/confidence tradeoff that one would expect if \(\xi \) were subgaussian. This different accuracy/confidence tradeoff is caused by the noise-related outliers one encounters and is the price one pays for using the squared loss in a problem involving a potentially heavy-tailed noise (\(\xi \in L_4\) rather than \(\xi \in L_{\psi _2}\)). As a result, the accuracy/confidence tradeoff has a polynomial dependence on \(1/\delta \) rather than the logarithmic dependence ERM exhibits when \(\xi \) is subgaussian.

6.3 The log-loss

It is straightforward to verify that for \(t \ge 0\), \(\ell ^\prime (t)=1-2/(\exp (t)+1)\), and in particular \(\ell ^\prime (t) \le \min \{2t,1\}\). Also, \(\ell ^{\prime \prime }(t)= 2\exp (t)/(\exp (t)+1)^2\), which is a decreasing function on \(\mathbb {R}_+\) and is upper-bounded by 1. Therefore, \(\ell \) satisfies Assumption 2.2, and one may verify that \(\theta \sim \rho (t_1,t_2) \ge \exp (-c_1\Vert \xi \Vert _{L_2})\).

The difference between the log-loss and the squared loss can be seen in the behaviour of \(r_M(\theta /16,\delta /2)\). While the multipliers for the squared loss are independent copies of \(\ell ^\prime (\xi )=\xi \), for the log-loss one has \(|\ell ^\prime (\xi )| \le \min \{2|\xi |,1\}\). This almost linear growth of the loss outside [0, 1] helps one overcome the issue of noise-related outliers and leads to an improved accuracy/confidence tradeoff—a logarithmic dependence in \(1/\delta \) rather than the polynomial one exhibited by the squared loss—but only when \(\Vert \xi \Vert _{L_2} \sim 1\). The tradeoff deteriorates when \(\Vert \xi \Vert _{L_2}\) is small (because of the multiplier component) and when \(\Vert \xi \Vert _{L_2}\) is large (because of the quadratic component).

The improved tradeoff is based on a contraction argument: since \(|\ell ^\prime (\xi )|\le 1\),

$$\begin{aligned}&Pr\left( \sup _{f \in F \cap r D_{f^*}} \left| \frac{1}{N}\sum _{i=1}^N \varepsilon _i \ell ^\prime (\xi _i) (f-f^*)(X_i)\right|> t \right) \\&\quad \le 2Pr\left( \sup _{f \in F \cap r D_{f^*}} \left| \frac{1}{N}\sum _{i=1}^N \varepsilon _i (f-f^*)(X_i)\right| > \frac{t}{4} \right) . \end{aligned}$$

Clearly, removing any dependence on the multipliers is a costly step when \(\Vert \xi \Vert _{L_2}\) is very small, but when \(\Vert \xi \Vert _{L_2} \sim 1\) the resulting estimate is a significant improvement over the estimate for the squared loss: if \(u(r,\delta )\) is as defined above,

$$\begin{aligned} r_M(\theta /16,\delta /2) \le \frac{1}{\sqrt{N}} + \inf \left\{ r>0: \mathbb {E}\left\| G\right\| _{F_r} \lesssim \theta u(r,\delta )\sqrt{N} r^2\right\} ; \end{aligned}$$

hence, if \(r_M \le \Vert \xi \Vert _{L_2}\), then with probability at least \(1-\delta -2\exp (-c_0\varepsilon ^2 N)\),

$$\begin{aligned} \left\| \hat{f}-f^*\right\| _{L_2} \lesssim 2\max \left\{ r_Q,r_M\right\} \ \ {\mathrm{and}} \ \ \mathbb {E}{\mathcal {L}}_{\hat{f}} \lesssim \max \left\{ r_Q^2,r_M^2\right\} . \end{aligned}$$

Again, let us consider the case in which \(\inf \{r>0: \mathbb {E}\Vert G\Vert _{F_r} \lesssim \theta u(r,\delta )\sqrt{N} r^2\}\) is attained by a value r for which \(u(r,\delta )=c(L)\). Then,

$$\begin{aligned} r_M(\theta /16,\delta /2) \le \frac{1}{\sqrt{N}} + \inf \left\{ r>0: \mathbb {E}\left\| G\right\| _{F_r} \lesssim \theta \sqrt{N} r^2\right\} , \end{aligned}$$

which is the same bound on \(r_M\) one has for the squared loss when \(\xi \) is a standard gaussian random variable that is independent of X, though we only know that \(\Vert \xi \Vert _{L_2} \sim 1\).

The improved accuracy/confidence tradeoff occurs when \(\Vert \xi \Vert _{L_2} \sim 1\) simply because the log-loss happens to be calibrated for that noise level. This is a mere coincidence rather than premeditation, and the accuracy/confidence tradeoff of ERM relative to the log-loss deteriorates when \(\Vert \xi \Vert _{L_2}\) is either very large or very small.

6.4 The Huber loss

Let \(r_Q\) be as above for suitable constants \(\zeta _1\) and \(\zeta _2\). For \(\zeta _3=\min \{\zeta _1,\zeta _2\}\) one has

$$\begin{aligned} r_Q=\inf \left\{ r>0: \mathbb {E}\left\| G\right\| _{F_r} \le c\zeta _3 r \sqrt{N}\right\} \end{aligned}$$

where \(c=c(L)\). Without loss of generality, one may assume that \(c \zeta _3\) is smaller than a fixed constant which will be the constant \(c_3=c_3(L)\) defined below.

Let \(\ell (t)\) be the Huber loss with parameter \(\gamma =c_0(L) \max \{\Vert \xi \Vert _{L_2},r_Q\}\) for a constant \(c_0\) to be specified later.

Observe that \(|\ell ^\prime (t)|=\min \{|t/2|,\gamma \}\), that \(\ell ^\prime \) is a Lipschitz function with constant 1 and that Assumption 2.2 is verified. Moreover, this setup falls within the scope of Theorem 5.8 for \(\theta =c_1(L)\).

Regarding the multiplier component, note that if \(\Vert f-f^*\Vert _{L_2} \le r\) then

$$\begin{aligned} \left\| \ell ^\prime (\xi )(f-f^*)\right\| _{L_2} \le \gamma \left\| f-f^*\right\| _{L_2} \le \gamma r \le (c_1(L)/16) r^2\sqrt{N}, \end{aligned}$$

provided that \(r \gtrsim _L \gamma /\sqrt{N}\). Moreover, a contraction argument shows that

$$\begin{aligned}&Pr\left( \sup _{f \in F \cap r D_{f^*}} \left| \frac{1}{N}\sum _{i=1}^N \varepsilon _i \ell ^\prime (\xi _i) (f-f^*)(X_i)\right|> t \right) \\&\quad \le 2Pr\left( \sup _{f \in F \cap r D_{f^*}} \left| \frac{1}{N}\sum _{i=1}^N \varepsilon _i (f-f^*)(X_i)\right| > \frac{t}{4\gamma } \right) \end{aligned}$$

and if \(u(r,\delta )\) is as defined above, one has

$$\begin{aligned} r_M \le c_2(L) \left( \frac{\max \left\{ \left\| \xi \right\| _{L_2},r_Q\right\} }{\sqrt{N}} + \inf \left\{ r>0: \gamma \mathbb {E}\left\| G\right\| _{F_r} \le c_1(L) u(r,\delta )\sqrt{N} r^2\right\} \right) . \end{aligned}$$

Again, let us consider the case in which \(\inf \{r>0: \gamma \mathbb {E}\Vert G\Vert _{F_r} \le c_2 u(r,\delta )\sqrt{N} r^2\}\) is attained for a value r for which \(u(r,\delta )=c^\prime (L)\). Hence, setting \(c_3=c_1(L)u\),

$$\begin{aligned} r_M \le c_2\left( c_0 \frac{\max \left\{ \left\| \xi \right\| _{L_2},r_Q\right\} }{\sqrt{N}} + \inf \left\{ r>0: \gamma \mathbb {E}\left\| G\right\| _{F_r} \le c_3 \sqrt{N} r^2\right\} \right) , \end{aligned}$$
(6.3)

which matches the estimate on \(r_M\) for the squared loss when \(\Vert \xi \Vert _{\psi _2} \sim \Vert \xi \Vert _{L_2}\).

If \(r_Q \le \Vert \xi \Vert _{L_2}\) then \(\gamma = c_0\Vert \xi \Vert _{L_2}\), and since \(c_3 \ge c\zeta _3\), \(r=\Vert \xi \Vert _{L_2}\) belongs to the set \(\{r>0: c_0\Vert \xi \Vert _{L_2} \mathbb {E}\Vert G\Vert _{F_r} \le c_3 \sqrt{N} r^2\}\) in (6.3). Therefore,

$$\begin{aligned} r_M \le c_2\left( 1+\frac{c_0}{\sqrt{N}}\right) \left\| \xi \right\| _{L_2} \le c_0\left\| \xi \right\| _{L_2}=\gamma \end{aligned}$$

as long as \(N \ge c_4(L)\) and for a well chosen \(c_0\). Hence, the assumption of Theorem 5.8 is verified.

Otherwise, \(\Vert \xi \Vert _{L_2} \le r_Q\), and in which case, \(r_Q\) belongs to the set in (6.3) implying that

$$\begin{aligned} r_M \le c_2\left( 1+\frac{c_0}{\sqrt{N}}\right) r_Q \le \gamma . \end{aligned}$$

Therefore, by Theorem 5.8, with probability at least \(1-\delta -2\exp (-c_0\varepsilon ^2N)\),

$$\begin{aligned} \left\| \hat{f}-f^*\right\| _{L_2} \lesssim \max \left\{ r_M,r_Q\right\} \ \ {\mathrm{and}} \ \ \mathbb {E}{\mathcal {L}}_{\hat{f}} \lesssim \max \left\{ r_M^2,r_Q^2\right\} . \end{aligned}$$

The right choice of \(\gamma \) in the Huber loss leads to the optimal interval of strong convexity \([0, c\max \{\Vert \xi \Vert _{L_2},r_Q\}]\), and to a far better accuracy/confidence tradeoff than exhibited by the squared loss for such a heavy-tailed problem. In fact, it matches the behaviour of the squared loss when \(\Vert \xi \Vert _{\psi _2} \sim \Vert \xi \Vert _{L_2}\).

Let us present a concrete example in which all these rates can be computed explicitly, and which shows how the rates are affected by the choice of the loss.

For \(T \subset \mathbb {R}^n\), set \(F_T=\{\bigl < t,\cdot \bigr > : t \in T\}\) and assume that \(\mu \) is an isotropic, L-subgaussian measure on \(\mathbb {R}^n\); that is, its covariance structure coincides with the standard \(\ell _2^n\) distance on \(\mathbb {R}^n\) (isotropicity), and for every \(t \in S^{n-1}\) and every \(p \ge 2\), \(\Vert \bigl< X,t \bigr>\Vert _{L_p} \le L\sqrt{p} \Vert \bigl < X,t \bigr >\Vert _{L_2}\) (L-subgaussian). In particular, \(F-F\) satisfies a small-ball property with constants that depend only on L.

We focus on the case \(T=\mathbb {R}^n\). Clearly, for every \(r>0\) and any possible \(f^* \in F\), \(F_r=rB_2^n\), implying that \(\mathbb {E}\Vert G\Vert _{F_r} \sim r\sqrt{n}\). Therefore, \(k_{F_r} \sim \sqrt{n}\) and \(\mathbb {E}\Vert G\Vert _{F_r} \le \alpha \sqrt{N} r^2\) when \(r \ge \alpha ^{-1}\sqrt{n/N}\).

  • The squared loss. It is straightforward to verify that if \(N \ge c_1(L)n\), then with probability at least \(1-2\exp (-c_2(L)N)\), \(r_Q=0\). Also,

    $$\begin{aligned} u(r,\delta )=c_3(L) \left( 1+\sqrt{\frac{\log (1/\delta )}{N}}\right) . \end{aligned}$$

    Using the definition of \(r_M\), it is evident that with probability at least \(1-\delta -2\exp (-c_4(L)N)\),

    $$\begin{aligned} \Vert \hat{f}-f^*\Vert _{L_2} \lesssim _L \max \left\{ \frac{1}{(N\delta )^{1/4}},1\right\} \cdot \left( 1+\sqrt{\frac{\log (1/\delta )}{n}}\right) \cdot \Vert \xi \Vert _{L_4} \sqrt{\frac{n}{N}}, \end{aligned}$$
    (6.4)

    exhibiting once again that the accuracy/confidence tradeoff has a polynomial dependence in \(1/\delta \).

  • The log-loss. Let \(t_2 \sim _{L} \Vert \xi \Vert _{L_2}\) and therefore, \(\theta \sim _L \exp (-c_1\Vert \xi \Vert _{L_2})\). One has to take \(N \ge c_2(\theta ) n\) to ensure a nontrivial bound on \(r_Q\), and in which case, \(r_Q=0\). Therefore, and in a similar way to the squared loss, with probability at least \(1-\delta -2\exp (-c_3(L)N)\)

    $$\begin{aligned} \left\| \hat{f}-f^*\right\| _{L_2} \lesssim _L \exp \left( c_1(L)\left\| \xi \right\| _{L_2}\right) \left( 1+\sqrt{\frac{\log (1/\delta )}{n}}\right) \cdot \sqrt{\frac{n}{N}}, \end{aligned}$$

    which is better than (6.4) in terms of the dependence on \(\delta \) when \(\Vert \xi \Vert _{L_2}\) is of the order of a constant and \(\delta \ll 1/N\), but does not scale correctly with \(\Vert \xi \Vert _{L_2}\) when the norm is either very small or very large. This was to be expected from the ‘calibration’ of the log-loss, which only fits a constant noise level, when \(\Vert \xi \Vert _{L_2} \sim 1\).

  • The Huber loss. As noted previously, for a nontrivial bound on \(r_Q\) one must take \(N \ge c_1(L)n\), and in which case, \(r_Q=0\).

    Fix \(\gamma =c_2(L)\max \{\Vert \xi \Vert _{L_2},r_Q\}=c_2(L) \Vert \xi \Vert _{L_2}\) and \(\theta =c_2(L)\). Using the definition of \(r_M\) (because \(|\ell ^\prime (\xi )| \le \gamma \)), it is evident that with probability at least \(1-\delta -2\exp (-c_3(L)N)\),

    $$\begin{aligned} \left\| \hat{f}-f^*\right\| _{L_2} \le c_{4}(L) \left( \sqrt{\frac{\log (1/\delta )}{n}}+1\right) \cdot \left\| \xi \right\| _{L_2} \sqrt{\frac{n}{N}}. \end{aligned}$$

    This is the optimal accuracy/confidence tradeoff for any choice of \(\Vert \xi \Vert _{L_2}\) and coincides with the optimal tradeoff for the squared loss when \(\xi \) is gaussian and independent of X (see, e.g. [10]).

    Unlike the squared loss and the log-loss, the optimal rate is obtained thanks to this choice of the Huber loss and despite the noise being heavy-tailed. This phenomenon occurs specifically because the Huber loss can be calibrated to fit the noise level of the problem and the intrinsic complexity of the class. It indicates that the right choice of loss can be used to effectively remove outliers caused by a heavy-tailed noise.

7 Final remarks

Because we study general loss functions, one should take care when comparing the results on the accuracy/confidence tradeoff to the tradeoff exhibited by ERM for the squared loss. Specifically, one should keep in mind that different loss functions may lead to different minimizers, and \(f^*={\mathrm{argmin}}_{f \in F} \mathbb {E}\ell (f(X)-Y)\) need not be the same as \({\mathrm{argmin}}_{f \in F} \mathbb {E}(f(X)-Y)^2\). As noted previously, one generic situation in which the ‘best in the class’ is the same for every convex loss is when \(Y=f_0(X)+W\) for \(f_0 \in F\) and W that is symmetric and independent of X.

Our results indicate that the choice of the Huber loss with parameter \(\gamma \sim \Vert \xi \Vert _{L_2}\) leads to the removal of outliers—with a dramatic effect when the class is ‘well-behaved’ and the cause of the outliers is the heavy-tailed noise. The effect is diminished when F itself happens to consist of heavy-tailed functions: although it is still true that

$$\begin{aligned}&Pr \left( \sup _{f \in F \cap rD_{f^*}} \left| \frac{1}{N}\sum _{i=1}^N \varepsilon _i \ell ^\prime (\xi _i)(f-f^*)(X_i) \right|> t \right) \nonumber \\&\quad \le 2 Pr \left( \sup _{f \in F \cap rD_{f^*}} \left| \frac{1}{N}\sum _{i=1}^N \varepsilon _i (f-f^*)(X_i) \right| > t/4\gamma \right) , \end{aligned}$$
(7.1)

the random variable \(\sup _{f \in F \cap rD_{f^*}} \left| \sum _{i=1}^N \varepsilon _i (f-f^*)(X_i) \right| \) is likely to be large with a non-negligible probability. As a result, the accuracy/confidence tradeoff exhibited by ERM is far from the conjectured performance of an optimal procedure, even when using a well-calibrated Huber loss.

Unfortunately, the suboptimal accuracy/confidence tradeoff of ERM is a fact of life and not an artifact of the argument we use. The fundamental instability of ERM is caused by the oscillations in (7.1) and cannot be overcome by a wise choice of a loss. Thus, the optimal tradeoff has to be based on a different procedure and not on ERM. Recently, a procedure that attains the optimal accuracy/confidence tradeoff for the squared loss was introduced in [14], and the optimal tradeoff performance was guaranteed under minimal assumptions on the class and on the target.

Finally, it is natural to ask whether it is possible to choose the right calibration for the Huber loss in a data-dependent manner—for example, to use the given data to identify a suitable upper estimate on \(\Vert \xi \Vert _{L_2}\). For the sake of brevity we will sketch the method one may use and only consider the case in which the true minimizer is the same as for the squared lossFootnote 4. Data dependent estimates on \(\Vert \xi \Vert _{L_2}\) are possible using arguments similar to those introduced in [21], where a high-probability, uniform, crude isomorphic estimator of \(L_2\) distances was constructed. More precisely, one may construct a data-dependent functional \(\phi \), which, given a function in a class H whose \(L_2\) norm is above a critical level \(r^*\), returns \(\phi (h)\) such that \(\alpha \Vert h\Vert _{L_2} \le \phi (h) \le \beta \Vert h\Vert _{L_2}\) for suitable constants \(\alpha \) and \(\beta \), and when \(\Vert h\Vert _{L_2} \le r^*\) then \(\phi (h) \le \beta r^*\). Such a functional may be constructed for the class \(H=\{f(X)-Y : f \in F\}\), and by selecting \(\bar{h}={\mathrm{argmin}}_{h \in H} \phi (h)\) and using a standard median-of-means estimator to bound \(\mathbb {E}h^2\), one may obtain a high probability estimator for \(\Vert \xi \Vert _{L_2}\).