1 INTRODUCTION

Consider \(n\in\mathbb{N}^{*}\) independent \(\mathbb{R}^{d}\times\mathbb{R}\)-valued (\(d\in\mathbb{N}^{*}\)) random variables \((X_{1},Y_{1}),\dots,(X_{n},Y_{n})\), having the same probability distribution assumed to be absolutely continuous with respect to Lebesgue’s measure, and

$$\widehat{s}_{K,\ell}(n;x):=\frac{1}{n}\sum_{i=1}^{n}K(X_{i},x)\ell(Y_{i});\quad x\in\mathbb{R}^{d},$$

where \(\ell:\mathbb{R}\rightarrow\mathbb{R}\) is a Borel function and \(K\) is a symmetric continuous map from \(\mathbb{R}^{d}\times\mathbb{R}^{d}\) into \(\mathbb{R}\). This is an estimator of the function \(s:\mathbb{R}^{d}\rightarrow\mathbb{R}\) defined by

$$s(x):=\mathbb{E}(\ell(Y_{1})|X_{1}=x)f(x);\quad\forall x\in\mathbb{R}^{d},$$

where \(f\) is a density of \(X_{1}\). For \(\ell=1\), \(\widehat{s}_{K,\ell}(n;.)\) coincides with the estimator of \(f\) studied in Lerasle et al. [13], covering Parzen-Rosenblatt’s and projection estimators already deeply studied in the literature (see Parzen [16], Rosenblatt [17], Tsybakov [18], etc.), but for \(\ell\not=1\), it covers estimators involved in nonparametric regression. Assume that for every \(i\in\{1,\dots,n\}\),

$$Y_{i}=b(X_{i})+\sigma(X_{i})\varepsilon_{i},$$
(1)

where \(\varepsilon_{i}\) is a centered random variable of variance \(1\), independent of \(X_{i}\), and \(b,\sigma:\mathbb{R}^{d}\rightarrow\mathbb{R}\) are Borel functions.

  • If \(\ell=\textrm{Id}_{\mathbb{R}}\), \(k\) is a symmetric kernel and

    $$K(x^{\prime},x)=\prod_{q=1}^{d}\frac{1}{h_{q}}k\left(\frac{x_{q}^{\prime}-x_{q}}{h_{q}}\right)\textrm{\ with\ }h_{1},\dots,h_{d}>0$$
    (2)

    for every \(x,x^{\prime}\in\mathbb{R}^{d}\), then \(\widehat{s}_{K,\ell}(n;.)\) is the numerator of the well-known Nadaraya-Watson estimator of the regression function \(b\) (see Nadaraya [15] and Watson [20]). Precisely, \(\widehat{s}_{K,\ell}(n;.)\) is an estimator of \(s=bf\) because \(\varepsilon_{1}\) is independent to \(X_{1}\) and \(\mathbb{E}(\varepsilon_{1})=0\). If \(\ell\not=\textrm{Id}_{\mathbb{R}}\), then \(\widehat{s}_{K,\ell}(n;.)\) is the numerator of the estimator studied in Einmahl and Mason [7, 8].

  • If \(\ell=\textrm{Id}_{\mathbb{R}}\), \(\mathcal{B}_{m_{q}}=\{\varphi_{1}^{m_{q}},\dots,\varphi_{m_{q}}^{m_{q}}\}\) (\(m_{q}\in\mathbb{N}^{*}\) and \(q\in\{1,\dots,d\}\)) is an orthonormal family of \(\mathbb{L}^{2}(\mathbb{R})\) and

    $$K(x^{\prime},x)=\prod_{q=1}^{d}\sum_{j=1}^{m_{q}}\varphi_{j}^{m_{q}}(x_{q})\varphi_{j}^{m_{q}}(x_{q}^{\prime})$$
    (3)

    for every \(x,x^{\prime}\in\mathbb{R}^{d}\), then \(\widehat{s}_{K,\ell}(n;.)\) is the projection estimator on \(\mathcal{S}=\textrm{span}(\mathcal{B}_{m_{1}}\otimes\dots\otimes\mathcal{B}_{m_{d}})\) of \(s=bf\).

Now, assume that \(b=0\) in Model (1): for every \(i\in\{1,\dots,n\}\),

$$Y_{i}=\sigma(X_{i})\varepsilon_{i}.$$
(4)

If \(\ell(x)=x^{2}\) for every \(x\in\mathbb{R}\), then \(\widehat{s}_{K,\ell}(n;.)\) is an estimator of \(s=\sigma^{2}f\).

These ten last years, several data-driven procedures have been proposed in order to select the bandwidth of Parzen–Rosenblatt’s estimator (\(\ell=1\) and \(K\) defined by (2)). First, Goldenshluger–Lepski’s method, introduced in [10], which reaches the adequate bias-variance compromise, but is not completely satisfactory on the numerical side (see Comte and Rebafka [5]). More recently, in [12], Lacour, Massart and Rivoirard proposed the PCO (Penalized Comparison to Overfitting) method and proved an oracle inequality for the associated adaptive Parzen-Rosenblatt’s estimator by using a concentration inequality for the U-statistics due to Houdré and Reynaud-Bouret [11]. Together with Varet, they established the numerical efficiency of the PCO method in Varet et al. [19]. Still in the density estimation framework, the PCO method has been extended to bandwidths selection for the recursive Wolverton-Wagner estimator in Comte and Marie [3].

Comte and Marie [4] deal with an oracle inequality and numerical experiments for an adaptive Nadaraya-Watson’s estimator with a numerator and a denominator having distinct bandwidths, both selected via the PCO method. Since the output variable in a regression model has no reason to be bounded, there were significant additional difficulties, bypassed in [4], to establish an oracle inequality for the numerator’s adaptive estimator. Via similar arguments, the present article deals with an oracle inequality for \(\widehat{s}_{\widehat{K},\ell}(n;.)\), where \(\widehat{K}\) is selected via the PCO method in the spirit of Lerasle et al. [13]. As in Comte and Marie [4], one can deduce an oracle inequality for the adaptive quotient estimator \(\widehat{s}_{\widehat{K},\ell}(n;.)/\widehat{s}_{\widehat{L},1}(n;.)\) of \(\mathbb{E}(\ell(Y_{1})|X_{1}=\cdot)\), where \(\widehat{K}\) and \(\widehat{L}\) are both selected via the PCO method.

In addition to the bandwidth selection for kernel-based estimators already studied in [12, 4], the present paper covers the dimension selection for projection estimators of \(f\), \(bf\) when \(Y_{1},\dots,Y_{n}\) are defined by Model (1) with \(\ell=\textrm{Id}_{\mathbb{R}}\), and \(\sigma^{2}f\) when \(Y_{1},\dots,Y_{n}\) are defined by Model (4) with \(\ell(x)=x^{2}\) for every \(x\in\mathbb{R}\). For projection estimators, when \(d=1\), the usual model selection method (see Comte [2, Chapter 2, Section 5]) seems hard to beat. However, when \(d>1\) and \(K\) is defined by (3), \(m_{1},\dots,m_{d}\) are selected via a Goldenshluger–Lepski type method (see Chagny [1]), which has the same numerical weakness than the Goldenshluger–Lepski method for bandwidth selection when \(K\) is defined by (2). So, for the dimension selection for anisotropic projection estimators, the PCO method is interesting.

In Section 2, some examples of kernels sets are provided and a risk bound on \(\widehat{s}_{K,\ell}(n;.)\) is established. Section 3 deals with an oracle inequality for \(\widehat{s}_{\widehat{K},\ell}(n;.)\), where \(\widehat{K}\) is selected via the PCO method.

2 RISK BOUND

Throughout the paper, \(s\in\mathbb{L}^{2}(\mathbb{R}^{d})\). Let \(\mathcal{K}_{n}\) be a set of symmetric continuous maps from \(\mathbb{R}^{d}\times\mathbb{R}^{d}\) into \(\mathbb{R}\), of cardinality less or equal than \(n\), fulfilling the following assumption.

Assumption 2.1. There exists a deterministic constant \(\mathfrak{m}_{\mathcal{K},\ell}>0\), not depending on \(n\), such that

(1) for every \(K\in\mathcal{K}_{n}\),

$$\sup_{x^{\prime}\in\mathbb{R}^{d}}||K(x^{\prime},.)||_{2}^{2}\leqslant\mathfrak{m}_{\mathcal{K},\ell}n;$$

(2) for every \(K\in\mathcal{K}_{n}\),

$$||s_{K,\ell}||_{2}^{2}\leqslant\mathfrak{m}_{\mathcal{K},\ell}$$

with

$$s_{K,\ell}:=\mathbb{E}(\widehat{s}_{K,\ell}(n;.))=\mathbb{E}(K(X_{1},.)\ell(Y_{1}));$$

(3) for every \(K,K^{\prime}\in\mathcal{K}_{n}\),

$$\mathbb{E}(\langle K(X_{1},.),K^{\prime}(X_{2},.)\ell(Y_{2})\rangle_{2}^{2})\leqslant\mathfrak{m}_{\mathcal{K},\ell}\overline{s}_{K^{\prime},\ell}$$

with

$$\overline{s}_{K^{\prime},\ell}:=\mathbb{E}(||K^{\prime}(X_{1},.)\ell(Y_{1})||_{2}^{2}).$$

(4) for every \(K\in\mathcal{K}_{n}\) and \(\psi\in\mathbb{L}^{2}(\mathbb{R}^{d})\),

$$\mathbb{E}(\langle K(X_{1},.),\psi\rangle_{2}^{2})\leqslant\mathfrak{m}_{\mathcal{K},\ell}||\psi||_{2}^{2}.$$

The elements of \(\mathcal{K}_{n}\) are called kernels. Let us provide two natural examples of kernels sets.

Proposition 2.2. Consider

$$\mathcal{K}_{k}(h_{\textrm{min}}):=\left\{(x^{\prime},x)\mapsto\prod_{q=1}^{d}\frac{1}{h_{q}}k\left(\frac{x_{q}^{\prime}-x_{q}}{h_{q}}\right);\ h_{1},\dots,h_{d}\in\mathcal{H}(h_{\textrm{min}})\right\},$$

where \(k\) is a symmetric kernel (in the usual sense), \(h_{\textrm{min}}\in[n^{-1/d},1]\) and \(\mathcal{H}(h_{\textrm{min}})\) is a finite subset of \([h_{\textrm{min}},1]\). The kernels set \(\mathcal{K}_{k}(h_{\textrm{min}})\) fulfills Assumption 2.1 and, for any \(K\in\mathcal{K}_{k}(h_{\textrm{min}})\) (i.e., defined by (2) with \(h_{1},\dots,h_{d}\in\mathcal{H}(h_{\textrm{min}})\)),

$$\overline{s}_{K,\ell}=||k||_{2}^{2d}\mathbb{E}(\ell(Y_{1})^{2})\prod_{q=1}^{d}\frac{1}{h_{q}}.$$

Proposition 2.3. Consider

$$\mathcal{K}_{\mathcal{B}_{1},\dots,\mathcal{B}_{n}}(m_{\textrm{max}}):=\left\{(x^{\prime},x)\mapsto\prod_{q=1}^{d}\sum_{j=1}^{m_{q}}\varphi_{j}^{m_{q}}(x_{q})\varphi_{j}^{m_{q}}(x_{q}^{\prime});\ m_{1},\dots,m_{d}\in\{1,\dots,m_{\textrm{max}}\}\right\},$$

where \(m_{\textrm{max}}^{d}\in\{1,\dots,n\}\) and, for every \(m\in\{1,\dots,n\}\), \(\mathcal{B}_{m}=\{\varphi_{1}^{m},\dots,\varphi_{m}^{m}\}\) is an orthonormal family of \(\mathbb{L}^{2}(\mathbb{R})\) such that

$$\sup_{x^{\prime}\in\mathbb{R}}\sum_{j=1}^{m}\varphi_{j}^{m}(x^{\prime})^{2}\leqslant\mathfrak{m}_{\mathcal{B}}m,$$

with \(\mathfrak{m}_{\mathcal{B}}>0\) not depending on \(m\) and \(n\), and such that one of the two following conditions is satisfied:

$$\mathcal{B}_{m}\subset\mathcal{B}_{m+1};\quad\forall m\in\{1,\dots,n-1\}$$
(5)

or

$$\overline{\mathfrak{m}}_{\mathcal{B}}:=\sup\{|\mathbb{E}(K(X_{1},x))|;\ K\in\mathcal{K}_{\mathcal{B}_{1},\dots,\mathcal{B}_{n}}(m_{\textrm{max}})\quad\textit{and}\quad x\in\mathbb{R}^{d}\}\textit{ is finite and doesn't depend on $n$.}$$
(6)

The kernels set \(\mathcal{K}_{\mathcal{B}_{1},\dots,\mathcal{B}_{n}}(m_{\textrm{max}})\) fulfills Assumption 2.1 and, for any \(K\in\mathcal{K}_{\mathcal{B}_{1},\dots,\mathcal{B}_{n}}(m_{\textrm{max}})\) (i.e., defined by (3) with \(m_{1},\dots,m_{n}\in\{1,\dots,m_{\textrm{max}}\}\)),

$$\overline{s}_{K,\ell}\leqslant\mathfrak{m}_{\mathcal{B}}^{d}\mathbb{E}(\ell(Y_{1})^{2})\prod_{q=1}^{d}m_{q}.$$

Remark 2.4. For the sake of simplicity, the present paper focuses on \(\mathcal{K}_{\mathcal{B}_{1},\dots,\mathcal{B}_{n}}(m_{\textrm{max}})\), but Proposition 2.3 is still true for the weighted projection kernels set

$$\mathcal{K}_{\mathcal{B}_{1},\dots,\mathcal{B}_{n}}(w_{1},\dots,w_{n};m_{\textrm{max}})$$
$${}:=\left\{(x^{\prime},x)\mapsto\prod_{q=1}^{d}\sum_{j=1}^{m_{q}}w_{j}\varphi_{j}^{m_{q}}(x_{q})\varphi_{j}^{m_{q}}(x_{q}^{\prime});\ m_{1},\dots,m_{d}\in\{1,\dots,m_{\textrm{max}}\}\right\},$$

where \(w_{1},\dots,w_{n}\in[0,1]\).

Remark 2.5. Note that Condition (5) is close, but more restrictive than Condition (19) of Lerasle et al. [13], Proposition 3.2, which is that the spaces \({\textrm{span}}(\mathcal{B}_{m})\), \(m\in\mathbb{N}\) are nested. See Massart [14], Subsection 7.5.2 for examples of nested spaces. Our condition (5) is fulfilled by the trigonometric basis, Hermite’s basis or Laguerre’s basis.

Note also that in the same proposition of Lerasle et al. [13], Condition (20) coincides with our condition (6). The regular histograms basis satisfies condition (6). Indeed, by taking \(\varphi_{j}^{m}=\psi_{j}^{m}:=\sqrt{m}\mathbf{1}_{[(j-1)/m,j/m[}\) for every \(m\in\{1,\dots,n\}\) and \(j\in\{1,\dots,m\}\),

$$\left|\mathbb{E}\left[\prod_{q=1}^{d}\sum_{j=1}^{m_{q}}\psi_{j}^{m_{q}}(X_{1,q})\psi_{j}^{m_{q}}(x_{q})\right]\right|=\sum_{j_{1}=1}^{m_{1}}\cdots\sum_{j_{d}=1}^{m_{d}}\left(\prod_{q=1}^{d}m_{q}\mathbf{1}_{[(j_{q}-1)/m_{q},j_{q}/m_{q}[}(x_{q})\right)$$
$${}\times\int\limits_{(j_{1}-1)/m_{1}}^{j_{1}/m_{1}}\cdots\int\limits_{(j_{d}-1)/m_{d}}^{j_{d}/m_{d}}f(x_{1}^{\prime},\dots,x_{d}^{\prime})dx_{1}^{\prime}\cdots dx_{d}^{\prime}$$
$${}\leqslant||f||_{\infty}\prod_{q=1}^{d}\sum_{j=1}^{m_{q}}\mathbf{1}_{[(j-1)/m_{q},j/m_{q}[}(x_{q})\leqslant||f||_{\infty}$$

for every \(m_{1},\dots,m_{d}\in\{1,\dots,n\}\) and \(x\in\mathbb{R}^{d}\).

The following proposition shows that Legendre’s basis also fulfills condition (6).

Proposition 2.6. For every \(m\in\{1,\dots,n\}\) and \(j\in\{1,\dots,m\}\), let \(\xi_{j}^{m}\) be the function defined on \([-1,1]\) by

$$\xi_{j}^{m}(x):=\sqrt{\frac{2j+1}{2}}Q_{j}(x);\quad\forall x\in[-1,1],$$

where

$$Q_{j}:x\in[-1,1]\longmapsto\frac{1}{2^{j}j!}\cdot\frac{d^{j}}{dx^{j}}(x^{2}-1)^{j}$$

is the \(j\)th Legendre’s polynomial. If \(f\in C^{2d}([0,1]^{d})\) and \(\mathcal{B}_{m}=\{\xi_{1}^{m},\dots,\xi_{m}^{m}\}\) for every \(m\in\{1,\cdots,n\}\), then \(\mathcal{K}_{\mathcal{B}_{1},\dots,\mathcal{B}_{m}}(m_{\textrm{max}})\) fulfills condition (6).

The following proposition provides a suitable control of the variance of \(\widehat{s}_{K,\ell}(n;.)\).

Proposition 2.7. Under Assumptions 2.1.(1)–2.1.(3), if \(s\in\mathbb{L}^{2}(\mathbb{R}^{d})\) and if there exists \(\alpha>0\) such that \(\mathbb{E}(\exp(\alpha|\ell(Y_{1})|))<\infty\), then there exists a deterministic constant \(\mathfrak{c}_{2.7}>0\), not depending on \(n\), such that for every \(\theta\in]0,1[\),

$$\mathbb{E}\left(\sup_{K\in\mathcal{K}_{n}}\left\{\left|||\widehat{s}_{K,\ell}(n;.)-s_{K,\ell}||_{2}^{2}-\frac{\overline{s}_{K,\ell}}{n}\right|-\frac{\theta}{n}\overline{s}_{K,\ell}\right\}\right)\leqslant\mathfrak{c}_{2.7}\frac{\log(n)^{5}}{\theta n}.$$

Finally, let us state the main result of this section.

Theorem 2.8. Under Assumption 2.1, if \(s\in\mathbb{L}^{2}(\mathbb{R}^{d})\) and if there exists \(\alpha>0\) such that \(\mathbb{E}(\exp(\alpha|\ell(Y_{1})|))<\infty\), then there exist deterministic constants \(\mathfrak{c}_{2.8},\overline{\mathfrak{c}}_{2.8}>0\), not depending on \(n\), such that for every \(\theta\in]0,1[\),

$$\mathbb{E}\left(\sup_{K\in\mathcal{K}_{n}}\left\{||\widehat{s}_{K,\ell}(n;.)-s||_{2}^{2}-(1+\theta)\left(||s_{K,\ell}-s||_{2}^{2}+\frac{\overline{s}_{K,\ell}}{n}\right)\right\}\right)\leqslant\mathfrak{c}_{2.8}\frac{\log(n)^{5}}{\theta n}$$

and

$$\mathbb{E}\left(\sup_{K\in\mathcal{K}_{n}}\left\{||s_{K,\ell}-s||_{2}^{2}+\frac{\overline{s}_{K,\ell}}{n}-\frac{1}{1-\theta}||\widehat{s}_{K,\ell}(n;.)-s||_{2}^{2}\right\}\right)\leqslant\overline{\mathfrak{c}}_{2.8}\frac{\log(n)^{5}}{\theta(1-\theta)n}.$$

Remark 2.9. Note that the first inequality in Theorem 2.8 gives a risk bound on the estimator \(\widehat{s}_{K,\ell}(n;.)\):

$$\mathbb{E}(||\widehat{s}_{K,\ell}(n;.)-s||_{2}^{2})\leqslant(1+\theta)\left(||s_{K,\ell}-s||_{2}^{2}+\frac{\overline{s}_{K,\ell}}{n}\right)+\mathfrak{c}_{2.8}\frac{\log(n)^{5}}{\theta n}$$

for every \(\theta\in]0,1[\). The second inequality is useful in order to establish a risk bound on the adaptive estimator defined in the next section (see Theorem 3.2).

Remark 2.10. In Proposition 2.7 and Theorem 2.8, the exponential moment condition may appear too strong. Nevertheless, this is de facto satisfied when

$$\ell(Y_{1}),\dots,\ell(Y_{n})\textit{ have a compactly supported distribution.}$$
(7)

This last condition is satisfied in the density estimation framework because \(\ell=1\), but even in the nonparametric regression framework, where \(\ell\) is not bounded, when \(Y_{1},\dots,Y_{n}\) have a compactly supported distribution. Moreover, note that under condition (7), the risk bounds of Theorem 2.8 can be stated in deviation, without additional steps in the proof. Precisely, under Assumption 2.1 and condition (7), if \(s\in\mathbb{L}^{2}(\mathbb{R}^{d})\), then there exists a deterministic constant \(\mathfrak{c}_{L}>0\), depending on \(L=\sup_{z\in{\textrm{supp}}(\mathbb{P}_{Y_{1}})}|\ell(z)|\) but not on \(n\), such that for every \(\vartheta\in]0,1[\) and \(\lambda>0\),

$$\sup_{K\in\mathcal{K}_{n}}\left|||s_{K,\ell}-s||_{2}^{2}+\frac{\overline{s}_{K,\ell}}{n}-\frac{1}{1-\vartheta}||\widehat{s}_{K,\ell}(n;.)-s||_{2}^{2}\right|\leqslant\frac{\mathfrak{c}_{L}}{\vartheta n}(1+\lambda)^{3}$$

with probability larger than \(1-9.4|\mathcal{K}_{n}|e^{-\lambda}\).

When condition (7) doesn’t hold true, one can replace the exponential moment condition of Proposition 2.7 and Theorem 2.8 by a \(q\)th order moment condition on \(\ell(Y_{1})\) (\(q\in\mathbb{N}^{*}\)), but with a damaging effect on the rate of convergence of \(\widehat{s}_{K,\ell}(n;.)\). For instance, at Remark B.5, it is established that under a \((12-4\varepsilon)/\beta\)th moment condition (\(\varepsilon\in]0,1[\) and \(0<\beta<\varepsilon/2\)), the rate of convergence is of order \(O(1/n^{1-\varepsilon})\) (instead of \(1/n\)) in Lemma B.2. This holds true for the three technical lemmas of Appendix B.1, and then for Proposition 2.7 and Theorem 2.8.

3 KERNEL SELECTION

This section deals with a risk bound on the adaptive estimator \(\widehat{s}_{\widehat{K},\ell}(n;.)\), where

$$\widehat{K}\in\arg\min_{K\in\mathcal{K}_{n}}\{||\widehat{s}_{K,\ell}(n;\cdot)-\widehat{s}_{K_{0},\ell}(n;\cdot)||_{2}^{2}+\textrm{pen}_{\ell}(K)\},$$

\(K_{0}\) is an overfitting proposal for \(K\) in the sense that

$$K_{0}\in\arg\max_{K\in\mathcal{K}_{n}}\left\{\sup_{x\in\mathbb{R}^{d}}|K(x,x)|\right\},$$

and

$$\textrm{pen}_{\ell}(K):=\frac{2}{n^{2}}\sum_{i=1}^{n}\langle K(.,X_{i}),K_{0}(.,X_{i})\rangle_{2}\ell(Y_{i})^{2};\quad\forall K\in\mathcal{K}_{n}.$$
(8)

Example. On the one hand, for any \(K\in\mathcal{K}_{k}(h_{\textrm{min}})\) (i.e., defined by (2) with \(h_{1},\dots,h_{d}\in\mathcal{H}(h_{\textrm{min}})\)),

$$\sup_{x\in\mathbb{R}^{d}}|K(x,x)|=|k(0)|^{d}\prod_{q=1}^{d}\frac{1}{h_{q}}.$$

Then, for \(\mathcal{K}_{n}=\mathcal{K}_{k}(h_{\textrm{min}})\),

$$K_{0}(x^{\prime},x)=\frac{1}{h_{\textrm{min}}^{d}}\prod_{q=1}^{d}k\left(\frac{x_{q}^{\prime}-x_{q}}{h_{\textrm{min}}}\right);\quad\forall x,x^{\prime}\in\mathbb{R}^{d}.$$

On the other hand, for any \(K\in\mathcal{K}_{\mathcal{B}_{1},\dots,\mathcal{B}_{n}}(m_{\textrm{max}})\) (i.e., defined by (3) with \(m_{1},\dots,m_{n}\in\{1,\dots,m_{\textrm{max}}\}\)),

$$\sup_{x\in\mathbb{R}^{d}}|K(x,x)|=\sup_{x\in\mathbb{R}^{d}}\prod_{q=1}^{d}\sum_{j=1}^{m_{q}}\varphi_{j}^{m_{q}}(x_{q})^{2}.$$

Then, for \(\mathcal{K}_{n}=\mathcal{K}_{\mathcal{B}_{1},\dots,\mathcal{B}_{n}}(m_{\textrm{max}})\), at least for the usual bases mentioned at Remark 2.5,

$$K_{0}(x^{\prime},x)=\prod_{q=1}^{d}\sum_{j=1}^{m_{\textrm{max}}}\varphi_{j}^{m_{\textrm{max}}}(x_{q})\varphi_{j}^{m_{\textrm{max}}}(x_{q}^{\prime});\quad\forall x,x^{\prime}\in\mathbb{R}^{d}.$$

In the sequel, in addition to Assumption 2.1, the kernels set \(\mathcal{K}_{n}\) fulfills the following assumption.

Assumption 3.1. There exists a deterministic constant \(\overline{\mathfrak{m}}_{\mathcal{K},\ell}>0\), not depending on \(n\), such that

$$\mathbb{E}\left(\sup_{K,K^{\prime}\in\mathcal{K}_{n}}\langle K(X_{1},.),s_{K^{\prime},\ell}\rangle_{2}^{2}\right)\leqslant\overline{\mathfrak{m}}_{\mathcal{K},\ell}.$$

The following theorem provides an oracle inequality for the adaptive estimator \(\widehat{s}_{\widehat{K},\ell}(n;.)\).

Theorem 3.2. Under Assumptions 2.1 and 3.1, if \(s\in\mathbb{L}^{2}(\mathbb{R}^{d})\) and if there exists \(\alpha>0\) such that \(\mathbb{E}(\exp(\alpha|\ell(Y_{1})|))<\infty\), then there exists a deterministic constant \(\mathfrak{c}_{3.2}>0\), not depending on \(n\), such that for every \(\vartheta\in]0,1[\),

$$\mathbb{E}(||\widehat{s}_{\widehat{K},\ell}(n;.)-s||_{2}^{2})\leqslant(1+\vartheta)\min_{K\in\mathcal{K}_{n}}\mathbb{E}(||\widehat{s}_{K,\ell}(n;.)-s||_{2}^{2})+\frac{\mathfrak{c}_{3.2}}{\vartheta}\left(||s_{K_{0},\ell}-s||_{2}^{2}+\frac{\log(n)^{5}}{n}\right).$$

Remark 3.3. As mentioned in Comte and Marie [4, p. 6], when \(\mathcal{K}_{n}=\mathcal{K}_{k}(h_{\textrm{min}})\), if \(s\) belongs to a Nikol’skii ball and \(h_{\textrm{min}}=1/n\), then Theorem 3.2 says that the PCO estimator has a performance of same order than \(O_{n}:=\min_{K\in\mathcal{K}_{n}}\mathbb{E}(||\widehat{s}_{K,\ell}(n;.)-s||_{2}^{2})\) up to a factor \(1+\vartheta\). When \(\mathcal{K}_{n}=\mathcal{K}_{\mathcal{B}_{1},\dots,\mathcal{B}_{n}}(m_{\textrm{max}})\), it depends on the bases \(\mathcal{B}_{1},\dots,\mathcal{B}_{n}\). For instance, with the same ideas than in Comte and Marie [4], thanks to DeVore and Lorentz [6, Theorem 2.3 p. 205], if \(s\) belongs to a Sobolev space and \(m_{\textrm{max}}=n\), then our Theorem 3.2 also says that the PCO estimator has a performance of same order than \(O_{n}\).

Notation. For any \(B\in\mathcal{B}(\mathbb{R}^{d})\), \(||.||_{2,f,B}\) is the norm on \(\mathbb{L}^{2}(B,f(x)\lambda_{d}(dx))\) defined by

$$||\varphi||_{2,f,B}:=\left(\int\limits_{B}\varphi(x)^{2}f(x)\lambda_{d}(dx)\right)^{1/2}\textrm{; }\forall\varphi\in\mathbb{L}^{2}(B,f(x)\lambda_{d}(dx)).$$

The following corollary provides an oracle inequality for \(\widehat{s}_{\widehat{K},\ell}(n;.)/\widehat{s}_{\widehat{L},1}(n;.)\), where \(\widehat{K}\) and \(\widehat{L}\) are both selected via the PCO method.

Corollary 3.4. Let \((\beta_{j})_{j\in\mathbb{N}}\) be a decreasing sequence of elements of \(]0,\infty[\) such that \(\lim_{\infty}\beta_{j}=0\) and, for every \(j\in\mathbb{N}\), consider

$$B_{j}:=\{x\in\mathbb{R}^{d}:f(x)\geqslant\beta_{j}\}.$$

Under Assumptions 2.1 and 3.1 for \(\ell\) and \(1\), if \(s,f\in\mathbb{L}^{2}(\mathbb{R}^{d})\) and if there exists \(\alpha>0\) such that \(\mathbb{E}(\exp(\alpha|\ell(Y_{1})|))<\infty\), then there exists a deterministic constant \(\mathfrak{c}_{3.2}>0\), not depending on \(n\), such that for every \(\vartheta\in]0,1[\),

$$\mathbb{E}\left[\left|\left|\frac{\widehat{s}_{\widehat{K},\ell}(n;.)}{\widehat{s}_{\widehat{L},1}(n;.)}-\frac{s}{f}\right|\right|_{2,f,B_{n}}^{2}\right]\leqslant\frac{\mathfrak{c}_{3.2}}{\beta_{n}^{2}}\left[(1+\vartheta)\min_{(K,L)\in\mathcal{K}_{n}^{2}}\{\mathbb{E}(||\widehat{s}_{K,\ell}(n;.)-s||_{2}^{2})+\mathbb{E}(||\widehat{s}_{L,1}(n;.)-f||_{2}^{2})\}\right.$$
$${}+\left.\frac{1}{\vartheta}\left(||s_{K_{0},\ell}-s||_{2}^{2}+||s_{K_{0},1}-f||_{2}^{2}+\frac{\log(n)^{5}}{n}\right)\right],$$

where

$$\widehat{K}\in\arg\min_{K\in\mathcal{K}_{n}}\{||\widehat{s}_{K,\ell}(n;\cdot)-\widehat{s}_{K_{0},\ell}(n;\cdot)||_{2}^{2}+{\textrm{pen}}_{\ell}(K)\}$$

and

$$\widehat{L}\in\arg\min_{L\in\mathcal{K}_{n}}\{||\widehat{s}_{L,1}(n;\cdot)-\widehat{s}_{K_{0},1}(n;\cdot)||_{2}^{2}+{\textrm{pen}}_{1}(L)\}.$$

The proof of Corollary 3.4 is the same than the proof of Comte and Marie [4], Corollary 4.3.

Finally, let us discuss about Assumption 3.1. This assumption is difficult to check in practice, then let us provide a sufficient condition.

Assumption 3.5. The function \(s\) is bounded and

$$\mathfrak{m}_{\mathcal{K}}:=\sup\{||K(x^{\prime},.)||_{1}^{2};\ K\in\mathcal{K}_{n}\quad\textrm{and}\quad x^{\prime}\in\mathbb{R}^{d}\}$$

doesn’t depend on \(n\).

Under Assumption 3.5, \(\mathcal{K}_{n}\) fulfills Assumption 3.1. Indeed,

$$\mathbb{E}\left(\sup_{K,K^{\prime}\in\mathcal{K}_{n}}\langle K(X_{1},.),s_{K^{\prime},\ell}\rangle_{2}^{2}\right)\leqslant\left(\sup_{K^{\prime}\in\mathcal{K}_{n}}||s_{K^{\prime},\ell}||_{\infty}^{2}\right)\mathbb{E}\left(\sup_{K\in\mathcal{K}_{n}}||K(X_{1},.)||_{1}^{2}\right)$$
$${}\leqslant\mathfrak{m}_{\mathcal{K}}\sup\left\{\left(\int\limits_{-\infty}^{\infty}|K^{\prime}(x^{\prime},x)s(x)|dx\right)^{2};\ K^{\prime}\in\mathcal{K}_{n}\quad\textrm{and}\quad x^{\prime}\in\mathbb{R}\right\}\leqslant\mathfrak{m}_{\mathcal{K}}^{2}||s||_{\infty}^{2}.$$

Note that in the nonparametric regression framework (see Model (1)), to assume \(s\) bounded means that \(bf\) is bounded. For instance, this condition is fulfilled by the linear regression models with Gaussian inputs.

Let us provide two examples of kernels sets fulfilling Assumption 3.5, the sufficient condition for Assumption 3.1:

  • Consider \(K\in\mathcal{K}_{k}(h_{\textrm{min}})\). Then, there exist \(h_{1},\dots,h_{d}\in\mathcal{H}(h_{\textrm{min}})\) such that

    $$K(x^{\prime},x)=\prod_{q=1}^{d}\frac{1}{h_{q}}k\left(\frac{x_{q}^{\prime}-x_{q}}{h_{q}}\right);\quad\forall x,x^{\prime}\in\mathbb{R}^{d}.$$

    Clearly, \(||K(x^{\prime},.)||_{1}=||k||_{1}^{d}\) for every \(x^{\prime}\in\mathbb{R}^{d}\). So, for \(\mathcal{K}_{n}=\mathcal{K}_{k}(h_{\textrm{min}})\), \(\mathfrak{m}_{\mathcal{K}}\leqslant||k||_{1}^{2d}\).

  • For \(\mathcal{K}_{n}=\mathcal{K}_{\mathcal{B}_{1},\dots,\mathcal{B}_{n}}(m_{\textrm{max}})\), the condition on \(\mathfrak{m}_{\mathcal{K}}\) seems harder to check in general. Let us show that it is satisfied for the regular histograms basis defined in Section 2. For every \(m_{1},\dots,m_{d}\in\{1,\dots,n\}\),

    $$\left|\left|\prod_{q=1}^{d}\sum_{j=1}^{m_{q}}\psi_{j}^{m_{q}}(x_{q}^{\prime})\psi_{j}^{m_{q}}(.)\right|\right|_{1}\leqslant\prod_{q=1}^{d}\left(m_{q}\sum_{j=1}^{m_{q}}\mathbf{1}_{[(j-1)/m_{q},j/m_{q}[}(x_{q}^{\prime})\int\limits_{(j-1)/m_{q}}^{j/m_{q}}dx\right)\leqslant 1.$$

Now, let us show that even if it doesn’t fulfill Assumption 3.5, the trigonometric basis fulfills Assumption 3.1.

Proposition 3.6. Consider \(\chi_{1}:=\mathbf{1}_{[0,1]}\) and, for every \(j\in\mathbb{N}^{*}\), the functions \(\chi_{2j}\) and \(\chi_{2j+1}\) defined on \(\mathbb{R}\) by

$$\chi_{2j}(x):=\sqrt{2}\cos(2\pi jx)\mathbf{1}_{[0,1]}(x)\quad{and}\quad\chi_{2j+1}(x):=\sqrt{2}\sin(2\pi jx)\mathbf{1}_{[0,1]}(x);\quad\forall x\in\mathbb{R}.$$

If \(s\in C^{2}(\mathbb{R}^{d})\) and \(\mathcal{B}_{m}=\{\chi_{1},\dots,\chi_{m}\}\) for every \(m\in\{1,\dots,n\}\), then \(\mathcal{K}_{\mathcal{B}_{1},\dots,\mathcal{B}_{n}}(m_{\textrm{max}})\) fulfills Assumption 3.1.