Log in

Adaptive Minimax Testing for Circular Convolution

  • Published:
Mathematical Methods of Statistics Aims and scope Submit manuscript

Abstract

Given observations from a circular random variable contaminated by an additive measurement error, we consider the problem of minimax optimal goodness-of-fit testing in a non-asymptotic framework. We propose direct and indirect testing procedures using a projection approach. The structure of the optimal tests depends on regularity and ill-posedness parameters of the model, which are unknown in practice. Therefore, adaptive testing strategies that perform optimally over a wide range of regularity and ill-posedness classes simultaneously are investigated. Considering a multiple testing procedure, we obtain adaptive i.e. assumption-free procedures and analyse their performance. Compared with the non-adaptive tests, their radii of testing face a deterioration by a log-factor. We show that for testing of uniformity this loss is unavoidable by providing a lower bound. The results are illustrated considering Sobolev spaces and ordinary or super smooth error densities.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Subscribe and save

Springer+ Basic
EUR 32.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or Ebook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Price includes VAT (France)

Instant access to the full article PDF.

Similar content being viewed by others

REFERENCES

  1. Y. Baraud, ‘‘Non-asymptotic minimax rates of testing in signal detection,’’ Bernoulli 8 (5), 577–606 (2002).

    MathSciNet  MATH  Google Scholar 

  2. C. Butucea, ‘‘Goodness-of-fit testing and quadratic functional estimation from indirect observations,’’ The Annals of Statistics 35 (5), 1907–1930 (2007).

    MathSciNet  MATH  Google Scholar 

  3. C. Butucea and K. Meziani, ‘‘Quadratic functional estimation in inverse problems,’’ Statistical Methodology 8 (1), 31–41 (2011).

    Article  MathSciNet  Google Scholar 

  4. C. Butucea, C. Matias, and C. Pouet, ‘‘Adaptive goodness-of-fit testing from indirect observations,’’ Annales de l’Institut Henri Poincaré, Probabilités et Statistiques 45, 352–372 (2009).

    Article  MathSciNet  Google Scholar 

  5. O. Collier, L. Comminges, and A. B. Tsybakov, ‘‘Minimax estimation of linear and quadratic functionals on sparsity classes,’’ The Annals of Statistics 45 (3), 923–958 (2017).

    Article  MathSciNet  Google Scholar 

  6. F. Comte and M.-L. Taupin, ‘‘Adaptive density deconvolution for circular data.’’ http://helios.mi.parisdescartes.fr/ comte/efrodens5.pdf (2003).

  7. S. Efromovich, ‘‘Density estimation for the case of supersmooth measurement error,’’ Journal of the American Statistical Association 92 (438), 526–535 (1997).

    Article  MathSciNet  Google Scholar 

  8. N. I. Fisher, Statistical Analysis of Circular Data (Cambridge University Press, 1995).

    Google Scholar 

  9. M. Fromont and B. Laurent, ‘‘Adaptive goodness-of-fit tests in a density model,’’ The Annals of Statistics 34 (2), 680–720 (2006).

    Article  MathSciNet  Google Scholar 

  10. J. Gill and D. Hangartner, ‘‘Circular data in political science and how to handle it,’’ Political Analysis 18 (3), 316–336 (2010).

    Article  Google Scholar 

  11. E. Gine and R. Nickl, Mathematical Foundations of Infinite-Dimensional Statistical Models (Cambridge University Press, 2015).

    MATH  Google Scholar 

  12. Y. Ingster, ‘‘Asymptotically minimax hypothesis testing for nonparametric alternatives I,’’ Mathematical Methods of Statistics 2 (2), 85–114 (1993).

    MathSciNet  MATH  Google Scholar 

  13. Y. Ingster, ‘‘Asymptotically minimax hypothesis testing for nonparametric alternatives II,’’ Mathematical Methods of Statistics 2 (2), 171–189 (1993).

    MathSciNet  MATH  Google Scholar 

  14. Y. Ingster, ‘‘Asymptotically minimax hypothesis testing for nonparametric alternatives III,’’ Mathematical Methods of Statistics 2 (2), 249—268 (1993).

    MathSciNet  MATH  Google Scholar 

  15. Y. Ingster and I. A. Suslina, Nonparametric Goodness-of-Fit Testing under Gaussian Models (Springer, New York, 2012).

    MATH  Google Scholar 

  16. J. Johannes and M. Schwarz, ‘‘Adaptive circular deconvolution by model selection under unknown error distribution,’’ Bernoulli 19 (5A), 1576–1611 (2013).

    Article  MathSciNet  Google Scholar 

  17. G. Kerkyacharian, T. M. Pham Ngoc, and D. Picard, ‘‘Localized spherical deconvolution,’’ The Annals of Statistics 39 (2), 1042–1068 (2011).

    Article  MathSciNet  Google Scholar 

  18. M. Kroll, ‘‘Rate optimal estimation of quadratic functionals in inverse problems with partially unknown operator and application to testing problems,’’ ESAIM: Probability and Statistics 23, 524–551 (2019).

    Article  MathSciNet  Google Scholar 

  19. C. Lacour and T. M. P. Ngoc, ‘‘Goodness-of-fit test for noisy directional data,’’ Bernoulli 20 (4), 2131–2168 (2014).

    Article  MathSciNet  Google Scholar 

  20. B. Laurent, ‘‘Adaptive estimation of a quadratic functional of a density by model selection,’’ ESAIM: Probability and Statistics 9, 1–18 (2005).

    Article  MathSciNet  Google Scholar 

  21. B. Laurent and P. Massart, ‘‘Adaptive estimation of a quadratic functional by model selection,’’ The Annals of Statistics 28(5), 1302–1338 (2000).

    Article  MathSciNet  Google Scholar 

  22. B. Laurent, S. Huet, and Y. Baraud, ‘‘Adaptive tests of linear hypotheses by model selection,’’ The Annals of Statistics 31 (1), 225–251 (2003).

    MathSciNet  MATH  Google Scholar 

  23. B. Laurent, J.-M. Loubes, and C. Marteau, ‘‘Testing inverse problems: a direct or an indirect problem?,’’ Journal of Statistical Planning and Inference 141 (5), 1849–1861 (2011).

    Article  MathSciNet  Google Scholar 

  24. B. Laurent, J.-M. Loubes, and C. Marteau, ‘‘Non asymptotic minimax rates of testing in signal detection with heterogeneous variances,’’ Electronic Journal of Statistics 6, 91–122 (2012).

    Article  MathSciNet  Google Scholar 

  25. K. V. Mardia, Statistics of Directional Data (Academic press, 1972).

    MATH  Google Scholar 

  26. K. V. Mardia and P. E. Jupp, Directional Statistics (John Wiley and Sons, 2009).

    MATH  Google Scholar 

  27. C. Marteau and T. Sapatinas, ‘‘A unified treatment for non-asymptotic and asymptotic approaches to minimax signal detection,’’ Statistics Surveys 9, 253–297 (2015).

    Article  MathSciNet  Google Scholar 

  28. C. Marteau and T. Sapatinas, ‘‘Minimax goodness-of-fit testing in ill-posed inverse problems with partially unknown operators,’’ Annales de l’Institut Henri Poincaré, Probabilités et Statistiques 53 (4), 1675–1718 (2017).

    Article  MathSciNet  Google Scholar 

  29. A. Meister, Deconvolution Problems in Nonparametric Statistics (Springer, New York, 2009).

    Book  Google Scholar 

  30. S. Schluttenhofer and J. Johannes, ‘‘Adaptive minimax testing in inverse Gaussian sequence space models,’’ ar**v preprint ar**v:2002.07623 (2020).

  31. S. Schluttenhofer and J. Johannes, ‘‘Minimax testing and quadratic functional estimation for circular convolution,’’ ar**v preprint ar**v:2004.12714 (2020).

  32. V. Spokoiny, ‘‘Adaptive hypothesis testing using wavelets,’’ The Annals of Statistics 24 (6), 2477–2498 (1996).

    Article  MathSciNet  Google Scholar 

  33. A. B. Tsybakov, Introduction to Nonparametric Estimation (Springer, New York, 2009).

    Book  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Sandra Schluttenhofer or Jan Johannes.

Appendices

AUXILIARY RESULTS USED IN SECTIONS 2 AND 4

The next two assertions, a concentration inequality for canonical U-statistics and a Bernstein inequality, provide our key arguments in order to control the deviation of the test statistics. The first assertion is a reformulation of Theorem 3.4.8 in [11].

Proposition A.1. Let \(\{Y_{j}\}_{j=1}^{n}\) be independent and identically distributed \([0,1)\)-valued random variables, and let \(h:[0,1)^{2}\to\mathbb{R}\) be a bounded symmetric kernel, i.e., \(h(y,\tilde{y})=h(\tilde{y},y)\) for all \(y,\tilde{y}\in[0,1)\), fulfilling in addition

$$\mathbb{E}(h(Y_{1},y_{2}))=0\quad\forall y_{2}\in[0,1).$$
(A.1)

Then there are finite constantsA, B, C, and \(\mathrm{d}\) such that

$$\sup_{y_{1},y_{2}\in[0,1)}|h(y_{1},y_{2})|\leq\textrm{A},\quad\sup_{y_{2}\in[0,1)}\mathbb{E}h^{2}(Y_{1},y_{2})\leq\textrm{B}^{2},\quad\mathbb{E}h^{2}(Y_{1},Y_{2})\leq\textrm{C}^{2},\quad{and}$$
$$\sup\{\mathbb{E}\left(h(Y_{1},Y_{2})\zeta(Y_{1})\xi(Y_{2})\right),\mathbb{E}\zeta^{2}(Y_{1})\leq 1,\mathbb{E}\xi^{2}(Y_{2})\leq 1\}\leq\mathrm{d}$$
(A.2)

and for all \(n\geq 2\) the real-valued canonical U-statistic

$$\textrm{U}_{n}=\frac{1}{n(n-1)}\sum_{l,m\in[\![n]\!]\atop l\neq m}h(Y_{l},Y_{m})$$

satisfies for all \(x\geq 0\)

$$\mathbb{P}(\textrm{U}_{n}\geq 8\textrm{C}n^{-1}x^{1/2}+13\textrm{D}n^{-1}x+261\textrm{B}n^{-3/2}x^{3/2}+343\textrm{A}n^{-2}x^{2})\leq\exp(1-x).$$

The following version of Bernstein’s inequality can directly be deduced from Theorem 3.1.7 in [11].

Proposition A.2. Let \(\{Z_{j}\}_{j=1}^{n}\) be independent with \(|Z_{j}|\leq\textrm{b}\) almost surely and \(\mathbb{E}(|Z_{j}|^{2})\leq\textrm{v}\) for all \(j\in[\![n]\!]\), then for all \(x>0\) and \(n\geq 1\), we have

$$\mathbb{P}\left(\frac{1}{n}\sum_{j\in[\![n]\!]}(Z_{j}-\mathbb{E}Z_{j})\geq\sqrt{\frac{2\textrm{v}x}{n}}+\frac{\textrm{b}x}{3n}\right)\leq\exp(-x).$$

Preliminaries.We eventually calculate first A, B, and C satisfying (A.2) and exploit that \(\textrm{D}:=\textrm{C}\) automatically also fulfils (A.2), which we briefly justify next. Throughout this section we assume that \(\{Y_{j}\}_{j=1}^{n}\) are independent and identically distributed with Lebesgue-density \(g=f{\bigcirc\hskip-8.5pt\ast}\varphi\in\mathcal{L}^{2}\). We denote by \(\mathcal{L}_{\mathbb{R}}^{2}(g)\) the set of (Borel-measurable) functions \(\zeta:[0,1)\to\mathbb{R}\) with \(||\zeta||_{\mathcal{L}_{\mathbb{R}}^{2}(g)}^{2}:=\int_{[0,1)}\zeta^{2}(x)g(x)dx<\infty\). Let \(h:[0,1)^{2}\to\mathbb{R}\) be a bounded kernel, i.e., \(||h||_{\mathcal{L}^{\infty}}:=\sup_{y_{1},y_{2}\in[0,1)}|h(y_{1},y_{2})|<\infty\), and define the integral operator \(H:\mathcal{L}_{\mathbb{R}}^{2}(g)\to\mathcal{L}_{\mathbb{R}}^{2}(g)\) with \(\zeta\mapsto H\zeta\) and \(H\zeta(s):=\int_{[0,1)}h(t,s)\zeta(t)g(t)dt\) for all \(s\in[0,1)\). Then \(H\) is linear and bounded, i.e., \(||H||_{\mathcal{L}_{\mathbb{R}}^{2}(g)\to\mathcal{L}_{\mathbb{R}}^{2}(g)}:=\sup\{||H\zeta||_{\mathcal{L}_{\mathbb{R}}^{2}(g)}:||\zeta||_{\mathcal{L}_{\mathbb{R}}^{2}(g)}\leq 1\}\leq\mathbb{E}h^{2}(Y_{1},Y_{2})\leq C^{2}\) due to the Cauchy–Schwarz inequality. This shows the claim since its operator norm satisfies

$$||H||_{\mathcal{L}_{\mathbb{R}}^{2}(g)\to\mathcal{L}_{\mathbb{R}}^{2}(g)}=\sup\{\mathbb{E}(h(Y_{1},Y_{2})\zeta(Y_{1})\xi(Y_{2})),\mathbb{E}\zeta^{2}(Y_{1})\leq 1,\mathbb{E}\xi^{2}(Y_{2})\leq 1\}.$$
(A.3)

However, an additional assumption allows us to determine a slightly different quantity D. For a symmetric kernel the operator \(H\) is self-adjoint, i.e. \(\langle H\zeta,\xi\rangle_{\mathcal{L}_{\mathbb{R}}^{2}(g)}=\langle\zeta,H\xi\rangle_{\mathcal{L}_{\mathbb{R}}^{2}(g)}\) for all \(\zeta,\xi\in\mathcal{L}_{\mathbb{R}}^{2}(g)\) using the inner product \(\langle\zeta,\xi\rangle_{\mathcal{L}_{\mathbb{R}}^{2}(g)}:=\int_{[0,1)}\zeta(s)\xi(s)g(s)ds\), and we have

$$||H||_{\mathcal{L}_{\mathbb{R}}^{2}(g)\to\mathcal{L}_{\mathbb{R}}^{2}(g)}=\sup\{|\langle H\zeta,\zeta\rangle_{\mathcal{L}_{\mathbb{R}}^{2}(g)}|:||\zeta||_{\mathcal{L}_{\mathbb{R}}^{2}(g)}\leq 1\}.$$
(A.4)

The last identity can further be reformulated in terms of a discrete convolution, which we briefly recall next. For \(p\geq 1\) we denote by \(\ell^{p}:=\ell^{p}(\mathbb{Z})\) the Banach space of complex sequences over \(\mathbb{Z}\) endowed with its usual \(\ell^{p}\)-norm given by \(||a_{\bullet}||_{\ell^{p}}:=\left(\sum_{j\in\mathbb{Z}}|a_{j}|^{p}\right)^{1/p}\) for \(a_{\bullet}:=(a_{j})_{j\in\mathbb{Z}}\in\mathbb{C}^{\mathbb{Z}}\). In the case \(p=2\), \(\ell^{2}\) is a Hilbert space and the \(\ell^{2}\)-norm is induced by the inner product \(\langle a_{\bullet},b_{\bullet}\rangle_{\ell^{2}}:=\sum_{j\in\mathbb{Z}}a_{j}\overline{b}_{j}\) for all \(a_{\bullet},b_{\bullet}\in\ell^{2}\). For each sequence \(a_{\bullet}\in\ell^{1}\) the discrete convolution operator \(a_{\bullet}\star:\ell^{2}\to\ell^{2}\) with \(b_{\bullet}\mapsto a_{\bullet}\star b_{\bullet}\) and \((a_{\bullet}\star b_{\bullet})_{j}:=\sum_{l\in\mathbb{Z}}a_{j-l}b_{l}\) for all \(j\in\mathbb{Z}\), is linear and bounded by \(||{a_{\bullet}}||_{\ell^{2}}\), i.e. \(||a_{\bullet}\star||_{\ell^{2}\to\ell^{2}}:=\sup\{||a_{\bullet}\star b_{\bullet}||_{\ell^{2}}:||b_{\bullet}||_{\ell^{2}}\leq 1\}\leq||{a_{\bullet}}||_{\ell^{2}}\). Particularly, it holds

$$\left|\sum_{j\in\mathbb{Z}}\overline{b}_{j}\sum_{l\in\mathbb{Z}}a_{j-l}b_{l}\right|=|\langle a_{\bullet}\star b_{\bullet},b_{\bullet}\rangle_{\ell^{2}}|\leq||{a_{\bullet}}||_{\ell^{2}}||b_{\bullet}||_{\ell^{2}}^{2}\quad\text{for all}\quad a_{\bullet}\in\ell^{1}\quad\text{and}\quad b_{\bullet}\in\ell^{2}.$$
(A.5)

Note that the adjoint of \(a_{\bullet}\star\) is a discrete convolution operator \(a_{\bullet}^{\star}\star\) with \(a_{j}^{\star}:=\bar{a}_{-j}\) for all \(j\in\mathbb{Z}\). Hence, if in addition \(a_{j}=\bar{a}_{-j}\) for all \(j\in\mathbb{Z}\), then \(a_{\bullet}\star\) is self-adjoint. Recall that the real density \(g\in\mathcal{L}^{2}\) admits Fourier coefficients \(g_{\bullet}=(g_{j})_{j\in\mathbb{Z}}\). The coefficients belong to both \(\ell^{2}\) by Parseval’s identity, i.e. \(||{g}||_{\mathcal{L}^{2}}=||g_{\bullet}||_{\ell^{2}}\), and also to \(\ell^{1}\) due to the convolution theorem. Indeed, \((g_{j}=f_{j}\varphi_{j})_{j\in\mathbb{Z}}\) with \(g=f{\bigcirc\hskip-8.5pt\star}\,\,\varphi\) and \(f,\varphi\in\mathcal{L}^{2}\) implies \(||{g_{\bullet}}||_{\ell^{1}}\leq||f_{\bullet}||_{\ell^{2}}||\varphi_{\bullet}||_{\ell^{2}}<\infty\) due to the Cauchy–Schwarz inequality. Consequently, the discrete convolution \(g_{\bullet}\star:\ell^{2}\to\ell^{2}\) is linear, bounded and self-adjoint. Moreover, for all \(\zeta\in\mathcal{L}^{2}\) with \(|\zeta|\in\mathcal{L}_{\mathbb{R}}^{2}(g)\) and Fourier coefficients \(\zeta_{\bullet}=(\zeta_{j})_{j\in\mathbb{Z}}\in\ell^{2}\) we note that \(\langle g_{\bullet}\star\zeta_{\bullet},\zeta_{\bullet}\rangle_{\ell^{2}}=\sum_{j\in\mathbb{Z}}\bar{\zeta}_{j}\sum_{l\in\mathbb{Z}}g_{j-l}\zeta_{l}=\sum_{j\in\mathbb{Z}}\bar{\zeta}_{j}\sum_{l\in\mathbb{Z}}\mathbb{E}(\textrm{e}_{l}(Y)\textrm{e}_{j}(-Y))\zeta_{l}=\mathbb{E}|\zeta(Y)|^{2}=||\zeta||_{\mathcal{L}_{\mathbb{R}}^{2}(g)}^{2}\geq 0\). Thereby, if for all \(\zeta=\sum_{j\in\mathbb{Z}}\zeta_{j}e_{l}\in\mathcal{L}^{2}\) we also have \(|\zeta|\in\mathcal{L}_{\mathbb{R}}^{2}(g)\), then \(g_{\bullet}\star\) is non-negative. As a result, there is a non-negative operator \((g_{\bullet}\star)^{1/2}\) with \(||(g_{\bullet}\star)^{1/2}\zeta_{\bullet}||_{\ell^{2}}^{2}=\langle g_{\bullet}\star\zeta_{\bullet},\zeta_{\bullet}\rangle_{\ell^{2}}=||\zeta||_{\mathcal{L}_{\mathbb{R}}^{2}(g)}^{2}\) for all \(\zeta_{\bullet}\in\ell^{2}\), which we use frequently in the proofs below.

A.1 Auxiliary Results Used in the Proof of Proposition 2.1

Lemma A.3. Consider \(\{Y_{l}\}_{l=1}^{n}\stackrel{{\scriptstyle{iid}}}{{\sim}}g\in\mathcal{L}^{2}\) and for \(k\in\mathbb{N}\) the kernel \(h:[0,1)^{2}\to\mathbb{R}\) given by

$$h(y_{1},y_{2})=\sum_{|j|\in[\![k]\!]}\frac{(\textrm{e}_{j}(-y_{1})-g_{j})(\textrm{e}_{j}(y_{2})-\bar{g}_{j})}{|\varphi_{j}|^{2}},\quad\forall y_{1},y_{2}\in[0,1),$$

which is real-valued, bounded, symmetric and fulfils (A.1). Let \(\nu_{k}\) and \(m_{k}\) as in (2.4) then

$$\textrm{A}=4\nu_{k}^{4},\quad\textrm{B}={3||g_{\bullet}||_{\ell^{2}}\nu_{k}^{3}}\quad{and}\quad\textrm{D}=\textrm{C}=2||g_{\bullet}||_{\ell^{2}}\nu_{k}^{2}$$
(A.6)

satisfy the condition (A.2) in Proposition A.1. If, in addition, \(\mathcal{L}_{\mathbb{R}}^{2}(g)=\mathcal{L}_{\mathbb{R}}^{2}\), then

$$\textrm{D}=4||{g_{\bullet}}||_{\ell^{1}}m_{k}^{2}$$
(A.7)

also satisfies the condition (A.2) in Proposition A.1.

Proof of Lemma A.3. We first calculate quantities \(\textrm{A},\textrm{B}\), and C satisfying (A.2), then by the above discussion \(\textrm{D}=\textrm{C}\) also satisfies (A.2). First, consider A. From

$$||(\textrm{e}_{j}-g_{j})(\textrm{e}_{-j}-\bar{g}_{l})||_{\mathcal{L}^{\infty}}\leq 4\quad\text{ and}\quad|\varphi_{j}|\leq 1\quad\text{for all}\quad j,l\in\mathbb{Z}$$
(A.8)

we immediately conclude that \(||h||_{\mathcal{L}^{\infty}}\leq 4\sum_{|j|\in[\![k]\!]}|\varphi_{j}|^{-2}\leq 4\nu_{k}^{4}=\textrm{A}\). Next, consider B. Since \(\mathbb{E}(\textrm{e}_{j}(-Y_{1})\textrm{e}_{l}(Y_{1}))=g_{j-l}\) for all \(j,l\in\mathbb{Z}\), we deduce for arbitrary \(y_{2}\in[0,1)\) that

$$\mathbb{E}\left(|h(Y_{1},y_{2})|^{2}\right)=\mathbb{V}\textrm{ar}\left(\sum_{|j|\in[\![k]\!]}\textrm{e}_{j}(-Y_{1})\frac{(\textrm{e}_{j}(y_{2})-\bar{g}_{j})}{|\varphi_{j}|^{2}}\right)\leq\mathbb{E}\left|\sum_{|j|\in[\![k]\!]}\textrm{e}_{j}(-Y_{1})\frac{(\textrm{e}_{j}(y_{2})-\bar{g}_{j})}{|\varphi_{j}|^{2}}\right|^{2}$$
$${}=\sum_{|j|\in[\![k]\!]}\frac{(\textrm{e}_{j}(y_{2})-\bar{g}_{j})}{|\varphi_{j}|^{2}}\sum_{|l|\in[\![k]\!]}g_{j-l}\frac{(\textrm{e}_{l}(-y_{2})-g_{l})}{|\varphi_{l}|^{2}}=\langle a_{\bullet}\star b_{\bullet},b_{\bullet}\rangle_{\ell^{2}},$$
(A.9)

where \(a_{l}:=g_{l}{\mathchoice{\rm 1\mskip-4.0mu l}{\rm 1\mskip-4.0mu l}{\rm 1\mskip-4.5mu l}{\rm 1\mskip-5.0mu l}}{\{|l|\in[\![{{2k}}]\!]\}}\) and \(b_{l}:=(\textrm{e}_{l}(-y_{2})-g_{l})|\varphi_{l}|^{-2}{\mathchoice{\rm 1\mskip-4.0mu l}{\rm 1\mskip-4.0mu l}{\rm 1\mskip-4.5mu l}{\rm 1\mskip-5.0mu l}}{\{|l|\in[\![k]\!]\}}\) for all \(l\in\mathbb{Z}\). Making use of (A.5), (A.8), \(||g_{\bullet}||_{\ell^{2}}\geq 1\) and \((2k)^{1/2}\leq\nu_{k}^{2}\) it follows

$$\langle a_{\bullet}\star b_{\bullet},b_{\bullet}\rangle_{\ell^{2}}$$
$${}\leq\left(\sum_{|j|\in[\![{{2k}}]\!]}|g_{j}|\right)\sum_{|j|\in[\![k]\!]}\frac{|\textrm{e}_{j}(y_{2})-\bar{g}_{j}|^{2}}{|\varphi_{j}|^{4}}\leq(4k)^{1/2}\left(\sum_{|j|\in[\![{2k}]\!]}|g_{j}|^{2}\right)^{1/2}4\nu_{k}^{4}\leq 9||g_{\bullet}||_{\ell^{2}}^{2}\nu_{k}^{6}=\textrm{B}^{2}.$$

Combining the last bound and (A.9) we see that \(\sup_{y_{2}\in[0,1)}\mathbb{E}|h(Y_{1},y_{2})|^{2}\leq\textrm{B}^{2}\). Next, consider C. Since \(\mathbb{E}(\textrm{e}_{j}(-Y_{1})-g_{j})(\textrm{e}_{l}(Y_{1})-\bar{g}_{l})=g_{j-l}-g_{j}\bar{g}_{l}\) for all \(j,l\in\mathbb{Z}\), applying the Cauchy–Schwarz inequality we obtain

$$\mathbb{E}|h(Y_{1},Y_{2})|^{2}=\sum_{|j|\in[\![k]\!]}\frac{1}{|\varphi_{j}|^{2}}\sum_{|l|\in[\![k]\!]}\frac{|g_{j-l}-g_{j}\bar{g}_{l}|^{2}}{|\varphi_{l}|^{2}}\leq 2\langle a_{\bullet}\star b_{\bullet},b_{\bullet}\rangle_{\ell^{2}}+2\nu_{k}^{4}||g_{\bullet}^{2}||_{\ell^{2}}^{2},$$
(A.10)

where \(a_{l}:=|g_{l}|^{2}{\mathchoice{\rm 1\mskip-4.0mu l}{\rm 1\mskip-4.0mu l}{\rm 1\mskip-4.5mu l}{\rm 1\mskip-5.0mu l}}{\{|l|\in[\![{{2k}}]\!]\}}\) and \(b_{l}:=|\varphi_{l}|^{-2}{\mathchoice{\rm 1\mskip-4.0mu l}{\rm 1\mskip-4.0mu l}{\rm 1\mskip-4.5mu l}{\rm 1\mskip-5.0mu l}}{\{|l|\in[\![k]\!]\}}\) for all \(l\in\mathbb{Z}\). Moreover, from \(\langle a_{\bullet}\star b_{\bullet},b_{\bullet}\rangle_{\ell^{2}}\leq||{a_{\bullet}}||_{\ell^{2}}||b_{\bullet}||_{\ell^{2}}^{2}\leq||g_{\bullet}||_{\ell^{2}}^{2}\nu_{k}^{4}\) due to (A.5) we conclude that (A.10) and \(|g_{j}|\leq 1\), \(j\in\mathbb{Z}\), together imply \(\mathbb{E}|h(Y_{1},Y_{2})|^{2}\leq 4\nu_{k}^{4}||g_{\bullet}||_{\ell^{2}}^{2}=\textrm{C}^{2}\). Finally, consider D and assume in addition \(\mathcal{L}_{\mathbb{R}}^{2}(g)=\mathcal{L}_{\mathbb{R}}^{2}\) which allows us to use the identities (A.3) and (A.4) formulated in terms of an operator \(H\). Let \(\zeta\in\mathcal{L}_{\mathbb{R}}^{2}(g)\), which implies \(\zeta=\sum_{j\in\mathbb{Z}}\zeta_{j}\textrm{e}_{j}\in\mathcal{L}^{2}\). Exploiting \(\mathbb{E}(\textrm{e}_{j}(-Y_{1})-g_{j})\zeta(Y_{1})=(g_{\bullet}\star\zeta_{\bullet})_{j}-g_{j}\mathbb{E}\zeta(Y_{1})\) and \(|g_{j}|\leq 1\) for all \(j\in\mathbb{Z}\) straightforward calculations show

$$|\langle H\zeta,\zeta\rangle_{\mathcal{L}_{\mathbb{R}}^{2}(g)}|=\sum_{|j|\in[\![k]\!]}\frac{1}{|\varphi_{j}|^{2}}|\mathbb{E}((\textrm{e}_{j}(-Y_{1})-g_{j})\zeta(Y_{1}))|^{2}\leq m_{k}^{2}\sum_{|j|\in[\![k]\!]}|(g_{\bullet}\star\zeta_{\bullet})_{j}-g_{j}\mathbb{E}\zeta(Y_{1})|^{2}$$
$${}\leq 2m_{k}^{2}\left(||g_{\bullet}\star\zeta_{\bullet}||_{\ell^{2}}^{2}+||{g_{\bullet}}||_{\ell^{1}}||\zeta||_{\mathcal{L}_{\mathbb{R}}^{2}(g)}^{2}\right).$$
(A.11)

Using the properties of the discrete convolution recalled above it follows

$$||g_{\bullet}\star\zeta_{\bullet}||_{\ell^{2}}^{2}\leq||(g_{\bullet}\star)^{1/2}||_{\ell^{2}\to\ell^{2}}^{2}||(g_{\bullet}\star)^{1/2}\zeta_{\bullet}||_{\ell^{2}}^{2}=||g_{\bullet}\star||_{\ell^{2}\to\ell^{2}}||\zeta||_{\mathcal{L}_{\mathbb{R}}^{2}(g)}^{2}\leq||{g_{\bullet}}||_{\ell^{1}}||\zeta||_{\mathcal{L}_{\mathbb{R}}^{2}(g)}^{2},$$

which together with (A.11) implies \(|\langle H\zeta,\zeta\rangle_{\mathcal{L}_{\mathbb{R}}^{2}(g)}|\leq 4m_{k}^{2}||{g_{\bullet}}||_{\ell^{1}}||\zeta||_{\mathcal{L}_{\mathbb{R}}^{2}(g)}^{2}\) for all \(\zeta\in\mathcal{L}_{\mathbb{R}}^{2}(g)\). We conclude from (A.4) that \(||H||_{\mathcal{L}_{\mathbb{R}}^{2}(g)\to\mathcal{L}_{\mathbb{R}}^{2}(g)}\leq 4m_{k}^{2}||{g_{\bullet}}||_{\ell^{1}}=\textrm{D}\), and finally that D satisfies (A.2), by (A.4), which completes the proof. \(\Box\)

Lemma A.4. Let \(\{Y_{l}\}_{l=1}^{n}\stackrel{{\scriptstyle\mathrm{iid}}}{{\sim}}g=f{\bigcirc\hskip-8.5pt\star}\,\,\varphi\in\mathcal{L}^{2}\) with joint distribution \(\mathbb{P}_{f}\) and \(g^{\circ}=f^{\circ}{\bigcirc\hskip-8.5pt\ast}\varphi\in\mathcal{L}^{2}\). For each \(k\in\mathbb{N}\) consider \(\mathrm{q}^{2}_{k}(f-f^{\circ})\) and \(m_{k}\) as in (2.1) and (2.4), respectively. Then the linear centred statistic \(\textrm{V}_{n}\) defined in (2.3) satisfies for all \(x\geq 1\) and \(n\geq 1\)

$$\mathbb{P}_{f}(2\textrm{V}_{n}\leq-cx^{2}(1\vee m_{k}^{2}n^{-1})m_{k}^{2}n^{-1}-\frac{1}{2}\textrm{q}^{2}_{k}(f-f^{\circ}))\leq\exp(-x),$$

where \(c=8||{g_{\bullet}}||_{\ell^{1}}+||\varphi_{\bullet}||_{\ell^{2}}^{2}\).

Proof of Lemma A.4. Introducing the real function \(\psi:=\sum_{|j|\in[\![k]\!]}(g_{j}-g_{j}^{\circ})|\varphi_{j}|^{-2}\textrm{e}_{j}\) and independent and identically distributed random variables \(Z_{j}:=2\psi(Y_{j})\), \(j\in[\![n]\!]\), we intend to apply Proposition A.2 to \(\textrm{V}_{n}=\frac{1}{n}\sum_{j\in[\![n]\!]}(Z_{j}-\mathbb{E}_{f}(Z_{j}))\). Therefore, we compute the required quantities v and b. First consider b. Using subsequently the identity \(g_{l}-g_{l}^{\circ}=(f_{l}-f_{l}^{\circ})\varphi_{l}\), \(l\in\mathbb{Z}\), and the Cauchy–Schwarz inequality we deduce that

$$|Z_{1}|\leq 2||{\psi}||_{\mathcal{L}^{\infty}}\leq 2m_{k}^{2}\sum_{|l|\in[\![k]\!]}|g_{l}-g_{l}^{\circ}|\leq 2m_{k}^{2}\textrm{q}_{k}(f-f^{\circ})||\varphi_{\bullet}||_{\ell^{2}}=:\textrm{b}.$$
(A.12)

Secondly, consider v. Since \(\mathbb{E}_{f}(\mathrm{e}_{j}(-Y_{1})e_{l}(Y_{1}))=g_{j-l}\) for all \(j,l\in\mathbb{Z}\), we see that

$$\mathbb{E}_{f}|Z_{1}|^{2}=4\mathbb{E}_{f}|\psi(Y_{1})|^{2}=4\sum_{|j|\in[\![k]\!]}\frac{\bar{g}_{j}-\bar{g}_{j}^{\circ}}{|\varphi_{j}|^{2}}\sum_{|l|\in[\![k]\!]}g_{j-l}\frac{g_{l}-g_{l}^{\circ}}{|\varphi_{l}|^{2}}=4\langle a_{\bullet}\star b_{\bullet},b_{\bullet}\rangle_{\ell^{2}},$$

where \(a_{l}:=g_{l}{\mathchoice{\rm 1\mskip-4.0mu l}{\rm 1\mskip-4.0mu l}{\rm 1\mskip-4.5mu l}{\rm 1\mskip-5.0mu l}}{\{|l|\in[\![{2k}]\!]\}}\) and \(b_{l}:=(g_{l}-g_{l}^{\circ})|\varphi_{l}|^{-2}{\mathchoice{\rm 1\mskip-4.0mu l}{\rm 1\mskip-4.0mu l}{\rm 1\mskip-4.5mu l}{\rm 1\mskip-5.0mu l}}{\{|l|\in[\![k]\!]\}}\) for all \(l\in\mathbb{Z}\). Successively exploiting further (A.5) and the identity \(g_{l}-g_{l}^{\circ}=(f_{l}-f_{l}^{\circ})\varphi_{l}\), \(l\in\mathbb{Z}\), we conclude that

$$\mathbb{E}_{f}|Z_{1}|^{2}\leq 4||b_{\bullet}||_{\ell^{2}}^{2}||{a_{\bullet}}||_{\ell^{1}}=4\sum_{|j|\in[\![k]\!]}|\varphi_{j}|^{-4}|g_{j}-g_{j}^{\circ}|^{2}\sum_{|j|\in[\![k]\!]}|g_{j}|$$
$${}\leq 4m_{k}^{2}\textrm{q}^{2}_{k}(f-f^{\circ})||{g_{\bullet}}||_{\ell^{1}}=:\textrm{v}.$$
(A.13)

The claim of Lemma A.4 now follows from Proposition A.2 with b and v as in (A.12) and (A.13), respectively. Indeed, making use of \(2ac\leq\frac{a^{2}}{\varepsilon}+c^{2}\varepsilon\) for any \(\varepsilon,a,c>0\), we have

$$\frac{\textrm{b}x}{3n}\leq\varepsilon_{1}\textrm{q}^{2}_{k}(f-f^{\circ})+\frac{x^{2}}{9\varepsilon_{1}}||\varphi_{\bullet}||_{\ell^{2}}^{2}\frac{m_{k}^{4}}{n^{2}}\quad\text{and}\quad\sqrt{\frac{2\textrm{v}x}{n}}\leq\varepsilon_{2}\textrm{q}^{2}_{k}(f-f^{\circ})+\frac{2x}{\varepsilon_{2}}\frac{m_{k}^{2}}{n}||{g_{\bullet}}||_{\ell^{1}}.$$

Combining both bounds (with \(\varepsilon_{1}=\varepsilon_{2}=\frac{1}{4}\)) yields for all \(x\geq 1\)

$$\sqrt{\frac{2\textrm{v}x}{n}}+\frac{\textrm{b}x}{3n}\leq\frac{1}{2}\textrm{q}^{2}_{k}(f-f^{\circ})+cx^{2}\left(1\vee\frac{m_{k}^{2}}{n}\right)\frac{m_{k}^{2}}{n}\quad\text{with}\quad c=8||{g_{\bullet}}||_{\ell^{1}}+||\varphi_{\bullet}||_{\ell^{2}}^{2}.$$

Hence, the assertion follows from Proposition A.2 by the usual symmetry argument. \(\Box\)

A.2 Auxiliary Results Used in the Proof of Proposition 4.1

Corollary A.5. Consider \(\{Y_{l}\}_{l=1}^{n}\stackrel{{\scriptstyle\mathrm{iid}}}{{\sim}}g\in\mathcal{L}^{2}\) and for \(k\in\mathbb{N}\) the kernel \(h:[0,1)^{2}\to\mathbb{R}\) given by

$$h(y_{1},y_{2})=\sum_{|j|\in[\![k]\!]}(\mathrm{e}_{j}(-y_{1})-g_{j})(\mathrm{e}_{j}(y_{2})-\bar{g}_{j}),\quad\forall y_{1},y_{2}\in[0,1),$$

which is real-valued, bounded, symmetric and fulfils (A.1). Then the quantities

$$\textrm{A}=8k,\quad\textrm{B}=3||g_{\bullet}||_{\ell^{2}}(2k)^{3/4},\quad\text{and}\quad\textrm{D}=\textrm{C}=2||g_{\bullet}||_{\ell^{2}}(2k)^{1/2}$$

satisfy the condition (A.2) in Proposition A.1. If, in addition, \(\mathcal{L}_{\mathbb{R}}^{2}(g)=\mathcal{L}_{\mathbb{R}}^{2}\), then

$$\textrm{D}=4||{g_{\bullet}}||_{\ell^{1}}$$

also satisfies the condition (A.2) in Proposition A.1.

Proof of Corollary A.5. Setting \(|\varphi_{j}|^{2}=1\) for all \(|j|\in[\![k]\!]\) the assertion immediately follows from Lemma A.3. \(\Box\)

Lemma A.6. Let \(\{Y_{l}\}_{l=1}^{n}\stackrel{{\scriptstyle mathrm{iid}}}{{\sim}}g=f{\bigcirc\hskip-8.5pt\star}\,\,\varphi\in\mathcal{L}^{2}\) with joint distribution \(\mathbb{P}_{f}\) and \(g^{\circ}=f^{\circ}{\bigcirc\hskip-8.5pt\ast}\varphi\in\mathcal{L}^{2}\). For each \(k\in\mathbb{N}\) consider \(\mathrm{q}^{2}_{k}(g-g^{\circ})\) as in (4.1). Then the linear centred statistic \(\textrm{V}_{n}^{\mathrm{d}}\) defined in (4.3) satisfies for all \(x\geq 1\) and \(n\geq 1\)

$$\mathbb{P}_{f}(2\textrm{V}_{n}^{\textrm{d}}\leq-cx^{2}(1\vee(2k)^{1/2}n^{-1})(2k)^{1/2}n^{-1}-\frac{1}{2}\textrm{q}^{2}_{k}(g-g^{\circ}))\leq\exp(-x),$$

where \(c=12||g_{\bullet}||_{\ell^{2}}+1\).

Proof of Lemma A.6. Introducing the real function \(\psi:=\sum_{|j|\in[\![k]\!]}(g_{j}-g_{j}^{\circ})\textrm{e}_{j}\) and independent and identically distributed random variables \(Z_{j}:=2\psi(Y_{j})\), \(j\in[\![n]\!]\), we intend to apply Proposition A.2 to \(\textrm{V}_{n}^{\textrm{d}}=\frac{1}{n}\sum_{j\in[\![n]\!]}(Z_{j}-\mathbb{E}_{f}(Z_{j}))\). Therefore, we compute the required quantities v and b. First consider b. Using the Cauchy–Schwarz inequality we see that

$$|Z_{1}|\leq 2||{\psi}||_{\mathcal{L}^{\infty}}\leq 2\sum_{|l|\in[\![k]\!]}|g_{l}-g_{l}^{\circ}|\leq 2(2k)^{1/2}\textrm{q}_{k}(g-g^{\circ})=:\textrm{b}.$$
(A.14)

Secondly, consider v. Since \(\mathbb{E}_{f}(\mathrm{e}_{j}(-Y_{1})e_{l}(Y_{1}))=g_{j-l}\) for all \(j,l\in\mathbb{Z}\), we deduce that

$$\mathbb{E}_{f}|Z_{1}|^{2}=4\mathbb{E}_{f}|\psi(Y_{1})|^{2}=4\sum_{|j|\in[\![k]\!]}(\bar{g}_{j}-\bar{g}_{j}^{\circ})\sum_{|l|\in[\![k]\!]}g_{j-l}(g_{l}-g_{l}^{\circ})=4\langle a_{\bullet}\star b_{\bullet},b_{\bullet}\rangle_{\ell^{2}},$$

where \(a_{l}:=g_{l}{\mathchoice{\rm 1\mskip-4.0mu l}{\rm 1\mskip-4.0mu l}{\rm 1\mskip-4.5mu l}{\rm 1\mskip-5.0mu l}}{\{|l|\in[\![{{2k}}]\!]\}}\) and \(b_{l}:=(g_{l}-g_{l}^{\circ}){\mathchoice{\rm 1\mskip-4.0mu l}{\rm 1\mskip-4.0mu l}{\rm 1\mskip-4.5mu l}{\rm 1\mskip-5.0mu l}}{\{|l|\in[\![k]\!]\}}\) for all \(l\in\mathbb{Z}\). Hence, (A.5) shows that

$$\mathbb{E}_{f}|Z_{1}|^{2}\leq 4||b_{\bullet}||_{\ell^{2}}^{2}||{a_{\bullet}}||_{\ell^{1}}=4\textrm{q}^{2}_{k}(g-g^{\circ})\sum_{|j|\in[\![{{2}k}]\!]}|g_{j}|=:\textrm{v}.$$
(A.15)

The claim of Lemma A.6 follows now from Proposition A.2 with b and v as in (A.14) and (A.15), respectively. Indeed, exploiting \(2ac\leq\frac{a^{2}}{\varepsilon}+c^{2}\varepsilon\) for any \(\varepsilon,a,c>0\) we see that

$$\frac{\textrm{b}x}{3n}\leq\varepsilon_{1}\textrm{q}^{2}_{k}(g-g^{\circ})+\frac{x^{2}}{9\varepsilon_{1}}\frac{2k}{n^{2}}\quad\text{and }$$
$$\sqrt{\frac{2\textrm{v}x}{n}}\leq\varepsilon_{2}\textrm{q}^{2}_{k}(g_{\bullet}-g^{\circ})+\frac{2x}{\varepsilon_{2}n}\sum_{|j|\in[\![{{2k}}]\!]}|g_{j}|\leq\varepsilon_{2}\textrm{q}^{2}_{k}(g-g^{\circ})+\frac{2x}{\varepsilon_{2}}\frac{(4k)^{1/2}}{n}||g_{\bullet}||_{\ell^{2}}.$$

Combining both bounds (with \(\varepsilon_{1}=\varepsilon_{2}=\frac{1}{4}\)) we get for all \(x\geq 1\)

$$\sqrt{\frac{2\textrm{v}x}{n}}+\frac{\textrm{b}x}{3n}\leq\frac{1}{2}\textrm{q}^{2}_{k}(g-g^{\circ})+cx^{2}(1\vee(2k)^{1/2}n^{-1})(2k)^{1/2}n^{-1}\quad\text{with}\quad c=12||g_{\bullet}||_{\ell^{2}}+1.$$

Hence, the assertion follows from Proposition A.2 by the usual symmetry argument. \(\Box\)

CALCULATIONS FOR THE ILLUSTRATIONS

B.1 Calculations for the Radius Bounds in Illustration 3.5

Firstly, we determine the order of the term \(\rho_{\mathcal{K},a_{\bullet}}^{2}(\delta n)\) by showing that \(\rho_{\mathcal{K},a_{\bullet}}^{2}(n)\sim\rho_{a_{\bullet}}^{2}(n)\) and replacing \(n\) with \(\delta n\). Indeed, we trivially have \(\rho_{a_{\bullet}}^{2}(n)\leq\rho_{\mathcal{K},a_{\bullet}}^{2}(n)\). By defining \(j_{\star}:=\lceil\frac{2}{4p+4s+1}\log_{2}n\rceil\lesssim\log(n^{2}/2)\) (in the ordinary smooth case) respectively \(j_{\star}:=\lceil\frac{1}{s}\log_{2}\log n\rceil\lesssim\frac{1}{s_{\star}}\log\log n\) (in the super smooth case), straightforward calculations then show that \(\rho_{\mathcal{K},a_{\bullet}}^{2}(n)\lesssim\rho_{2^{j\star},a_{\bullet}}^{2}(n)\lesssim\rho_{a_{\bullet}}^{2}(n)\). Next, we determine the order of the remainder term \(r_{\mathcal{K},a_{\bullet}}^{2}(\delta^{2}n)\) by first calculating \(r_{\mathbb{N},a_{\bullet}}^{2}(n):=\min_{k\in\mathbb{N}}a_{k}^{2}\vee\frac{m_{k}^{2}}{n}\), showing \(r_{\mathbb{N},a_{\bullet}}^{2}(n)\sim r_{\mathcal{K}_{g},a_{\bullet}}^{2}(n)\) and then replacing \(n\) with \(\delta^{2}n\). The variance term \(\frac{m_{k}^{2}}{n}\) is of order \(\frac{k^{2p}}{n}\). In the ordinary smooth case the bias term \(a_{k}^{2}\) is of order \(k^{-2s}\). Hence, the minimising \(k_{\star}\) satisfies \(k_{\star}\sim n^{\frac{1}{2s+2p}}\), which yields \(r_{\mathbb{N},a_{\bullet}}^{2}(n)\sim n^{-\frac{s}{s+p}}\). We define \(j_{\star}:=\lceil\frac{1}{2p+2s}\log_{2}n\rceil\lesssim\log(n^{2}/2)\). Straightforward calculations show that \(r_{\mathcal{K}_{2},a_{\bullet}}^{2}(n)\lesssim r_{2^{j\star},a_{\bullet}}^{2}(n)\lesssim r_{\mathbb{N},a_{\bullet}}^{2}(n)\). Since, trivially \(r_{\mathbb{N},a_{\bullet}}^{2}(n)\leq r_{\mathcal{K}_{g},a_{\bullet}}^{2}(n)\), we obtain the assertion. In the super smooth case the bias term \(a_{k}^{2}\) is of order \(e^{-2k^{s}}\). Hence, the minimising \(k_{\star}\) satisfies \(k_{\star}\sim\log(n)^{\frac{1}{s}}\), which yields \(r_{\mathbb{N},a_{\bullet}}^{2}(n)\sim\frac{1}{n}(\log n)^{\frac{2p}{s}}\). We define \(j_{\star}:=\lceil\frac{1}{s}\log_{2}\log n\rceil\lesssim\frac{1}{s_{\star}}\log\log n\). Straightforward calculations show that \(r_{\mathcal{K}_{s_{\star},a_{\bullet}}}^{2}(n)\lesssim r_{2^{j\star},a_{\bullet}}^{2}(n)\lesssim r_{\mathbb{N},a_{\bullet}}^{2}(n)\). Since, trivially \(r_{\mathbb{N},a_{\bullet}}^{2}(n)\leq r_{\mathcal{K}_{s_{\star},a_{\bullet}}}^{2}(n)\), we obtain the assertion.

B.2 Calculations for the Radius Bounds in Illustration 4.5

Ordinary smooth—mildly ill-posed. Since \(\frac{(2k)^{1/2}}{n}m_{k}^{2}\sim\frac{1}{n}k^{2p+1/2}\) and \(a_{k}^{2}\sim k^{-2s}\), the optimal \(k_{a_{\bullet}}^{\textrm{d}}\) satisfies \(k_{a_{\bullet}}^{\textrm{d}}\sim n^{\frac{2}{4p+4s+1}}\), which yields an upper bound of order \((\rho_{a_{\bullet}}^{\textrm{d}})^{2}\sim(k_{a_{\bullet}}^{\textrm{d}})^{-2s}\sim n^{-\frac{4s}{4p+4s+1}}\). Ordinary smooth—severely ill-posed. Since \(\frac{(2k)^{1/2}}{n}m_{k}^{2}\sim\frac{1}{n}k^{1/2}e^{2k^{p}}\) and \(a_{k}^{2}\sim k^{-2s}\), we obtain \(k_{a_{\bullet}}^{\textrm{d}}\sim(\log n)^{\frac{1}{p}}\), which yields an upper bound of order \((\rho_{a_{\bullet}}^{\textrm{d}})^{2}\sim(k_{a_{\bullet}}^{\textrm{d}})^{-2s}\sim(\log n)^{-\frac{2s}{p}}\). Super smooth—mildly ill-posed. Since \(\frac{(2k)^{1/2}}{n}m_{k}^{2}\sim\frac{1}{n}k^{2p+1/2}\) and \(a_{k}^{2}\sim e^{-2k^{s}}\), we obtain \(k_{a_{\bullet}}^{\textrm{d}}\sim(\log n)^{\frac{1}{s}}\), which yields an upper bound of order \((\rho_{a_{\bullet}}^{\textrm{d}})^{2}\sim\frac{1}{n}(k_{a_{\bullet}}^{\textrm{d}})^{2p+1/2}\sim\frac{1}{n}(\log n)^{\frac{2p+1/2}{s}}\).

B.3 Calculations for the Radius Bounds in Illustration 5.4

Firstly, we determine the order of the terms \((\rho_{\mathcal{K},a_{\bullet}}^{\textrm{d}}(\delta n))^{2}\) by showing that \((\rho_{\mathcal{K}_{2},a_{\bullet}}^{\textrm{d}}(n))^{2}\sim(\rho_{a_{\bullet}}^{\textrm{d}}(n))^{2}\) and replacing \(n\) with \(\delta n\). Indeed, we trivially have \((\rho_{\mathcal{K},a_{\bullet}}^{\textrm{d}}(\delta n))^{2}\leq(\rho_{a_{\bullet}}^{\textrm{d}}(n))^{2}\). Define \(j_{\star}:=\lceil\frac{2}{4p+4s+1}\log_{2}n\rceil\lesssim\log(n^{2}/2)\) (ordinary smooth—mildly ill-posed case), \(j_{\star}:=\lceil\frac{1}{s}\log_{2}\log n\rceil\lesssim\frac{1}{s_{\star}}\log\log n\) (super smooth—mildly ill-posed case) respectively \(j_{\star}:=\lceil\frac{1}{s}\log_{2}\log n\rceil\lesssim\frac{1}{s_{\star}}\log\log n\) (ordinary smooth—severely ill-posed case). Straightforward calculations then show that \(\rho_{\mathcal{K},a_{b}ullet}^{2}(n)\lesssim\rho_{2^{j\star},a_{b}ullet}^{2}(n)\lesssim\rho_{a_{b}ullet}^{2}(n)\). Next, we determine the order of the remainder term \(r_{\mathcal{K},a_{\bullet}}^{2}(\delta^{2}n)\) by first calculating \(r_{\mathbb{N},a_{\bullet}}^{2}(\delta^{2}n):=\min_{k\in\mathbb{N}}a_{k}^{2}\vee\frac{m_{k}^{2}}{n}\) and then showing that minimisation over \(\mathcal{K}_{g}\) approximates the minimisation over \(\mathbb{N}\) well enough. The calculations in the (ordinary smooth—mildly ill-posed) and (super smooth—mildly ill-posed) cases have already been done in Illustration 3.5. It remains to consider the third case (ordinary smooth—severely ill-posed). Since \(\frac{m_{k}^{2}}{n}\sim\frac{e^{2k^{p}}}{n}\) and \(a_{k}^{2}\sim k^{-2s}\), the minimising \(k_{\star}\) satisfies \(k_{\star}^{-2s}\sim\frac{e^{2k_{\star}^{p}}}{n}\) and thus \(k_{\star}\sim(\log n)^{\frac{1}{p}}\), which yields \(r_{\mathbb{N},a_{\bullet}}^{2}(n)\sim(\log n)^{\frac{2s}{p}}\). Next, we show \(r_{\mathcal{K}_{g},a_{\bullet}}^{2}(n)\sim r_{\mathbb{N},a_{\bullet}}^{2}(n)\). We define \(j_{\star}:=\lceil\frac{1}{p}\log_{2}\log n\rceil\lesssim\log(n^{2}/2)\). Straightforward calculations show that \(r_{\mathcal{K}_{g},a_{\bullet}}^{2}(n)\lesssim r_{2^{j\star},a_{\bullet}}^{2}(n)\lesssim r_{\mathbb{N},a_{\bullet}}^{2}(n)\). Since, trivially \(r_{\mathbb{N},a_{\bullet}}^{2}(n)\leq r_{\mathcal{K}_{g},a_{\bullet}}^{2}(n)\), we obtain the assertion by replacing \(n\) with \(\delta^{2}n\).

CALCULATIONS FOR THE \(\chi^{2}\)-DIVERGENCE

In the proof of Lemma C.2 below we apply the following assertion due to [31] (Lemma A.1 in the Appendix A).

Lemma C.1. For \(k\in\mathbb{N}\) and for each sign vector \(\tau\in\{\pm\}^{k}\) let \(J^{\tau}=(J_{j}^{\tau_{j}})_{j\in[\![k]\!]}\in\mathbb{R}^{k}\). Then,

$$\frac{1}{2^{k}}\sum_{\tau\in\{\pm\}^{k}}\prod_{j\in[\![k]\!]}J_{j}^{\tau_{j}}=\prod_{j\in[\![k]\!]}\frac{J_{j}^{-}+J_{j}^{+}}{2}.$$

Lemma C.2 ( \(\boldsymbol{\chi}^{\mathbf{2}}\) -divergence for mixtures over hypercubes over several classes). Let \(\mathcal{S}\) be an arbitrary index set of finite cardinality \(|\mathcal{S}|\subset\mathbb{N}\) . For each \(s\in\mathcal{S}\) assume \(k^{s}\in\mathbb{N}\) and \(\theta_{\bullet}^{s}\in\ell^{2}(\mathbb{N})\subseteq\mathbb{R}^{\mathbb{N}}\) . For \(\tau\in\{\pm\}^{k^{s}}\) define coefficients \(\theta_{\bullet}^{s,\tau}\in\ell^{2}(\mathbb{N})\) and functions \(g^{s,\tau}\in\mathcal{L}^{2}\) by setting

$$\theta_{j}^{s,\tau_{j}}=\begin{cases}\tau_{j}\theta_{j}^{s}\quad j\in[\![{k^{s}}]\!]\\ 0\quad{otherwise}\end{cases}\quad{and}\quad g^{s,\tau}=\textrm{e}_{0}+\sum_{|j|\in[\![{k^{s}}]\!]}\theta_{|j|}^{s,\tau_{|j|}}\textrm{e}_{j}.$$

Assuming \(g^{s,\tau}\in\mathcal{D}\) for each \(s\in\mathcal{S}\) and \(\tau\in\{\pm\}^{k^{s}}\) , we consider the mixture \(\mathbb{P}_{1}\) with probability density \(\frac{1}{|\mathcal{S}|}\sum_{s\in\mathcal{S}}(\frac{1}{2^{k^{s}}}\sum_{\tau\in\{\pm\}^{k^{s}}}\prod_{j\in[\![n]\!]}g^{s,\tau}(z_{j}))\) , \(z_{j}\in[0,1),j\in[\![n]\!]\) and denote \(\mathbb{P}_{0}:=\mathbb{P}_{{\mathchoice{\rm 1\mskip-4.0mu l}{\rm 1\mskip-4.0mu l}{\rm 1\mskip-4.5mu l}{\rm 1\mskip-5.0mu l}}_{[0,1)}}\) . Then, the \(\chi^{2}\) -divergence satisfies

$$\chi^{2}(\mathbb{P}_{1},\mathbb{P}_{0})\leq\frac{1}{|\mathcal{S}|^{2}}\sum_{s,t\in\mathcal{S}}\exp\left(2n^{2}\sum_{j\in[\![{k^{s}\wedge k^{t}}]\!]}(\theta_{j}^{s}\theta_{j}^{t})^{2}\right)-1.$$

Proof of Lemma C.2. Recall that \(\chi^{2}(\mathbb{P}_{1},\mathbb{P}_{0})=\mathbb{E}_{0}\left(\frac{\textrm{d}\mathbb{P}_{1}}{\textrm{d}\mathbb{P}_{0}}(Z_{j},_{j\in[\![n]\!]})\right)^{2}-1\) where \((Z_{j})_{j\in[\![n]\!]}\) are independent with identical marginal density \(\textrm{e}_{0}={\mathchoice{\rm 1\mskip-4.0mu l}{\rm 1\mskip-4.0mu l}{\rm 1\mskip-4.5mu l}{\rm 1\mskip-5.0mu l}}_{[0,1)}\) under \(\mathbb{P}_{0}\). Let \(z_{j}\in[0,1)\), \(j\in[\![n]\!]\), then

$$\frac{\textrm{d}\mathbb{P}_{1}}{\textrm{d}\mathbb{P}_{0}}(z_{j},_{j\in[\![n]\!]})=\frac{1}{|\mathcal{S}|}\sum_{s\in\mathcal{S}}\left(\frac{1}{2^{k^{s}}}\sum_{\tau\in\{\pm\}^{k^{s}}}\prod_{j\in[\![n]\!]}g^{s,\tau}(z_{j})\right).$$

Squaring, taking the expectation under \(\mathbb{P}_{0}\) and exploiting the independence yields

$$\mathbb{E}_{0}\left(\frac{\textrm{d}\mathbb{P}_{1}}{\textrm{d}\mathbb{P}_{0}}(Z_{j},_{j\in[\![n]\!]})\right)^{2}=\frac{1}{|\mathcal{S}|^{2}}\sum_{s,t\in\mathcal{S}}\frac{1}{2^{k^{s}}}\frac{1}{2^{k^{t}}}\sum_{\tau\in\{\pm\}^{k^{s}}}\sum_{\eta\in\{\pm\}^{k^{t}}}\prod_{j\in[\![n]\!]}\mathbb{E}_{0}(g^{s,\tau}(Z_{j})g^{t,\eta}(Z_{j}))$$
$${}=\frac{1}{|\mathcal{S}|^{2}}\sum_{s,t\in\mathcal{S}}\frac{1}{2^{k^{s}}}\frac{1}{2^{k^{t}}}\sum_{\tau\in\{\pm\}^{k^{s}}}\sum_{\eta\in\{\pm\}^{k^{t}}}(\mathbb{E}_{0}(g^{s,\tau}(Z_{1})g^{t,\eta}(Z_{1})))^{n}.$$

Exploiting the orthonormality of \((\textrm{e}_{j})_{j\in\mathbb{Z}}\) we calculate

$$\mathbb{E}_{0}(g^{s,\tau}(Z_{1})g^{t,\eta}(Z_{1}))=\int g^{s,\tau}(z)g^{t,\eta}(z)\textrm{d}z=1+2\sum_{j\in[\![{k^{s}\wedge k^{t}}]\!]}\theta_{j}^{s,\tau_{j}}\theta_{j}^{t,\eta_{j}}.$$

Applying the inequality \(1+x\leq\exp(x)\) for all \(x\in\mathbb{R}\) we obtain

$$\mathbb{E}_{0}(g^{s,\tau}(Z_{1})g^{t,\eta}(Z_{1}))$$
$${}=1+2\sum_{j\in[\![{k^{s}\wedge k^{t}}]\!]}\theta_{j}^{s,\tau_{j}}\theta_{j}^{t,\eta_{j}}\leq\exp\left(2\sum_{j\in[\![{k^{s}\wedge k^{t}}]\!]}\theta_{j}^{s,\tau_{j}}\theta_{j}^{t,\eta_{j}}\right)$$
$$=\prod_{j\in[\![{k^{s}\wedge k^{t}}]\!]}\exp\left(2\theta_{j}^{s,\tau_{j}}\theta_{j}^{t,\eta_{j}}\right).$$

Hence,

$$\mathbb{E}_{0}\left(\frac{\textrm{d}\mathbb{P}_{1}}{\textrm{d}\mathbb{P}_{0}}(Z_{j},_{j\in[\![n]\!]})\right)^{2}\leq\frac{1}{|\mathcal{S}|^{2}}\sum_{s,t\in\mathcal{S}}\frac{1}{2^{k^{s}}}\frac{1}{2^{k^{t}}}\sum_{\tau\in\{\pm\}^{k^{s}}}\sum_{\eta\in\{\pm\}^{k^{t}}}\prod_{j\in[\![{k^{s}\wedge k^{t}}]\!]}\exp\left(2n\theta_{j}^{s,\tau_{j}}\theta_{j}^{t,\eta_{j}}\right),$$

where we apply Lemma C.1 to the \(\eta\)-summation with \(J_{j}^{\eta_{j}}=\exp(2n\theta_{j}^{s,\tau_{j}}\theta_{j}^{t,\eta_{j}})\) and obtain

$$\mathbb{E}_{0}\left(\frac{\textrm{d}\mathbb{P}_{1}}{\textrm{d}\mathbb{P}_{0}}(Z_{j},_{j\in[\![n]\!]})\right)^{2}\leq\frac{1}{|\mathcal{S}|^{2}}\sum_{s,t\in\mathcal{S}}\frac{1}{2^{k^{s}}}\sum_{\tau\in\{\pm\}^{k^{s}}}\prod_{j\in[\![{k^{s}\wedge k^{t}}]\!]}\frac{\exp(-2n\theta_{j}^{s,\tau_{j}}\theta_{j}^{t})+\exp(2n\theta_{j}^{s,\tau_{j}}\theta_{j}^{t})}{2}.$$

Again applying Lemma C.1 to the \(\tau\)-summation with \(J_{j}^{\tau_{j}}=\frac{\exp(-2n\theta_{j}^{s,\tau_{j}}\theta_{j}^{t})+\exp(2n\theta_{j}^{s,\tau_{j}}\theta_{j}^{t})}{2}\) yields

$$\mathbb{E}_{0}\left(\frac{\textrm{d}\mathbb{P}_{1}}{\textrm{d}\mathbb{P}_{0}}(Z_{j},_{j\in[\![n]\!]})\right)^{2}\leq\frac{1}{|\mathcal{S}|^{2}}\sum_{s,t\in\mathcal{S}}\prod_{j\in[\![{k^{s}\wedge k^{t}}]\!]}\frac{\exp(-2n\theta_{j}^{s}\theta_{j}^{t})+\exp(2n\theta_{j}^{s}\theta_{j}^{t})}{2}$$
$${}=\frac{1}{|\mathcal{S}|^{2}}\sum_{s,t\in\mathcal{S}}\prod_{j\in[\![{k^{s}\wedge k^{t}}]\!]}\cosh(2n\theta_{j}^{s}\theta_{j}^{t}).$$

Since \(\cosh(x)\leq\exp(x^{2}/2)\), \(x\in\mathbb{R}\), we obtain

$$\mathbb{E}_{0}\left(\frac{\textrm{d}\mathbb{P}_{1}}{\textrm{d}\mathbb{P}_{0}}(Z_{j},_{j\in[\![n]\!]})\right)^{2}\leq\frac{1}{|\mathcal{S}|^{2}}\sum_{s,t\in\mathcal{S}}\prod_{j\in[\![{k^{s}\wedge k^{t}}]\!]}\exp\left(2n^{2}(\theta_{j}^{s}\theta_{j}^{t})^{2}\right)=\frac{1}{|\mathcal{S}|^{2}}\sum_{s,t\in\mathcal{S}}\exp\left(2n^{2}\sum_{j\in[\![{k^{s}\wedge k^{t}}]\!]}\hskip-4.3pt(\theta_{j}^{s}\theta_{j}^{t})^{2}\right),$$

which completes the proof. \(\Box\)

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Schluttenhofer, S., Johannes, J. Adaptive Minimax Testing for Circular Convolution. Math. Meth. Stat. 29, 106–133 (2020). https://doi.org/10.3103/S1066530720020039

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.3103/S1066530720020039

Keywords:

Navigation