Skip to main content
Log in

An expectile computation cookbook

  • Original Paper
  • Published:
Statistics and Computing Aims and scope Submit manuscript

Abstract

A substantial body of work in the last 15 years has shown that expectiles constitute an excellent candidate for becoming a standard tool in probabilistic and statistical modeling. Surprisingly, the question of how expectiles may be efficiently calculated has been left largely untouched. We fill this gap by, first, providing a general outlook on the computation of expectiles that relies on the knowledge of analytic expressions of the underlying distribution function and mean residual life function. We distinguish between discrete distributions, for which an exact calculation is always feasible, and continuous distributions, where a Newton–Raphson approximation algorithm can be implemented and a list of exceptional distributions whose expectiles are in analytic form can be given. When the distribution function and/or the mean residual life is difficult to compute, Monte-Carlo algorithms are introduced, based on an exact calculation of sample expectiles and on the use of control variates to improve computational efficiency. We discuss the relevance of our findings to statistical practice and provide numerical evidence of the performance of the considered methods.

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

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9

Similar content being viewed by others

Notes

  1. Available on GitHub at https://github.com/AntoineUC/Expectrem.

  2. Available at https://www.kaggle.com/ranja7/vehicle-insurance-customer-data and on file with the authors.

  3. Available in the R package Expectrem as the dataset commerzbank.

References

  • Artzner, P., Delbaen, F., Eber, J.-M., Heath, D.: Coherent measures of risk. Math. Financ. 9(3), 203–228 (1999)

    Article  MathSciNet  Google Scholar 

  • Asmussen, S., Albrecher, H.: Ruin Probabilities. Advanced Series on Statistical Science and Applied Probability, vol. 14. World Scientific, Singapore (2010)

    Google Scholar 

  • Bellini, F.: Isotonicity properties of generalized quantiles. Stat. Probab. Lett. 82(11), 2017–2024 (2012)

    Article  MathSciNet  Google Scholar 

  • Bellini, F., Di Bernardino, E.: Risk management with expectiles. Eur. J. Finance 23(6), 487–506 (2017)

    Article  Google Scholar 

  • Bellini, F., Klar, B., Müller, A., Gianin, E.R.: Generalized quantiles as risk measures. Insurance Math. Econom. 54, 41–48 (2014)

    Article  MathSciNet  Google Scholar 

  • Bellini, F., Klar, B., Müller, A.: Expectiles, omega ratios and stochastic ordering. Methodol. Comput. Appl. Probab. 20(3), 855–873 (2016)

    Article  MathSciNet  Google Scholar 

  • Birnbaum, Z.W.: An inequality for Mill’s ratio. Ann. Math. Stat. 13(2), 245–246 (1942)

    Article  MathSciNet  Google Scholar 

  • Black, F., Scholes, M.: The pricing of options and corporate liabilities. J. Polit. Econ. 81(3), 637–654 (1973)

    Article  MathSciNet  Google Scholar 

  • Breckling, J., Chambers, R.: M-quantiles. Biometrika 75(4), 761–772 (1988)

    Article  MathSciNet  Google Scholar 

  • Brockwell, P.J., Davis, R.A.: Time Series: Theory and Methods, 2nd edn. Springer, Berlin (1991)

    Book  Google Scholar 

  • Cascos, I., Ochoa, M.: Expectile depth: theory and computation for bivariate datasets. J. Multivar. Anal. 184, 104757 (2021)

    Article  MathSciNet  Google Scholar 

  • Chambers, R., Dreassi, E., Salvati, N.: Disease mapping via negative binomial regression M-quantiles. Stat. Med. 33(27), 4805–4824 (2014)

    Article  MathSciNet  Google Scholar 

  • Chambers, R., Salvati, N., Tzavidis, N.: Semiparametric small area estimation for binary outcomes with application to unemployment estimation for local authorities in the UK. J. R. Stat. Soc. Ser. A 179(2), 453–479 (2016)

    Article  MathSciNet  Google Scholar 

  • Daouia, A., Girard, S., Stupfler, G.: Estimation of tail risk based on extreme expectiles. J. R. Stat. Soc. B 80(2), 263–292 (2018)

    Article  MathSciNet  Google Scholar 

  • Daouia, A., Girard, S., Stupfler, G.: Tail expectile process and risk assessment. Bernoulli 26(1), 531–556 (2020)

    Article  MathSciNet  Google Scholar 

  • Daouia, A., Girard, S., Stupfler, G.: ExpectHill estimation, extreme risk and heavy tails. J. Econom. 221(1), 97–117 (2021)

    Article  MathSciNet  Google Scholar 

  • Daouia, A., Stupfler, G., Usseglio-Carleve, A.: Inference for extremal regression with dependent heavy-tailed data. Ann. Stat. 51(5), 2040–2066 (2023)

    Article  MathSciNet  Google Scholar 

  • Dawber, J., Salvati, N., Fabrizi, E., Tzavidis, N.: Expectile regression for multicategory outcomes with application to small area estimation of labour force participation. J. R. Stat. Soc.: Ser. A 185(Supplement_2), 590–619 (2022)

  • de Haan, L., Ferreira, A.: Extreme Value Theory: An Introduction. Springer, New York (2006)

    Book  Google Scholar 

  • del Castillo, J., Soler, D.M., Serra, I.: Ercv: Fitting Tails by the Empirical Residual Coefficient of Variation. R package version 1.0.1 (2019). https://CRAN.R-project.org/package=ercv

  • Dickson, D.C.: Insurance Risk and Ruin. Cambridge University Press, Cambridge (2016)

    Book  Google Scholar 

  • Eberl, A., Klar, B.: Stochastic orders and measures of skewness and dispersion based on expectiles. Stat. Pap. 64(2), 509–527 (2023)

    Article  MathSciNet  Google Scholar 

  • Francq, C., Zakoïan, J.-M.: Maximum likelihood estimation of pure GARCH and ARMA-GARCH processes. Bernoulli 10(4), 605–637 (2004)

    Article  MathSciNet  Google Scholar 

  • Galanos, A., Kley, T.: Rugarch: Univariate GARCH Models. R package version 1.4-9 (2022). https://CRAN.R-project.org/package=rugarch

  • Girard, S., Stupfler, G., Usseglio-Carleve, A.: Extreme conditional expectile estimation in heavy-tailed heteroscedastic regression models. Ann. Stat. 49(6), 3358–3382 (2021)

    Article  MathSciNet  Google Scholar 

  • Girard, S., Stupfler, G., Usseglio-Carleve, A.: Nonparametric extreme conditional expectile estimation. Scand. J. Stat. 49(1), 78–115 (2022a)

  • Girard, S., Stupfler, G., Usseglio-Carleve, A.: On automatic bias reduction for extreme expectile estimation. Stat. Comput. 32(4), 64 (2022b)

  • Gneiting, T.: Making and evaluating point forecasts. J. Am. Stat. Assoc. 106(494), 746–762 (2011)

    Article  MathSciNet  Google Scholar 

  • Hankin, R.K.S., Clausen, A., Murdoch, D.: Gsl: Wrapper for the Gnu Scientific Library. R package version 2.1-8 (2023). https://cran.r-project.org/package=gsl

  • Heston, S.L.: A closed-form solution for options with stochastic volatility with applications to bond and currency options. Rev. Financ. Stud. 6(2), 327–343 (1993)

    Article  MathSciNet  Google Scholar 

  • Holzmann, H., Klar, B.: Expectile asymptotics. Electron. J. Stat. 10(2), 2355–2371 (2016)

    Article  MathSciNet  Google Scholar 

  • Jones, M.C.: Expectiles and M-quantiles are quantiles. Stat. Probab. Lett. 20(2), 149–153 (1994)

    Article  MathSciNet  Google Scholar 

  • Koenker, R.: When are expectiles percentiles? Econom. Theor. 9(3), 526–527 (1993)

    Article  Google Scholar 

  • Koenker, R.: Quantile Regression. Cambridge University Press, Cambridge (2005)

    Book  Google Scholar 

  • Krätschmer, V., Zähle, H.: Statistical inference for expectile-based risk measures. Scand. J. Stat. 44(2), 425–454 (2017)

    Article  MathSciNet  Google Scholar 

  • Kuan, C.-M., Yeh, J.-H., Hsu, Y.-C.: Assessing value at risk with care, the conditional autoregressive expectile models. J. Econom. 150(2), 261–270 (2009)

    Article  MathSciNet  Google Scholar 

  • Laudagé, C., Turkalj, I.: Calculating Expectiles and Range Value-at-Risk using Quantum Computers. arXiv:2211.04456 (2022)

  • Mező, I., Baricz, Á.: On the generalization of the Lambert W function. Trans. Am. Math. Soc. 369(11), 7917–7934 (2017)

    Article  MathSciNet  Google Scholar 

  • Newey, W.K., Powell, J.L.: Asymmetric least squares estimation and testing. Econometrica 55(4), 819–847 (1987)

  • Otto-Sobotka, F., Spiegel, E., Schnabel, S., Schulze Waltrup, L., Eilers, P., Kauermann, G., Kneib, T.: Expectreg: Expectile and Quantile Regression. R package version 0.52 (2022). https://CRAN.R-project.org/package=expectreg

  • Padoan, S., Stupfler, G.: ExtremeRisks: Extreme Risk Measures. R package version 0.0.4 (2020). https://cran.r-project.org/package=ExtremeRisks

  • Racine, J.S., Hayfield, T.: Np: Nonparametric Kernel Smoothing Methods for Mixed Data Types. R package version 0.60-17 (2023). https://CRAN.R-project.org/package=np

  • Sobotka, F., Kneib, T.: Geoadditive expectile regression. Comput. Stat. Data Anal. 56(4), 755–767 (2012)

    Article  MathSciNet  Google Scholar 

  • Sobotka, F.O., Kneib, T., Salvati, N., Ranalli, M.G.: Adaptive semiparametric M-quantile regression. Econom. Stat. 11, 116–129 (2019)

    MathSciNet  Google Scholar 

  • Stupfler, G., Usseglio-Carleve, A.: Composite bias-reduced \(L^p-\)quantile-based estimators of extreme quantiles and expectiles. Can. J. Stat. 51(2), 704–742 (2023)

    Article  Google Scholar 

  • Taylor, J.W.: Estimating value at risk and expected shortfall using expectiles. J. Financ. Econom. 6(2), 231–252 (2008)

    Google Scholar 

  • Wuertz, D., Chalabi, Y., Setz, T., Maechler, M., Boudt, C., Chausse, P., Miklovac, M., Boshnakov, G.N.: fGarch: Rmetrics—Autoregressive Conditional Heteroskedastic Modelling. R package version 4031.90 (2023). https://CRAN.R-project.org/package=fGarch

  • Yee, T., Moler, C.: VGAM: Vector Generalized Linear And Additive Models. R package version 1.1-9. (2023) https://cran.r-project.org/package=VGAM

  • Zhu, L., Dong, Y., Li, R.: Semiparametric estimation of conditional heteroscedasticity via single-index modeling. Stat. Sin. 23(3), 1235–1255 (2013)

    MathSciNet  Google Scholar 

  • Ziegel, J.F.: Coherence and elicitability. Math. Financ. 26(4), 901–918 (2016)

    Article  MathSciNet  Google Scholar 

  • Zou, H.: Generalizing Koenker’s distribution. J. Stat. Plan. Inference 148, 123–127 (2014)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

The authors acknowledge two anonymous reviewers for their helpful comments that led to an improved version of this article, as well as Mikael Escobar-Bach for a useful chat about bounds for hazard ratios. This research was supported by the French National Research Agency under the Grants ANR-19-CE40-0013 (ExtremReg project), ANR-17-EURE-0010 (EUR CHESS) and ANR-11-LABX-0020-01 (Centre Henri Lebesgue). A. Daouia and G. Stupfler acknowledge financial support from the TSE-HEC ACPR Chair “Regulation and systemic risks”, and G. Stupfler acknowledges further support from the Chair Stress Test, RISK Management and Financial Steering of the Foundation Ecole Polytechnique.

Author information

Authors and Affiliations

Authors

Contributions

All authors contributed equally to the work, and read and approved the final manuscript.

Corresponding author

Correspondence to Antoine Usseglio-Carleve.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Proofs of the theoretical results

Proof of Theorem 2.1

Clearly

$$\begin{aligned}{} & {} \forall i\in I, \ \forall x\in [a_i,a_{i+1}), \ \varphi (x)\\{} & {} \quad =\int _{{\mathbb {R}}} (t-x)\mathbb {1}_{\{ t>x \}} \, \mu (\textrm{d}t) =\psi (a_i)-x\overline{F}(a_i) \end{aligned}$$

where \(\psi (x)=\int _{{\mathbb {R}}} t \mathbb {1}_{\{ t>x \}} \, \mu (\textrm{d}t)\) and \(\overline{F}(x)=1-F(x)\) with \(F(x)=\mu ((-\infty ,x])\). Consequently, for any \(i\in I\) and \(x\in [a_i,a_{i+1})\),

$$\begin{aligned} (1-\tau ) g_{\tau }(x)&= -x\{ (2\tau -1) \overline{F}(a_i)+1-\tau \} \\&\quad + (2\tau -1) \psi (a_i) + (1-\tau )m \\&= -x( \tau \overline{F}(a_i)+(1-\tau )F(a_i) )\\&\quad + \tau \psi (a_i) + (1-\tau ) (m-\psi (a_i)) \end{aligned}$$

where \(m=\int _{{\mathbb {R}}} x \, \mu (\textrm{d}x)=\sum _{k\in I} p_k a_k\). In other words, the function \(g_{\tau }\) is continuous, piecewise linear and decreasing, and tends to \(+\infty \) (resp. \(-\infty \)) as \(x\rightarrow -\infty \) (resp. \(x\rightarrow +\infty \)). It follows that there is a unique index \(i=i(\tau )\) such that the two inequalities

$$\begin{aligned}&\tau \psi (a_i) + (1-\tau ) (m-\psi (a_i)) \\&\quad \ge a_i( \tau \overline{F}(a_i)+(1-\tau )F(a_i) ) \text{ and } \\&\tau \psi (a_{i+1}) + (1-\tau ) (m-\psi (a_{i+1})) \\&\quad < a_{i+1}( \tau \overline{F}(a_{i+1})+(1-\tau )F(a_{i+1}) ) \end{aligned}$$

hold. With this index i, the expectile \(\xi _{\tau }\) is the unique root of the linear function \(g_{\tau }\) on the interval \([a_i,a_{i+1})\), namely:

$$\begin{aligned} \xi _{\tau }=\frac{\tau \psi (a_i) + (1-\tau ) (m-\psi (a_i))}{\tau \overline{F}(a_i)+(1-\tau )F(a_i)}. \end{aligned}$$
(A1)

Now \(\psi (a_i)=\sum _{k>i} p_k a_k\) and \(\overline{F}(a_i)=\sum _{k>i} p_k\), so that, for any i,

$$\begin{aligned}{} & {} \tau \psi (a_i) + (1-\tau ) (m-\psi (a_i)) \\{} & {} \qquad - a_i( \tau \overline{F}(a_i)+(1-\tau )F(a_i) ) \\{} & {} \quad = \tau \left( \sum _{k<i} p_k(a_i-a_k) \right. \\{} & {} \qquad \left. + \sum _{k>i} p_k(a_k-a_i) \right) - \sum _{k<i} p_k(a_i-a_k). \end{aligned}$$

The above pair of inequalities is therefore equivalent to

$$\begin{aligned}{} & {} \frac{\sum _{k<i} p_k (a_i-a_k)}{\sum _{k<i} p_k (a_i-a_k) + \sum _{k>i} p_k (a_k-a_i)} \\{} & {} \quad \le \tau< \frac{\sum _{k<i+1} p_k (a_{i+1}-a_k)}{\sum _{k<i+1} p_k (a_{i+1}-a_k) + \sum _{k>i+1} p_k (a_k-a_{i+1})} \end{aligned}$$

as announced. The two identities involving \(\xi _{\tau }\) follow immediately from (A1). \(\square \)

Proof of Theorem 2.2

Define a continuous function \(G=G_{\tau }\) by setting

$$\begin{aligned} G(x)=x-\frac{g_{\tau }(x)}{g_{\tau }'(x)}, \text{ so } \text{ that } x_{n+1}=G(x_n). \end{aligned}$$

Recall that \(g_{\tau }:x\mapsto \frac{2\tau -1}{1-\tau }\varphi (x) + m-x\) is convex. It is also clear that the derivative \(g_{\tau }'\) is negative. In particular, \(g_{\tau }\) is decreasing, and since \(\xi _{\tau }\) is the unique solution of the equation \(g_{\tau }(x)=0\), one has \(g_{\tau }(x_0)>0\) for any \(x_0<\xi _{\tau }\). Then

$$\begin{aligned}&\forall x_0<\xi _{\tau }, \ G(x_0)-x_0 =-\frac{g_{\tau }(x_0)}{g_{\tau }'(x_0)}>0 \\&\quad \text{ and } \ G(x_0)-\xi _{\tau } {=}\left\{ \frac{\xi _{\tau }-x_0}{g_{\tau }(\xi _{\tau }){-}g_{\tau }(x_0)} {-} \frac{1}{g_{\tau }'(x_0)} \right\} g_{\tau }(x_0) \le 0 \end{aligned}$$

using the convexity property of \(g_{\tau }\). It follows that for any \(x_0<\xi _{\tau }\), \(x_0<G(x_0)\le \xi _{\tau }\), and therefore, for any starting point \(x_0<\xi _{\tau }\), the Newton–Raphson sequence of iterates \((x_n)\) is nondecreasing and bounded and hence convergent. The limit must be a root of \(g_{\tau }\) by taking limits in the equation \(x_{n+1}=G(x_n)\), meaning that \((x_n)\) converges to \(\xi _{\tau }\).

A Taylor expansion of \(g_{\tau }\) on the interval \([x_n,\xi _{\tau }]\subset [x_0,\xi _{\tau }]\) (on which \(g_{\tau }\) is twice continuously differentiable) with remainder in integral form entails

$$\begin{aligned} 0= & {} g_{\tau }(\xi _{\tau }) =g_{\tau }(x_n)+(\xi _{\tau }-x_n)g_{\tau }'(x_n)\\{} & {} +\int _{x_n}^{\xi _{\tau }} (\xi _{\tau }-u) g_{\tau }''(u) \, \textrm{d}u. \end{aligned}$$

This is readily rewritten as

$$\begin{aligned} x_{n+1}-\xi _{\tau } = \frac{1}{g_{\tau }'(x_n)} \int _{x_n}^{\xi _{\tau }} (\xi _{\tau }-u) g_{\tau }''(u) \, \textrm{d}u. \end{aligned}$$

Then

$$\begin{aligned}&|x_{n+1}-\xi _{\tau }| \le \frac{\max _{[x_n,\xi _{\tau }]} g_{\tau }''}{2|g_{\tau }'(x_n)|} (\xi _{\tau }-x_n)^2 \\&\quad = \frac{1}{2} \frac{(2\tau -1)\max _{[x_n,\xi _{\tau }]} f}{1-\tau +(2\tau -1) \overline{F}(x_n)} |x_n-\xi _{\tau }|^2 \\&\quad \le \frac{1}{2} \frac{(2\tau -1)\max _{[x_n,\xi _{\tau }]} f}{1-\tau +(2\tau -1) \overline{F}(\xi _{\tau })} |x_n-\xi _{\tau }|^2 \\&\quad \le \frac{1}{2} \frac{(2\tau -1)\max _{[x_k,\xi _{\tau }]} f}{1-\tau +(2\tau -1) \overline{F}(\xi _{\tau })} |x_n-\xi _{\tau }|^2 \end{aligned}$$

for any \(k<n\). The proof is complete. \(\square \)

Proof of Theorem 2.3

With the notation of the proof of Theorem 2.2, the function G is twice continuously differentiable in a neighborhood of \(\xi _{\tau }\) with

$$\begin{aligned} G'(x)= & {} \frac{g_{\tau }(x) g_{\tau }''(x)}{\{ g_{\tau }'(x) \}^2}, \\ G''(x)= & {} \frac{g_{\tau }''(x)}{g_{\tau }'(x)} + \frac{g_{\tau }(x)}{\{ g_{\tau }'(x) \}^2} \left( g_{\tau }^{(3)}(x) - 2\frac{\{ g_{\tau }''(x) \}^2}{g_{\tau }'(x)} \right) . \end{aligned}$$

In particular, \(G'(\xi _{\tau })=0\) and

$$\begin{aligned} G''(\xi _{\tau })= & {} \frac{g_{\tau }''(\xi _{\tau })}{g_{\tau }'(\xi _{\tau })}\\= & {} -\frac{(2\tau -1)f(\xi _{\tau })}{(2\tau -1)\overline{F}(\xi _{\tau })+1-\tau }. \end{aligned}$$

Use a Taylor expansion to get, for any n,

$$\begin{aligned} x_{n+1}-\xi _{\tau }=G(x_n)-G(\xi _{\tau })=\frac{1}{2} (x_n-\xi _{\tau })^2 G''(c_n) \end{aligned}$$

for some \(c_n\in \left]x_n,\xi _{\tau }\right[\). It follows from Theorem 2.2 that \(c_n\rightarrow \xi _{\tau }\) and therefore, using the continuity of \(G''\),

$$\begin{aligned}{} & {} \lim _{n\rightarrow \infty } \frac{\xi _{\tau }-x_{n+1}}{(x_n-\xi _{\tau })^2}\\{} & {} \quad = -\frac{1}{2} G''(\xi _{\tau })\\{} & {} \quad =\frac{1}{2} \times \frac{(2\tau -1)f(\xi _{\tau })}{(2\tau -1)\overline{F}(\xi _{\tau })+1-\tau } \end{aligned}$$

as required. \(\square \)

Proof of Theorem 2.4

From the last chain of inequalities in the proof of Theorem 2.2, we get, for any n,

$$\begin{aligned} |x_{n+1}-\xi _{\tau }|\le & {} \frac{1}{2} \frac{(2\tau -1)\max _{[x_0,\xi _{\tau }]} f}{1-\tau +(2\tau -1) \overline{F}(\xi _{\tau })} |x_n-\xi _{\tau }|^2 \\\le & {} \frac{1}{2} \frac{(2\tau -1)\max _{[(1-\varepsilon )\xi _{\tau },\xi _{\tau }]} f}{1-\tau +(2\tau -1) \overline{F}(\xi _{\tau })} |x_n-\xi _{\tau }|^2. \end{aligned}$$

Take then \(c>0\) sufficiently large and write

$$\begin{aligned} \overline{F}(x)= & {} x^{-1/\gamma } \left\{ c^{1/\gamma } \overline{F}(c) \exp \left( \int _c^x \eta (t) \frac{\textrm{d}t}{t} \right) \right\} \\{} & {} \text{ with } \eta (t) = \frac{1}{\gamma } - \frac{t f(t)}{\overline{F}(t)}. \end{aligned}$$

The function \(\eta \) is continuous on \([c,\infty )\) and converges to 0 at infinity. By the representation theorem for regularly varying functions (see Theorem B.1.6 p.365 in de Haan and Ferreira 2006), \(\overline{F}\) is regularly varying with index \(-1/\gamma \), and then the von Mises condition ensures that f is regularly varying with index \(-1/\gamma -1\). By the uniform convergence theorem for regularly varying functions (see Theorem B.1.4 p.363 in de Haan and Ferreira 2006),

$$\begin{aligned} \frac{\xi _{\tau }}{\overline{F}(\xi _{\tau })} \max _{[(1-\varepsilon )\xi _{\tau },\xi _{\tau }]} f= & {} \frac{\xi _{\tau } f(\xi _{\tau })}{\overline{F}(\xi _{\tau })} \max _{x\in [1-\varepsilon ,1]} \frac{f(\xi _{\tau } x)}{f(\xi _{\tau })}\\{} & {} \rightarrow \frac{(1-\varepsilon )^{-1/\gamma -1}}{\gamma } \text{ as } \tau \uparrow 1. \end{aligned}$$

Moreover (see Bellini et al. 2014; Daouia et al. 2018)

$$\begin{aligned} \frac{\overline{F}(\xi _{\tau })}{1-\tau } \rightarrow \frac{1}{\gamma } - 1 \text{ as } \tau \uparrow 1. \end{aligned}$$

Conclude from these two convergences that

$$\begin{aligned}{} & {} \frac{1}{2} \frac{(2\tau -1)\max _{[(1-\varepsilon )\xi _{\tau },\xi _{\tau }]} f}{1-\tau +(2\tau -1) \overline{F}(\xi _{\tau })}\\{} & {} \quad = \frac{1}{\xi _{\tau }} \left( (1-\varepsilon )^{-1/\gamma -1} \frac{1/\gamma -1}{2} + {\text {o}}(1) \right) \text{ as } \tau \uparrow 1. \end{aligned}$$

The proof is complete. \(\square \)

Proof of Theorem 3.1

We first prove the asymptotic normality statement, and for this, it is enough to show that \(\widehat{\Sigma }_{12,n}/\widehat{\sigma }_n^2\rightarrow {\Sigma _{12}/\sigma ^2=\Sigma _{12}/\Sigma _{22}}\) in probability. By the law of large numbers, \(\widehat{\sigma }_n^2\rightarrow \sigma ^2\) in probability, so it suffices to show that \(\widehat{\Sigma }_{12,n}\rightarrow \Sigma _{12}\) in probability. Finally, since \(\widehat{\xi }_{\tau ,n}\rightarrow \xi _{\tau }\) in probability (for example by Theorem 2 in Holzmann and Klar 2016), it is enough to prove that \(\widehat{\overline{F}}_n(\widehat{\xi }_{\tau ,n})\rightarrow \overline{F}(\xi _{\tau })\) and \(\widehat{\varphi }_n^{(2)}(\widehat{\xi }_{\tau ,n})\rightarrow \varphi ^{(2)}(\xi _{\tau })\) in probability.

Fix \(\varepsilon >0\). Then \(\widehat{\xi }_{\tau ,n}\in [\xi _{\tau }-\varepsilon ,\xi _{\tau }+\varepsilon ]\) with arbitrarily high probability as \(n\rightarrow \infty \). Then clearly

$$\begin{aligned}{} & {} \left| \widehat{\overline{F}}_n(\widehat{\xi }_{\tau ,n})-\widehat{\overline{F}}_n(\xi _{\tau }) \right| \le \frac{1}{n} \sum _{i=1}^n | \mathbb {1}_{\{ X_i>\widehat{\xi }_{\tau ,n} \}} - \mathbb {1}_{\{ X_i>\xi _{\tau } \}} | \\{} & {} \quad \le \frac{1}{n} \sum _{i=1}^n \mathbb {1}_{\{ X_i\in [\xi _{\tau }-\varepsilon ,\xi _{\tau }+\varepsilon ] \}} \end{aligned}$$

with arbitrarily high probability as \(n\rightarrow \infty \). The upper bound converges in probability to \(\mu ([\xi _{\tau }-\varepsilon ,\xi _{\tau }+\varepsilon ])\), by the law of large numbers, which is arbitrarily small as \(\varepsilon \downarrow 0\). Conclude that \(\widehat{\overline{F}}_n(\widehat{\xi }_{\tau ,n}) = (\widehat{\overline{F}}_n(\widehat{\xi }_{\tau ,n})-\widehat{\overline{F}}_n(\xi _{\tau }))+\widehat{\overline{F}}_n(\xi _{\tau })\rightarrow \overline{F}(\xi _{\tau })\) in probability, by the law of large numbers again. We now prove that \(\widehat{\varphi }_n(\widehat{\xi }_{\tau ,n})=\widehat{\varphi }_n^{(1)}(\widehat{\xi }_{\tau ,n})\rightarrow \varphi ^{(1)}(\xi _{\tau })=\varphi (\xi _{\tau })\) before turning to the convergence of \(\widehat{\varphi }_n^{(2)}(\widehat{\xi }_{\tau ,n})\). Write

$$\begin{aligned}{} & {} (X-x) \mathbb {1}_{\{ X>x \}} - (X-x') \mathbb {1}_{\{ X>x' \}} \\{} & {} \quad =(x'-x) \mathbb {1}_{\{ X>x \}} + (X-x') ( \mathbb {1}_{\{ X>x \}} - \mathbb {1}_{\{ X>x' \}} ) \end{aligned}$$

to obtain

$$\begin{aligned}&\left| \widehat{\varphi }_n^{(1)}(\widehat{\xi }_{\tau ,n})-\widehat{\varphi }_n^{(1)}(\xi _{\tau }) \right| \\&\quad \le |\widehat{\xi }_{\tau ,n}-\xi _{\tau }| \widehat{\overline{F}}_n(\widehat{\xi }_{\tau ,n}) \\&\qquad + \frac{1}{n} \sum _{i=1}^n |X_i-\xi _{\tau }| | \mathbb {1}_{\{ X_i>\widehat{\xi }_{\tau ,n} \}} - \mathbb {1}_{\{ X_i>\xi _{\tau } \}} | \\&\quad \le \varepsilon \left( \widehat{\overline{F}}_n(\widehat{\xi }_{\tau ,n}) + \frac{1}{n} \sum _{i=1}^n \mathbb {1}_{\{ X_i\in [\xi _{\tau }-\varepsilon ,\xi _{\tau }+\varepsilon ] \}} \right) \le 2\varepsilon \end{aligned}$$

with arbitrarily high probability as \(n\rightarrow \infty \). The upper bound is arbitrarily small as \(\varepsilon \downarrow 0\), so that the law of large numbers yields \(\widehat{\varphi }_n^{(1)}(\widehat{\xi }_{\tau ,n})\rightarrow \varphi ^{(1)}(\xi _{\tau })\) in probability. Finally

$$\begin{aligned} (X-x)^2 - (X-x')^2 = 2(x'-x)(X-x)-(x'-x)^2 \end{aligned}$$

so that, with arbitrarily high probability as \(n\rightarrow \infty \),

$$\begin{aligned}&\left| \widehat{\varphi }_n^{(2)}(\widehat{\xi }_{\tau ,n})-\widehat{\varphi }_n^{(2)}(\xi _{\tau }) \right| \\&\quad \le |\widehat{\xi }_{\tau ,n}-\xi _{\tau }|^2\widehat{\overline{F}}_n(\widehat{\xi }_{\tau ,n}) + 2 |\widehat{\xi }_{\tau ,n}-\xi _{\tau }| \widehat{\varphi }_n^{(1)}(\widehat{\xi }_{\tau ,n})\\&\qquad + \frac{1}{n} \sum _{i=1}^n (X_i-\xi _{\tau })^2 | \mathbb {1}_{\{ X_i>\widehat{\xi }_{\tau ,n} \}} - \mathbb {1}_{\{ X_i>\xi _{\tau } \}} | \\&\quad \le \varepsilon \left( \varepsilon \widehat{\overline{F}}_n(\widehat{\xi }_{\tau ,n}) + 2 \widehat{\varphi }_n^{(1)}(\widehat{\xi }_{\tau ,n}) + \varepsilon \right. \\&\qquad \left. \times \frac{1}{n} \sum _{i=1}^n \mathbb {1}_{\{ X_i\in [\xi _{\tau }-\varepsilon ,\xi _{\tau }+\varepsilon ] \}} \right) \\&\quad \le 2\varepsilon \left( \varepsilon + \widehat{\varphi }_n^{(1)}(\widehat{\xi }_{\tau ,n}) \right) . \end{aligned}$$

Conclude in a similar fashion that \(\widehat{\varphi }_n^{(2)}(\widehat{\xi }_{\tau ,n})\rightarrow \varphi ^{(2)}(\xi _{\tau })\) in probability, as required. The fact that \(\widetilde{\xi }_{\tau ,n}\) has the lowest asymptotic variance among all asymptotically unbiased linear combinations of \(\widehat{\xi }_{\tau ,n}\) and \(\overline{X}_n-m\) is then obvious since

$$\begin{aligned} \widehat{\xi }_{\tau ,n}- \frac{\Sigma _{12}}{\sigma ^2} (\overline{X}_n-m) = \widetilde{\xi }_{\tau ,n} + {\text {o}}_{{\mathbb {P}}}(1/\sqrt{n}) \end{aligned}$$

has this property.

It remains to prove the assertions about the variance reduction factor \(1-R(\tau ,\mu )\). This function is clearly zero at \(\tau =1/2\). Note also that

$$\begin{aligned} \varphi ^{(2)}(x)={\mathbb {E}}((X-x)^2 \mathbb {1}_{\{ X>x \}})=2\int _x^{\infty } (t-x) \overline{F}(t) \, \textrm{d}t. \end{aligned}$$

As a result, and since \(\tau \mapsto \xi _{\tau }\) is continuously differentiable on \(I=(\tau _1,\tau _2)\) (see Proposition 1(iii) in Holzmann and Klar 2016), the function \(\tau \mapsto 1-R(\tau ,\mu )\) is continuously differentiable on this interval. It remains to prove the statements about monotonicity. Set

$$\begin{aligned} u(\tau )&=u(\tau ,\mu ) =(1-\tau ) {\mathbb {E}}((X-\xi _{\tau })^2) \\&\qquad + (2 \tau -1) \varphi ^{(2)}(\xi _{\tau }) \\ \text{ and } v(\tau )&=v(\tau ,\mu ) =(1-\tau )^2 {\mathbb {E}}((X-\xi _{\tau })^2) \\&\qquad + (2 \tau -1) \varphi ^{(2)}(\xi _{\tau }) \end{aligned}$$

so that \(R(\tau ,\mu )=(u(\tau ))^2/(\sigma ^2 v(\tau ))\) and therefore

$$\begin{aligned} \frac{\partial R}{\partial \tau }(\tau ,\mu ) = \frac{u(\tau )}{\sigma ^2 (v(\tau ))^2} (2 u'(\tau ) v(\tau ) - v'(\tau ) u(\tau )). \end{aligned}$$

Writing \(u(\tau )=(1-\tau ) {\mathbb {E}}((X-\xi _{\tau })^2 \mathbb {1}_{\{ X<\xi _{\tau } \}}) + \tau {\mathbb {E}}((X-\xi _{\tau })^2 \mathbb {1}_{\{ X>\xi _{\tau } \}})\) yields in particular that \(u(\tau )>0\) for any \(\tau \), meaning that the partial derivative \(\frac{\partial R}{\partial \tau }(\tau ,\mu )\) has the same sign as \(2 u'(\tau ) v(\tau ) - v'(\tau ) u(\tau )\). Now

$$\begin{aligned} u'(\tau )&=2\varphi ^{(2)}(\xi _{\tau })-{\mathbb {E}}((X-\xi _{\tau })^2)\\&\quad -2\frac{\textrm{d}\xi _{\tau }}{\textrm{d}\tau } ( (2\tau -1)\varphi (\xi _{\tau })+(1-\tau )(m-\xi _{\tau }) ) \\&= 2\varphi ^{(2)}(\xi _{\tau })-{\mathbb {E}}((X-\xi _{\tau })^2) \text{(using } \text{(3)) } \\ \text{ and } v'(\tau )&=2\varphi ^{(2)}(\xi _{\tau })-2(1-\tau ){\mathbb {E}}((X-\xi _{\tau })^2)\\&\quad -2\frac{\textrm{d}\xi _{\tau }}{\textrm{d}\tau } ( (2\tau -1)\varphi (\xi _{\tau })+(1-\tau )^2(m-\xi _{\tau }) ) \\&=2\varphi ^{(2)}(\xi _{\tau })-2(1-\tau ){\mathbb {E}}((X-\xi _{\tau })^2)\\&\quad +2\tau (1-\tau )\frac{\textrm{d}\xi _{\tau }}{\textrm{d}\tau } (m-\xi _{\tau }) \text{(from }~(3) \text{ again). } \end{aligned}$$

Straightforward calculations yield

$$\begin{aligned}{} & {} 2 u'(\tau ) v(\tau ) - v'(\tau ) u(\tau ) \\{} & {} \quad =2\left( (1-2\tau ) {\mathbb {E}}((X-\xi _{\tau })^2 \mathbb {1}_{\{ X<\xi _{\tau } \}}) {\mathbb {E}}((X-\xi _{\tau })^2 \mathbb {1}_{\{ X>\xi _{\tau } \}})\right. \\{} & {} \qquad \left. + \tau (1-\tau ) u(\tau ) \frac{\textrm{d}\xi _{\tau }}{\textrm{d}\tau } (m-\xi _{\tau }) \right) . \end{aligned}$$

This quantity is positive on (0, 1/2) and negative on (1/2, 1) because \(\tau \mapsto \xi _{\tau }\) is strictly increasing (see Proposition 1(ii) in Holzmann and Klar 2016) and \(\xi _{1/2}=m\). The proof is complete. \(\square \)

Proof of Theorem 3.2

Let , where c is to be found so as to minimize the relative asymptotic variance of . Write

Applying Proposition 3.1 entails that the desired value of c satisfies

$$\begin{aligned} c \frac{q_{\alpha _n}}{\xi _{\tau _n}} {\rightarrow } -\frac{V_{12}}{V_{22}}= & {} - \left( \left\{ \min \left( \frac{\lambda }{1/\gamma -1}, 1 \right) \right\} ^{1-\gamma } \right. \\{} & {} \left. + \lambda \left\{ \min \left( \frac{\lambda }{1/\gamma -1}, 1 \right) \right\} ^{-\gamma } - \lambda \right) { \text{ as } n\rightarrow \infty .} \end{aligned}$$

Apply Proposition 1(i) in Daouia et al. (2020) to obtain

$$\begin{aligned} \frac{q_{\alpha _n}}{\xi _{\tau _n}} = \frac{q_{\alpha _n}}{q_{\tau _n}} \frac{q_{\tau _n}}{\xi _{\tau _n}} \rightarrow \lambda ^{-\gamma } (1/\gamma -1)^{\gamma } \text{ as } n\rightarrow \infty , \end{aligned}$$

leading to a choice of c minimizing the relative asymptotic variance of as

$$\begin{aligned} c= & {} - \left( \frac{\lambda }{1/\gamma -1} \right) ^{\gamma } \left( \left\{ \min \left( \frac{\lambda }{1/\gamma -1}, 1 \right) \right\} ^{1-\gamma } \right. \\{} & {} \left. + \lambda \left\{ \min \left( \frac{\lambda }{1/\gamma -1}, 1 \right) \right\} ^{-\gamma } - \lambda \right) . \end{aligned}$$

The statement on the asymptotic normality of with this choice of c is then immediate. Finding the maximum of \(\lambda \mapsto C(\gamma ,\lambda )\) is done by noting that this function is decreasing past \(1/\gamma -1\), and

$$\begin{aligned}{} & {} \forall \lambda \in (0,1/\gamma -1), \ C(\gamma ,\lambda )\\{} & {} \qquad =\frac{1-2\gamma }{2\gamma } \left( \frac{\lambda ^{1/2-\gamma }}{\gamma (1/\gamma -1)^{1-\gamma }} - \sqrt{\lambda } \right) ^2. \end{aligned}$$

Maximizing this function over \((0,1/\gamma -1)\) is straightforward and leads to the value \(\lambda ^{\star }\) specified in the statement of Theorem 3.2. The last asymptotic normality result follows by plugging \(\lambda ^{\star }\) into \(C(\gamma ,\lambda )\). \(\square \)

Detailed calculations related to the examples

Distribution supported on a set with three elements (Example 2.2)

Let \(\mu \) be the probability distribution on a set \(\{a,b,c\}\) with \(a<b<c\) characterized by \(\mu (\{b\})=p\) and \(\mu (\{c\})=q\), with \(p,q>0\) and \(p+q<1\). Then, from Corollary 2.1,

$$\begin{aligned} \xi _{\tau }= & {} \frac{\tau (pb+qc) + (1-\tau )(1-p-q)a}{(2\tau -1)(p+q)+1-\tau }, \\{} & {} \text{ for } \tau \le 1-\frac{q(c-b)}{(1-p)(b-a)+q(a+c-2b)}, \end{aligned}$$

and

$$\begin{aligned} \xi _{\tau } = \frac{\tau qc + (1-\tau )\{(1-p-q)a+pb\}}{(2\tau -1)q+1-\tau } \text{ otherwise }. \end{aligned}$$

In particular, for the distribution \(\mu \) on \(\{0,1,2\}\) with \(\mu (\{1\})=p\) and \(\mu (\{2\})=q\),

$$\begin{aligned} \xi _{\tau } = \left\{ \begin{array}{ll} \dfrac{\tau (p+2q)}{(2\tau -1)(p+q)+1-\tau } &{} \text{ for } \tau \le 1-\dfrac{q}{1-p}, \\ \dfrac{2\tau q + (1-\tau )p}{(2\tau -1)q+1-\tau } &{} \text{ otherwise. } \end{array} \right. \end{aligned}$$

Taking \(p=p_1(1-p_2)+p_2(1-p_1)\) and \(q=p_1 p_2\), for \(p_1,p_2\in (0,1)\), yields the expectile of the sum of two independent random variables having Bernoulli distributions with parameters \(p_1\) and \(p_2\):

$$\begin{aligned} \xi _{\tau } = \left\{ \begin{array}{ll} \dfrac{\tau (p_1+p_2)}{(2\tau -1)(p_1+p_2-p_1p_2)+1-\tau }\\ \qquad \qquad \qquad \qquad \text{ for } \tau \le \dfrac{1-p_1-p_2+p_1p_2}{1-p_1-p_2+2p_1p_2}, \\ \dfrac{(2\tau -1) 2p_1p_2 + (1-\tau )(p_1+p_2)}{(2\tau -1)p_1p_2+1-\tau } \text{ otherwise. } \end{array} \right. \end{aligned}$$

Uniform distribution on \(\{1,\ldots ,n\}\) (Example 2.3)

Fix \(n\ge 2\). For the uniform distribution on \(\{1,\ldots ,n\}\), solving the inequalities of Corollary 2.2 is equivalent to finding the unique index \(i\in \{1,\ldots ,n-1\}\) such that

$$\begin{aligned}{} & {} \frac{i(i-1)}{i(i-1)+(n-i)(n-i+1)} \le \tau \\{} & {} \quad < \frac{i(i+1)}{i(i+1)+(n-i)(n-i-1)}. \end{aligned}$$

This is equivalent to finding the unique solution (which we already know to exist, by Corollary 2.2) to the inequalities \(P_{\tau }(i+1)<0\le P_{\tau }(i)\) for \(i\in \{1,\ldots ,n-1\}\), where \(P_{\tau }\) is the polynomial

$$\begin{aligned} P_{\tau }(x)=(2\tau -1)x^2 - \{ 2\tau (n+1)-1 \} x + \tau n(n+1). \end{aligned}$$

This polynomial has discriminant \(4\tau (1-\tau )(n+1)(n-1)+1>0\) and then (for \(\tau \ne 1/2\)) two real roots \(x_{\tau ,-}\) and \(x_{\tau ,+}\) defined as

$$\begin{aligned} x_{\tau ,\pm } = \frac{2\tau (n+1)-1 \pm \sqrt{4\tau (1-\tau )(n+1)(n-1)+1}}{2(2\tau -1)}. \end{aligned}$$

A straightforward calculation yields \(P_{\tau }(1)=\tau n(n-1)>0\) and \(P_{\tau }(n)=-(1-\tau )n(n-1)<0\). It follows that when \(\tau >1/2\) (resp. \(\tau <1/2\)), only the lowest (resp. largest) of the two roots \(x_{\tau ,-}\) and \(x_{\tau ,+}\) belongs to the interval [1, n]. Conclude that, in both cases, \(P_{\tau }(i+1)<0\le P_{\tau }(i) \Leftrightarrow i\le x_{\tau ,-}<i+1 \Leftrightarrow i=\lfloor x_{\tau ,-} \rfloor \). With this index i,

$$\begin{aligned} \xi _{\tau }=\frac{\tau n(n+1) - (2\tau -1) i(i+1)}{2\tau n-2(2\tau -1)i}. \end{aligned}$$

Consequently

$$\begin{aligned} \xi _{\tau } = \left\{ \begin{array}{l} \dfrac{n(n+1)}{2} \text{ when } \tau =1/2, \\ \dfrac{\tau n(n+1) - (2\tau -1) \lfloor x_{\tau } \rfloor (\lfloor x_{\tau } \rfloor +1)}{2\tau n-2(2\tau -1)\lfloor x_{\tau } \rfloor } \text { otherwise, } \end{array} \right. \end{aligned}$$

with

$$\begin{aligned} x_{\tau }=\dfrac{2\tau (n+1)-1 - \sqrt{4\tau (1-\tau )(n+1)(n-1)+1}}{2(2\tau -1)}. \end{aligned}$$

Geometric distribution (Example 2.4)

For the geometric distribution with success probability \(p\in (0,1)\), namely, \(\mu (\{k\})=p(1-p)^{k-1}\) for any positive integer k, the inequalities of Theorem 2.1 read as

$$\begin{aligned} \frac{(1-p)^i - (1-pi)}{2(1-p)^i - (1-pi)}\le \tau <\frac{(1-p)^{i+1} - (1-p(i+1))}{2(1-p)^{i+1} - (1-p(i+1))}. \end{aligned}$$

Solving these inequalities is equivalent to finding the index \(i\ge 1\) such that \(h_{\tau }(i+1)<0\le h_{\tau }(i)\), where

$$\begin{aligned} h_{\tau }(x)=(2\tau -1)(1-p)^x - (1-\tau )(px-1). \end{aligned}$$

Straightforward calculations reveal that the unique root \(x_{\tau }\) of \(h_{\tau }\) over \([1,+\infty )\) satisfies the equation

$$\begin{aligned}{} & {} -\log (1-p)\left( x_{\tau }-\frac{1}{p} \right) \exp \left( -\log (1-p)\left( x_{\tau }-\frac{1}{p} \right) \right) \\{} & {} \quad =\frac{2\tau -1}{1-\tau } \left( -\frac{\log (1-p)}{p} (1-p)^{1/p} \right) . \end{aligned}$$

This is a transcendental equation (unless \(\tau \ne 1/2\), for which \(x_{1/2}=1/p\)). Nevertheless, since by construction

$$\begin{aligned} -\log (1-p)\left( x_{\tau }-\frac{1}{p} \right) \ge -\log (1-p)\left( 1-\frac{1}{p} \right) >-1 \end{aligned}$$

for any \(p\in (0,1)\), and

$$\begin{aligned}{} & {} \frac{2\tau -1}{1-\tau } \left( -\frac{\log (1-p)}{p} (1-p)^{1/p} \right) \\{} & {} \quad = \frac{2\tau -1}{1-\tau } \{ - (1-p)^{1/p} \log ((1-p)^{1/p}) \} > -e^{-1} \end{aligned}$$

for any \(\tau ,p\in (0,1)\), one may express \(x_{\tau }\) using the main branch of Lambert’s W function, that is

$$\begin{aligned} x_{\tau } = \frac{1}{p} - \frac{1}{\log (1-p)} W\left( -\frac{(1-p)^{1/p}\log (1-p)}{p} \frac{2\tau -1}{1-\tau } \right) \end{aligned}$$

where, for \(x>0\), W(x) is the unique (positive) solution to the equation \(w e^w=x\). The main branch of the Lambert function is available numerically in R using (for instance) the gsl package (Hankin et al. 2023), acting as a wrapper for the GNU Scientific Library.

Note further that when \(\tau >1/2\), the function \(h_{\tau }\) is obviously decreasing, so the inequalities \(h_{\tau }(i+1)<0\le h_{\tau }(i)\) are equivalent to \(i\le x_{\tau }<i+1\), i.e. \(i=\lfloor x_{\tau } \rfloor \). When \(\tau <1/2\), it is readily shown that \(h_{\tau }'\) is decreasing and \(h_{\tau }'(1)=-(1-2\tau ) (1-p)\log (1-p)-(1-\tau )p<0\) for any \(p\in (0,1)\), so again \(h_{\tau }\) is decreasing and \(h_{\tau }(i+1)<0\le h_{\tau }(i) \Leftrightarrow i=\lfloor x_{\tau } \rfloor \). Conclude that

$$\begin{aligned} \xi _{\tau }= & {} \frac{(2\tau -1)(1-p)^{\lfloor x_{\tau } \rfloor } (1+p\lfloor x_{\tau } \rfloor )+1-\tau }{p\{ (2\tau -1)(1-p)^{\lfloor x_{\tau } \rfloor }+1-\tau \}} \text{ with } \\ x_{\tau }= & {} \frac{1}{p} - \frac{1}{\log (1-p)} W\left( -\frac{(1-p)^{1/p}\log (1-p)}{p} \frac{2\tau -1}{1-\tau } \right) . \end{aligned}$$

Cardano and Ferrari formulae (relevant to Sect. 2.3.1)

To solve a real cubic polynomial equation of the form \(x^3+bx^2+cx+d=0\), Cardano’s method consists in letting \(p=c-b^2/3\) and \(q=d+b(2b^2-9c)/27\), and in computing the discriminant \(\Delta =-4p^3-27q^2\) of the so-called depressed cubic \(X^3+pX+q=0\). If \(\Delta \le 0\), then the unique real root of the equation is given by

$$\begin{aligned} x= \root 3 \of {\frac{-q+\sqrt{\frac{-\Delta }{27}}}{2}} + \root 3 \of {\frac{-q-\sqrt{\frac{-\Delta }{27}}}{2}} -\frac{b}{3}. \end{aligned}$$

If on the contrary \(\Delta >0\), then necessarily \(p<0\) and there are 3 real solutions, given by Viète’s formula:

$$\begin{aligned} x= 2 \sqrt{\frac{-p}{3}} \cos \left( \frac{1}{3} \arccos \left( \frac{3q}{2p} \sqrt{\frac{3}{-p}} \right) + \frac{2k \pi }{3} \right) - \frac{b}{3}, \end{aligned}$$

\(k \in \{ 0,1,2 \}\). To solve a real quartic polynomial equation \(x^4+bx^3+cx^2+dx+e=0\), Ferrari’s method first finds a root \(\lambda \) to the cubic equation \(8 \lambda ^3 -4c \lambda ^2+(2bd-8e) \lambda -b^2 e+4ce-d^2=0\). The four (possibly complex) solutions to the quartic equation are then

$$\begin{aligned}{} & {} \frac{\varepsilon _1 \sqrt{2 \lambda -c+\frac{b^2}{4}}}{2} - \frac{b}{4} \\{} & {} \quad + \frac{\varepsilon _2 \sqrt{-2 \lambda -c -\varepsilon _1 \left( \frac{2(d-b) \lambda }{\sqrt{2 \lambda -c+\frac{b^2}{4}}}+ b \sqrt{2 \lambda -c+\frac{b^2}{4}} \right) + \frac{b^2}{2}}}{2} \end{aligned}$$

where \(\varepsilon _1, \varepsilon _2 \in \{ -1,1 \}\).

Student distribution with \(\nu =4\) degrees of freedom (Example 2.13)

Consider the Student distribution with \(\nu \) degrees of freedom, having probability density function

$$\begin{aligned} f(x)=\frac{\Gamma ((\nu +1)/2)}{\sqrt{\nu \pi } \Gamma (\nu /2)} \left( 1 + \frac{x^2}{\nu } \right) ^{-(\nu +1)/2}, \ x\in {\mathbb {R}}. \end{aligned}$$

Here \(\Gamma \) is Euler’s Gamma function. When \(\nu =4\), the probability density function simplifies to

$$\begin{aligned} f(x)=\frac{3}{8} \left( 1+\frac{x^2}{4} \right) ^{-5/2}, \ x\in {\mathbb {R}}. \end{aligned}$$

The change of variables \(t=2\tan (\theta )\) combined with the trigonometric identities \(\cos (3\phi )=4\cos ^3(\phi )-3\cos (\phi )\), \(\sin (3\phi )=3\sin (\phi )-4\sin ^3(\phi )\) and \(\sin (\arctan \theta )=\theta /\sqrt{1+\theta ^2}\) then yield, after straightforward calculations, the following closed form for the survival function:

$$\begin{aligned} \overline{F}(x)=\int _x^{\infty } f(t) \, \textrm{d}t=\frac{1}{2} - \frac{x}{8} \frac{3+x^2/2}{(1+x^2/4)^{3/2}}. \end{aligned}$$

Further straightforward calculations based on the change of variables \(u=t^2\) then provide

$$\begin{aligned} \varphi (x)=\int _x^{\infty } \overline{F}(t)\,\textrm{d}t= \frac{1}{2} \left( \frac{x^2+2}{\sqrt{x^2+4}} - x \right) . \end{aligned}$$

Since the Student distribution is centered, \(m=0\) and Eq. (3) is

$$\begin{aligned} \xi _{\tau }^4 +4 \xi _{\tau }^2- \frac{(2 \tau -1)^2}{\tau (1-\tau )}=0. \end{aligned}$$

This is a biquadratic equation, leading to \( \xi _{\tau }^2 = -2+1/\sqrt{\tau (1-\tau )} \) because \(\xi _{\tau }^2\ge 0\), and then

$$\begin{aligned} \xi _{\tau }= {\text {sign}}(2 \tau -1) \sqrt{\frac{1}{\sqrt{\tau (1-\tau )}}-2}. \end{aligned}$$

In general, the distribution function and mean residual life function of the Student distribution involve the hypergeometric function. It is not hard to see that, while the distribution function and mean residual life function can in fact be written in closed form when \(\nu \) is an even integer, resulting in a polynomial equation characterizing \(\xi _{\tau }\), only the cases \(\nu \in \{2,4,6\}\) result in an equation of degree 4 or lower.

Fisher distribution with (4, 4) degrees of freedom (Example 2.14)

The Fisher distribution with degrees of freedom \(\nu _1>0\) and \(\nu _2>0\) has density function

$$\begin{aligned} f(x)=\frac{(\nu _1/\nu _2)^{\nu _1/2}}{B(\nu _1/2,\nu _2/2)} x^{\nu _1/2-1} (1+\nu _1 x/\nu _2)^{-(\nu _1+\nu _2)/2}, \end{aligned}$$

\(x>0\), where B is the Beta function. In the specific case \(\nu _1=\nu _2=4\), one finds \(\varphi (x)=(3x+2)/(x+1)^2\) for \(x>0\), and \(m=2\). Equation (3) is thus equivalent to the cubic equation

$$\begin{aligned} \xi _{\tau }^3-\frac{3 \tau }{1-\tau } \xi _{\tau }-\frac{2 \tau }{1-\tau }=0. \end{aligned}$$

The discriminant of this equation is \(\Delta =108 \tau ^2 (2 \tau -1)/(1-\tau )^3\). If \(\tau \le 1/2\), then \(\Delta \le 0\) and the unique solution is

$$\begin{aligned} \xi _{\tau }= \root 3 \of {\frac{\tau }{1-\tau }} \left( \root 3 \of {1+\sqrt{\frac{1-2\tau }{1-\tau }}} + \root 3 \of {1-\sqrt{\frac{1-2\tau }{1-\tau }}} \right) . \end{aligned}$$

If now \(\tau > 1/2\), then \(\Delta > 0\) and the 3 possible solutions are

$$\begin{aligned} \xi _{\tau }= 2 \sqrt{\frac{\tau }{1-\tau }} \cos \left( \frac{1}{3} \arccos \left( \sqrt{\frac{1-\tau }{\tau }} \right) + \frac{2 k \pi }{3} \right) , \end{aligned}$$

\(k \in \{ 0,1,2 \}\). Since \(\tau > 1/2\), \(\arccos ( \sqrt{(1-\tau )/\tau } )\in [0,\pi /2]\), and therefore (taking the constraint \(\xi _{\tau } \ge 0\) into account) \(k=0\) is the only admissible solution, namely

$$\begin{aligned} \xi _{\tau }= 2 \sqrt{\frac{\tau }{1-\tau }} \cos \left( \frac{1}{3} \arccos \left( \sqrt{\frac{1-\tau }{\tau }} \right) \right) . \end{aligned}$$
Table 3 Table of expectiles of the standard Gaussian distribution, computed via the Newton–Raphson algorithm (see Example 2.12)
Table 4 Table of expectiles of the standard log-normal distribution, computed via the Newton–Raphson algorithm

Pareto distribution with extreme value index 1/4 (Example 2.15)

The Pareto distribution with extreme value index \(\gamma >0\) has survival function \(\overline{F}(x)=x^{-1/\gamma }\) for \(x>1\). This distribution has a finite first moment when \(\gamma <1\), and since \(\varphi (x)=\gamma x^{1-1/\gamma }/(1-\gamma )\) for \(x>1\) and \(m=1/(1-\gamma )\), Eq. (3) leads to

$$\begin{aligned} (1-\gamma )(1-\tau ) \xi _{\tau }^{1/\gamma } - (1-\tau ) \xi _{\tau }^{1/\gamma -1} - \gamma (2\tau -1)=0. \end{aligned}$$

When \(\gamma =1/4\), this is the quartic equation \(\xi _{\tau }^4+b \xi _{\tau }^3 + c \xi _{\tau }^2 + d \xi _{\tau }+ e=0\), where \(b=-4/3\), \(c=0\), \(d=0\) and \(e=(1-2 \tau )/(3(1-\tau ))\). Ferrari’s method leads to finding \(\lambda =\lambda _{\tau }\) which is a root of the cubic equation

$$\begin{aligned} \lambda _{\tau }^3 - \frac{1-2 \tau }{3(1-\tau )} \lambda _{\tau } -\frac{2}{9} \frac{1-2 \tau }{3(1-\tau )}=0. \end{aligned}$$

The discriminant of this equation is

$$\begin{aligned} \Delta = -\frac{4}{27} \frac{(1-2 \tau )^2}{(1-\tau )^2} \frac{\tau }{1-\tau } \le 0 \end{aligned}$$

for all \(\tau \in (0,1)\), from which the unique solution of the equation involving \(\lambda \) is

$$\begin{aligned} \lambda _{\tau }= & {} \root 3 \of {\frac{1-2 \tau +|1-2 \tau | \sqrt{\frac{\tau }{1-\tau }}}{27 (1-\tau )}} \\{} & {} + \root 3 \of {\frac{1-2 \tau -|1-2 \tau | \sqrt{\frac{\tau }{1-\tau }}}{27 (1-\tau )}}. \end{aligned}$$

Ferrari’s method yields four possible solutions. The only real-valued solution greater than 1 is obtained with \(\varepsilon _1=\varepsilon _2=1\), leading to the solution

$$\begin{aligned} \xi _{\tau }= & {} \frac{1}{2} \sqrt{-2 \lambda _{\tau }-\frac{8}{3} \frac{\lambda _{\tau }}{\sqrt{2 \lambda _{\tau }+\frac{4}{9}}} + \frac{4}{3} \sqrt{2 \lambda _{\tau }+\frac{4}{9}} + \frac{8}{9}}\\{} & {} + \frac{1}{2} \sqrt{2 \lambda _{\tau }+\frac{4}{9}} + \frac{1}{3}. \end{aligned}$$
Table 5 Table of expectiles of the Student distribution with \(\nu \) degrees of freedom, computed via the Newton–Raphson algorithm
Fig. 10
figure 10

Quantiles (black) and expectiles (red) of the standard Gaussian distribution, as functions of \(\tau \in (0,1)\). (Color figure online)

Fig. 11
figure 11

Quantiles (black) and expectiles (red) of the log-normal distribution, as functions of \(\tau \in (0,1)\). (Color figure online)

Fig. 12
figure 12

Quantiles (black) and expectiles (red) of the Student distribution with 2 (solid curves), 4 (dashed curves) and 10 (dotted curves) degrees of freedom, as functions of \(\tau \in (0,1)\). The Student distribution with 2 degrees of freedom is Koenker’s distribution (Koenker 1993), for which quantiles and expectiles are identical. (Color figure online)

Fig. 13
figure 13

Quantiles (black) and expectiles (red) of the chi-squared distribution with 1 (dotted curves), 5 (dashed curves) and 10 (solid curves) degrees of freedom, as functions of \(\tau \in (0,1)\). (Color figure online)

Catalog of expectile functions of continuous distributions

This section provides a catalog of expectile functions of continuous distributions, obtained either via numerical means or, in exceptional cases, in closed or analytic form. Tables 345 and 6 give reference values for the expectiles of the standard Gaussian, log-normal, Student and chi-squared distributions, respectively. Figures 101112 and 13 provide graphical representations of the corresponding expectile functions over (0, 1). Table 7 gathers closed-form expressions for expectiles of certain bounded continuous distributions. Table 8 gives analytic-form expressions for expectiles of some unbounded continuous distributions. Tables 9 and 10 list closed-form expressions for expectiles of the Hall–Weiss distribution and the Pareto distribution, respectively, with particular parameters.

Table 6 Table of expectiles of the chi-squared distribution with \(\nu \) degrees of freedom, computed via the Newton–Raphson algorithm
Table 7 Closed-form expressions for expectiles in exceptional cases of certain bounded continuous distributions
Table 8 Analytic-form expressions for expectiles in exceptional cases of certain unbounded continuous distributions
Table 9 Closed-form expressions for expectiles in exceptional cases of the Hall–Weiss distribution with parameters \(\alpha >0\) and \(\beta \ge 0\), having distribution function \(F(x)=1-(x^{-\alpha } + x^{-\alpha -\beta })/2\), \(x>1\)
Table 10 Closed-form expressions for expectiles in exceptional cases of the Pareto distribution with extreme value index \(\gamma >0\), having distribution function \(F(x)=1-x^{-1/\gamma }\), \(x>1\)

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Daouia, A., Stupfler, G. & Usseglio-Carleve, A. An expectile computation cookbook. Stat Comput 34, 103 (2024). https://doi.org/10.1007/s11222-024-10403-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11222-024-10403-x

Keywords

Navigation