Skip to main content

Advertisement

Log in

Efficient parameter estimation and variable selection in partial linear varying coefficient quantile regression model with longitudinal data

  • Regular Article
  • Published:
Statistical Papers Aims and scope Submit manuscript

Abstract

Efficient estimation and variable selection in partial linear varying coefficient quantile regression model with longitudinal data is concerned in this paper. To improve estimation efficiency in quantile regression, based on B-spline basis approximation for nonparametric parts, we propose a new estimating function, which can incorporate the correlation structure between repeated measures. In order to reduce computational burdens, the induced smoothing method is used. The new method is empirically shown to be much more efficient and robust than the popular generalized estimating equations based methods. Under mild conditions, the asymptotically normal distribution of the estimators for the parametric components and the optimal convergence rate of the estimators for the nonparametric functions are established. Furthermore, to do variable selection, a smooth-threshold estimating equation is proposed, which can use the correlation structure and select the nonparametric and parametric parts simultaneously. Theoretically, the variable selection procedure works beautifully, including consistency in variable selection and oracle property in estimation. Simulation studies and real data analysis are included to show the finite sample performance.

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

Similar content being viewed by others

References

  • Brown B, Wang Y (2005) Standard errors and covariance matrices for smoothed rank estimators. Biometrika 92:149–158

    Article  MathSciNet  Google Scholar 

  • Fan J, Huang T (2005) Profile likelihood inferences on semiparametric varying coefficient partially linear models. Bernoulli 11:1031–1057

    Article  MathSciNet  Google Scholar 

  • Fan J, Li R (2001) Variable selection via nonconcave penalized likelihood and its oracle properties. J Am Stat Assoc 96:1348–1360

    Article  MathSciNet  Google Scholar 

  • Fu L, Wang Y (2012) Quantile regression for longitudinal data with a working correlation model. Comput Stat Data Anal 56:2526–2538

  • Fu L, Wang Y (2016) Efficient parameter estimation via Gaussian copulas for quantile regression with longitudinal data. J Multivar Anal 143:492–502

    Article  MathSciNet  Google Scholar 

  • Fu L, Wang Y, Zhu M (2016) A Gaussian pseudolikelihood approach for quantile regression with repeated measurements. Comput Stat Data Anal 84:41–53

    Article  MathSciNet  Google Scholar 

  • Huang J, Horowitz J, Wei F (2010) Variable selection in nonparametric additive models. Ann Stat 38:2282–2313

    Article  MathSciNet  Google Scholar 

  • Huang J, Breheny P, Ma S (2012) A selective review of group selection in high-dimensional models. Stat Sci 27:481–499

    Article  MathSciNet  Google Scholar 

  • Huang J, Wei F, Ma S (2012) Semiparametric regression pursuit. Stat Sin 22:1403–1426

    Google Scholar 

  • Jung S (1996) Quasi-likelihood for median regression models. J Am Stat Assoc 91:251–257

    Article  MathSciNet  Google Scholar 

  • Li R, Liang H (2008) Variable selection in semiparametric regression model. Ann Stat 36:261–286

    Article  Google Scholar 

  • Li J, Li Y, Zhang R (2017) B spline variable selection for the single index models. Stat Pap 58:691–706

    Article  MathSciNet  Google Scholar 

  • Ma S, Song Q, Wang L (2013) Simultaneous variable selection and estimation in semiparametric modeling of longitudinal/clustered data. Bernoulli 19:252–274

    Article  MathSciNet  Google Scholar 

  • Pollard D (1990) Empirical processes: theory and applications. In: NSF-CBMS regional conference series in probability and statistics. Institute of Mathematical Statistics, Hayward

  • Schumaker L (1981) Spline functions: basic theory. Wiley, New York

    MATH  Google Scholar 

  • Stone C (1982) Optimal global rates of convergence for nonparametric regression. Ann Stat 10:1040–1053

  • Stone C (1985) Additive regression and other nonparametric models. Ann Stat 13:689–705

  • Tang Y, Wang H, Zhu Z (2013) Variable selection in quantile varying coefficient models with longitudinal data. Comput Stat Data Anal 57:435–449

    Article  MathSciNet  Google Scholar 

  • Tibshirani RJ (1996) Regression shrinkage and selection via the lasso. J R Stat Soc B 58:267–288

    MathSciNet  MATH  Google Scholar 

  • Ueki M (2009) A note on automatic variable selection using smooth-threshold estimating equations. Biometrika 96:1005–1011

    Article  MathSciNet  Google Scholar 

  • Wang N (2003) Marginal nonparametric kernel regression accounting for within-subject correlation. Biometrika 90:43–52

    Article  MathSciNet  Google Scholar 

  • Wang Y, Carey V (2003) Working correlation structure misspecification, estimation and covariate design: implications for GEE performance. Biometrika 90:29–41

  • Wang H, Xia Y (2009) Shrinkage estimation of the varying coefficient model. J Am Stat Assoc 104:747–757

    Article  MathSciNet  Google Scholar 

  • Wang K, Lin L (2015) Variable selection in semiparametric quantile modeling for longitudinal data. Commun Stat Theory Methods 44:2243–2266

    Article  MathSciNet  Google Scholar 

  • Wang K, Lin L (2016) Robust and efficient estimator for simultaneous model structure identification and variable selection in generalized partial linear varying coefficient models with longitudinal data. Stat Pap. https://doi.org/10.1007/s00362-017-0890-z

  • Wang L, Li H, Huang J (2008) Variable selection in nonparametric varying coefficient models for analysis of repeated measurements. J Am Stat Assoc 103:1556–1569

  • Wang H, Zhu Z, Zhou J (2009) Quantile regression in partially linear varying coefficient models. Ann Stat 37:3841–3866

    Article  MathSciNet  Google Scholar 

  • Wang L, Xue L, Qu A, Liang H (2014) Estimation and model selection in generalized additive partial linear models for correlated data with diverging number of covariates. Ann Stat 42:592–624

    Article  MathSciNet  Google Scholar 

  • Wei F, Huang J, Li H (2011) Variable selection and estimation in high-dimensional varying-coefficient models. Stat Sin 21:1515–1540

    Article  MathSciNet  Google Scholar 

  • Xue L (2009) Variable selection in additive models. Stat Sin 19:1281–1296

    MathSciNet  MATH  Google Scholar 

  • Xue L, Qu A (2012) Variable selection in high-dimensional varying coefficient models with global optimality. J Mach Learn Res 13:1973–1998

    MathSciNet  MATH  Google Scholar 

  • Yang H, Guo C, Lv J (2016) Variable selection for generalized varying coefficient models with longitudinal data. Stat Pap 57:115–132

    Article  MathSciNet  Google Scholar 

  • Zhang R, Zhao W, Liu J (2013) Robust estimation and variable selection for semiparametric partially linear varying coefficient model based on modal regression. J Nonparametric Stat 25:523–544

    Article  MathSciNet  Google Scholar 

  • Zhao P, Xue L (2009) Variable selection in semiparametric regression analysis for longitudinal data. Ann Inst Stat Math 64:213–231

    Article  MathSciNet  Google Scholar 

  • Zhao P, Xue L (2010) Variable selection for semiparametric varying coefficient partially linear errors-in-variables models. J Multivar Anal 101:1872–1883

    Article  MathSciNet  Google Scholar 

  • Zhao W, Zhang R, Liu J, Lv Y (2014) Robust and efficient variable selection for semiparametric partially linear varying coefficient model based on modal regression. Ann Inst Stat Math 66:165–191

    Article  MathSciNet  Google Scholar 

  • Zhao W, Lian H, Liang H (2017) GEE analysis for longitudinal single-index quantile regression. J Stat Plan Inference 187:78–102

    Article  MathSciNet  Google Scholar 

  • Zhou Y, Liang H (2009) Statistical inference for semiparametric varying-coefficient partially linear models with error-prone linear covariates. Ann Stat 37:427–458

    Article  MathSciNet  Google Scholar 

  • Zhu Z, Fung W, He X (2008) On the asymptotics of marginal regression splines with longitudinal data. Biometrika 95:907–917

    Article  MathSciNet  Google Scholar 

  • Zou H (2006) The Adaptive LASSO and Its Oracle Properties. J Am Stat Assoc 101:1418–1429

    Article  MathSciNet  Google Scholar 

  • Zou H, Li R (2008) One-step sparse estimates in nonconcave penalized likelihood models. Ann Stat 36:1509–1533

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

Kangning Wang was supported by NSF Project (ZR2017BA002) of Shandong Province of China and NNSF Project (71673171) of China.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Xiaofei Sun.

Appendix

Appendix

Proof of Theorem 2.1. Let \(\bar{\varvec{U}}_{\tau }^o(\varvec{\zeta })=-\sum _{i=1}^n\varvec{D}_{i}^T \varvec{\Lambda }_i\varvec{R}_i^{-1} \varvec{P}_i(\varvec{\zeta })\) with \(\varvec{P}_i(\varvec{\zeta })=\left( \tau -\Pr (y_{i1}-\varvec{d}_{i1}^T \varvec{\zeta }<0),\ldots ,\tau -\Pr (y_{im_i}-\varvec{d}_{im_i}^T \varvec{\zeta }<0)\right) ^T\). Then we can get that

$$\begin{aligned}&\frac{1}{n}\left[ \bar{\varvec{U}}_{\tau }^o(\varvec{\zeta }) -{\varvec{U}}_{\tau }^o(\varvec{\zeta })\right] \\&\quad =\frac{1}{n}\sum _{i=1}^n\varvec{D}_{i}^T\varvec{\Lambda }_i \varvec{R}_i^{-1}\left[ \varvec{S}_i(\varvec{\zeta })-\varvec{P}_i(\varvec{\zeta })\right] \\&\quad =\frac{1}{n}\sum _{i=1}^n\varvec{D}_{i}^T\varvec{\Lambda }_i \varvec{R}_i^{-1}\begin{bmatrix} \Pr (y_{i1}-\varvec{d}_{i1}^T\varvec{\zeta }<0)-I(y_{i1} -\varvec{d}_{i1}^T\varvec{\zeta }<0) \\ \vdots \\ \Pr (y_{im_i}-\varvec{d}_{im_i}^T\varvec{\zeta }<0)-I(y_{im_i} -\varvec{d}_{im_i}^T\varvec{\zeta }<0) \end{bmatrix}\\&\quad =\frac{1}{n}\sum _{i=1}^n\sum _{j=1}^{m_i} \varvec{a}_{ij}\left[ \Pr (y_{ij}-\varvec{d}_{ij}^T\varvec{\zeta }<0) -I(y_{ij}-\varvec{d}_{ij}^T\varvec{\zeta }<0) \right] , \end{aligned}$$

where \(\varvec{a}_{ij}\) is a \((p(K_{n}+\hbar +1)+q)\times 1\) vector with \(\varvec{D}_{i}^T\varvec{\Lambda }_i \varvec{R}_i^{-1}=(\varvec{a}_{i1},\ldots ,\varvec{a}_{im_i})\). Under condition (C4) and from the law of large numbers (Pollard 1990), we have that

$$\begin{aligned} \sup _{\varvec{\zeta }}\left| \frac{1}{n}\sum _{i=1}^n\sum _{j=1}^{m_i}\varvec{a}_{ij}\left[ \Pr (y_{ij}-\varvec{d}_{ij}^T\varvec{\zeta }<0)-I(y_{ij}-\varvec{d}_{ij}^T\varvec{\zeta }<0)\right] \right| =o\left( \frac{1}{\sqrt{n}}\right) .\qquad \end{aligned}$$
(6.1)

Therefore, \(\sup _{\varvec{\zeta }}\left\| \frac{1}{n} \left[ \bar{\varvec{U}}_{\tau }^o(\varvec{\zeta })-{\varvec{U}}_{\tau }^{o} (\varvec{\zeta })\right] \right\| =o\left( \frac{1}{\sqrt{n}}\right) \). By condition (C3), we know that \(\varvec{\zeta }_0\) is the solution of \(\bar{\varvec{U}}_{\tau }^o(\varvec{\zeta })=\varvec{0}\). Due to \(\tilde{\varvec{\zeta }}^o\) is the solution of the equation \({\varvec{U}}_{\tau }^o(\varvec{\zeta })=\varvec{0}\) together with condition (C4), hence \(\tilde{\varvec{\zeta }}^o\rightarrow _{p}\varvec{\zeta }_0\) as \(n\rightarrow \infty \). For any \(\varvec{\zeta }\) satisfying \(\Vert \varvec{\zeta }-\varvec{\zeta }_0\Vert =O(n^{-1/3})\),

$$\begin{aligned}&{\varvec{U}}_{\tau }^o(\varvec{\zeta })-{\varvec{U}}_{\tau }^o(\varvec{\zeta }_0)\\&\quad =-\sum _{i=1}^n\varvec{D}_{i}^T\varvec{\Lambda }_i \varvec{R}_i^{-1}\left[ \varvec{S}_i(\varvec{\zeta })-\varvec{S}_i(\varvec{\zeta }_0)\right] \\&\quad =-\left\{ \sum _{i=1}^n\varvec{D}_{i}^T\varvec{\Lambda }_i \varvec{R}_i^{-1}\varvec{P}_i(\varvec{\zeta })+\sum _{i=1}^n\varvec{D}_{i}^T\varvec{\Lambda }_i \varvec{R}_i^{-1}\left[ \varvec{S}_i(\varvec{\zeta })-\varvec{S}_i(\varvec{\zeta }_0)-\varvec{P}_i(\varvec{\zeta })\right] \right\} \\&\quad =-\sum _{i=1}^n\varvec{D}_{i}^T\varvec{\Lambda }_i \varvec{R}_i^{-1}\varvec{P}_i(\varvec{\zeta })-\sum _{i=1}^n\sum _{j=1}^{m_i}\varvec{a}_{ij}\left[ \Pr (y_{ij} -\varvec{d}_{ij}^T\varvec{\zeta }<0)-I(y_{ij}-\varvec{d}_{ij}^T\varvec{\zeta }<0)\right. \\&\qquad +\left. I(y_{ij}-\varvec{d}_{ij}^T\varvec{\zeta }<0)-\tau \right] . \end{aligned}$$

According to Lemma 3 of Jung (1996), we have that

(6.2)

Therefore, \({\varvec{U}}_{\tau }^o(\varvec{\zeta })-{\varvec{U}}_{\tau }^o(\varvec{\zeta }_0) =\bar{\varvec{U}}_{\tau }^o(\varvec{\zeta })+o_{p}(\sqrt{n})\). By Taylor’s expansion of \(\bar{\varvec{U}}_{\tau }^o(\varvec{\zeta })\) together with \(\bar{\varvec{U}}_{\tau }^o(\varvec{\zeta }_0)=\varvec{0}\), we can obtain

$$\begin{aligned} {\varvec{U}}_{\tau }^o(\varvec{\zeta })-{\varvec{U}}_{\tau }^o (\varvec{\zeta }_0)={\varvec{D}}_{\tau }(\varvec{\zeta }_0)(\varvec{\zeta } -\varvec{\zeta }_0)+o_{p}(\sqrt{n}), \end{aligned}$$
(6.3)

where

$$\begin{aligned} {\varvec{D}}_{\tau }(\varvec{\zeta }_0)=\frac{\partial \bar{{\varvec{U}}}_{\tau }^o(\varvec{\zeta })}{\partial \varvec{\zeta }}|_{\varvec{\zeta }=\varvec{\zeta }_0}=\sum _{i=1}^n\varvec{D}_{i}^T\varvec{\Lambda }_i\varvec{R}_i^{-1}\varvec{\Lambda }_i\varvec{D}_{i}. \end{aligned}$$

Note that \(\tilde{\varvec{\zeta }}^o\) is in the \(n^{-1/3}\) neighborhood of \(\varvec{\zeta }_0\) and \({\varvec{U}}_{\tau }^o(\tilde{\varvec{\zeta }}^o)=\varvec{0}\), we have

$$\begin{aligned} \tilde{\varvec{\zeta }}^o-\varvec{\zeta }_0=-{\varvec{D}}_{\tau }^{-1}(\varvec{\zeta }_0){\varvec{U}}_{\tau }^o(\varvec{\zeta }_0)+o_{p}(\sqrt{n}). \end{aligned}$$
(6.4)

To obtain the closed form expression of \(\tilde{\varvec{\beta }}^o\), similar to Ma et al. (2013), we write the inverse of \({\varvec{D}}_{\tau }(\varvec{\zeta }_0)\) as the following block form

$$\begin{aligned} {\varvec{D}}_{\tau }^{-1}(\varvec{\zeta }_0)= & {} \begin{bmatrix} \sum _{i=1}^n\varvec{X}_i^T\varvec{\Lambda }_i\varvec{R}_i^{-1}\varvec{\Lambda }_i\varvec{X}_i&\quad \sum _{i=1}^n\varvec{X}_i^T\varvec{\Lambda }_i\varvec{R}_i^{-1}\varvec{\Lambda }_i\varvec{\Pi }_{i}\\ \sum _{i=1}^n\varvec{\Pi }_{i}^T\varvec{\Lambda }_i\varvec{R}_i^{-1}\varvec{\Lambda }_i\varvec{X}_i&\quad \sum _{i=1}^n\varvec{\Pi }_{i}^T\varvec{\Lambda }_i\varvec{R}_i^{-1}\varvec{\Lambda }_i \varvec{\Pi }_{i} \end{bmatrix}^{-1} \\= & {} \begin{bmatrix} \varvec{D}_{\varvec{X}\varvec{X}}&\quad \varvec{D}_{\varvec{X}\varvec{\Pi }} \\ \varvec{D}_{\varvec{\Pi }\varvec{X}}&\quad \varvec{D}_{\varvec{\Pi }\varvec{\Pi }} \end{bmatrix}^{-1}=\begin{bmatrix} \varvec{D}^{11}&\quad \varvec{D}^{12} \\ \varvec{D}^{21}&\quad \varvec{D}^{22} \end{bmatrix}, \end{aligned}$$

where \(\varvec{D}^{11}=\left( \varvec{D}_{\varvec{X}\varvec{X}}-\varvec{D}_{\varvec{X} \varvec{\Pi }}\varvec{D}^{-1}_{\varvec{\Pi }\varvec{\Pi }}\varvec{D}_{\varvec{\Pi }\varvec{X}}\right) ^{-1}\), \(\varvec{D}^{22}=\left( \varvec{D}_{\varvec{\Pi }\varvec{\Pi }} -\varvec{D}_{\varvec{\Pi }\varvec{X}}\varvec{D}^{-1}_{\varvec{X}\varvec{X}}\varvec{D}_{\varvec{X} \varvec{\Pi }}\right) ^{-1}\), \(\varvec{D}^{12}=-\varvec{D}^{11}\varvec{D}_{\varvec{X}\varvec{\Pi }} \varvec{D}_{\varvec{\Pi }\varvec{\Pi }}^{-1}\) and \(\varvec{D}^{21}=-\varvec{D}^{22}\varvec{D}_{\varvec{\Pi } \varvec{X}}\varvec{D}_{\varvec{X}\varvec{X}}^{-1}\). Furthermore, let

$$\begin{aligned} {\varvec{U}}_{\tau }^o(\varvec{\zeta }_0)= & {} ({\varvec{U}}_{\tau }^{o,1}(\varvec{\zeta }_0)^T,{\varvec{U}}_{\tau }^{o,2}(\varvec{\zeta }_0)^T)^T, \\ {\varvec{U}}_{\tau }^{o,1}(\varvec{\zeta }_0)= & {} -\sum _{i=1}^n\varvec{X}_i^T\varvec{\Lambda }_i\varvec{R}_i^{-1}\varvec{S}_i(\varvec{\zeta }_0), \\ {\varvec{U}}_{\tau }^{o,2}(\varvec{\zeta }_0)= & {} -\sum _{i=1}^n\varvec{\Pi }_i^T\varvec{\Lambda }_i\varvec{R}_i^{-1}\varvec{S}_i(\varvec{\zeta }_0). \end{aligned}$$

Then

$$\begin{aligned} \tilde{\varvec{\beta }}^o-\varvec{\beta }_0= & {} -\left[ \varvec{D}^{11} {\varvec{U}}_{\tau }^{o,1}(\varvec{\zeta }_0)+\varvec{D}^{12}{\varvec{U}}_{\tau }^{o,2} (\varvec{\zeta }_0)\right] +o_{p}\left( \frac{1}{\sqrt{n}}\right) \\= & {} \varvec{D}^{11}\sum _{i=1}^n\left[ \varvec{X}_i -\varvec{\Pi }_{i}\varvec{D}_{\varvec{\Pi }\varvec{\Pi }}^{-1}\varvec{D}_{\varvec{\Pi } \varvec{X}}\right] ^T\varvec{\Lambda }_i \varvec{R}_i^{-1}\varvec{S}_i(\varvec{\zeta }_0)+o_{p}\left( \frac{1}{\sqrt{n}}\right) . \end{aligned}$$

Thus,

$$\begin{aligned} \sqrt{n}(\tilde{\varvec{\beta }}^o-\varvec{\beta }_0) =(n\varvec{D}^{11})\frac{1}{\sqrt{n}}\sum _{i=1}^n\left[ \varvec{X}_i-\varvec{\Pi }_{i} \varvec{D}_{\varvec{\Pi }\varvec{\Pi }}^{-1}\varvec{D}_{\varvec{\Pi }\varvec{X}}\right] ^T\varvec{\Lambda }_i \varvec{R}_i^{-1}\varvec{S}_i(\varvec{\zeta }_0)+o_{p}\left( 1\right) . \end{aligned}$$

Because \(\varvec{S}_i(\varvec{\zeta }_0)\) are independent random variables with mean zero, and

$$\begin{aligned} \text{ var }\left( \frac{1}{\sqrt{n}}\sum _{i=1}^n\left[ \varvec{X}_i-\varvec{\Pi }_{i} \varvec{D}_{\varvec{\Pi }\varvec{\Pi }}^{-1}\varvec{D}_{\varvec{\Pi }\varvec{X}}\right] ^T\varvec{\Lambda }_i \varvec{R}_i^{-1}\varvec{S}_i(\varvec{\zeta }_0)\right) =\varvec{\Xi }. \end{aligned}$$

The multivariate central limit theorem implies that

$$\begin{aligned} \frac{1}{\sqrt{n}}\sum _{i=1}^n\left[ \varvec{X}_i-\varvec{\Pi }_{i} \varvec{D}_{\varvec{\Pi }\varvec{\Pi }}^{-1}\varvec{D}_{\varvec{\Pi }\varvec{X}}\right] ^T\varvec{\Lambda }_i \varvec{R}_i^{-1}\varvec{S}_i(\varvec{\zeta }_0)\rightarrow _{d} N(\varvec{0},\varvec{\Xi }), \end{aligned}$$

furthermore,

$$\begin{aligned} n\varvec{D}^{11}=\left[ \frac{1}{n}\left( \varvec{D}_{\varvec{X} \varvec{X}}-\varvec{D}_{\varvec{X}\varvec{\Pi }}\varvec{D}_{\varvec{X} \varvec{\Pi }}^{-1}\varvec{D}_{\varvec{\Pi }\varvec{X}}\right) \right] ^{-1} \rightarrow _{p}\varvec{\Sigma }, \end{aligned}$$

by the law of large numbers. Then, by using Slutskys theorem, it follows that \(\sqrt{n}(\tilde{\varvec{\beta }}^o-\varvec{\beta }_0)\rightarrow _{d} N(\varvec{0}, \varvec{\Sigma }^{-1}\varvec{\Xi }\varvec{\Sigma }^{-1})\). This complete the proof of part (a). Furthermore, by using the same arguments of proving the part (a), we can get

$$\begin{aligned} \frac{1}{n}\sum _{l=1}^p\sum _{i=1}^n\sum _{j=1}^{m_i} \left[ \varvec{\Pi }_{ij}^T\left( \hat{\varvec{\theta }}_l -\varvec{\theta }_{0l}\right) \right] = O_{p}\left( \frac{K_n}{n}\right) . \end{aligned}$$

The triangular inequality implies that

$$\begin{aligned}&\frac{1}{n}\sum _{l=1}^p\sum _{i=1}^n\sum _{j=1}^{m_i}(\tilde{\alpha }_{l}^o(t_{ij})-\alpha _{0l}(t_{ij}))^2\\&\quad \le \frac{2}{n}\sum _{l=1}^p\sum _{i=1}^n\sum _{j=1}^{m_i}\left[ \varvec{\Pi }_{ij}^T\left( \hat{\varvec{\theta }}_l-\varvec{\theta }_{0l}\right) \right] ^2+ \frac{2}{n}\sum _{l=1}^p\sum _{i=1}^n\sum _{j=1}^{m_i}\left[ \varvec{\Pi }_{ij}^T\varvec{\theta }_{0l}-\alpha _{0l}(t_{ij})\right] ^2\\&\quad =O_{p}\left( \frac{K_n}{n}+K_n^{-2r}\right) . \end{aligned}$$

The proof is completed.

Proof of Theorem 2.2

$$\begin{aligned} \varvec{\varsigma }\left( \varvec{\zeta }\right) =\left( \begin{aligned}&\varvec{\Xi }^{-1/2}\varvec{\Sigma }(\varvec{\beta }-\varvec{\beta }_{0}) \\&K_{n}^{-1/2}\varvec{H}_{n}(\varvec{\Theta }-\varvec{\Theta }_0) +K_{n}^{1/2}\varvec{H}_{n}^{-1}\varvec{\Pi }^T\varvec{\Lambda } \varvec{Z}(\varvec{\beta }-\varvec{\beta }_{0}) \end{aligned} \right) , \end{aligned}$$
(6.5)

where \(\varvec{H}_{n}^2=K_{n}\varvec{\Pi }^T\varvec{\Lambda }\varvec{\Pi }\). Furthermore, we standardize \(\widetilde{\varvec{z}}_{ij}=\varvec{\Xi }^{1/2}\varvec{\Sigma }^{-1}\varvec{z}_{ij}\) and \(\widetilde{\varvec{\Pi }}_{ij}=K_{n}^{1/2}\varvec{H}_{n}^{-1}\varvec{\Pi }_{ij}\). Note that \(\tilde{{S}}_{ij}({\varvec{\zeta }})-{{S}}_{ij}({\varvec{\zeta }}) =I(-\Delta _{ij})\Phi (-|\Delta _{ij}|)\), where \(\Delta _{ij}=(\epsilon _{ij}+u_{ij})/r_{ij}\) with \(u_{ij}=-(\varvec{\varsigma }\left( \varvec{\zeta }\right) ^T (\widetilde{\varvec{z}}_{ij}^T,\widetilde{\varvec{\Pi }}_{ij}^T)^T+R_{nij})\) and \(R_{nij}=\varvec{\Pi }_{ij}^T\varvec{\Theta }_0-\sum _{l=1}^px_{ij}^l \alpha _{0l}(t_{ij})\). We can obtain

$$\begin{aligned} \frac{1}{\sqrt{n}}\left[ {\varvec{U}}_{\tau }^o(\varvec{\zeta }) -\tilde{\varvec{U}}_{\tau }^o(\varvec{\zeta })\right]= & {} \frac{1}{\sqrt{n}} \sum _{i=1}^n\varvec{D}_{i}^T\varvec{\Lambda }_i \varvec{R}_i^{-1}\begin{bmatrix} I(-\Delta _{i1})\Phi (-|\Delta _{i1}|) \\ \vdots \\ I(-\Delta _{im_i})\Phi (-|\Delta _{im_i}|) \end{bmatrix}\\= & {} \frac{1}{\sqrt{n}}\sum _{i=1}^n \sum _{j=1}^{m_i}\varvec{a}_{ij}I(-\Delta _{ij})\Phi (-|\Delta _{ij}|). \end{aligned}$$

Note that

$$\begin{aligned}&E\left( \tilde{{S}}_{ij}({\varvec{\zeta }})-{{S}}_{ij}({\varvec{\zeta }})\right) \\&\quad =\int _{-\infty }^{\infty }I(-\Delta _{ij})\Phi (-|\Delta _{ij}|)f_{ij}(\epsilon )d\epsilon \\&\quad =\int _{-\infty }^{\infty }\Phi (-|\epsilon _{ij}+u_{ij}|/r_{ij})\{2I(\epsilon _{ij}+u_{ij}<0)-1\}f_{ij}(\epsilon )d\epsilon \\&\quad =r_{ij}\int _{-\infty }^{\infty }\Phi (-|t|)\{2I(t<0)-1\}\left[ f_{ij}(0)+f_{ij}^T(\eta (t))(r_{ij}t-u_{ij})\right] dt, \end{aligned}$$

where \(\eta (t)\) is between 0 and \(r_{ij}t-u_{ij}\). Because \(\int _{-\infty }^{\infty }\Phi (-|t|)\{2I(t<0)-1\}dt=0\), and by condition (C3), there exists a constant C such that \(\sup _{i,j}|f_{ij}^T(\eta (t))|\le C\). Then note that \(\int _{-\infty }^{\infty }\Phi (-|t|)|t|dt=1/2\), we have that

$$\begin{aligned} \left| E\left( \tilde{{S}}_{ij}({\varvec{\zeta }})-{{S}}_{ij}({\varvec{\zeta }})\right) \right| \le r_{ij}^2\int _{-\infty }^{\infty }\Phi (-|t|)|t||f_{ij}^T(\eta (t))|dt\le \frac{Cr_{ij}^2}{2}. \end{aligned}$$

Under conditions (C4) and (C5), as \(n\rightarrow \infty \), we can obtain

$$\begin{aligned} \left\| \frac{1}{\sqrt{n}}\left[ {\varvec{U}}_{\tau }^o (\varvec{\zeta })-\tilde{\varvec{U}}_{\tau }^o(\varvec{\zeta })\right] \right\| =\frac{1}{\sqrt{n}}\sup _{ij}\Vert \varvec{a}_{ij}\Vert \sum _{i=1}^n\frac{Cr_{ij}^2}{2}=o(1). \end{aligned}$$

In addition, by Cauchy–Schwartz inequality,

$$\begin{aligned}&\frac{1}{{n}}\text{ var }\left[ {\varvec{U}}_{\tau }^o(\varvec{\zeta })-\tilde{\varvec{U}}_{\tau }^o(\varvec{\zeta })\right] \\&\quad =\frac{1}{{n}}\sum _{i=1}^n\text{ var }\left[ \sum _{j=1}^{m_i}\varvec{a}_{ij}I(-\Delta _{ij})\Phi (-|\Delta _{ij}|)\right] \\&\quad \le \frac{1}{{n}}\sum _{i=1}^n\sum _{j=1}^{m_i}\varvec{a}_{ij}^T\varvec{a}_{ij}\text{ var }\left[ \tilde{{S}}_{ij}({\varvec{\zeta }})-{{S}}_{ij}({\varvec{\zeta }})\right] \\&\qquad +\, \frac{1}{{n}}\sum _{i=1}^n\sum _{j=1}^{m_i}\sum _{j^*=1,j^*\ne j}^{m_i}\varvec{a}_{ij}^T\varvec{a}_{ij^*}\left\{ \text{ var }\left[ \tilde{{S}}_{ij} ({\varvec{\zeta }})-{{S}}_{ij}({\varvec{\zeta }})\right] \text{ var }\left[ \tilde{{S}}_{ij^*} ({\varvec{\zeta }})-{{S}}_{ij^*}({\varvec{\zeta }})\right] \right\} ^{1/2}. \end{aligned}$$

For \(j=1,\ldots ,m_i\),

$$\begin{aligned} \text{ var }\left[ \tilde{{S}}_{ij}({\varvec{\zeta }})-{{S}}_{ij}({\varvec{\zeta }})\right]\le & {} E\left[ \tilde{{S}}_{ij}({\varvec{\zeta }})-{{S}}_{ij}({\varvec{\zeta }})\right] ^2 \\= & {} \int _{-\infty }^{\infty }\left\{ I(-\Delta _{ij})\Phi (-|\Delta _{ij}|)\right\} ^2f_{ij}(\epsilon )d\epsilon \\= & {} r_{ij}\int _{-\infty }^{\infty }\Phi ^2(-|t|)f_{ij}(r_{ij}t-u_{ij})dt \\= & {} r_{ij}\int _{|t|>c}\Phi ^2(-|t|)f_{ij}(r_{ij}t-u_{ij})dt \\&+\, r_{ij}\int _{|t|\le c}\Phi ^2(-|t|)f_{ij}(r_{ij}t-u_{ij})dt \\\le & {} \Phi ^2(-c)+r_{ij}cf_{ij}(\eta ), \end{aligned}$$

where \(\eta \) is a positive value, and \(\eta \) lies between \((-r_{ij}c-u_{ij},r_{ij}c-u_{ij})\). Let \(c=n^{1/3}\), under condition (C5), since \(r_{ij}=O(n^{-1/2})\), then \(r_{ij}c=O(n^{-1/6})\). Note that \(\Phi ^2(-c)\rightarrow 0\) and \(r_{ij}cf_{ij}(\eta )\rightarrow 0\), as \(n\rightarrow \infty \). By conditions (C4) and (C6), it is easy to obtain \(\frac{1}{{n}}\text{ var }\left[ {\varvec{U}}_{\tau }^o (\varvec{\zeta })-\tilde{\varvec{U}}_{\tau }^o(\varvec{\zeta })\right] =o(1)\). Therefore, we have \(\frac{1}{\sqrt{n}}\left[ {\varvec{U}}_{\tau }^o (\varvec{\zeta })-\tilde{\varvec{U}}_{\tau }^o(\varvec{\zeta })\right] \rightarrow 0\) as \(n\rightarrow \infty \) for any \(\varvec{\zeta }\). The proof is completed.

Proof of Theorem 2.3

Note that \(\sup _{\varvec{\zeta }}\Vert \frac{1}{n}\left( \bar{{\varvec{U}}}_{\tau }^o (\varvec{\zeta })-{\varvec{U}}_{\tau }^o(\varvec{\zeta })\right) \Vert =o \left( \frac{1}{\sqrt{n}}\right) \), and by Theorem 2.2, we know that \(\sup _{\varvec{\zeta }}\Vert \frac{1}{n}\left( \bar{{\varvec{U}}}_{\tau }^o (\varvec{\zeta })-\tilde{{\varvec{U}}}_{\tau }^o(\varvec{\zeta })\right) \Vert =o \left( \frac{1}{\sqrt{n}}\right) \). Because that \(\varvec{\zeta }_0\) is the unique solution of equation of \(\bar{{\varvec{U}}}_{\tau }^o(\varvec{\zeta })=\varvec{0}\). This together with the definition of \(\tilde{\varvec{\zeta }}\) implies that \(\tilde{\varvec{\zeta }}\rightarrow _{p}\varvec{\zeta }_0\) as \(n\rightarrow \infty \). In order to prove the asymptotic normality of \(\tilde{\varvec{\zeta }}\), we first prove that \(\frac{1}{n}\{\tilde{\varvec{D}}_{\tau }(\varvec{\zeta }_0) -{\varvec{D}}_{\tau }(\varvec{\zeta }_0)\}\rightarrow _{p} 0\). Note that \(E\left[ \tilde{\varvec{D}}_{\tau }(\varvec{\zeta }_0)\right] -{\varvec{D}}_{\tau }(\varvec{\zeta }_0)=(D_{ij})_{i,j=1,\ldots ,n}\), with \(D_{ij}=\frac{1}{r_{ij}}E\phi \left( \frac{\epsilon _{ij} +u_{ij}}{r_{ij}}\right) -f_{ij}(0)\). Because that

$$\begin{aligned}&\left| \frac{1}{r_{ij}}E\phi \left( \frac{\epsilon _{ij}+u_{ij}}{r_{ij}}\right) -f_{ij}(0)\right| \\&\quad =\left| \frac{1}{r_{ij}}\int _{-\infty }^{+\infty }\phi \left( \frac{\epsilon +u_{ij}}{r_{ij}}\right) f_{ij}(\epsilon )d\epsilon -f_{ij}(0)\right| \\&\quad =\left| \int _{-\infty }^{+\infty }\phi (t)\left\{ f_{ij}(0)+(r_{ij}t-u_{ij})f_{ij}^T(\eta (t))\right\} dt-f_{ij}(0)\right| \\&\quad =\left| \int _{-\infty }^{+\infty }\phi (t)(r_{ij}t-u_{ij})f_{ij}^T(\eta (t))dt\right| \\&\quad \le r_{ij}\int _{-\infty }^{+\infty }\left| \phi (t)tf_{ij}^T(\eta (t))\right| dt+\int _{-\infty }^{+\infty }\left| \phi (t)u_{ij}f_{ij}^T(\eta (t))\right| dt, \end{aligned}$$

where \(\eta (t)\) lies between 0 and \(r_{ij}t-u_{ij}\). By condition (C3), \(f_{ij}^T(\cdot )\) is uniformly bounded, hence there exists a constant C satisfying \(|f_{ij}^T(\eta (t))|\le C\), and by condition (C5), we have \(|\frac{1}{r_{ij}}E\phi (\frac{\epsilon _{ij}+u_{ij}}{r_{ij}})-f_{ij}(0)|\rightarrow 0\). By the strong law of large number, we know that \(\frac{1}{n}\tilde{\varvec{D}}_{\tau }(\varvec{\zeta }_0)\rightarrow E[\frac{1}{n}\tilde{\varvec{D}}_{\tau }(\varvec{\zeta }_0)]\). Using the triangle inequality, we have

$$\begin{aligned}&\left| \frac{1}{n}\left\{ \tilde{\varvec{D}}_{\tau }(\varvec{\zeta }_0)-{\varvec{D}}_{\tau }(\varvec{\zeta }_0)\right\} \right| \\&\quad \le \left| \frac{1}{n}\left\{ \tilde{\varvec{D}}_{\tau }(\varvec{\zeta }_0)-E\left[ \tilde{\varvec{D}}_{\tau }(\varvec{\zeta }_0)\right] \right\} \right| + \left| \frac{1}{n}\left\{ E\left[ \tilde{\varvec{D}}_{\tau }(\varvec{\zeta }_0)\right] -{\varvec{D}}_{\tau }(\varvec{\zeta }_0)\right\} \right| \\&\quad =o(1), \end{aligned}$$

which implies that \(\frac{1}{n}\left\{ \tilde{\varvec{D}}_{\tau }(\varvec{\zeta }_0) -{\varvec{D}}_{\tau }(\varvec{\zeta }_0)\right\} \rightarrow _{p}0\). By Taylor series expansion of \(\tilde{{\varvec{U}}}_{\tau }^o(\varvec{\zeta })\) around \(\varvec{\zeta }_0\), we have \(\tilde{{\varvec{U}}}_{\tau }^o(\varvec{\zeta })=\tilde{{\varvec{U}}}_{\tau }^o (\varvec{\zeta }_0)+\tilde{\varvec{D}}_{\tau }(\varvec{\zeta }^*)(\varvec{\zeta }-\varvec{\zeta }_0)\), where \(\varvec{\zeta }^*\) lies between \(\varvec{\zeta }\) and \(\varvec{\zeta }_0\). Because \(\tilde{{\varvec{U}}}_{\tau }^o(\tilde{\varvec{\zeta }})=\varvec{0}\) and \(\tilde{\varvec{\zeta }}\rightarrow \varvec{\zeta }_0\), we therefore obtain \(\varvec{\zeta }^*\rightarrow \varvec{\zeta }_0\) and \(\tilde{\varvec{D}}_{\tau }(\varvec{\zeta }^*)\rightarrow \tilde{\varvec{D}}_{\tau }(\varvec{\zeta }_0)\). Then by the same arguments used in the proof of Theorem 2.2, we can complete the proof of Theorem 2.3.

Proof (a) of Theorem 3.2. Let \(\delta _n=n^{-r/(2r+1)}\), \(\varvec{\beta }=\varvec{\beta }_0+\delta _n\varvec{T}_1\), \(\varvec{\Theta }=\varvec{\Theta }_0+\delta _n\varvec{T}_2\) and \(\varvec{T}=(\varvec{T}_1^T,\varvec{T}_2^T)^T\). Let \(\varvec{S}_n(\varvec{\beta },\varvec{\Theta })=\left( \varvec{I}_{p(K_{n}+\hbar +1)+q}-\hat{\varvec{\Psi }}\right) \tilde{\varvec{U}}_{\tau }^o(\varvec{\zeta })+\hat{\varvec{\Psi }}\varvec{\zeta }\). Our aim is to show that for \(\varepsilon >0\), there exists a constant \(C>0\), such that

$$\begin{aligned} \Pr \left( \sup _{\Vert \varvec{T}\Vert =C}\delta _n\varvec{T}^{T} \varvec{S}_n(\varvec{\beta }_0+\varvec{T}_1,\varvec{\Theta }_0+\varvec{T}_2)>0\right) \ge 1-\varepsilon , \end{aligned}$$

for n large enough. This will imply with probability at least \(1-\varepsilon \) that there exists a local minimum value of the equation \(\varvec{S}_n(\varvec{\beta },\varvec{\Theta })=\varvec{0}\) such that \(\Vert \left( \hat{\varvec{\beta }}^T,\hat{\varvec{\Theta }}^T\right) ^T -\left( \varvec{\beta }_0^T,\varvec{\Theta }_0^T\right) ^T\Vert =O_{p}(\delta _n)\). We will evaluate the sign of \(\delta _n\varvec{T}^T\varvec{S}_n(\varvec{\beta }_0+\varvec{T}_1,\varvec{\Theta }_0+\varvec{T}_2)\) in the ball \(\{\varvec{\beta }_0+\varvec{T}_1,\varvec{\Theta }_0+\varvec{T}_2:\Vert \varvec{T}\Vert =C\}\). By the Taylor approximation, we have that

$$\begin{aligned} \delta _n\varvec{T}^T\varvec{S}_n(\varvec{\beta }_0+\varvec{T}_1, \varvec{\Theta }_0+\varvec{T}_2)= & {} \delta _n\varvec{T}^T \varvec{S}_n(\varvec{\beta }_0,\varvec{\Theta }_0)+ \delta _n^2\varvec{T}^T \frac{\partial \varvec{S}_n(\tilde{\varvec{\beta }}, \tilde{\varvec{\Theta }})}{\partial \left( \varvec{\beta }^T, \varvec{\Theta }^T\right) ^T}\varvec{T}\nonumber \\= & {} I_{n1}+I_{n2}, \end{aligned}$$
(6.6)

where \((\tilde{\varvec{\beta }}^T,\tilde{\varvec{\Theta }}^T)^T\) lies between \(\left( \varvec{\beta }_0^T,\varvec{\Theta }_0^T\right) ^T\) and \(\left( \varvec{\beta }_0^T,\varvec{\Theta }_0^T\right) ^T+\delta _n\varvec{T}\). Next we will consider \(I_{n1}\) and \(I_{n2}\) respectively. For \(I_{n1}\), by some elementary calculations, we have

$$\begin{aligned} I_{n1}= & {} \delta _n\varvec{T}^T\left( \varvec{I}_{p(K_{n}+\hbar +1)+q} -\hat{\varvec{\Psi }}\right) \tilde{\varvec{U}}_{\tau }^o (\varvec{\zeta }_0)+\delta _n\varvec{T}^T\hat{\varvec{\Psi }}\varvec{\zeta }_0\\= & {} I_{n11}+I_{n12}. \end{aligned}$$

By Cauchy–Schwarz inequality, we can derive that

$$\begin{aligned} |I_{n11}|\le & {} \delta _n\left\| \varvec{T}^T\left( \varvec{I}_{p(K_{n}+\hbar +1)+q}-\hat{\varvec{\Psi }}\right) \right\| \left\| \tilde{\varvec{U}}_{\tau }^o(\varvec{\zeta }_0)\right\| \\\le & {} \delta _n\left( 1-\min \left[ \min \{\hat{\delta }_{1,k}:\hat{\delta }_{1,k}\ne 1\},\min \{\hat{\delta }_{2,k}:\hat{\delta }_{2,k}\ne 1 \}\right] \right) \Vert \varvec{T}\Vert \left\| \tilde{\varvec{U}}_{\tau }^o(\varvec{\zeta }_0)\right\| . \end{aligned}$$

Since \(\min \{\hat{\delta }_{1,k}:\hat{\delta }_{1,k}\ne 1\} \le \min \{\hat{\delta }_{1,k}:k=1,\ldots ,v\}\) and \(\min \{\hat{\delta }_{2,k}:\hat{\delta }_{2,k}\ne 1\}\le \min \{\hat{\delta }_{2,k}:k=1,\ldots ,c\}\). We only need to obtain the convergence rate of \(\min \{\hat{\delta }_{1,k}:k=1,\ldots ,v\}\) and \(\min \{\hat{\delta }_{2,k}:k=1,\ldots ,c\}\). By Theorem 2.3, we know that the initial estimator \((\tilde{\varvec{\beta }},\tilde{\varvec{\Theta }})\) satisfy \(\Vert (\tilde{\varvec{\beta }},\tilde{\varvec{\Theta }})-(\varvec{\beta }_0,\varvec{\Theta }_0) \Vert =O_{p}(n^{-r/(2r+1)})\). By using the condition \(n^{r/(2r+1)}\lambda _{\max }\rightarrow 0\), for any \(\varepsilon >0\) and \(k\in \{1,\ldots ,v\}\), we can derive that

$$\begin{aligned} \Pr \left( \hat{\delta }_{1,k}>n^{\frac{-r}{1+2r}}\varepsilon \right)= & {} \Pr \left( \frac{\lambda _{\max }}{|\tilde{\varvec{\theta }}_k|^{1+\tau }}>n^{\frac{-r}{1+2r}}\varepsilon \right) \\= & {} \Pr \left( (\lambda _{\max } n^{\frac{r}{1+2r}}/\varepsilon )^{1/(1+\tau )}>|\tilde{\varvec{\theta }}_k|\right) \\\le & {} \Pr \left( (\lambda _{\max } n^{\frac{r}{1+2r}}/\varepsilon )^{1/(1+\tau )} >\min _{k\in \{1,\ldots ,v\}}\Vert \tilde{\varvec{\theta }}_{0k}\Vert -O_{p}(n^{-r/(2r+1)})\right) \\&\rightarrow 0, \end{aligned}$$

which implies that for each \(k\in \{1,\ldots ,v\}\), \(\hat{\delta }_{1,k}=o_{p}(n^{-r/(2r+1)})\). Therefore, we can get that \(\min _{k=1,\ldots ,v}\hat{\delta }_{1,k}=o_{p}(n^{-r/(2r+1)})\). Similarly, we can prove that \(\hat{\delta }_{2,k}=o_{p}(n^{-r/(2r+1)})\), for each \(k\in \{1,\ldots ,c\}\). Therefore, we have that

$$\begin{aligned} \min \left[ \min _{k\in \mathcal {A}_1}\hat{\delta }_{1,k},\min _{k\in \mathcal {A}_2}\hat{\delta }_{2,k}\right] =o_{p}(n^{-r/(2r+1)}). \end{aligned}$$

Thus, we can obtain that

$$\begin{aligned} |I_{n11}|=O_{p}(\sqrt{n}\delta _n)\Vert \varvec{T}\Vert -o_{p}(\sqrt{n} \delta _n^2)\Vert \varvec{T}\Vert . \end{aligned}$$

Furthermore, for the \(I_{n12}\), we have \(I_{n12}\le \delta _n\Vert \varvec{T}\Vert \Vert \varvec{\zeta }_0\Vert =O_{p}(\delta _n\Vert \varvec{T}\Vert )\). Therefore, \(I_{n1}=O_{p}(\sqrt{n}\delta _n^2\Vert \varvec{T}\Vert )\). For \(I_{n2}\), we can obtain that

$$\begin{aligned} I_{n2}= & {} \delta _n^2\varvec{T}^T\frac{\partial \varvec{S}_n(\tilde{\varvec{\beta }},\tilde{\varvec{\Theta }})}{\partial \left( \varvec{\beta }^T,\varvec{\Theta }^T\right) ^T}\varvec{T}\\= & {} n\delta _n^2\varvec{T}^T\left( \varvec{I}_{p(K_{n}+\hbar +1)+q}-\hat{\varvec{\Psi }}\right) \left\{ \frac{1}{\sqrt{n}}\sum _{i=1}^n\varvec{D}_{i}^T\varvec{\Lambda }_i \varvec{R}_i^{-1}\varvec{\Lambda }_i\varvec{D}_{i}\right\} \varvec{T}\\&+\,\delta _n^2\varvec{T}^T\hat{\varvec{\Psi }}\varvec{T}+o_{p}(n\delta _n^2)\\= & {} I_{n21}+I_{n22}+o_{p}(n\delta _n^2). \end{aligned}$$

With the same argument, it is easy to prove that \(I_{n22}=O_{p}(\delta _n^2)\Vert \varvec{T}\Vert =o(n\delta _n^2\Vert \varvec{T}\Vert )\). Thus, \(\delta _n\varvec{T}^T\varvec{S}_n(\varvec{\beta }_0+\varvec{T}_1,\varvec{\Theta }_0+\varvec{T}_2)\) is asymptotically dominated in probability by \(I_{n21}\) on \(\{\varvec{\beta }_0+\varvec{T}_1,\varvec{\Theta }_0+\varvec{T}_2:\Vert \varvec{T}\Vert =C\}\), which is positive for the sufficiently large C. This implies, with probability at least \(1-\varepsilon \), that there exists a local minimizer \(\left( \hat{\varvec{\beta }}^T,\hat{\varvec{\Theta }}^T\right) ^T\) such that \(\Vert \left( \hat{\varvec{\beta }}^T,\hat{\varvec{\Theta }}^T\right) ^T -\left( \varvec{\beta }_0^T,\varvec{\Theta }_0^T\right) ^T\Vert =O_{p}(\delta _n)\). Then by the same arguments used in the proof of Theorem 2.1, the proof can be completed.

Proof of Theorem 3.1

For any given \(k\in \{v+1,\ldots ,p\}\), we have \(\Vert \tilde{\varvec{\theta }}_k\Vert =O_p(-r/(2r+1))\), together with \(n^{(1+\tau )r/(2r+1)}\lambda _{\min }\rightarrow \infty \), we can derive that

$$\begin{aligned} \Pr \left( \frac{\lambda _2}{\Vert \tilde{\varvec{\theta }}_k\Vert ^{1+\tau }}<1\right)= & {} \Pr \left( \Vert \tilde{\varvec{\theta }}_k\Vert ^{1+\tau }>\lambda _2\right) \\\le & {} \frac{1}{\lambda _2}E\left( \Vert \tilde{\varvec{\theta }}_k\Vert ^{1+\tau }\right) \\= & {} \frac{1}{\lambda _2}n^{-(1+\tau )r/(2r+1)}\\\le & {} \frac{1}{\lambda _{\min }}n^{-(1+\tau )r/(2r+1)}\rightarrow 0, \end{aligned}$$

This implies that \(\lim _{n\rightarrow \infty }\Pr \left( \hat{\delta }_{1,k}=1, \text{ for } \text{ all }~~k\in \{v+1,\ldots ,p\} \right) =1\). On the other hand, by the condition \(n^{r/(2r+1)}\lambda _{\max }\rightarrow 0\), for \(\varepsilon > 0\) and \(k\in \{1,\ldots ,v\}\), we have that \( \Pr \left( \hat{\delta }_{1,k}>n^{\frac{-r}{1+2r}}\varepsilon \right) \rightarrow 0\), which implies that for each \(k\in \{1,\ldots ,v\}\), \(\hat{\delta }_{1,k}=o_{p}(n^{-r/(2r+1)})\). Therefore, we can get that \(\lim _{n\rightarrow \infty }\Pr \left( \hat{\delta }_{1,k}<1, \text{ for } \text{ all } k\in \{1,\ldots ,v\} \right) =1\). The proof of (a) is completed. For part (b), apply the similar techniques as in part (a), we have, with probability tending to 1, that \(\hat{\delta }_{2,k}=1\) for \(k\in \{c+1,\ldots ,q\}\) and \(\hat{\delta }_{2,k}<1\) for \(k\in \{1,\ldots ,c\}\).

Proof (b) of Theorem 3.2. It can be proved by using the same method used in the proof of Theorem 2.2 in Wang and Lin (2016), we omit the detail for saving space.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wang, K., Sun, X. Efficient parameter estimation and variable selection in partial linear varying coefficient quantile regression model with longitudinal data. Stat Papers 61, 967–995 (2020). https://doi.org/10.1007/s00362-017-0970-0

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00362-017-0970-0

Keywords

Mathematics Subject Classification

Navigation