Skip to main content
Log in

Testing high-dimensional mean vector with applications

A normal reference approach

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

Abstract

A centered \(L^2\)-norm based test statistic is used for testing if a high-dimensional mean vector equals zero where the data dimension may be much larger than the sample size. Inspired by the fact that under some regularity conditions the asymptotic null distributions of the proposed test are the same as the limiting distributions of a chi-square-mixture, a three-cumulant matched chi-square-approximation is suggested to approximate this null distribution. The asymptotic power of the proposed test under a local alternative is established and the effect of data non-normality is discussed. A simulation study under various settings demonstrates that in terms of size control, the proposed test performs significantly better than some existing competitors. Several real data examples are presented to illustrate the wide applicability of the proposed test to a variety of high-dimensional data analysis problems, including the one-sample problem, paired two-sample problem, and MANOVA for correlated samples or independent samples.

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.

Similar content being viewed by others

References

  • Ahmad MR, Werner C, Brunner E (2008) Analysis of high-dimensional repeated measures designs: the one sample case. Comput Stat Data Anal 53(2):416–427

    Article  MathSciNet  MATH  Google Scholar 

  • Allen GI, Tibshirani R (2010) Transposable regularized covariance models with an application to missing data imputation. Ann Appl Stat 4(2):764–790

    Article  MathSciNet  MATH  Google Scholar 

  • Alon U, Barkai N, Notterman D, Gish K, Ybarra S, Mack D, Levine A (1999) Broad patterns of gene expression revealed by clustering analysis of tumor and normal colon tissues probed by oligonucleotide arrays. Proc Natl Acad Sci 96(12):6745–6750

    Article  Google Scholar 

  • Anderson TW (1963) A test for equality of means when covariance matrices are unequal. Ann Math Stat 34(2):671–672

    Article  MathSciNet  MATH  Google Scholar 

  • Anderson TW (2003) An introduction to multivariate statistical analysis. Wiley, New York

    MATH  Google Scholar 

  • Bai ZD, Saranadasa H (1996) Effect of high dimension: by an example of a two sample problem. Stat Sin 6(2):311–329

    MathSciNet  MATH  Google Scholar 

  • Bai Z, Hu J, Wang C, Zhang C (2021) Test on the linear combinations of covariance matrices in high-dimensional data. Stat Pap 62:701–719

    Article  MathSciNet  MATH  Google Scholar 

  • Bennett BM (1950) Note on a solution of the generalized Behrens-Fisher problem. Ann Inst Stat Math 2(1):87–90

    Article  MathSciNet  MATH  Google Scholar 

  • Box GE (1954) Some theorems on quadratic forms applied in the study of analysis of variance problems, I. Effect of inequality of variance in the one-way classification. Ann Math Stat 25(2):290–302

    Article  MathSciNet  MATH  Google Scholar 

  • Burczynski ME, Peterson RL, Twine NC, Zuberek KA, Brodeur BJ, Casciotti L, Maganti V, Reddy PS, Strahs A, Immermann F, Spinelli W, Schwertschlag U, Slager AM, Cotreau MM, Dorner AJ (2006) Molecular classification of Crohn’s disease and ulcerative colitis patients using transcriptional profiles in peripheral blood mononuclear cells. J Mol Diagn 8(1):51–61

  • Chen SX, Qin YL (2010) A two-sample test for high-dimensional data with applications to gene-set testing. Ann Stat 38(2):808–835

    Article  MathSciNet  MATH  Google Scholar 

  • Chen LS, Paul D, Prentice RL, Wang P (2011) A regularized Hotelling’s \(T^2\) test for pathway analysis in proteomic studies. J Am Stat Assoc 106(496):1345–1360

    Article  MATH  Google Scholar 

  • Dempster AP (1958) A high dimensional two sample significance test. Ann Math Stat 29(4):995–1010

    Article  MathSciNet  MATH  Google Scholar 

  • Dong K, Pang H, Tong T, Genton MG (2016) Shrinkage-based diagonal Hotelling’s tests for high-dimensional small sample size data. J Multivar Anal 143:127–142

    Article  MathSciNet  MATH  Google Scholar 

  • Feng L, Sun F (2016) Spatial-sign based high-dimensional location test. Electron J Stat 10(2):2420–2434

    Article  MathSciNet  MATH  Google Scholar 

  • Feng L, Zou C, Wang Z, Zhu L (2017) Composite \(T^2\) test for high-dimensional data. Stat Sin 27:1419–1436

    MATH  Google Scholar 

  • Hall P (1983) Chi squared approximations to the distribution of a sum of independent random variables. Ann Probab 11(4):1028–1036

    Article  MathSciNet  MATH  Google Scholar 

  • Henze N (2002) Invariant tests for multivariate normality: a critical review. Stat Pap 43:467–506

    Article  MathSciNet  MATH  Google Scholar 

  • Hu J, Bai Z, Wang C, Wang W (2017) On testing the equality of high dimensional mean vectors with unequal covariance matrices. Ann Inst Stat Math 69:365–387

    Article  MathSciNet  MATH  Google Scholar 

  • Hu Z, Tong T, Genton MG (2019) Diagonal likelihood ratio test for equality of mean vectors in high-dimensional data. Biometrics 75:256–267

    Article  MathSciNet  MATH  Google Scholar 

  • Katayama S, Kano Y, Srivastava MS (2013) Asymptotic distributions of some test criteria for the mean vector with fewer observations than the dimension. J Multivar Anal 116:410–421

    Article  MathSciNet  MATH  Google Scholar 

  • Li H, Aue A, Paul D (2020) High-dimensional general linear hypothesis tests via non-linear spectral shrinkage. Bernoulli 26(4):2541–2571

    Article  MathSciNet  MATH  Google Scholar 

  • Nishiyama T, Hyodo M, Seo T, Pavlenko T (2013) Testing linear hypotheses of mean vectors for high-dimension data with unequal covariance matrices. J Stat Plann Inference 143(11):1898–1911

    Article  MathSciNet  MATH  Google Scholar 

  • Paindaveine D, Verdebout T (2016) On high-dimensional sign tests. Bernoulli 22(3):1745–1769

    Article  MathSciNet  MATH  Google Scholar 

  • Park J, Ayyala DN (2013) A test for the mean vector in large dimension and small samples. J Stat Plann Inference 143(5):929–943

    Article  MathSciNet  MATH  Google Scholar 

  • Pauly M, Ellenberger D, Brunner E (2015) Analysis of high-dimensional one group repeated measures designs. Statistics 49(6):1243–1261

    Article  MathSciNet  MATH  Google Scholar 

  • Peng L, Qi Y, Wang F (2014) Test for a mean vector with fixed or divergent dimension. Stat Sci 29(1):113–127

    Article  MathSciNet  MATH  Google Scholar 

  • Satterthwaite FE (1946) An approximate distribution of estimates of variance components. Biometrics Bull 2(6):110–114

    Article  Google Scholar 

  • Scheffé H (1943) On solutions of the Behrens-Fisher problem, based on the \(t\)-distribution. Ann Math Stat 14(1):35–44

    Article  MathSciNet  MATH  Google Scholar 

  • Schott JR (2007) Some high-dimensional tests for a one-way MANOVA. J Multivar Anal 98(9):1825–1839

    Article  MathSciNet  MATH  Google Scholar 

  • Shen Y, Lin Z (2015) An adaptive test for the mean vector in large-p-small-n problems. Comput Stat Data Anal 89:25–38

    Article  MathSciNet  MATH  Google Scholar 

  • Shen Y, Lin Z, Zhu J (2011) Shrinkage-based regularization tests for high-dimensional data with application to gene set analysis. Comput Stat Data Anal 55(7):2221–2233

    Article  MathSciNet  MATH  Google Scholar 

  • Silva IR, Zhuang Y, da Silva Junior JCA (2021) Kronecker delta method for testing independence between two vectors in high-dimension. Stat Pap (In Press)

  • Sottoriva A, Spiteri I, Piccirillo SGM, Touloumis A, Collins VP, Marioni JC, Curtis C, Watts C, Tavaré S (2013) Intratumor heterogeneity in human glioblastoma reflects cancer evolutionary dynamics. Proc Natl Acad Sci 110(10):4009–4014

    Article  Google Scholar 

  • Srivastava MS, Du M (2008) A test for the mean vector with fewer observations than the dimension. J Multivar Anal 99(3):386–402

    Article  MathSciNet  MATH  Google Scholar 

  • Srivastava MS, Kubokawa T (2013) Tests for multivariate analysis of variance in high dimension under non-normality. J Multivar Anal 115:204–216

    Article  MathSciNet  MATH  Google Scholar 

  • Srivastava MS, Yanagihara H (2010) Testing the equality of several covariance matrices with fewer observations than the dimension. J Multivar Anal 101(6):1319–1329

    Article  MathSciNet  MATH  Google Scholar 

  • Touloumis A, Tavaré S, Marioni JC (2015) Testing the mean matrix in high-dimensional transposable data. Biometrics 71(1):157–166

    Article  MathSciNet  MATH  Google Scholar 

  • Wang R, Xu X (2019) A feasible high dimensional randomization test for the mean vector. J Stat Plann Inference 199:160–178

    Article  MathSciNet  MATH  Google Scholar 

  • Wang L, Peng B, Li R (2015) A high-dimensional nonparametric multivariate test for mean vector. J Am Stat Assoc 110(512):1658–1669

    Article  MathSciNet  MATH  Google Scholar 

  • Welch BL (1947) The generalization of ‘Student’s’ problem when several different population variances are involved. Biometrika 34(1/2):28–35

    Article  MathSciNet  MATH  Google Scholar 

  • Yamada T, Himeno T (2015) Testing homogeneity of mean vectors under heteroscedasticity in high-dimension. J Multivar Anal 139:7–27

    Article  MathSciNet  MATH  Google Scholar 

  • Zhang JT (2005) Approximate and asymptotic distributions of chi-squared-type mixtures with applications. J Am Stat Assoc 100(469):273–285

    Article  MathSciNet  MATH  Google Scholar 

  • Zhang JT (2013) Analysis of variance for functional data. CRC Press, Boca Raton

    Book  Google Scholar 

  • Zhang JT, Xu J (2009) On the k-sample Behrens-Fisher problem for high-dimensional data. Sci China Ser A 52(6):1285–1304

    Article  MathSciNet  MATH  Google Scholar 

  • Zhang JT, Guo J, Zhou B (2017) Linear hypothesis testing in high-dimensional one-way MANOVA. J Multivar Anal 155:200–216

    Article  MathSciNet  MATH  Google Scholar 

  • Zhang JT, Guo J, Zhou B, Cheng MY (2020) A simple two-sample test in high dimensions based on \(L^2\)-norm. J Am Stat Assoc 115(530):1011–1027

    Article  MATH  Google Scholar 

  • Zhang T, Wang Z, Wan Y (2021) Functional test for high-dimensional covariance matrix, with application to mitochondrial calcium concentration. Stat Pap 62:1213–1230

    Article  MathSciNet  MATH  Google Scholar 

  • Zhao J (2017) A new test for the mean vector in large dimension and small samples. Commun Stat-Simul Comput 46(8):6115–6128

    Article  MathSciNet  MATH  Google Scholar 

  • Zhou B, Guo J, Chen J, Zhang JT (2019) An adaptive spatial-sign-based test for mean vectors of elliptically distributed high-dimensional data. Stat Interface 12:93–106

    Article  MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Bu Zhou.

Additional information

Publisher's Note

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

Jin-Ting Zhang was financially supported by the National University of Singapore Academic Research Grant R-155-000-187-114. Bu Zhou was supported by the National Natural Science Foundation of China under Grant No. 11901520, the Zhejiang Provincial Natural Science Foundation of China under Grant No. LY21A010007, and the Characteristic & Preponderant Discipline of Key Construction Universities in Zhejiang Province (Zhejiang Gongshang University—Statistics). Jia Guo was supported by the National Natural Science Foundation of China under Grant No. 11901522. We are grateful to two reviewers for their valuable comments and suggestions which help to improve the presentation of the paper.

Supplementary Information

Below is the link to the electronic supplementary material.

Supplementary material 1 (pdf 332 KB)

Appendix: Technical proofs

Appendix: Technical proofs

We first prove the following useful lemma.

Lemma 1

Let \(\varvec{{W}}\sim W_{p}(v,\varvec{{\varSigma }}/v)\) denote a Wishart distribution with v degrees of freedom and a covariance matrix \(\varvec{{\varSigma }}/v\). Assume that \(p/v^2\longrightarrow 0\) as \(v, p\rightarrow \infty \). Then the unbiased and ratio-consistent estimator of \({\text {tr}}(\varvec{{\varSigma }}^{3})\) is given by

$$\begin{aligned} \widehat{{\text {tr}}(\varvec{{\varSigma }}^{3})}=\frac{v^{4}}{(v-1)(v+4)(v^{2}-4)}\left[ {\text {tr}}(\varvec{{W}}^{3})-\frac{3}{v}{\text {tr}}(\varvec{{W}}){\text {tr}}(\varvec{{W}}^{2})+\frac{2}{v^{2}}{\text {tr}}^{3}(\varvec{{W}})\right] . \end{aligned}$$
(A.33)

Proof of Lemma 1

Let \(\varvec{{V}}=v\varvec{{W}}\). Then \(\varvec{{V}}\sim W_{p}(v,\varvec{{\varSigma }})\). To show (A.33) is equivalent to show

$$\begin{aligned} \widehat{{\text {tr}}(\varvec{{\varSigma }}^{3})}=\frac{v}{(v-1)(v+4)(v^{2}-4)}\left[ {\text {tr}}(\varvec{{V}}^{3})-\frac{3}{v}{\text {tr}}(\varvec{{V}}){\text {tr}}(\varvec{{V}}^{2})+\frac{2}{v^{2}}{\text {tr}}^{3}(\varvec{{V}})\right] \end{aligned}$$
(A.34)

is unbiased and ratio consistent for \({\text {tr}}(\varvec{{\varSigma }}^{3})\). By the proof of Theorem 2.1 of Srivastava and Yanagihara (2010), we can express the above expression as \(\widehat{{\text {tr}}(\varvec{{\varSigma }}^{3})}=I_{1}+I_{2}+I_{3}+I_{4},\) where

$$\begin{aligned} \begin{array}{rcl} I_{1} &{} = &{} \frac{1}{v(v+2)(v+4)}\sum _{i=1}^{p}\lambda _{i}^{3}w_{ii}^{3},\\ I_{2} &{} = &{} \frac{3}{(v-1)(v+2)(v+4)}\sum _{i\ne j}\lambda _{i}^{2}\lambda _{j}\left( w_{ii}w_{ij}^{2}-v^{-1}w_{ii}^{2}w_{jj}\right) ,\\ I_{3} &{} = &{} \frac{v}{(v-1)(v+4)(v^{2}-4)}\sum _{i\ne j\ne k}\lambda _{i}\lambda _{j}\lambda _{k}\left( w_{ij}w_{jk}w_{ki}-v^{-2}w_{ii}w_{jj}w_{kk}\right) ,\\ I_{4} &{} = &{} \frac{-3}{(v-1)(v+4)(v^{2}-4)}\sum _{i\ne j\ne k}\lambda _{i}\lambda _{j}\lambda _{k}\left( w_{ij}w_{jk}^{2}-v^{-1}w_{ii}w_{jj}w_{kk}\right) , \end{array} \end{aligned}$$

where \(w_{ij}=\varvec{{u}}_{i}^{\top }\varvec{{u}}_{j},\ i,j=1,\ldots ,p\) and \(\varvec{{u}}_{1},\ldots ,\varvec{{u}}_{p}{\mathop {\sim }\limits ^{\text {i.i.d.}}}\mathcal {N}_{v}(\varvec{{0}},\varvec{{I}}_{v})\). By Lemma 6.2 (a) and (e) of Srivastava and Yanagihara (2010), we have \({\text {E}}(w_{ii}^{3})=v(v+2)(v+4)\) and \({\text {Var}}(w_{ii}^{3})=6v(v+2)(v+4)(3v^{2}+30v+80)\). Therefore,

$$\begin{aligned} \begin{array}{rcl} {\text {E}}(I_{1}) &{} = &{} \frac{1}{v(v+2)(v+4)}\sum _{i=1}^{p}\lambda _{i}^{3}{\text {E}}(w_{ii}^{3})={\text {tr}}(\varvec{{\varSigma }}^{3}),\\ {\text {Var}}(I_{1}) &{} = &{} [\frac{1}{v(v+2)(v+4)}]^{2}\sum _{i=1}^{p}\lambda _{i}^{6}{\text {Var}}(w_{ii}^{3})=v^{-1}{\text {tr}}(\varvec{{\varSigma }}^{6})[1+o(1)]. \end{array} \end{aligned}$$

Noting that

$$\begin{aligned} {\text {tr}}(\varvec{{A}}^{r})/{\text {tr}}^{r}(\varvec{{A}})\le 1,\ r=1,2,\ldots \end{aligned}$$
(A.35)

hold for any nonnegative matrix \(\varvec{{A}}\), we have \({\text {Var}}[I_{1}/{\text {tr}}(\varvec{{\varSigma }}^{3})]<v^{-1}[1+o(1)]\). It follows that \(I_{1}/{\text {tr}}(\varvec{{\varSigma }}^{3})\longrightarrow 1\) as \(v\rightarrow \infty \).

Following the proof of Theorem 2.1 of Srivastava and Yanagihara (2010) closely, let \(r_{ij}=w_{ii}w_{ij}^{2}-v^{-1}w_{ii}^{2}w_{jj}\). Then we have \({\text {E}}(r_{ij})=0\) and \({\text {Cov}}(r_{ij},r_{kl})=0\) where \((i\ne j),\ (k\ne l)\) and \((i,j)\ne (k,l)\). In addition, by Lemma 6.2 (f), (g) and (h) of Srivastava and Yanagihara (2010), we have

$$\begin{aligned} \begin{array}{rcl} {\text {Var}}(r_{ij}) &{} = &{} {\text {E}}\left( w_{ii}w_{ij}^{2}-v^{-1}w_{ii}^{2}w_{jj}\right) ^{2}={\text {E}}\left( w_{ii}^{2}w_{ij}^{4}-2v^{-1}w_{ii}^{3}w_{ij}^{2}w_{jj}+v^{-2}w_{ii}^{4}w_{jj}^{2}\right) \\ &{} = &{} O(v^{4})-O(v^{-1}v^{5})+O(v^{-2}v^{6})=O(v^{4}). \end{array} \end{aligned}$$

It follows that \({\text {E}}(I_{2})=0\) and

$$\begin{aligned} \begin{array}{rcl} {\text {Var}}(I_{2})&{}=&{}9\left[ (v-1)(v+2)(v+4)\right] ^{-2}\sum _{i\ne j}\lambda _{i}^{4}\lambda _{j}^{2}{\text {Var}}(r_{ij})\\ &{}\le &{} 9v^{-2}{\text {tr}}(\varvec{{\varSigma }}^{4}){\text {tr}}(\varvec{{\varSigma }}^{2})[1+o(1)]. \end{array} \end{aligned}$$

It follows from (A.35) that \({\text {Var}}\left[ I_{2}/{\text {tr}}(\varvec{{\varSigma }}^{3})\right] \le [{\text {tr}}(\varvec{{\varSigma }}^{4})(9\kappa )]/[{\text {tr}}^{2}(\varvec{{\varSigma }}^{2})v^{2}][1+o(1)]\le {9\kappa }/{v^{2}}[1+o(1)],\) where

$$\begin{aligned} \kappa =\frac{{\text {tr}}^{3}(\varvec{{\varSigma }}^{2})}{{\text {tr}}^{2}(\varvec{{\varSigma }}^{3})}. \end{aligned}$$
(A.36)

By Theorem 5 of Zhang et al. (2020), we have \(\kappa /p\le 1\). Since \(p/v^2\longrightarrow 0\) as \(v,p\rightarrow \infty \), we have \(\kappa /v^{2}\longrightarrow 0\) as \(v,p\rightarrow \infty \). Therefore, we have that \(I_{2}=o_{p}[{\text {tr}}(\varvec{{\varSigma }}^{3})]\).

Similarly, we have \({\text {E}}(I_{3})=0,\ {\text {E}}(I_{4})=0\) and

$$\begin{aligned} {\text {Var}}[I_{3}/{\text {tr}}(\varvec{{\varSigma }}^{3})]\le \kappa v^{-3}[1+o(1)]\longrightarrow 0,\;\;{\text {Var}}[I_{4}/{\text {tr}}(\varvec{{\varSigma }}^{3})]\le \kappa v^{-4}[1+o(1)]\longrightarrow 0, \end{aligned}$$

showing that \(I_{3}=o_{p}[{\text {tr}}(\varvec{{\varSigma }}^{3})]\) and \(I_{4}=o_{p}[{\text {tr}}(\varvec{{\varSigma }}^{3})]\). It follows that \(\widehat{{\text {tr}}(\varvec{{\varSigma }}^{3})}\) is an unbiased and ratio-consistent estimator of \({\text {tr}}(\varvec{{\varSigma }}^{3})\). The lemma is then proved. \(\square \)

We now prove the main results.

Proof of Theorem 1

We first prove (a). We shall use the characteristic function (\(\psi _{X}(t)={\text {E}}(e^{itX})\) for a random variable X) method. Set \(\varvec{{x}}_i=\varvec{{y}}_i-\varvec{{\mu }},\ i=1,\ldots ,n\) and hence \(\bar{\varvec{{x}}}=\bar{\varvec{{y}}}-\varvec{{\mu }}\). Set \(\varvec{{w}}_{n,p}=\sqrt{n}\bar{\varvec{{x}}}\). By (10), we have

$$\begin{aligned} \tilde{T}_{n,p,0}=\left[ \Vert \varvec{{w}}_{n,p}\Vert ^2-{\text {tr}}(\varvec{{\varSigma }})\right] /\sqrt{2{\text {tr}}(\varvec{{\varSigma }}^2)}[1+o(1)], \end{aligned}$$
(A.37)

since \({\text {tr}}(\hat{\varvec{{\varSigma }}})/{\text {tr}}(\varvec{{\varSigma }})\longrightarrow 1\) as \(n,p\rightarrow \infty \) (See Proof of Theorem 9 in Zhang et al. 2020). Further, we have \({\text {E}}(\varvec{{w}}_{n,p})=\varvec{{0}}\) and \({\text {Cov}}(\varvec{{w}}_{n,p})=\varvec{{\varSigma }}\). Write \(\varvec{{w}}_{n,p}=\sum _{r=1}^{p}\xi _{n,p,r}\varvec{{u}}_{p,r}\), where \(\xi _{n,p,r}=\varvec{{w}}_{n,p}^{\top }\varvec{{u}}_{p,r}\), and \(\varvec{{u}}_{p,1},\ldots ,\varvec{{u}}_{p,p}\) denote the eigenvectors associated with the eigenvalues \(\lambda _{p,1},\ldots ,\lambda _{p,p}\) of \(\varvec{{\varSigma }}\) in the descending order. We have \({\text {E}}(\xi _{n,p,r})=0\), and \({\text {Var}}(\xi _{n,p,r})=\lambda _{p,r},\ r=1,2,\ldots \), and \(\xi _{n,p,r},\ r=1,\ldots ,p\), are uncorrelated.

By Lemma S.4 of Zhang et al. (2020), we have

$$\begin{aligned} {\text {Var}}(\xi _{n,p,r}^{2})=2\lambda _{p,r}^{2}+[{\text {E}}(\varvec{{x}}_{1}^{\top }\varvec{{u}}_{p,r})^{4}-3\lambda _{p,r}^{2}]/n. \end{aligned}$$

Under Conditions C1 and C2, by some simple algebra, we have \({\text {E}}(\varvec{{x}}_{1}^{\top }\varvec{{u}}_{p,r})^{4}\le (3+\varDelta )\lambda _{p,r}^{2}\). Thus, we have

$$\begin{aligned} {\text {Var}}(\xi _{n,p,r}^{2})\le (2+\varDelta /n)\lambda _{p,r}^{2},\ r=1,2,\ldots . \end{aligned}$$
(A.38)

It follows from (A.37) that

$$\begin{aligned} \tilde{T}_{n,p,0}=\sum _{r=1}^p (\xi _{n,p,r}^2-\lambda _{p,r})/\sqrt{2{\text {tr}}(\varvec{{\varSigma }}^2)}[1+o(1)]. \end{aligned}$$

Set \(\tilde{T}_{n,p,0}^{(q)}=\sum _{r=1}^q (\xi _{n,p,r}^2-\lambda _{p,r})/\sqrt{2{\text {tr}}(\varvec{{\varSigma }}^2)}\) where \(q<p\). Then

$$\begin{aligned}&{\text {E}}\big (\tilde{T}_{n,p,0}-\tilde{T}_{n,p,0}^{(q)}\big )^{2}={\text {E}}\bigg [\sum _{r=q+1}^{p}(\xi _{n,p,r}^{2}-\lambda _{p,r})/\sqrt{2{\text {tr}}(\varvec{{\varSigma }}^2)}\bigg ]^{2}[1+o(1)]\\&\qquad \qquad \qquad \qquad \qquad \qquad ={\text {Var}}\bigg (\sum _{r=q+1}^{p}\xi _{n,p,r}^{2}\bigg )/\left[ 2{\text {tr}}(\varvec{{\varSigma }}^2)\right] [1+o(1)]\\&\qquad \qquad \qquad \qquad \qquad \qquad \le \bigg [\sum _{r=q+1}^{p}\sqrt{{\text {Var}}(\xi _{n,p,r}^{2})}\bigg ]^{2}/\left[ 2{\text {tr}}(\varvec{{\varSigma }}^2)\right] [1+o(1)]. \end{aligned}$$

By (A.38), we have

$$\begin{aligned} \bigg [\sum _{r=q+1}^p \sqrt{{\text {Var}}(\xi _{n,p,r}^{2})} \bigg ]^2/\left[ 2{\text {tr}}(\varvec{{\varSigma }}^2)\right] \le \left( 1+\varDelta /n\right) \bigg (\sum _{r=q+1}^p \rho _{p,r}\bigg )^2. \end{aligned}$$

It follows that

$$\begin{aligned} \begin{array}{rcl} |\psi _{\tilde{T}_{n,p,0}}(t)-\psi _{\tilde{T}_{n,p,0}^{(q)}}(t)| &{}\le &{} |t|\big [{\text {E}}(\tilde{T}_{n,p,0}-\tilde{T}_{n,p,0}^{(q)})^{2}\big ]^{1/2}\\ &{}\le &{} |t|\left( 1+\varDelta /n\right) ^{1/2} \bigg (\sum _{r=q+1}^{p}\rho _{p,r}\bigg )[1+o(1)]. \end{array} \end{aligned}$$
(A.39)

Set \(\zeta ^{(q)}=\sum _{r=1}^{q}\rho _{r}(A_{r}-1)/\sqrt{2}\), we have

$$\begin{aligned} \big |\psi _{\tilde{T}_{n,p,0}}(t)-\psi _{\zeta }(t)\big |\le & {} \big |\psi _{\tilde{T}_{n,p,0}}(t)-\psi _{\tilde{T}_{n,p,0}^{(q)}}(t)\big | + \big |\psi _{\tilde{T}_{n,p,0}^{(q)}}(t)-\psi _{\tilde{T}_{p,0}^{(q)}}(t)\big | \\&+\big |\psi _{\tilde{T}_{p,0}^{(q)}}(t)-\psi _{\zeta ^{(q)}}(t)\big |+\big |\psi _{\zeta ^{(q)}}(t)-\psi _{\zeta }(t)\big |. \end{aligned}$$

By similar arguments in Proof of Theorem 2 of Zhang et al. (2020), we can show under Conditions C1–C3 all the four terms on the right hand side of the previous inequality converge to zero as \(n, p\rightarrow \infty \), so the first expression of (11) follows. In particular, the convergence of the first term can be derived from (A.39) and Condition C3, the convergence of the second term is ensured by the standard central limit theorem, and the convergence of the last two terms is due to Condition C3.

Notice that when the data (2) are normally distributed, Conditions C1–C2 are automatically satisfied so that under Condition C3, the second expression of (11) follows immediately since under the normality assumption, we have \(T_{n,p,0}{\mathop {=}\limits ^{d}}T_{n,p,0}^*\).

We now prove (b). By Lemma 8.1 of Pauly et al. (2015), Condition C5 is equivalent to the condition “\({\text {tr}}(\varvec{{\varSigma }}^4)/{\text {tr}}^2(\varvec{{\varSigma }}^2)=o(1)\)” imposed by Chen and Qin (2010). Therefore, under Conditions C1, C2 and C5, the proofs of the asymptotic normality of \(\tilde{T}_{n,p,0}\) and \(\tilde{T}_{n,p,0}^*\) are along the same lines as the one given by Chen and Qin (2010) for the asymptotic normality of their test statistic.

Finally, we prove (13). We have \(\sup _{x} \big |\Pr (T_{n,p,0}\le x)-\Pr (T_{n,p,0}^*\le x)\big |\le \sup _{x} \big |\Pr (\tilde{T}_{n,p,0}\le \tilde{x})-\Pr (\zeta \le \tilde{x})\big |+ \sup _{x} \big |\Pr (\tilde{T}_{n,p,0}^*\le \tilde{x})-\Pr (\zeta \le \tilde{x})\big |\), where \(\tilde{x}=[x-{\text {tr}}(\hat{\varvec{{\varSigma }}})]/\sqrt{\frac{2n}{(n-1)}{\text {tr}}(\varvec{{\varSigma }}^2)}\). Under the conditions of (a), the two terms on the right hand side of previous inequality both converge to zero because both \(\tilde{T}_{n,p,0}\) and \(\tilde{T}_{n,p,0}^*\) converge to \(\zeta \) in distribution. Under the conditions of (b), the proof of (13) are along the same lines as the above and hence are omitted for space saving. \(\square \)

Proof of Theorem 2

Under the local alternative (21), \(T_{n,p}=\left( T_{n,p,0}+n\Vert \varvec{{\mu }}\Vert ^2\right) [1+o_p(1)]\). In addition, under the given conditions, we have \(\hat{\beta }_0/\beta _0{\mathop {\longrightarrow }\limits ^{P}}1,\ \hat{\beta }_1/\beta _1{\mathop {\longrightarrow }\limits ^{P}}1\) and \(\hat{d}/d{\mathop {\longrightarrow }\limits ^{P}}1\) as \(n, p\rightarrow \infty \). We first prove (a). Under Conditions C1, C2 and C3, Theorem 1(a) indicates that as \(n,p\rightarrow \infty \), we have \(\tilde{T}_{n,p,0}=T_{n,p,0}/\sqrt{\frac{2n}{(n-1)}{\text {tr}}(\varvec{{\varSigma }}^2)}{\mathop {\longrightarrow }\limits ^{\mathcal {L}}}\zeta \). It follows that as \(n, p\rightarrow \infty \), we have

$$\begin{aligned} \begin{array}{rcl} \Pr \left[ T_{n,p}\ge \hat{\beta }_0+\hat{\beta }_1\chi _{\hat{d}}^2(\alpha )\right] &{}=&{} \Pr \left[ \tilde{T}_{n,p,0} \ge \frac{\beta _0+ \beta _1 \chi ^2_{d}(\alpha )}{\sqrt{\frac{2n}{(n-1)}{\text {tr}}(\varvec{{\varSigma }}^2)}}-\frac{n\Vert \varvec{{\mu }}\Vert ^2}{\sqrt{\frac{2n}{(n-1)}{\text {tr}}(\varvec{{\varSigma }}^2)}} \right] [1+o(1)] \\ &{}= &{} \Pr \left[ \zeta \ge \frac{\chi _{d}^2(\alpha )-d}{\sqrt{2d}}-\frac{n\Vert \varvec{{\mu }}\Vert ^2}{\sqrt{2{\text {tr}}(\varvec{{\varSigma }}^2)}} \right] [1+o(1)]. \end{array} \end{aligned}$$
(A.40)

We now prove (b). Under the given conditions, Theorem 1(b) indicates that as \(n\rightarrow \infty \), we have \(\tilde{T}_{n,p,0} {\mathop {\longrightarrow }\limits ^{\mathcal {L}}}\mathcal {N}(0,1)\) and by Theorem 5, as \(n,p\rightarrow \infty \), we have \(d\longrightarrow \infty \) and \([\chi _d^2(\alpha )-d]/\sqrt{2d}\longrightarrow z_{\alpha }\) where \(z_{\alpha }\) denotes the upper \(100\alpha \)-percentile of \(\mathcal {N}(0,1)\). Then by (A.40), as \(n,p\rightarrow \infty \), we have

$$\begin{aligned} \begin{array}{rcl} \Pr \left[ T_{n,p}\ge \hat{\beta }_0+\hat{\beta }_1\chi _{\hat{d}}^2(\alpha )\right] &{}=&{} \Pr \left[ \tilde{T}_{n,p,0} \ge \frac{\beta _0+ \beta _1 \chi ^2_{d}(\alpha )}{\sqrt{\frac{2n}{(n-1)}{\text {tr}}(\varvec{{\varSigma }}^2)}}-\frac{n\Vert \varvec{{\mu }}\Vert ^2}{\sqrt{\frac{2n}{(n-1)}{\text {tr}}(\varvec{{\varSigma }}^2)}} \right] [1+o(1)] \\ &{}= &{} \varPhi \left[ -z_{\alpha }+\frac{n\Vert \varvec{{\mu }}\Vert ^2}{\sqrt{2{\text {tr}}(\varvec{{\varSigma }}^2)}} \right] [1+o(1)], \end{array} \end{aligned}$$

where \(\varPhi (\cdot )\) denotes the cumulative distribution function of \(\mathcal {N}(0,1)\). The proof is complete. \(\square \)

Proof of Theorem 3

Let \(\varvec{{x}}_{i}=\varvec{{y}}_{i}-\varvec{{\mu }},\ i=1,\ldots ,n\). Then \(\varvec{{x}}_{i},\ i=1,\ldots ,n\) are i.i.d. with \({\text {E}}(\varvec{{x}}_{i})=\varvec{{0}}\) and \({\text {Cov}}(\varvec{{x}}_{i})=\varvec{{\varSigma }}\). It is easy to verify that \(T_{n,p,0}=2(n-1)^{-1}\sum _{i<j}\varvec{{x}}_{i}^{\top }\varvec{{x}}_{j}\). It follows that \({\text {E}}(T_{n,p,0})=0\) and

$$\begin{aligned} {\text {Var}}(T_{n,p,0})={\text {E}}(T_{n,p,0}^{2})=4(n-1)^{-2}\sum _{i<j}{\text {E}}(\varvec{{x}}_{i}^{\top }\varvec{{x}}_{j})^{2}=\frac{2n}{(n-1)}{\text {tr}}(\varvec{{\varSigma }}^{2}). \end{aligned}$$

Furthermore, we have

$$\begin{aligned} {\text {E}}(T_{n,p,0}^3)= & {} 8(n-1)^{-3}{\text {E}}\left[ \sum _{i<j}(\varvec{{x}}_{i}^{\top }\varvec{{x}}_{j})\right] ^{3}\\= & {} 8(n-1)^{-3}{\text {E}}\left[ \sum _{i<j}(\varvec{{x}}_{i}^{\top }\varvec{{x}}_{j})^{3}+3\sum _{*}(\varvec{{x}}_{i}^{\top }\varvec{{x}}_{j})^{2}(\varvec{{x}}_{u}^{\top }\varvec{{x}}_{v})\right. \\&\left. +6\sum _{**}(\varvec{{x}}_{i}^{\top }\varvec{{x}}_{j})(\varvec{{x}}_{u}^{\top }\varvec{{x}}_{v})(\varvec{{x}}_{\alpha }^{\top }\varvec{{x}}_{\beta })\right] \\= & {} 8(n-1)^{-3}\left\{ \frac{n(n-1)}{2}{\text {E}}(\varvec{{x}}_{1}^{\top }\varvec{{x}}_{2})^{3}+n(n-1)(n-2){\text {E}}[(\varvec{{x}}_{1}^{\top }\varvec{{x}}_{2})(\varvec{{x}}_{2}^{\top }\varvec{{x}}_{3})(\varvec{{x}}_{3}^{\top }\varvec{{x}}_{1})]\right\} \\= & {} \frac{8n(n-2)}{(n-1)^{2}}{\text {tr}}(\varvec{{\varSigma }}^{3})+\frac{4n\varUpsilon }{(n-1)^{2}}, \end{aligned}$$

where \(\varUpsilon ={\text {E}}(\varvec{{x}}_{1}^{\top }\varvec{{x}}_{2})^{3}={\text {E}}[(\varvec{{y}}_{1}-\varvec{{\mu }})^{\top }(\varvec{{y}}_{2}-\varvec{{\mu }})]^{3}\), \(*\) means “\(i<j,\ u<v\)” and “\((i,j)\ne (u,v)\)” while \(**\) means “\(i<j,\ u<v,\ \alpha <\beta \)” and “\((i,j),\ (u,v),\ (\alpha ,\beta )\) are not mutually equal to each other.” The proof is complete. \(\square \)

Proof of Theorem 4

We first show (a). Under Condition C1, we have \(\varvec{{y}}_i=\varvec{{\mu }}+\varvec{{\varGamma }}\varvec{{z}}_i,\ i=1,\ldots , n\) where \(\varvec{{z}}_i,\ i=1,\ldots , n\) are i.i.d. with \({\text {E}}(\varvec{{z}}_i)=\varvec{{0}}\) and \({\text {Cov}}(\varvec{{z}}_i)=\varvec{{I}}_p\) and \(\varvec{{\varSigma }}=\varvec{{\varGamma }}\varvec{{\varGamma }}^{\top }\). It follows that \(\varUpsilon ={\text {E}}[(\varvec{{y}}_{1}-\varvec{{\mu }})^{\top }(\varvec{{y}}_{2}-\varvec{{\mu }})]^{3}={\text {E}}(\varvec{{z}}_1^{\top }\varvec{{\varOmega }}\varvec{{z}}_2)^3\) where \(\varvec{{\varOmega }}=\varvec{{\varGamma }}^{\top }\varvec{{\varGamma }}\). By Jensen’s inequality, we have

$$\begin{aligned} \varUpsilon ={\text {E}}(\varvec{{z}}_1^{\top }\varvec{{\varOmega }}\varvec{{z}}_2)^3\le \left[ {\text {E}}(\varvec{{z}}_1^{\top }\varvec{{\varOmega }}\varvec{{z}}_2)^4\right] ^{3/4}. \end{aligned}$$
(A.41)

Denote the (ij)-th entry of \(\varvec{{\varOmega }}\) as \(w_{ij},\ i,j,=1,\ldots ,p\). Under Conditions C1 and C2, from the proof of Lemma 6.2 of Srivastava and Kubokawa (2013) (p. 215), we have

$$\begin{aligned} {\text {E}}\left( \varvec{{z}}_1^{\top }\varvec{{\varOmega }}\varvec{{z}}_2 \right) ^4=\varDelta ^2\sum _{i=1}^p\sum _{j=1}^p w_{ij}^4+6\varDelta \sum _{i=1}^p\left( \sum _{j=1}^p w_{ij}^2 \right) ^2 +6{\text {tr}}(\varvec{{\varOmega }}^4)+3{\text {tr}}^2(\varvec{{\varOmega }}^2),\qquad \end{aligned}$$
(A.42)

where \(\varDelta \) is given in Condition C2. Notice that we have

$$\begin{aligned} \begin{array}{rcl} \sum _{i=1}^p\sum _{j=1}^p w_{ij}^4&{}\le &{} \left( \sum _{i=1}^p\sum _{j=1}^p w_{ij}^2 \right) ^2={\text {tr}}^2(\varvec{{\varOmega }}^2),\\ \sum _{i=1}^p \left( \sum _{j=1}^p w_{ij}^2 \right) ^2 &{}\le &{} \left( \sum _{i=1}^p \sum _{j=1}^p w_{ij}^2 \right) ^2={\text {tr}}^2(\varvec{{\varOmega }}^2),\\ {\text {tr}}(\varvec{{\varOmega }}^4) &{}\le &{}{\text {tr}}^2(\varvec{{\varOmega }}^2). \end{array} \end{aligned}$$

These, together with (A.42), imply that \({\text {E}}\left( \varvec{{z}}_1^{\top }\varvec{{\varOmega }}\varvec{{z}}_2\right) ^4\le (\varDelta ^2+6\varDelta +9){\text {tr}}^2(\varvec{{\varOmega }}^2)\). Then by (A.41), we have

$$\begin{aligned} \varUpsilon \le (\varDelta ^2+6\varDelta +9)^{3/4}{\text {tr}}^{3/2}(\varvec{{\varOmega }}^2)=(\varDelta ^2+6\varDelta +9)^{3/4}{\text {tr}}^{3/2}(\varvec{{\varSigma }}^2), \end{aligned}$$
(A.43)

where we use the fact that \( {\text {tr}}(\varvec{{\varOmega }}^2)={\text {tr}}(\varvec{{\varGamma }}^{\top }\varvec{{\varGamma }}\varvec{{\varGamma }}^{\top }\varvec{{\varGamma }})={\text {tr}}(\varvec{{\varGamma }}\varvec{{\varGamma }}^{\top }\varvec{{\varGamma }}\varvec{{\varGamma }}^{\top })={\text {tr}}(\varvec{{\varSigma }}^2). \)

We now show (b). First of all, under Conditions C1 and C2, by (22) and (A.43), we have

$$\begin{aligned} \delta \le 1+\frac{(\varDelta ^2+6\varDelta +9)^{3/4}{\text {tr}}^{3/2}(\varvec{{\varSigma }}^2)}{2(n-2){\text {tr}}(\varvec{{\varSigma }}^3)} = 1+(\varDelta ^2+6\varDelta +9)^{3/4} \frac{\sqrt{\kappa }}{2(n-2)},\nonumber \\ \end{aligned}$$
(A.44)

where \(\kappa \) is given in (A.36). By Theorem 5 of Zhang et al. (2020), we have

$$\begin{aligned} 1\le \kappa \le \frac{{\text {tr}}^2(\varvec{{\varSigma }})}{{\text {tr}}(\varvec{{\varSigma }}^2)}\le p. \end{aligned}$$
(A.45)

Under Condition C3, as \(p\rightarrow \infty \), we have \( {{\text {tr}}^2(\varvec{{\varSigma }})}/{{\text {tr}}(\varvec{{\varSigma }}^2)}=(\sum _{r=1}^p \rho _{p,r})^2\longrightarrow (\sum _{r=1}^{\infty } \rho _{r})^2<\infty . \) It follows that under Conditions C1, C2 and C3, \(\kappa \) is bounded as \(p\rightarrow \infty \). This, together with (A.44), implies that under Conditions C1, C2 and C3, as \(n, p\rightarrow \infty \), we have \(\delta =1+o(1)\) as \(n, p\rightarrow \infty \).

Under Conditions C1, C2 and C4, by (A.45), as \(n, p\rightarrow \infty \), we have \( \sqrt{\kappa }/[2(n-2)]\le \sqrt{p/[2(n-2)]^2}=o(1). \) This, together with (A.44), implies that under Conditions C1, C2 and C4, as \(n, p\rightarrow \infty \), we always have \(\delta =1+o(1)\). The proof is complete. \(\square \)

Proof of Theorem 5

By (23), the skewness of \(T_{n,p,0}\) is \(\sqrt{8/f}\) where f is defined in (22) and under Conditions C1, C2 and C4, by Theorem 4, we have \(f=d[1+o(1)]\). On the one hand, when \(\tilde{T}_{n,p,0}{\mathop {\longrightarrow }\limits ^{\mathcal {L}}}\mathcal {N}(0,1)\), the skewness of \(T_{n,p,0}\) must tend to 0 as \(n,p\rightarrow \infty \), it follows that we must have \(d\longrightarrow \infty \) as \(n,p\rightarrow \infty \). On the other hand, by (15), as \(n\rightarrow \infty \), we have \(d=\kappa [1+o(1)]\) where \(\kappa \) is defined in (A.36). We have \(\kappa ={{\text {tr}}^{3}(\varvec{{\varSigma }}^{2})}/{{\text {tr}}^{2}(\varvec{{\varSigma }}^{3})}\ge {{\text {tr}}(\varvec{{\varSigma }}^{2})}/{\lambda _{p,\max }^{2}}\) where \(\lambda _{p,\max }\) is the largest eigenvalue of \(\varvec{{\varSigma }}\). Therefore, as \(d\rightarrow \infty \), we have \(\kappa \longrightarrow \infty \) and \({\lambda _{p,\max }^{2}}/{{\text {tr}}(\varvec{{\varSigma }}^{2})}\ge \kappa ^{-1}\longrightarrow 0\) as \(p\rightarrow \infty \). That is, Condition C5 holds. This, together with Conditions C1, C2 and C4, implies that by Theorem 1(b), we have \(\tilde{T}_{n,p,0}{\mathop {\longrightarrow }\limits ^{\mathcal {L}}}\mathcal {N}(0,1)\). The proof is then complete. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhang, JT., Zhou, B. & Guo, J. Testing high-dimensional mean vector with applications. Stat Papers 63, 1105–1137 (2022). https://doi.org/10.1007/s00362-021-01270-z

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00362-021-01270-z

Keywords

Navigation