Top Banner
Multiplier Bootstrap for Quantile Regression: Non-Asymptotic Theory under Random Design Xiaoou Pan * and Wen-Xin Zhou Abstract This paper establishes non-asymptotic concentration bound and Bahadur representa- tion for the quantile regression estimator and its multiplier bootstrap counterpart in the random design setting. The non-asymptotic analysis keeps track of the impact of the parameter dimension d and sample size n in the rate of convergence, as well as in nor- mal and bootstrap approximation errors. These results represent a useful complement to the asymptotic results under fixed design, and provide theoretical guarantees for the validity of Rademacher multiplier bootstrap in the problems of confidence construction and goodness-of-fit testing. Numerical studies lend strong support to our theory, and highlight the eectiveness of Rademacher bootstrap in terms of accuracy, reliability and computational eciency. Keywords: Quantile regression, multiplier bootstrap, robustness, concentration inequality, Bahadur representation, confidence interval, goodness-of-fit test. 1 Introduction 1.1 Quantile regression Since Koenker and Bassett’s seminal work (Koenker and Bassett, 1978), quantile regression has attracted enormous attention in statistics, econometrics and related fields primarily due to two ad- vantages over the (conditional) mean regression: (i) robustness against outliers in the response or heavy-tailed errors, and (ii) the ability to explore heterogeneity in the response that are associated with the covariates. We refer to the monograph by Koenker (2005) for an overview of the statistical theory and methods and computational aspects of quantile regression. Classical theory of quantile regression includes statistical consistency (see, e.g., Zhao, Rao and Chen (1993) for weak consistency and Bassett and Koenker (1986) for strong consistency), asymp- totic normality (Bassett and Koenker, 1978; Pollard, 1991) and Bahadur representation (Portnoy and Koenker, 1989; He and Shao, 1996; Arcones, 1996). A common thread of the previous work is that the regression estimators are studied under the fixed design setting, that is, the covariates { x i } n i=1 are deterministic vectors and satisfy some (asymptotic and non-asymptotic) conditions, and the only * Department of Mathematics, University of California, San Diego, La Jolla, CA 92093, USA. E-mail: [email protected]. Department of Mathematics, University of California, San Diego, La Jolla, CA 92093, USA. E-mail: [email protected]. 1
40

Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

Aug 16, 2020

Download

Documents

dariahiddleston
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
Page 1: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

Multiplier Bootstrap for Quantile Regression:Non-Asymptotic Theory under Random Design

Xiaoou Pan∗ and Wen-Xin Zhou†

Abstract

This paper establishes non-asymptotic concentration bound and Bahadur representa-tion for the quantile regression estimator and its multiplier bootstrap counterpart in therandom design setting. The non-asymptotic analysis keeps track of the impact of theparameter dimension d and sample size n in the rate of convergence, as well as in nor-mal and bootstrap approximation errors. These results represent a useful complementto the asymptotic results under fixed design, and provide theoretical guarantees for thevalidity of Rademacher multiplier bootstrap in the problems of confidence constructionand goodness-of-fit testing. Numerical studies lend strong support to our theory, andhighlight the effectiveness of Rademacher bootstrap in terms of accuracy, reliabilityand computational efficiency.

Keywords: Quantile regression, multiplier bootstrap, robustness, concentration inequality, Bahadurrepresentation, confidence interval, goodness-of-fit test.

1 Introduction

1.1 Quantile regression

Since Koenker and Bassett’s seminal work (Koenker and Bassett, 1978), quantile regression hasattracted enormous attention in statistics, econometrics and related fields primarily due to two ad-vantages over the (conditional) mean regression: (i) robustness against outliers in the response orheavy-tailed errors, and (ii) the ability to explore heterogeneity in the response that are associatedwith the covariates. We refer to the monograph by Koenker (2005) for an overview of the statisticaltheory and methods and computational aspects of quantile regression.

Classical theory of quantile regression includes statistical consistency (see, e.g., Zhao, Rao andChen (1993) for weak consistency and Bassett and Koenker (1986) for strong consistency), asymp-totic normality (Bassett and Koenker, 1978; Pollard, 1991) and Bahadur representation (Portnoy andKoenker, 1989; He and Shao, 1996; Arcones, 1996). A common thread of the previous work is thatthe regression estimators are studied under the fixed design setting, that is, the covariates xi

ni=1 are

deterministic vectors and satisfy some (asymptotic and non-asymptotic) conditions, and the only

∗Department of Mathematics, University of California, San Diego, La Jolla, CA 92093, USA. E-mail:[email protected].

†Department of Mathematics, University of California, San Diego, La Jolla, CA 92093, USA. E-mail:[email protected].

1

Page 2: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

randomness arises from the regression errors εini=1. A comprehensive review of the asymptotic

theory under fixed design can be found in Sections 4.1–4.3 of Koenker (2005).In contrast to fixed designs, more recent work in statistics has emphasized non-asymptotic re-

sults in the random design setting, where the covariates xini=1 are treated as random vectors (Hsu,

Kakade and Zhang, 2014; Wainwright, 2019). This additional randomness increases the complexityof the model, and makes theoretical analysis more subtle because the empirical processes involvednow depend on the random covariates with dimensionality possibly growing with the sample size.As stated in Hsu, Kakade and Zhang (2014), a major difference between fixed and random designsis that the fixed design setting does not directly address out-of-sample prediction. Specifically, afixed design analysis assesses the accuracy of the estimator on the observed data, while the predic-tive performance on unseen data is of primary concern of a random design analysis. Even thoughextensive studies have been carried out on ordinary and regularized least squares estimators (Hsu,Kakade and Zhang, 2014; Wainwright, 2019), it is not naturally clear whether similar results remainvalid for quantile regression. A main difficulty is that the quantile loss is piecewise linear, and henceits “curvature energy” is concentrated in a single point. This is substantially different from otherpopular regression loss functions, such as the squared loss and Huber loss, which are at least locallystrongly convex. The lack of smoothness and strong convexity makes it much more challenging toestablish non-asymptotic theory for quantile regression under random designs.

In Section 2.1 of this paper, we will establish non-asymptotic concentration bound (Theorem 1)and Bahadur representation (Theorem 2) of the quantile regression estimator under mild conditionson the random predictor and noise variable. To prove Theorem 1, we propose a new device to provea local restricted strong convexity (RSC) property of the empirical quantile loss; see Proposition 2.The notion of RSC was introduced by Negahban et al. (2012) to analyze convex regularized M-estimators, and extended by Loh and Wainwright (2015) to the case of nonconvex functions. Thusfar the RSC property has only been established for locally strongly convex and twice differentiableloss functions (Loh and Wainwright, 2015; Pan, Sun and Zhou, 2019). New techniques are thereforerequired to deal with piecewise linear functions, typified by the quantile loss and hinge loss. Theproof of Theorem 2, the Bahadur representation, builds on the concentration bound in Theorem 1along with techniques from empirical process theory. These results are non-asymptotic with expliciterrors, which allow to track the impact of the parameter dimension d and of the sample size n inquantile regression. These non-asymptotic results, to the best of our knowledge, are new to theprevious asymptotic results under fixed designs.

1.2 Statistical inference for quantile regression

In addition to the finite sample theory of standard quantile regression, we are also interested intwo fundamental statistical inference problems: (i) the construction of confidence intervals, and(ii) goodness-of-fit test. Broadly speaking, inference of quantile regression can be categorized intotwo classes: normal calibration and bootstrap calibration (resampling) methods. Normal calibra-tion heavily depends on either the estimation of 1/ fε|x(0), also known as the sparsity, where fε|x(·)is the conditional density function of ε given x, or the regression rank scores (Gutenbrunner andJureckova, 1992). Resampling, or bootstrap calibration methods (Efron, 1979), are commonly usedfor quantile regression inference because they are more robust against heteroscedastic errors andbypass the estimation of sparsity although at the cost of computing time. Over the past two decades,

2

Page 3: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

various bootstrap calibration methods have been developed for constructing confidence intervals,including the residual bootstrap and pairwise bootstrap (see Section 9.5 of Efron and Tibshirani(1994)), bootstrapping pivotal estimation functions method (Parzen, Wei and Ying, 1994), Markovchain marginal bootstrap (He and Hu, 2002; Kocherginsky, He and Mu, 2005) and wild bootstrap(Feng, He and Hu, 2011). For relatively small samples or in the presence of heteroscedastic er-rors, resampling methods have proven to outperform calibration through the normal approximation.Therefore, in this paper we only focus on the resampling method.

Among a variety of bootstrap methods, we are primarily interested in the multiplier bootstrap,also known as the weighted bootstrap, which is one of the most widely used inference tools for con-structing confidence intervals and measuring the significance of a test. The theoretical validity ofthe empirical bootstrap (Efron, 1979) is typically guaranteed by the bootstrapped law of large num-bers and central limit theorem; see, for example, Gine and Zinn (1990), Arcones and Gine (1992),Praestgaard and Wellner (1993) and Wellner and Zhan (1996), among others. Rigorous theoreticalguarantees of the multiplier bootstrap for M-estimation can be found in Chatterjee and Bose (2005)and Ma and Kosorok (2005), in which

√n-consistency and asymptotic normality are established.

See also Cheng and Huang (2010) for extensions to general semi-parametric models. It has sincebecome an effective and nearly universal inference tool for both parametric and semi-parametricM-estimations. We refer to Spokoiny and Zhilova (2015) for the use of multiplier bootstrap onconstructing likelihood-based confidence sets, and Chen and Zhou (2019) for a systematic study ofmultiplier bootstrap for adaptive Huber regression (Sun, Zhou and Fan, 2019) with applications tolarge-scale multiple testing for heavy-tailed data.

As stated in the previous section, the major theoretical challenge arises from the lack of smooth-ness and strong convexity of the quantile loss. New techniques are in demand. In Section 2.2, wewill first revisit the multiplier bootstrap in the problem of confidence estimation for quantile re-gression. Next, we will provide new non-asymptotic theory for bootstrap estimators, including theconditional deviation bound (Theorem 4) and Bahadur representation (Theorem 5) conditioned ondata already seen. We justify the validity of the multiplier bootstrap via a distributional approxima-tion result (Theorem 6), which characterizes the difference in distribution between the regressionestimator and its bootstrap counterpart. In Section 2.3, we further disucss the use of multiplier boot-strap on goodness-of-fit testing, extending the special case of median regression studied by Chenet al. (2008).

1.3 Notation

Let us summarize our notation. For every integer k ≥ 1, we use Rk to denote the the k-dimensionalEuclidean space. The inner product of any two vectors u = (u1, . . . , uk)ᵀ, v = (v1, . . . , vk)ᵀ ∈ Rk

is defined by uᵀv = 〈u, v〉 =∑k

i=1 uivi. We use ‖ · ‖p (1 ≤ p ≤ ∞) to denote the `p-norm in Rk:‖u‖p = (

∑ki=1 |ui|

p)1/p and ‖u‖∞ = max1≤i≤k |ui|. For k ≥ 2, Sk−1 = u ∈ Rk : ‖u‖2 = 1 denotes theunit sphere in Rk.

Throughout this paper, we use bold capital letters to represent matrices. For k ≥ 2, Ik representsthe identity/unit matrix of size k. For any k×k symmetric matrix A ∈ Rk×k, ‖A‖2 is the operator normof A, and we use λA and λA to denote the minimal and maximal eigenvalues of A, respectively. For apositive semidefinite matrix A ∈ Rk×k, ‖·‖A denotes the norm linked to A given by ‖u‖A = ‖A1/2u‖2,u ∈ Rk. Moreover, given r ≥ 0, define the Euclidean ball and ellipse as Bk(r) = u ∈ Rk : ‖u‖2 ≤ r

3

Page 4: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

and BA(r) = u ∈ Rk : ‖u‖A ≤ r, respectively. For any integer d ≥ 1, we write [d] = 1, . . . , d. Forany set S, we use |S| to denote its cardinality, i.e. the number of elements in S.

2 Random Design Quantile regression

2.1 Finite sample theory under random design

We consider a response variable y and d-dimensional covariates x = (x1, . . . , xd)ᵀ such that theτ-th (0 < τ < 1) conditional quantile of y given x is given by F−1

y|x(τ|x) = 〈x,β∗〉, where β∗ =

(β∗1, . . . , β∗d)ᵀ ∈ Rd. Here we assume x1 ≡ 1 so that β∗1 represents the intercept. Let (yi, xi)ni=1 be

independent and identically distributed (iid) data vectors from (y, x). The preceding model assump-tion is equivalent to

yi = 〈xi,β∗〉 + εi, (1)

where εi’s are independent noise variables that satisfy P(εi ≤ 0 | xi) = τ. The quantile regressionestimator of β∗ is then defined as

β = β(τ) ∈ argminβ∈Rd

Qn(β), (2)

where

Qn(β) =1n

n∑i=1

ρτ(yi − 〈xi,β〉) with ρτ(u) = uτ − I(u < 0) (3)

is the empirical loss. The loss function ρτ is known as the “check function” or “pinball loss”.This section presents two non-asymptotic results, the concentration inequality and Bahadur rep-

resentation, for the quantile regression estimator under random design. We refer to Chapter 4 ofKoenker (2005) for the classical fixed design and asymptotic analysis of quantile regression. Seealso Remark 2 and Table 1 below for a comparison of quantile regression and smooth robust regres-sion in terms of the scalings of the pair (n, d).

First, we specify the conditions on the random pair (x, ε) under which the analysis applies.

Condition 1 (Random design). The random predictor x ∈ Rd is sub-Gaussian: there exists υ0 ≥ 1such that P(|〈u, x〉| ≥ υ0‖u‖Σ · t) ≤ 2e−t2/2 for all u ∈ Rd and t ≥ 0, where Σ = E(xxᵀ).

Condition 1 is satisfied for a class of multivariate distributions. Typical examples include:(i) Multivariate Gaussian and (symmetric) Bernoulli distributions, (ii) uniform distribution on thesphere in Rd with center at the origin and radius

√d, (iii) uniform distribution on the Euclidean ball,

and (iv) uniform distribution on the unit cube [−1, 1]d. The constant υ0 is dimension-free, and thuscan be viewed as an absolute constant. See Chapter 6 in Wainwright (2019) and references thereinfor further discussion of sub-Gaussian distributions in higher dimensions.

Condition 2 (Regularity condition on error distribution). Let fε|x(·) be the conditional probabilitydensity function of ε given x, which is continuous on its support. Moreover, there exist constantsf ≥ f > 0 and L0 > 0 such that

f ≤ fε|x(0) ≤ f and | fε|x(u) − fε|x(0)| ≤ L0|u| for all u ∈ R, almost surely.

4

Page 5: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

Condition 2 on the conditional density function of ε given x is standard and routinely used inthe study of quantile regression.

Throughout this paper, “.” stands for “≤” up to constants that are independent of (n, d) butmay depend on the constants in Conditions 1 and 2. Our first main result characterizes the non-asymptotic deviation of the quantile regression estimator.

Theorem 1. Assume Conditions 1 and 2 hold. Then, for any t ≥ 0, the quantile regression estimatorβ = β(τ) (0 < τ < 1) given in (2) satisfies

‖β − β∗‖Σ ≤c1

f

√d + t

n(4)

with probability at least 1 − 2e−t as long as n ≥ c2L20 f −4(d + t), where c1, c2 > 0 are constants

depending only on υ0.

The following theorem provides a non-asymptotic version of the Bahadur representation for thequantile regression estimator; see Section 4.3 in Koenker (2005).

Theorem 2. Suppose that, in addition to the conditions in Theorem 1, supu∈R | fε|x(u)| ≤ M0 almostsurely for some M0 > 0. Then, for any t ≥ 0,∥∥∥∥∥S1/2(β − β∗) + S−1/2 1

n

n∑i=1

xiI(εi ≤ 0) − τ

∥∥∥∥∥2

≤ c3

(d + t)1/4(d log n + t)1/2

n3/4 + (d + log n)1/2 d log nn

+ (d log n)1/2 tn

(5)

with probability at least 1− 4e−t whenever n ≥ c2L20 f −4(d + t), where S = E fε|x(0)xxᵀ, and c3 > 0

is a constant depending only on (υ0, f , f , L0,M0).

Remark 1. With some basic analysis, the property that supu∈R | fε|x(u)| ≤ M0 almost surely is a con-sequence of Condition 2 with M0 depending implicitly on ( f , L0). Hence introducing the constantM0 is not to initiate an additional assumption, but to simplify the theorem and its proof.

The significance of Bahadur representation lies in expression of a complicated nonlinear es-timator as a normalized sum of independent random variables from which asymptotically normalbehavior follows. To validate this point, the following result provides a Berry-Esseen bound for anylinear contrast of the quantile regression estimator.

Theorem 3. Let λ ∈ Rd be a deterministic vector that defines a linear contrast of interest. Underthe conditions of Theorem 2, it holds that

supx∈R

∣∣∣P(n1/2〈λ, β − β∗〉 ≤ x)− Φ(x/στ)

∣∣∣ . (d + log n)1/4(d log n)1/2n−1/4, (6)

where σ2τ = τ(1 − τ)‖S−1λ‖2

Σand Φ(·) denotes the standard normal distribution function.

Remark 2 (Large-d asymptotics). A broader view of classical asymptotics recognizes that the para-metric dimension of appropriate model sequences may tend to infinity with the sample size; that isd = dn → ∞ as n → ∞. Such considerations, however, are rarely found in the quantile regression

5

Page 6: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

literature. In the standard quantile regression setting, Welsh (1989) shows that d3(log n)2/n → 0suffices for a normal approximation, which provides some support to the viability of observed ratesof parametric growth in the applied literature (Koenker, 1988).

In the (sub-Gaussian) random design setting, the obtained non-asymptotic Bahadur representa-tion (5) with t = log n reads:

n1/2(β − β∗) = S−1 1√

n

n∑i=1

τ − I(εi ≤ 0)

xi

+ OP

d3/4(log n)1/2 + d1/2(log n)3/4

n1/4 +d3/2 log n + d(log n)3/2

n1/2

.

Combined with a multivariate central limit theorem (Portnoy, 1986) or Theorem 3, this shows thatthe normal approximation holds as long as d3(log n)2/n → 0, which matches the scaling underfixed design although the proofs are entirely different. For smooth robust regression estimators, thescaling conditions required for asymptotic normality can be weakened. A prototypical example isHuber’s M-estimator. Note that the Huber loss has an absolutely continuous derivative, and is twicedifferentiable except at two points. Portnoy (1985) obtains the scaling condition (d log n)3/2/n→ 0that validates asymptotic normality when the predictors x1, . . . , xn form a sample from a mixedmultivaraite normal distribution in Rd. In the case of random, non-Gaussian predictors and ofsymmetric noise, d2/n is necessary for normal approximation; see Portnoy (1985, 1986).

Table 1: Summary of scaling conditions required for normal approximation under the Huber andpinball loss functions.

Loss function Design Scaling conditionHuber loss (Portnoy, 1985) Mixed Gaussian (with symmetric noise) (d log n)3/2 = o(n)Huber loss (Portnoy, 1986) Fixed design (with symmetric noise) d2 = o(n)Huber loss (Chen and Zhou, 2019) Sub-Gaussian (with asymmetric noise) d2 = o(n)Pinball loss (Welsh, 1989) Fixed design d3(log n)2 = o(n)Pinball loss (this work) Sub-Gaussian d3(log n)2 = o(n)

2.2 Multiplier bootstrap and confidence estimation

Let Rn = e1, . . . , en be a sequence of independent Rademacher random variables that are indepen-dent of the observed dataDn = (yi, xi)ni=1. Specifically, ei ∈ −1, 1 and satisfies P(ei = 1) = P(ei =

−1) = 1/2. Randomly perturb the empirical loss Qn(β) = (1/n)∑n

i=1 ρτ(yi − 〈xi,β〉) by multiplyingits summands with wi := ei + 1, we obtain the bootstrapped loss function

Q[n(β) :=

1n

n∑i=1

wi ρτ(yi − 〈xi,β〉), β ∈ Rd. (7)

Note that wi ∈ 0, 2 satisfies E(wi) = 1 and var(wi) = 1. Moreover, the bootstrapped loss Q[n :

Rd 7→ [0,∞) is also convex.Let E∗(·) = E(· |Dn) and P∗(·) = P(· |Dn) be the conditional expectation and probability given

Dn, respectively. Then we have E∗Q[n(β) = Qn(β) for any β ∈ Rd. This indicates that the quantile

6

Page 7: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

estimator β(τ) = (β1, . . . , βd)ᵀ in theDn-world is the target parameter in the bootstrap world:

argminβ∈Rd

E∗Q[n(β) = argmin

β∈RdQn(β) = β(τ).

This simple observation motivates the following multiplier bootstrap estimator:

β[ = β[(τ) ∈ argminβ∈Rd

Q[n(β). (8)

Let 1 − α ∈ (0, 1) be a prespecified confidence level. Based on the bootstrap statistic β[ =

(β[1, , . . . , β[d)ᵀ, we consider three methods to construct bootstrap confidence intervals.

(i) (Efron’s percentile method). For every 1 ≤ j ≤ d and q ∈ (0, 1), let ζ j,q be the (conditional)upper q-quantile of β[j, that is,

ζ j,q = infz ∈ R : P∗(β[j > z) ≤ q

. (9)

Efron’s percentile interval is of the form

Iperj = [ζ j,1−α/2, ζ j,α/2], j = 1, . . . , d. (10)

(ii) (Normal interval). The second method is the normal interval:

Inormj = [β j − zα/2seboot

j , β j + zα/2sebootj ], j = 1, . . . , d, (11)

where sebootj is the conditional standard deviation of β[j given Dn, and zα/2 is the upper α/2-

quantile of the standard normal distribution.

(iii) (Pivotal interval). The third method, which uses the conditional distribution of β[(τ) − β(τ)to approximate the distribution of the pivot β(τ) − β∗, is the pivotal interval. Specifically, the1 − α bootstrap pivotal confidence intervals for β∗j’s are

Ipivj = [2β j − ζ j,α/2, 2β j − ζ j,1−α/2], j = 1, . . . , d. (12)

In fact, there is a simple connection between the bootstrap pivotal interval and the percentileinterval: the percentile interval is the pivotal interval reflected about the point β j.

Before we formally investigate the theoretical properties of the bootstrap estimator β[(τ), recallthe Bahadur representation of β(τ):

β(τ) = β∗ +1n

n∑i=1

τ − I(εi ≤ 0)

S−1xi + rn,

where rn is the higher-order remainder term. Heuristically, the bootstrap estimator β[(τ) can beviewed as the quantile regression estimator of β(τ) in the bootstrap world under the model yi =

〈xi, β(τ)〉 + ε[i . According to the Bahadur representation, it can be written as yi ≈ 〈xi,β∗〉 +

(1/n)∑n

i=1〈xi,S−1xi〉τ − I(εi ≤ 0). The accuracy of the percentile interval, however, relies onthe property that β[τ is randomly concentrated around β∗. Motivated by this observation and the

7

Page 8: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

finite-sample correction method used in Feng, He and Hu (2011), for practical implementation wereplace the original response yi in the multiplier bootstrap by yi = yi − fε(0)−1hiτ − I (εi ≤ 0),where hi = xᵀi (

∑nj=1 x jx

ᵀj )−1xi and fε(0) is estimated from the fitted residuals εi = yi − 〈xi, β(τ)〉. In

particular, the density estimate fε employs the adaptive kernel method (Silverman, 1986), which isimplemented in the quantreg package as function akj (Koenker, 2019).

Back to β[ defined in (8), the following result provides a conditional deviation inequality, con-ditioned on some event that occurs with high probability.

Theorem 4. Assume Conditions 1 and 2 hold. For any t ≥ 0, there exists some event Et withPE(t) ≥ 1 − 2e−t such that the bound (4) holds on E(t), and with P∗-probability at least 1 − e−t

conditioned on E(t), the bootstrap estimator β[ = β[(τ) (0 < τ < 1) given in (8) satisfies

‖β[ − β∗‖Σ ≤ c4

√d + t

n(13)

as long as n ≥ c5(d + t), where c4, c5 > 0 are constants depending only on (υ0, f , L0).

To characterize the distribution of β[ conditional on the initial sample Dn = (yi, xi)ni=1, weestablish in the following result a conditional Bahadur representation under P∗.

Theorem 5. Suppose that the conditions in Theorem 2 hold. Under the scaling n & d + log n, thereexists some event En with P(En) ≥ 1−4n−1 such that, with P∗-probability at least 1−n−1 conditionedon En,

S1/2(β[ − β) = S−1/2 1n

n∑i=1

eixiτ − I(εi ≤ 0)

+ r[n, (14)

where r[n = r[n((ei, yi, xi)ni=1) satisfies ‖r[n‖2 = OP∗(χn), and χn = χn((yi, xi)ni=1) is such thatχn = OP(d + log n)1/4(d log n)1/2n−3/4 + (d + log n)1/2d log(n) n−1.

We end this section with a distributional approximation result, which establishes the validity ofthe (Rademacher) multiplier bootstrap for approximating the distributions of linear contrasts of thequantile regression estimator.

Theorem 6. Let λ ∈ Rd be an arbitrary d-vector defining a linear contrast of interest. AssumeConditions 1 and 2 hold, and that the parameter dimension d = dn, as a function of the sample sizen, satisfies the scaling d3(log n)2 = o(n). Then, as n→ ∞,

supx∈R

∣∣∣P(n1/2〈λ, β − β∗〉 ≤ x)− P∗

(n1/2〈λ, β[ − β〉 ≤ x

)∣∣∣ P−→ 0. (15)

2.3 Goodness-of-fit testing

The multiplier bootstrap method can also be applied to goodness-of-fit testing for quantile regres-sion. Under model (1), consider a subset Ω0 ⊆ Rd, and we wish to test

H0 : β∗ ∈ Ω0 versus H1 : β∗ ∈ Rd \Ω0. (16)

8

Page 9: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

We first construct the test statistics based on the empirical loss Qn(β) defined in (3). Let β bequantile estimator under the full model (2), and set β0 ∈ argminβ∈Ω0

Qn(β). The test statistic isdefined as

Tn = Qn(β0) − Qn(β).

In the bootstrap world, we intend to mimic the distribution of Tn using that of Q[n(β) defined in

(7). Let β[ ∈ argminβ∈Rd Q[n(β) and β[0 ∈ argminβ∈Ω0

Q[n(β) be the bootstrap statistics in the full

model and null model, respectively. Motivated by Chen et al. (2008), we consider the bootstrap teststatistic

T [n =

Q[

n(β[0) − Q[n(β[)

Q[

n(β0) − Q[n(β)

.

See Remark 2 therein for the intuition behind this construction. The conditional distribution of T [n

given the data then serves as an approximation of the distribution of Tn. For every q ∈ (0, 1), let γq

be the (conditional) upper q-quantile of T [n, that is,

γq = infz ∈ R : P∗(T [

n > z) ≤ q,

Consequently, for significance level α ∈ (0, 1), we reject H0 in (16) whenever Tn > γα.It is worth noticing that the above method was first proposed and studied by Chen et al. (2008)

using standard exponential weights in the case of median regression, and can be implemented bythe R package quantreg (Koenker, 2019). As discussed earlier, the Rademacher multiplier bootstrapis computationally more attractive and also has provable finite-sample guarantees. See Sections 3.2and B.2 for a thorough numerical comparison.

3 Numerical Experiments

In this section, we conduct numerical experiments to compare the multiplier bootstrap on con-structing confidence intervals and goodness-of-fit testing with some well-known existing methodsfor quantile regression. Our computational results are reproducible using codes available fromhttps://github.com/XiaoouPan/mbQuantile.

3.1 Confidence estimation

We first consider the problem of confidence estimation. The limiting distribution of the quantile re-gression estimator involves the density of the errors, making the non-resampling (plug-in) inferenceprocedure unstable and unreliable. We refer to Kocherginsky, He and Mu (2005) for an overviewand numerical comparisons between plug-in and resampling methods. In this paper, we focus onthe following bootstrap calibration methods:

• pair: pairwise bootstrap by resampling (yi, xi)ni=1 in pairs with replacement (Section 9.5 ofEfron and Tibshirani (1994));

• pwy: a resampling method based on pivotal estimating functions (Parzen, Wei and Ying,1994);

• wild: wild bootstrap with Rademacher weights (Feng, He and Hu, 2011);

• mb-per: multiplier bootstrap percentile method defined in (10);

9

Page 10: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

• mb-norm: multiplier bootstrap normal-based method defined in (11).

The first three methods can be directly implemented using the R package quantreg (Koenker, 2019).To better evaluate the performance of these methods under various environments, we generate

data vectors (yi, xi)ni=1 from two types of linear models:

1. (Homoscedastic model):

yi = β∗0 + 〈xi,β∗〉 + εi, i = 1, . . . , n; (17)

2. (Heteroscedastic model):

yi = β∗0 + 〈xi,β∗〉 +

2 exp(xi1)1 + exp(xi1)

εi, i = 1, . . . , n. (18)

Here we use separate notations to differentiate the intercept β∗0 and coefficient vector β∗ ∈ Rd. Foreach model, we consider three error distributions as follows.

1. t2: εi ∼ t2;

2. Normal mixture type I: εi = az1 + (1 − a)z2, where a ∼ Ber(0.5), z1 ∼ N(−1, 1) and z2 ∼

N(1, 1);

3. Normal mixture type II: εi = az1 + (1 − a)z2, where a ∼ Ber(0.9), z1 ∼ N(0, 1) and z2 ∼

N(0, 52).

Moreover, we generate random predictors with three different covariance structures:

1. Independent design: xi ∼ N(0, Id) for i = 1, . . . , n;

2. Weakly correlated design: first generate a covariance matrix Σ = (σ jk)1≤ j,k≤d with diagonalentries σ j j independently drawn from Unif(0.5, 1) and σ jk = 0.5| j−k|(σ j jσkk)1/2 if j , k, andthen generate xi’s independently from N(0,Σ);

3. Equally correlated design: first generate a covariance matrix Σ = (σ jk)1≤ j,k≤d with diagonalentries σ j j independently drawn from Unif(0.5, 1) and σ jk = 0.5(σ j jσkk)1/2 if j , k, and thengenerate xi’s independently from N(0,Σ).

We set β∗0 = 2, β∗ = (2, . . . , 2)ᵀ and (n, d) = (200, 10). The confidence level is taken to be 1 − α ∈80%, 90%, 95%. All of the five methods are carried out using B = 1000 bootstrap samples. Tables2, 3, and 7-10 in Section B.1 of the Appendix display the average coverage probabilities and averageinterval widths over all the regression coefficients based on 200 Monte Carlo simulations.

10

Page 11: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

Independent Gaussian designCoverage probability Width

α pair pwy wild mb-per mb-norm pair pwy wild mb-per mb-norm0.05: 0.963 0.966 0.930 0.967 0.935 0.620 0.635 0.554 0.542 0.5400.1: 0.922 0.930 0.873 0.925 0.873 0.520 0.533 0.465 0.451 0.4530.2: 0.828 0.844 0.776 0.824 0.769 0.405 0.415 0.362 0.347 0.353

Weakly correlated Gaussian designCoverage probability Width

α pair pwy wild mb-per mb-norm pair pwy wild mb-per mb-norm0.05: 0.962 0.966 0.921 0.964 0.926 0.920 0.941 0.815 0.806 0.8020.1: 0.915 0.921 0.867 0.917 0.873 0.772 0.790 0.684 0.670 0.6730.2: 0.821 0.835 0.769 0.821 0.767 0.601 0.615 0.533 0.515 0.525

Equally correlated Gaussian designCoverage probability Width

α pair pwy wild mb-per mb-norm pair pwy wild mb-per mb-norm0.05: 0.964 0.968 0.925 0.967 0.930 0.980 1.004 0.868 0.860 0.8560.1: 0.913 0.926 0.861 0.921 0.867 0.823 0.842 0.729 0.714 0.7180.2: 0.826 0.831 0.766 0.816 0.767 0.641 0.656 0.568 0.550 0.559

Table 2: Average coverage probabilities and confidence interval (CI) widths over all the coefficientsunder homoscedastic model (17) with type I mixture normal error.

Independent Gaussian designCoverage probability Width

α pair pwy wild mb-per mb-norm pair pwy wild mb-per mb-norm0.05: 0.972 0.974 0.946 0.966 0.945 0.542 0.555 0.481 0.478 0.4740.1: 0.936 0.938 0.898 0.920 0.905 0.454 0.466 0.404 0.395 0.3980.2: 0.861 0.870 0.811 0.828 0.805 0.354 0.363 0.315 0.303 0.310

Weakly correlated Gaussian designCoverage probability Width

α pair pwy wild mb-per mb-norm pair pwy wild mb-per mb-norm0.05: 0.968 0.970 0.941 0.966 0.938 0.820 0.840 0.729 0.722 0.7160.1: 0.932 0.933 0.885 0.913 0.886 0.688 0.705 0.612 0.597 0.6010.2: 0.849 0.859 0.791 0.816 0.785 0.536 0.549 0.476 0.458 0.468

Equally correlated Gaussian designCoverage probability Width

α pair pwy wild mb-per mb-norm pair pwy wild mb-per mb-norm0.05: 0.968 0.974 0.938 0.964 0.941 0.877 0.898 0.778 0.772 0.7650.1: 0.928 0.932 0.881 0.917 0.883 0.736 0.754 0.653 0.638 0.6420.2: 0.839 0.847 0.787 0.804 0.786 0.573 0.587 0.509 0.490 0.500

Table 3: Average coverage probabilities and CI widths over all the coefficients under heteroscedasticmodel (18) with type I mixture normal error.

From Tables 2, 3, and 7-10 (in the Appendix) we find that all the bootstrap methods preservenominal levels, while pairwise bootstrap and bootstrap based on estimating functions (pwy) tendto be more conservative with wider intervals, and wild bootstrap loses coverage probability undersome cases; see Table 2. Across all the settings, the multiplier bootstrap methods (percentile andnormal-based) provide desirable results in terms of both accuracy (narrow width) and reliability(high confidence). It is worth noticing that the normal-based confidence interval (mb-norm) tendsto have lower coverage probabilities compared with the percentile method. As the sample sizeincreases, the coverage probability of mb-norm approaches the nominal level gradually; see Table 4.After taking into account the interval width, we recommend the multiplier bootstrap percentilemethod that has the best overall performance.

11

Page 12: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

Independent Gaussian designn = 200 n = 500 n = 1000

α mb-per mb-norm mb-per mb-norm mb-per mb-norm0.05: 0.967 (0.542) 0.935 (0.540) 0.950 (0.346) 0.923 (0.346) 0.960 (0.247) 0.948 (0.247)0.1: 0.925 (0.451) 0.873 (0.453) 0.904 (0.289) 0.871 (0.290) 0.923 (0.206) 0.895 (0.207)0.2: 0.824 (0.347) 0.769 (0.353) 0.817 (0.224) 0.768 (0.226) 0.824 (0.160) 0.792 (0.161)

Weakly correlated Gaussian designn = 200 n = 500 n = 1000

α mb-per mb-norm mb-per mb-norm mb-per mb-norm0.05: 0.964 (0.806) 0.926 (0.802) 0.954 (0.512) 0.933 (0.512) 0.966 (0.364) 0.948 (0.364)0.1: 0.917 (0.670) 0.873 (0.673) 0.905 (0.428) 0.875 (0.430) 0.913 (0.305) 0.899 (0.306)0.2: 0.821 (0.515) 0.767 (0.525) 0.798 (0.331) 0.770 (0.335) 0.824 (0.236) 0.799 (0.238)

Equally correlated Gaussian designn = 200 n = 500 n = 1000

α mb-per mb-norm mb-per mb-norm mb-per mb-norm0.05: 0.967 (0.860) 0.930 (0.856) 0.960 (0.547) 0.941 (0.546) 0.961 (0.389) 0.944 (0.389)0.1: 0.921 (0.714) 0.867 (0.718) 0.912 (0.456) 0.873 (0.458) 0.909 (0.326) 0.888 (0.327)0.2: 0.816 (0.550) 0.767 (0.559) 0.804 (0.353) 0.773 (0.357) 0.818 (0.253) 0.792 (0.255)

Table 4: Average coverage probabilities and CI widths (in brackets) over all the coefficients underhomoscedastic model (17) with type I mixture normal error.

Regarding computational complexity, for each bootstrap sample, pairwise and wild bootstrapssolve a quantile regression on a sample of size n, bootstrap based on estimating functions (pwy)solves a quantile regression of size n + 1, while multiplier bootstrap solves a quantile regressionessentially on a subsample of size n/2 on average. In summary, the multiplier bootstrap provides acomputationally efficient way to construct confidence intervals with high precision and reliability.

3.2 Goodness-of-fit testing

In this section, we compare the multiplier bootstrap with classical non-resampling methods ongoodness-of-fit testing for quantile regression. Specifically, we consider the following methods:

• Wald: Wald test based on unrestricted estimator (Koenker and Bassett, 1982);

• rank: rank score test (Gutenbrunner et al., 1993);

• mb-exp: multiplier bootstrap with exponential weights (Chen et al., 2008);

• mb-Rad: multiplier bootstrap with Rademacher weights.

The first three methods are included in the R package quantreg (Koenker, 2019).We generate data vectors the same way as in Section 3.1. Moreover, we set (n, d) = (200, 15),

and the confidence level is taken to be 1 − α ∈ 90%, 95%, 99%. We consider testing

H0 : β∗j = 0, for j = 1, . . . , 15 versus H1 : β∗j , 0, for some j.

To assess the overall performance, we employ the following three measurements:

1. Type I error under null model: β∗ = 0;

2. Power under sparse and strong signal: β∗1 = 0.5, and β∗j = 0 for j = 2, 3, . . . , 15;

3. Power under dense and weak signal: β∗j = 0.1 for j = 1, 2, . . . , 10, and β∗j = 0 for j =

11, 12, . . . , 15.

12

Page 13: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

The two resampling methods (mb-exp and mb-Rad) are carried out using B = 1000 bootstrapsamples. Tables 5, 6 and 11-14 in Section B.2 of the Appendix display the average type I error andpower over 200 Monte Carlo simulations.

Independent Gaussian designType I error under null model Power under sparse model Power under dense model

α Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad0.01 0.370 0.000 0.000 0.005 0.805 0.185 0.295 0.330 0.580 0.035 0.045 0.0750.05 0.490 0.025 0.055 0.050 0.915 0.460 0.570 0.540 0.725 0.150 0.315 0.3000.1 0.615 0.080 0.140 0.125 0.945 0.625 0.750 0.695 0.775 0.290 0.390 0.360

Weakly correlated Gaussian designType I error under null model Power under sparse model Power under dense model

α Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad0.01 0.300 0.010 0.005 0.010 0.650 0.115 0.230 0.250 0.710 0.160 0.210 0.2300.05 0.450 0.060 0.060 0.055 0.790 0.350 0.465 0.435 0.820 0.380 0.500 0.4850.1 0.555 0.095 0.120 0.090 0.850 0.515 0.605 0.575 0.870 0.535 0.640 0.600

Equally correlated Gaussian designType I error under null model Power under sparse model Power under dense model

α Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad0.01 0.300 0.010 0.010 0.010 0.660 0.135 0.205 0.225 0.915 0.470 0.595 0.6150.05 0.450 0.060 0.060 0.055 0.790 0.325 0.400 0.385 0.960 0.755 0.825 0.8000.1 0.555 0.095 0.120 0.090 0.870 0.460 0.575 0.515 0.970 0.860 0.860 0.860

Table 5: Average type I error and power under homoscedastic model (17) with type I mixture normalerror.

Independent Gaussian designType I error under null model Power under sparse model Power under dense model

α Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad0.01 0.315 0.005 0.000 0.000 0.815 0.410 0.475 0.510 0.590 0.085 0.095 0.1100.05 0.435 0.035 0.030 0.030 0.930 0.685 0.755 0.725 0.705 0.275 0.305 0.3050.1 0.510 0.065 0.065 0.050 0.955 0.785 0.840 0.810 0.780 0.415 0.415 0.380

Weakly correlated Gaussian designType I error under null model Power under sparse model Power under dense model

α Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad0.01 0.380 0.010 0.005 0.005 0.810 0.200 0.330 0.365 0.790 0.235 0.260 0.2950.05 0.480 0.060 0.055 0.050 0.885 0.525 0.610 0.565 0.865 0.510 0.595 0.5650.1 0.565 0.110 0.115 0.090 0.905 0.655 0.740 0.700 0.910 0.680 0.725 0.690

Equally correlated Gaussian designType I error under null model Power under sparse model Power under dense model

α Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad0.01 0.350 0.010 0.005 0.005 0.690 0.205 0.270 0.300 0.960 0.610 0.715 0.7350.05 0.470 0.060 0.045 0.040 0.815 0.450 0.550 0.520 0.990 0.850 0.900 0.8800.1 0.535 0.125 0.115 0.100 0.865 0.610 0.695 0.640 0.990 0.900 0.935 0.935

Table 6: Average type I error and power under heteroscedastic model (18) with type I mixturenormal error.

From Tables (5) and (6) we see that the Wald test suffers from severe size distortion by rejectingmuch more often than it should, while the other three methods have type I errors close to the nominallevel. Under both sparse and dense alternatives, the multiplier bootstrap outperforms the rank scoretest with higher power throughout all the combinations of design and error distributions.

To further compare the power of the last three methods, we draw the power curve with graduallyincreasing signal strength under sparse and dense settings. Figure 1 is a visualization of Table 5and Table 6 with type I mixture normal error and independent design. The advantage of multiplier

13

Page 14: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

bootstrap over rank test is conspicuous under homoscedastic model, and multiplier bootstrap revealsperceptible advantage as signal gets stronger under heteroscedastic model.

0.2 0.4 0.6 0.8 1.0

0.0

0.2

0.4

0.6

0.8

1.0

Signal strength

Pow

er

rankmb-expmb-Rad

(a) Homoscedastic model (17) with sparse signal.

0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.350.0

0.2

0.4

0.6

0.8

1.0

Signal strength

Pow

er

rankmb-expmb-Rad

(b) Homoscedastic model (17) with dense signal.

0.0 0.2 0.4 0.6 0.8

0.0

0.2

0.4

0.6

0.8

1.0

Signal strength

Pow

er

rankmb-expmb-Rad

(c) Heteroscedastic model (18) with sparse signal.

0.00 0.05 0.10 0.15 0.20 0.25 0.30

0.0

0.2

0.4

0.6

0.8

1.0

Signal strength

Pow

er

rankmb-expmb-Rad

(d) Heteroscedastic model (18) with dense signal.

Figure 1: Power curves of the three methods under independent design and type I mixture normalerror with α = 0.05.

4 Proofs of Main Results

All the probabilistic bounds presented in the proof are non-asymptotic with explicit errors. Thevalues of the constants involved are obtained with the goal of making the proof transparent, andmay be improved by more careful calculations or under less general distributional assumptions onthe covariates and noise variables.

14

Page 15: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

4.1 Preliminaries

Recall that Qn(β) = (1/n)∑n

i=1 ρτ(yi − 〈xi,β〉) is the empirical quantile loss function. Since Qn :Rd → R is convex, we define its subdifferential ∂Qn by

∂Qn(β) =ξ ∈ Rd : Qn(β′) ≥ Qn(β) + 〈ξ,β′ − β〉 for all β ∈ Rd. (19)

A vector ξ ∈ ∂Qn(β) is called a subgradient of Qn in β. More specifically, the subdifferential ∂Qn isthe collection of vectors ξβ = (ξβ,1, . . . , ξβ,d)ᵀ satisfying, for j = 1 . . . , d,

ξβ, j = −τ

n

n∑i=1

xi jI(yi > 〈xi,β〉)

+1 − τ

n

n∑i=1

xi jI(yi < 〈xi,β〉) −1n

n∑i=1

xi jviI(yi = 〈xi,β〉), (20)

where vi ∈ [τ − 1, τ].Of particular interest is the subdifferential ∂Qn(β∗) under model (1). By (20), every vector

ξ = (ξ1, . . . , ξd)ᵀ ∈ ∂Qn(β∗) can be written as

ξ j = −τ

n

n∑i=1

xi jI(εi > 0) − (1 − τ)

+1 − τ

n

n∑i=1

xi jI(εi < 0) − τ −1n

n∑i=1

xi jviI(εi = 0), j = 1, . . . , d, (21)

where vi ∈ [τ − 1, τ].

Proposition 1. Assume Conditions 1 and 2 hold. Then, every subgradient ξβ∗ ∈ ∂Qn(β∗) satisfies

P

(‖Σ−1/2ξβ∗‖2 ≥ 3υ0

√2d + x

n

)≤ e−x, valid for any x ≥ 0.

The following proposition provides a form of the restricted strong convexity (RSC) for theempirical quantile loss function.

Proposition 2. Assume Conditions 1 and 2 hold. Then, for any t ≥ 0, it holds with probability atleast 1 − e−t/2 that

〈ξβ − ξβ∗ ,β − β∗〉 ≥

18

f ‖β − β∗‖2Σ − 4υ20‖β − β

∗‖Σ

√2(d + t)

n(22)

uniformly over β ∈ Rd satisfying 0 ≤ ‖β − β∗‖Σ ≤ f /(6L0υ20).

Propositions 1 and 2 provide the key ingredients to prove Theorems 1 and 2. Similarly, the finitesample performance of the multiplier bootstrap estimator relies on the corresponding properties ofthe weighted quantile loss function, which are given by Propositions 3 and 4 below.

Recall that P∗ and E∗ denote, respectively, the probability measure and expectation (over Rn =

eini=1) conditioning onDn = (yi, xi)ni=1. For i = 1, . . . , n, define

ζi = I(εi ≤ 0) − τ and zi = Σ−1/2xi, (23)

which satisfy E(ζi|xi) = 0, E(ζ2i |xi) = τ(1 − τ) and E(zi z

ᵀi ) = Id.

15

Page 16: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

Proposition 3. Assume Conditions 1 and 2 hold, and let ξ[ ∈ ∂Q[n(β∗). For any t > 0, there exists

some event G1(t) = G1(t;Dn) with PG1(t) ≥ 1− e−2t such that, with P∗-probability at least 1− e−2t

conditioned on G1(t),

‖Σ−1/2(ξ[ − E∗ξ[)‖2 ≤ 2

√d + t

n(24)

as long as n & d + t.

Similarly to Proposition 2, the following result establishes the restricted strong convexity for theweighted quantile loss function.

Proposition 4. Assume Conditions 1 and 2 hold. For any t ≥ 0, there exists some event G2(t) =

G2(t;Dn) such that PG2(t) ≥ 1 − e−t, and with P∗-probability at least 1 − e−t/2 conditioned onG2(t),

〈ξ[β − ξ[β∗ ,β − β

∗〉 ≥18

f ‖β − β∗‖2Σ − 8υ20‖β − β

∗‖Σ

√2(d + t)

n(25)

uniformly over β ∈ Rd satisfying 0 ≤ ‖β − β∗‖Σ ≤ f /(6L0υ20) as long as n & log(d) + t.

Proofs of Propositions 1–4 are placed in the Appendix.

4.2 Proof of Theorem 1

By the convexity of β 7→ Qn(β), β satisfies the first-order condition that ξβ = 0 for some ξβ ∈

∂Qn(β). The proof builds on the symmetrized Bregman divergence associated with Qn, defined as

D(β1,β2) = 〈ξβ1 − ξβ2 ,β1 − β2〉, for ξβ1 ∈ ∂Qn(β1), ξβ2 ∈ ∂Qn(β2).

By convexity, D(β1,β2) ≥ 0 for any subdifferentials ξβ1 and ξβ2 . Taking (β1,β2) = (β,β∗), we have

0 ≤ 〈ξβ − ξβ∗ , β − β∗〉 = 〈−ξβ∗ , β − β

∗〉 ≤ ‖Σ−1/2ξβ∗‖2‖β − β∗‖Σ, (26)

for any ξβ∗ ∈ ∂Qn(β∗). Starting with (26), we bound the left- and right-hand sides of (27) separately.To establish the lower bound, we use a localized argument (Sun, Zhou and Fan, 2019) and a newrestricted strong convexity property for the empirical quantile loss (Proposition 2).

Define the rescaled `2-ball BΣ(t) = β ∈ Rd : ‖β‖Σ ≤ t, t > 0. For some 0 < r ≤ f /(6L0υ20) to

be determined, define

η = supu ∈ [0, 1] : u(β − β∗) ∈ BΣ(r)

and β = β∗ + η(β − β∗).

By the above definition, η = 1 if β ∈ β∗ + BΣ(r), and η < 1 if β < β∗ + BΣ(r). In the latter case, wehave β ∈ β∗ + ∂BΣ(r). Applying Lemma C.1 in Sun, Zhou and Fan (2019) with slight modificationsyields the bound D(β,β∗) ≤ ηD(β,β∗), leading to

〈ξβ − ξβ∗ , β − β∗〉 ≤ η〈ξβ − ξβ∗ , β − β

∗〉, (27)

where ξβ∗ ∈ ∂Qn(β∗) and ξβ ∈ ∂Qn(β). This, together with the fact ξβ = 0 and Cauchy-Schwarzinequality, implies

〈ξβ − ξβ∗ , β − β∗〉 ≤ η〈−ξβ∗ , β − β

∗〉 ≤ ‖Σ−1/2ξβ∗‖2‖β − β∗‖Σ. (28)

16

Page 17: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

Note that (28) is a localized version of (26) because β falls in a local neighborhood of β∗.Setting δ = β − β∗ ∈ BΣ(r), it follows from Proposition 2 that

〈ξβ − ξβ∗, β − β∗〉 ≥

18

f ‖δ‖2Σ − 4υ20‖δ‖Σ

√2(d + t)

n

with probability at least 1 − e−t/2. Combining this with (27) and (28), and taking x = t > 0 inProposition 1, we obtain

18

f ‖δ‖2Σ <(4υ2

0 + 3υ0)‖δ‖Σ

√2(d + t)

n

with probability at least 1 − 2e−t. Canceling ‖δ‖Σ on both sides yields

‖δ‖Σ < r := 8 f −1(4υ20 + 3υ0)

√2(d + t)

n

with probability at least 1−2e−t as long as n ≥ CL20 f −4(d+t) for some constant C > 0 depending only

on υ0. Consequently, β falls in the interior of β∗ + BΣ(r), enforcing η = 1 and β = β ∈ β∗ + BΣ(r).Otherwise if β < β∗ + BΣ(r), we must have β on the boundary, i.e. ‖β − β∗‖Σ = r, which leads tocontradiction. This completes the proof.

4.3 Proof of Theorem 2

To begin with, define the “gradient” function ∇Qn : Rd → Rd as

∇Qn(β) =1n

n∑i=1

xiI(yi ≤ 〈xi,β〉) − τ

, β ∈ Rd. (29)

Recall from Condition 2 that the conditional distribution of ε given x is continuous. Lemma A.1 inRuppert and Carroll (1980) states that with probability one, there is no vector δ ∈ Rd and 1 ≤ i ≤ nsuch that εi = 〈xi, δ〉. It follows that with probability one, ξβ = ∇Qn(β) for any ξβ ∈ ∂Qn(β). Hence,we will treat ∇Qn as the gradient of Qn throughout the proof. Moreover, consider the populationloss EQn(β) = Eρτ(y− 〈x,β〉), whose gradient vector and Hessian matrix are given, respectively, by

∇EQn(β) = E[xI(ε ≤ 〈x,β − β∗〉) − τ

]and ∇2EQn(β) = E

fε|x(〈x,β − β∗〉)xxᵀ

.

Next, define the vector-valued random process

∆(β) = S−1/2∇Qn(β) − ∇Qn(β∗)− S1/2(β − β∗), (30)

where S = ∇2EQn(β∗) = E fε|x(0)xxᵀ. The goal is to bound ‖∆(β)‖2 uniformly over β in a localneighborhood of β∗. To this end, we deal with E∆(β) and ∆(β) − E∆(β) separately, starting withE∆(β). Applying the mean value theorem for vector-valued functions yields

E∆(β) = S−1/2⟨ ∫ 1

0∇2EQn(β∗t )dt,β − β∗

⟩− S1/2(β − β∗)

=

⟨S−1/2

∫ 1

0∇2EQn(β∗t )dt S−1/2 − Id,S1/2(β − β∗)

⟩, (31)

17

Page 18: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

where β∗t = (1 − t)β∗ + tβ and ∇2EQn(β∗t ) = E fε|x(t〈x,β − β∗〉)xxᵀ. For r > 0, define the localelliptic neighborhood of β∗ as ΘΣ(r) :=

β ∈ Rd : ‖β − β∗‖Σ ≤ r

. By Conditions 1 and 2, Σ is

positive definite and f ≤ fε|x(0) ≤ f , so that f Σ S f Σ. For δ = β − β∗ with β ∈ ΘΣ(r), theLipschitz continuity of fε|x ensures that∥∥∥S−1/2∇2EQn(β∗t )S−1/2 − Id

∥∥∥2 =

∥∥∥S−1/2E[

fε|x(t〈x, δ〉) − fε|x(0)xxᵀ

]S−1/2

∥∥∥2

≤ L0t · supu∈Bd(1)

E〈S−1/2x,u〉2|〈x, δ〉|

≤ f −1L0t ·

(sup

u∈Bd(1)E|〈Σ−1/2x,u〉|3

)2/3(E|〈x, δ〉|3

)1/3

≤ L0 f −1m3rt,

where mk := supu∈Bd(1) E|〈Σ−1/2x,u〉|k (for k ≥ 1) depends only on υ0 and k. Combining this with

(31), we obtain

supβ∈ΘΣ(r)

‖E∆(β)‖2 ≤12

L0 f −1 f 1/2m3r2. (32)

Turning to the stochastic term ∆(β) − E∆(β), define the centered gradient function

Rn(β) =1n

n∑i=1

(1 − E)I(〈xi,β − β

∗〉 ≥ εi) − τxi,

so that ∆(β) − E∆(β) = S−1/2Rn(β) − Rn(β∗). By a change of variable v = Σ1/2(β − β∗), we have

supβ∈ΘΣ(r)

‖∆(β) − E∆(β)‖2 ≤ f −1/2 supβ∈ΘΣ(r)

‖Σ−1/2Rn(β) − Rn(β∗)‖2

= f −1/2 supv∈Bd(r)

‖Σ−1/2Rn(β∗ + Σ−1/2v) − Rn(β∗)‖2

= f −1/2r−1 supu,v∈Bd(r)

〈Σ−1/2Rn(β∗ + Σ−1/2v) − Rn(β∗),u〉︸ ︷︷ ︸n−1/2∆0(u,v)

, (33)

where ∆0(u, v) = n−1/2 ∑ni=1(1 − E)〈zi,u〉I(εi ≤ 〈zi, v〉) − I(εi ≤ 0). To bound supu,v∈Bd(r) ∆0(u, v),

we first show its concentration around the mean, and then bound the mean via a maximal inequalityspecialized to VC type classes (see, e.g., Chapter 2.6 in van der Vaart and Wellner (1996)). Considerthe following two classes of real-valued functions on R × Rd:

F1 = (z0, z) 7→ 〈z,u〉 : u ∈ Bd(r) and F2 = (z0, z) 7→ I(〈z, v〉 − z0 ≥ 0) : v ∈ Bd(r). (34)

Moreover, define the function f0 : (z0, z) 7→ I(z0 ≤ 0), and write zi = (εi, zi) ∈ R×Rd for i = 1, . . . , n.Then, the supremum supu,v∈Bd(r) ∆0(u, v) can be written as the supremum of an empirical process:

supv,u∈Bd(r)

∆0(u, v) = supf∈F

1√

n

n∑i=1

f (zi) − E f (zi)︸ ︷︷ ︸Gn f

, (35)

where F = F1 · (F2 − f0) is the pointwise product of F1 and F2 − f0. Under the assumptionthat supu | fε|x(u)| ≤ M0 almost surely, we have, for each i ∈ [n], sup f∈F f ( zi) ≤ r‖zi‖2 and

18

Page 19: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

sup f∈F E f ( zi)2 ≤ M0 supu,v∈Bd(r) E〈zi,u〉2|〈zi, v〉| ≤ M0m3r3. By Lemma 2.2.2 in van der Vaartand Wellner (1996),∥∥∥∥∥max

1≤i≤nsupf∈F| f ( zi)|

∥∥∥∥∥ψ1

≤ r∥∥∥∥∥max

1≤i≤n‖zi‖2

∥∥∥∥∥ψ1

≤ rd1/2∥∥∥∥∥ max

1≤i≤n,1≤ j≤d|zi j|

∥∥∥∥∥ψ1

≤ (log 2)1/2rd1/2∥∥∥∥∥ max

1≤i≤n,1≤ j≤d|zi j|

∥∥∥∥∥ψ2

≤ c0(d log n)1/2r,

where c0 > 0 depends only on υ0, and ‖ · ‖ψq (1 ≤ q ≤ 2) denotes the ψq-Orlicz norm. ApplyingTheorem 4 in Adamczak (2008) with α = 1 and δ = η = 1/2, we obtain that for any x ≥ 0,

supf∈FGn f ≤

32E

(supf∈FGn f

)+ x

with probability at least 1 − e−x2/(3M0m3r3) − 3e−x√

n/c1(d log n)1/2r, where c1 > 0 depends only on c0.Given t ≥ 0 such that 4e−t ≤ 1, taking

x = max(3M0m3)1/2r3/2t1/2, 2c1rt(d log n)1/2n−1/2

in the above bound yields that, with probability at least 1 − e−t − 3e−2t ≥ 1 − 2e−t,

supf∈FGn f ≤

32E

(supf∈FGn f

)+ max

(3M0m3)1/2r3/2t1/2, 2c1rt

√d log n

n

. (36)

To bound E(sup f∈F Gn f ), the key is to control the covering numbers N(F , L2(Q), ε‖F‖Q,2) forall finitely supported probability measures Q on R × Rd and 0 < ε < 1, where F( z) = r‖z‖2 is ameasurable envelope of F . Respectively, for the function classes F1 and F2 that have envelopesF1( z) = r‖z‖2 and F2( z) = 1, using Theorem B in Dudley (1979) and Theorem 2.6.7 in van derVaart and Wellner (1996) we have

supQ

N(F1, L2(Q), ε‖F1‖Q,2) ≤ (A1/ε)2(d+2) and supQ

N(F2, L2(Q), ε) ≤ (A1/ε)2(d+2)

for some A1 > e, where the suprema are taken over all finitely discrete probability measures Qon R × Rd. Combining the above bounds with Corollary A.1 in the supplement of Chernozhukov,Chetverikov and Kato (2014) shows that

supQ

N(F , L2(Q), ε‖F‖Q,2)

≤ supQ

N(F1, L2(Q), 2−1/2ε‖F1‖Q,2) · supQ

N(F2, L2(Q), 2−1/2ε) ≤ (A2/ε)4(d+2),

where A2 = 21/2A1. For the envelop function F : R × Rd → R+, we have EF(z)2 = r2d. Conse-quently, it follows from Corollary 5.1 in Chernozhukov, Chetverikov and Kato (2014) that

E

(supf∈FGn f

).

√M0m3r3d log

(A2

2d/(M0m3r))

+ rMnd

n1/2 log(A2

2d/(M0m3r)), (37)

where Mn := (Emax1≤i≤n ‖zi‖22)1/2. To bound Mn, we will reply on an exponential-type tail inequal-

ity for X := max1≤i≤n ‖zi‖22. Assume there exist constants A, a > 0 such that P(X ≥ A + au) ≤ e−u

19

Page 20: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

for every u ∈ R. Then

E(X) =

∫ ∞

0P(X ≥ t)dt ≤ A +

∫ ∞

AP(X ≥ t)dt

= A +

∫ ∞

0P(X ≥ A + t)dt = A + a

∫ ∞

0P(X ≥ A + au)du ≤ A + a.

Given ε ∈ (0, 1), there exits a finite subsetNε ⊆ Sd−1 with |Nε | ≤ (1+2/ε)d such that max1≤i≤n ‖zi‖2 ≤

(1 − ε)−1 max1≤i≤n maxu∈Nε 〈u,wi〉. For every i ∈ [n] and u ∈ Nε , Condition 1 indicates thatP(|〈u,wi〉| ≥ υ0u) ≤ 2e−u2/2 for any u ∈ R. Taking the union bound over i ∈ [n] and u ∈ Nε ,and setting u =

√2v + 2 log(2n) + 2d log(1 + 2/ε) (v > 0), we obtain that with probability at least

1 − 2n(1 + 2/ε)de−u2/2 = 1 − e−v, max1≤i≤n ‖zi‖2 ≤ (1 − ε)−1υ0√

2v + 2 log(2n) + 2d log(1 + 2/ε).Minimizing this upper bound with respect to ε ∈ (0, 1), we conclude that

P[max1≤i≤n

‖zi‖22 ≥ 2υ2

03.7d + log(2n) + v

]≤ e−v, valid for every v > 0.

Taking A = 2υ203.7d + log(2n) and a = 2υ2

0 in the earlier analysis yields the bound M2n =

E(max1≤i≤n ‖zi‖22) ≤ 2υ2

03.7d + log(2en). Plugging this into (37) gives

E

(supf∈FGn f

).

√M0m3r3d log

(A2

2d/(M0m3r))

+ r(d + log n)1/2 dn1/2 log

(A2

2d/(M0m3r)). (38)

Together, (33), (35), (36) and (38) imply that with probability at least 1 − 2e−t,

supβ∈ΘΣ(r)

‖∆(β) − E∆(β)‖2

≤ C1

√rtn

+

√log(C2d/r)

rdn

+ (d + log n)1/2 log(C2d/r)dn

+ (d log n)1/2 tn

. (39)

Thus far, we have established a high probability bound on the `2-norm of ∆(β) = S−1/2∇Qn(β)−∇Qn(β∗)

− S1/2(β− β∗) uniformly over β ∈ ΘΣ(r), a local neighborhood of β∗, for any prespecified

r > 0. By Theorem 1, we have β ∈ ΘΣ(rt) with probability at least 1 − 2e−t as long as n ≥CL2

0 f −4(d+t), where rt = C3√

(d + t)/n. Setting r = rt in (32) and (39), we find that with probabilityat least 1 − 2e−t,

supβ∈ΘΣ(rt)

‖∆(β)‖2 .(d + t)1/4(d log n + t)1/2

n3/4 + (d + log n)1/2 d log nn

+ (d log n)1/2 tn.

Recalling that ∇Qn(β) = 0, this completes the proof.

4.4 Proof of Theorem 3

Let λ ∈ Rd be an arbitrary vector defining a linear contrast. Define the normalized partial sumS n = n−1/2 ∑n

i=1 γiζi of independent zero-mean random variables, where ζi = I(εi ≤ 0) − τ andγi = −〈S−1λ, xi〉. Moreover, write δn = (d + log n)1/4(d log n)1/2n−1/4 + (d + log n)1/2d log(n) n−1/2.Applying Theorem 2 with t = log n yields that, under the scaling n & d + log n,∣∣∣n1/2〈λ, β − β∗〉 − S n

∣∣∣= n1/2

∣∣∣∣∣⟨S−1/2λ,S1/2(β − β∗) + S−1/2 1n

n∑i=1

I(εi ≤ 0) − τ

xi

⟩∣∣∣∣∣ ≤ c1‖S−1/2λ‖2 δn (40)

20

Page 21: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

with probability at least 1 − 4n−1 for some constant c1 > 0.For the partial sum S n, note that var(S n) = σ2

τ = τ(1 − τ)‖S−1λ‖2Σ

. Then it follows from theBerry-Esseen inequality (see, e.g., Tyurin (2011)) that

supx∈R

∣∣∣PS n ≤ var(S n)1/2x− Φ(x)

∣∣∣≤E|I(ε ≤ 0) − τ〈S−1λ, x〉|3

2n1/2σ3τ

≤1 − 2(τ − τ2)2(τ − τ2)1/2

m3

n1/2 = c2n−1/2. (41)

Moreover, for any a ≤ b, Φ(b/στ) − Φ(a/στ) ≤ (2π)−1/2(b − a)/στ. Combining this with (40) and(41), for any x ∈ R, we obtain

P(n1/2〈λ, β − β∗〉 ≤ x

)≤ P

(S n ≤ x + c1‖S−1/2λ‖2 δn

)+ 4n−1

≤ Pvar(S n)1/2G ≤ x + c1‖S−1/2λ‖2 δn

+ c2n−1/2 + 4n−1

≤ P(στG ≤ x

)+ c12πτ(1 − τ)−1/2δn + c2n−1/2 + 4n−1,

where G ∼ N(0, 1). A similar argument leads to the reverse inequality. Putting together the piecesestablished the Berry-Esseen bound (6).

4.5 Proof of Theorem 4

Without loss of generality, we assume t > 0 is such that 2e−t ≤ 1 throughout the proof. By theconvexity of β 7→ Q[

n(β), β[ satisfies the first-order condition that ξ[β[

= 0 for some ξ[β[∈ ∂Q[

n(β[).

Again, we follow the same localized analysis as in the proof of Theorem 1. For some 0 < r ≤f /(6L0υ

20) to be determined, if β[ < β∗+BΣ(r), there exists η ∈ (0, 1) such that β := β∗+η(β[−β∗) ∈

β∗ + ∂BΣ(r); otherwise if β[ ∈ β∗ + BΣ(r), we take η = 1 so that β = β[.Similar to (27) and (28), we have that for any ξ[

β∗∈ ∂Q[

n(β∗) and ξ[β∈ ∂Q[

n(β),

〈ξ[β− ξ[β∗ , β − β

∗〉 ≤ ‖Σ−1/2ξ[β∗‖2‖β − β∗‖Σ.

For the right-hand side, Proposition 3 implies that there exists some event G1(t) with PG1(t) ≥1 − e−2t such that, conditioned on G1(t),

‖Σ−1/2ξ[β∗‖2 ≤ 2

√d + t

n+ ‖Σ−1/2E∗ξ[β∗‖2

with P∗-probability at least 1 − e−2t as long as n & d + t. On the other hand, since ‖β − β∗‖Σ ≤ r,by Proposition 4, there exists some event G2(t) = G2(t;Dn) with PG2(t) ≥ 1 − e−t such that,conditioned on G2(t),

〈ξ[β− ξ[β∗ , β − β

∗〉 ≥18

f ‖δ‖2Σ − 8υ20‖δ‖Σ

√2(d + t)

n

with P∗-probability at least 1 − e−t/2 as long as n & log(d) + t, where δ = β − β∗. Together, the lastthree displays imply

‖δ‖Σ ≤ 8 f −1(21/2 + 8υ20)√2(d + t)

n+ 8 f −1‖Σ−1/2E∗ξ[β∗‖2 (42)

21

Page 22: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

with P∗-probability at least 1 − e−t conditioned on G1(t) ∩ G2(t).For ‖Σ−1/2E∗ξ[

β∗‖2, it follows from (21) and Proposition 1 that

‖Σ−1/2E∗ξ[β∗‖2 < 3υ0

√2(d + t)

n(43)

with probability at least 1 − e−2t. Let G3(t) be the event that (43) holds so that PG3(t) ≥ 1 − e−2t.Combining (42) and (43), we conclude that conditioned on G1(t) ∩ G2(t) ∩ G3(t), ‖δ‖Σ < r :=

C4 f −1 √(d + t)/n with P∗-probability at least 1 − e−t as long as n ≥ C5L20 f −4(d + t), and PG1(t) ∩

G2(t) ∩ G3(t) ≥ 1 − 2e−t, where the constants C4,C5 > 0 depend only on υ0. This enforces β = β[.Finally, taking E(t) = G1(t) ∩ G2(t) ∩ G3(t) establishes the claim.

4.6 Proof of Theorem 5

Following the proof of Theorem 2, we treat ∇Q[n(β) := (1/n)

∑ni=1 wixiI(yi ≤ 〈xi,β〉) − τ as the

gradient of Q[n(β). Under this notation, define the vector-valued random process

∆[(β) = S−1/2∇Q[n(β) − ∇Q[

n(β∗)− S1/2(β − β∗) for β ∈ Rd.

Recalling E(wi) = 1, we have E∗∇Q[n(β) = ∇Qn(β) = (1/n)

∑ni=1 xiI(yi ≤ 〈xi,β〉) − τ. Define

R[n(β) = ∇Q[n(β) − ∇Qn(β), so that

∆[(β) = S−1/2R[n(β) − R[n(β∗) + ∇Qn(β) − ∇Qn(β∗) − S(β − β∗)

and E∗∆[(β) = ∆(β) with ∆(β) defined in (30). By the triangle inequality, for any r > 0 we have

supβ∈ΘΣ(r)

‖∆[(β)‖2 ≤ supβ∈ΘΣ(r)

‖∆[(β) − E∗∆[(β)‖2 + supβ∈ΘΣ(r)

‖∆(β)‖2, (44)

where ΘΣ(r) = β ∈ Rd : ‖β − β∗‖Σ ≤ r.The last term supβ∈ΘΣ(r) ‖∆(β)‖2 in (44), which only depends on the data Dn = (yi, xi)ni=1, has

been dealt with in the proof of Theorem 2. Hence, it remains to bound the random fluctuation∆[(β) − E∗∆[(β) = S−1/2R[n(β) − R[n(β∗) over β ∈ ΘΣ(r), givenDn. As before, we use a change ofvariable v = Σ1/2(β − β∗) and obtain

supβ∈ΘΣ(r)

‖∆[(β) − E∗∆[(β)‖2 = supβ∈ΘΣ(r)

‖S−1/2R[n(β) − R[n(β∗)‖2

≤ f −1/2 supβ∈ΘΣ(r),u∈Bd(1)

〈R[n(β) − R[n(β∗),Σ−1/2u〉

= f −1/2r−1 supu,v∈Bd(r)

〈Σ−1/2R[n(β∗ + Σ−1/2v) − R[n(β∗), v〉︸ ︷︷ ︸n−1/2∆[0(u,v)

, (45)

where ∆[0(u, v) = n−1/2 ∑n

i=1 ei〈zi,u〉I(εi ≤ 〈zi, v〉) − I(εi ≤ 0). Let F1 and F2 be the func-tion classes defined in (34), and let F = F1 · (F2 − f0) be the pointwise product between F1

and F2 − f0 with f0 : (z0, z) 7→ I(z0 ≤ 0). With this notation, we have supu,v∈Bd(r) ∆[0(u, v) =

sup f∈F n−1/2 ∑ni=1 ei f ( zi). Recall that E∗ denotes the conditional expectation given Dn. By Theo-

rem 13 in Boucheron et al. (2005) and the bound sup1≤i≤n, f∈F f ( zi) ≤ r max1≤i≤n ‖zi‖2, we obtainthat, with Z := E∗sup f∈F |(1/n)

∑ni=1 ei f ( zi)| denoting the conditional Rademacher average,

E(Z − EZ)2k

+

1/(2k)≤ 2

√EZ · kκr

Mn,k

n+ 2kκr

Mn,k

n≤ EZ + 3kκr

Mn,k

n, valid for any k ≥ 1,

22

Page 23: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

where κ =√

e/(2√

e− 2) < 1.271 and Mn,k := (Emax1≤i≤n ‖zi‖2k2 )1/(2k). By (45), Markov’s inequal-

ity and the bound Z ≤ (Z − EZ)+ + EZ, we obtain that

supβ∈ΘΣ(r)

‖∆[(β) − E∗∆[(β)‖2 = OP∗(r−1Z

)and Z = OP

(EZ + rMn,1/n

). (46)

For EZ, by a similar argument to (38) and (39), we get

EZ . r3/2

√log(C2d/r)

dn

+ r(d + log n)1/2 log(C2d/r)dn. (47)

With the above preparations, we are ready to prove the claim. Together, Theorems 1–4 implythat under the scaling n & d + log n, there exists some event En, satisfying P(En) ≥ 1 − 4n−1, onwhich ‖β − β∗‖Σ ≤ rn = C3

√(d + log n)/n and

χ1n :=

∥∥∥∥∥∥S1/2(β − β∗) + S−1/2 1n

n∑i=1

xiI(εi ≤ 0) − τ

∥∥∥∥∥∥2

≤ supβ∈ΘΣ(rn)

‖∆(β)‖2 .(d + log n)1/4(d log n)1/2

n3/4 + (d + log n)1/2 d log nn︸ ︷︷ ︸

:=∆n,d

.

Moreover, with P∗-probability at least 1 − n−1 conditioned on En, ‖β[ − β∗‖Σ ≤ rn so that∥∥∥∥∥∥S1/2(β[ − β∗) + S−1/2 1n

n∑i=1

wixiI(εi ≤ 0) − τ

∥∥∥∥∥∥2≤ supβ∈ΘΣ(rn)

‖∆[(β)‖2.

By (44), (46), (47) and (39), χ2n = χ2n(Dn) := E∗supβ∈ΘΣ(r[n) ‖∆[(β)‖2 satisfies χ2n = OP(∆n,d).

Let r[n = S1/2(β[ − β) − S−1/2(1/n)∑n

i=1 eixiτ − I(εi ≤ 0). Then, with P∗-probability at least1 − n−1 conditioned on En, ‖r[n‖2 ≤ χ1n + supβ∈ΘΣ(r) ‖∆

[(β)‖2 with supβ∈ΘΣ(r) ‖∆[(β)‖2 = OP∗(χ2n)

and χ1n + χ2n = OP(∆n,d). This establishes the claim (14).

4.7 Proof of Theorem 6

Let λ ∈ Rd be an arbitrary vector defining a linear contrast of interest. Write γi = 〈S−1λ, xi〉 andζi = I(εi ≤ 0) − τ for i = 1, . . . , n, and define

S n =1√

n

n∑i=1

γiζi and S [n =

1√

n

n∑i=1

eiγiζi.

To begin with, it follows from Theorem 2 that under the scaling n & d+log n, there exists a sequenceof events En with P(En) ≥ 1−4n−1 such that, |n1/2〈λ, β−β∗〉−S n| ≤ c1‖S−1/2λ‖2δn,d on En, whereδn,d := (d + log n)1/4(d log n)1/2n−1/4 + (d + log n)1/2d log(n)n−1/2. By Theorems 4 and 5, we furtherhave |n1/2〈λ, β[ − β〉 − S [

n| ≤ ‖S−1/2λ‖2‖n1/2r[n‖2 with P∗-probability at least 1 − n−1 conditioned onEn. For the remainder r[n = r[n((ei, yi, xi)ni=1), using Markov’s inequality with the bounds (46) and(47), there exits some event Gn with P(Gc

n) . (δn,d/δ2)2 such that, conditioned on En ∩ Gn,

P∗(‖n1/2r[n‖2 ≥ δ1

). δ−1

1 (δn,d + δ2),

23

Page 24: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

valid for any δ1, δ2 > 0. Taking δ1 = δ2/5n,d and δ2 = δ4/5

n,d yields that P(Gcn) ≤ c2δ

2/5n,d and

P∗(‖n1/2r[n‖2 ≥ δ

2/5n,d

)≤ c3δ

2/5n,d , conditioned on En ∩ Gn.

Next we establish the closeness in distribution between S n and S [n. Note that γiζi are inde-

pendent random variables with mean zero and var(γiζi) = τ(1 − τ)‖S−1λ‖2Σ

. Thus, var(S n) =

τ(1 − τ)‖S−1λ‖2Σ≥ τ(1 − τ) f −1‖S−1/2λ‖22. Moreover, under Condition 1,

E(|γiζi|

3) ≤ τ(1 − τ)E|〈S−1λ, xi〉|3 ≤ τ(1 − τ)m3‖S−1λ‖3Σ.

Let Φ(·) be the standard normal distribution function. By the Berry-Esseen inequality (see, e.g.,Tyurin (2011)),

supx∈R

∣∣∣PS n ≤ var(S n)1/2x− Φ(x)

∣∣∣ ≤ m3

2√τ(1 − τ)n

. (48)

For S [n, using a conditional version of the Berry-Esseen inequality for sums of independent random

variables (Tyurin, 2011), we have

supx∈R

∣∣∣P∗S [n ≤ var∗(S [

n)1/2x− Φ(x)

∣∣∣ ≤ (1/n)∑n

i=1 |ζiγi|3

2√

n var∗(S [n)3/2

, (49)

where var∗(S [n) = (1/n)

∑ni=1(γiζi)2. Recall that zi = Σ−1/2xi, and let u = Σ1/2S−1λ/‖S−1λ‖Σ ∈ S

d−1

be a unit vector. For the two data-dependent quantities var∗(S [n) and (1/n)

∑ni=1 |γiζi|

3, we have

∣∣∣var∗(S [n)/var(S n) − 1

∣∣∣ =1

τ(1 − τ)

∣∣∣∣∣1nn∑

i=1

ζ2i 〈u, zi〉

2 − τ(1 − τ)∣∣∣∣∣ (50)

and

1n

n∑i=1

|γiζi|3 ≤ max

1≤i≤n|γiζi| ·

1n

n∑i=1

ζ2i 〈S−1λ, xi〉

2 ≤ max1≤i≤n

|γiζi| · ‖S−1λ‖2Σ ·1n

n∑i=1

ζ2i 〈u, zi〉

2. (51)

For independent zero-mean sub-Gaussian random variables γiζi, it can be shown that with prob-ability at least 1 − e−x, max1≤i≤n |γiζi| . ‖S−1λ‖Σ

√log(n) + x. Furthermore, following the proof of

Proposition 3, it can be similarly shown that∣∣∣∣∣1nn∑

i=1

ζ2i 〈u, zi〉

2 − τ(1 − τ)∣∣∣∣∣ ≤ 2υ2

0

√2x3n

+ 2υ20

xn

with probability at least 1 − 2e−x. Putting together the pieces, it follows from (50) that there existsan event E′n, satisfying P(E′n) ≥ 1 − n−1, on which max1≤i≤n |γiζi| . ‖S−1λ‖Σ(log n)1/2,

1n

n∑i=1

|γiζi|3 . ‖S−1λ‖3Σ (log n)1/2 and

∣∣∣var∗(S [n)/var(S n) − 1

∣∣∣ . √log n

n(52)

as long as n & log n.For the normal distribution function, we have the following property derived from Pinsker’s

inequality (see Lemma A.7 in the supplement of Spokoiny and Zhilova (2015)):

supx∈R

∣∣∣Φ(x/var(S n)1/2) − Φ(x/var∗(S [n)1/2)

∣∣∣ ≤ 12

∣∣∣var∗(S [n)/var(S n) − 1

∣∣∣ (53)

24

Page 25: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

as long as |var∗(S [n)/var(S n) − 1| ≤ 1/2. Moreover, for any a ≤ b,

Φ(b/var(S n)1/2) − Φ(a/var(S n)1/2) ≤b − a

√2πvar(S n)

≤f 1/2(b − a)

‖S−1/2λ‖2√

2πτ(1 − τ). (54)

Combining the ingredients, we derive that for any x ∈ R,

P(n1/2〈λ, β − β∗〉 ≤ x

)≤ P

(S n ≤ x + c1‖S−1/2λ‖2δn,d

)+ 4n−1

(i)≤ P

var(S n)1/2G ≤ x + c1‖S−1/2λ‖2δn,d

+

m3

2√τ(1 − τ)n

+ 4n−1

(ii)≤ P

var(S n)1/2G ≤ x − ‖S−1/2λ‖2δ

2/5n,d

+ f 1/2

c1δn,d + δ2/5n,d

√2πτ(1 − τ)

+m3

2√τ(1 − τ)n

+ 4n−1

(iii)≤ P∗

var∗(S [

n)1/2G ≤ x − ‖S−1/2λ‖2δ2/5n,d

+

12

∣∣∣∣∣var∗(S [n)

var(S n)− 1

∣∣∣∣∣ + f 1/2c1δn,d + δ2/5

n,d√

2πτ(1 − τ)+

m3

2√τ(1 − τ)n

+ 4n−1

(iv)≤ P∗

(S [

n ≤ x − ‖S−1/2λ‖2δ2/5n,d

)+

(1/n)∑n

i=1 |γiζi|3

2√

n var∗(S [n)3/2

+12

∣∣∣∣∣var∗(S [n)

var(S n)− 1

∣∣∣∣∣ + f 1/2c1δn,d + δ2/5

n,d√

2πτ(1 − τ)+

m3

2√τ(1 − τ)n

+ 4n−1,

where steps (i) and (iv) follow respectively from the Berry-Esseen inequalities (48) and (49), step(ii) uses the anti-concentration inequality (54), and step (iii) is due to the Gaussian comparisoninequality (53). Conditioned on En ∩ Gn,

P∗(S [

n ≤ x − ‖S−1/2λ‖2δ2/5n,d

)≤ P∗

(S [

n ≤ x − ‖S−1/2λ‖2‖n1/2r[n‖2)

+ P∗(‖n1/2r[n‖2 ≥ δ

2/5n,d

)≤ P∗

(n1/2〈λ, β[ − β〉 ≤ x

)+ n−1 + c3δ

2/5n,d .

Moreover, on the event E′n, the bounds in (52) imply

(1/n)∑n

i=1 |γiζi|3

2√

n var∗(S [n)3/2

+12

∣∣∣∣∣var∗(S [n)

var(S n)− 1

∣∣∣∣∣ .√

log nn

as long as n & log n. A similar argument leads to a series of reverse inequalities.Putting together the pieces, we conclude that conditioned on the event En ∩ E

′n ∩ Gn,

supx∈R

∣∣∣P(n1/2〈λ, β − β∗〉 ≤ x)− P∗

(n1/2〈λ, β[ − β〉 ≤ x

)∣∣∣ . δ2/5n,d .

Under the scaling d3(log n)2 = o(n), δn,d = o(1) as n → ∞. Combined with the above bound, thisestablishes the claim (15).

Acknowledgements

The authors are grateful to the Associate Editor and reviewers for thoughtful feedback and construc-tive comments. The second author would also like to thank Lan Wang and Jelena Bradic for helpfuldiscussions and encouragement. The authors further acknowledge the support of NSF Award DMS-1811376.

25

Page 26: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

References

Adamczak, R. (2008). A tail inequality for suprema of unbounded empirical processes with appli-cations to Markov chains. Electron. J. Probab. 13 1000–1034.

Arcones, M. A. (1996). The Bahadur-Kiefer representation of Lp regression estimators. Econom.Theory 12 257–283.

Arcones, M. A. and Gine, E. (1992). On the bootstrap of M-estimators and other statistical func-tionals. In Exploring the Limits of Bootstrap (R. LePage and L. Billard, ed.) 14–47. Wiley, NewYork.

Bassett, G. and Koenker, R. (1978). Asymptotic theory of least absolute error regression. J. Amer.Statist. Assoc. 73 618–622.

Bassett, G. and Koenker, R. (1986). Strong consistency of regression quantiles and related empiri-cal processes. Econom. Theory 2 191–201.

Boucheron, S., Bousquet, O., Lugosi, G. and Massart, P. (2005). Moment inequalities for functionsof independent random variables. Ann. Probab. 33 514–560.

Chatterjee, S. and Bose, A. (2005). Generalized bootstrap for estimating equations. Ann. Statist. 33414–436.

Chen, K., Ying, Z., Zhang, H. and Zhao, L. (2008). Analysis of least absolute deviation. Biometrika95 107–122.

Chen, X. and Zhou, W.-X. (2019). Robust inference via multiplier bootstrap. Ann. Statist., to appear.Preprint arXiv:1903.07208.

Cheng, G. and Huang, J. Z. (2010). Bootstrap consistency for general semiparametric M-estimation.Ann. Statist. 38 2884–2915.

Chernozhukov, V., Chetverikov, D. and Kato, K. (2014). Gaussian approximation of suprema ofempirical processes. Ann. Statist. 42 1564–1597.

Dudley, R. M. (1979). Balls in rk do not cut all subsets of k + 2 points. Adv. Math. 31 306–308.

Efron, B. (1979). Bootstrap methods: Another look at the jackknife. Ann. Statist. 7 1–26.

Efron, B. and Tibshirani, R. J. (1994). An Introduction to the Bootstrap. Chapman Hall, New York.

Feng, X., He, X. and Hu, J. (2011). Wild bootstrap for quantile regression. Biometrika 98 995–999.

Gine, E. and Zinn, J. (1990). Bootstrapping general empirical measures. Ann. Probab. 18 851–869.

Gutenbrunner, C. and Jureckova, J. (1992). Regression rank scores and regression quantiles. Ann.Statist. 20 305–330.

Gutenbrunner, C., Jureckova, J., Koenker, R. and Portnoy, S. (1993). Tests of linear hypothesesbased on regression rank scores. J. Nonparametr. Stat. 2 307–331.

26

Page 27: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

He, X. and Hu, F. (2002). Markov chain marginal bootstrap. J. Amer. Statist. Assoc. 97 783–795.

He, X. and Shao, Q.-M. (1996). A general Bahadur representation of M-estimators and its applica-tion to linear regression with nonstochastic designs. Ann. Statist. 24 2608–2630.

Hsu, D., Kakade, S. M. and Zhang, T. (2014). Random design analysis of ridge regression. Found.Comput. Math. 14 569–600.

Kocherginsky, M., He, X. and Mu, Y. (2005). Practical confidence intervals for regression quantiles.J. Comp. Graph. Statist. 14 41–55.

Koenker, R. (1988). Asymptotic theory and econometric practice. J. Appl. Econom. 3 139–147.

Koenker, R. (2005). Quantile Regression. Cambridge Univ. Press, Cambridge.

Koenker, R. (2019). Package ‘quantreg’, version 5.54. Reference manual available at R-CRAN:https://cran.r-project.org/web/packages/quantreg/quantreg.pdf.

Koenker, R. and Bassett, G. (1978). Regression quantiles. Econometrica 46 33–50.

Koenker, R. and Bassett, G. (1982). Tests of linear hypotheses and `1 estimation. Econometrica 501577–1583.

Ledoux, M. and Talagrand, M. (1991). Probability in Banach Spaces: Isoperimetry and Processes.Ergebnisse der Mathematik und Ihrer Grenzgebiete (3) 23. Springer, Berlin.

Loh, P.-L. and Wainwright, M. J. (2015). Regularized M-estimators with non-convexity: Statisticaland algorithmic theory for local optima. J. Mach. Learn. Res. 16 559–616.

Ma, S. and Kosorok, M. R. (2005). Robust semiparametric M-estimation and the weighted boot-strap. J. Multivariate Anal. 96 190–217.

McDiarmid, C. (1989). On the method of bounded differences. In Surveys in Combinatorics, LondonMath. Soc. Lecture Note Ser., 141 148–188. Cambridge Univ. Press, Cambridge.

Negahban, S. N., Ravikumar, P., Wainwright, M. J. and Yu, B. (2012). A unified framework forhigh-dimensional analysis of M-estimators with decomposable regularizers. Statist. Sci. 27 538–557.

Pan, X., Sun, Q. and Zhou, W.-X. (2019). Nonconvex regularized robust regression with oracleproperties in polynomial time. Preprint arXiv:1907.04027.

Parzen, M. I., Wei, L. J. and Ying, Z. (1994). A resampling method based on pivotal estimatingfunctions. Biometrika 81 341–350.

Pollard, D. (1991). Asymptotics for least absolute deviation regression estimators. Econom. Theory7 186–199.

Portnoy, S. (1985). Asymptotic behavior of M-estimators of p regression parameters when p2/n islarge; II. Normal approximation. Ann. Statist. 13 1403–1417.

27

Page 28: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

Portnoy, S. (1986). On the central limit theorem in Rp when p → ∞. Probab. Theory Relat. Fields73 571–583.

Portnoy, S. and Koenker, R. (1989). Adaptive L-estimation for linear models. Ann. Statist. 17 362–381.

Praestgaard, J. and Wellner, J. (1993). Exchangeably weighted bootstraps of the general empiricalprocess. Ann. Probab. 21 2053–2086.

Ruppert, D. and Carroll, R. J. (1980). Trimmed least squares estimation in the linear model. J.Amer. Statist. Assoc. 75 828–838.

Silverman, B. (1986). Density Estimation for Statistics and Data Analysis. Chapman Hall, NewYork.

Spokoiny, V. and Zhilova, M. (2015). Bootstrap confidence sets under model misspecification. Ann.Statist. 43 2653–2675.

Sun, Q., Zhou, W.-X. and Fan, J. (2019). Adaptive Huber regression. J. Amer. Statist. Assoc., toappear. Preprint arXiv:1706.06991.

Tyurin, I. S. (2011). On the convergence rate in Lyapunov’s theorem. Theory Probab. Appl. 55253–270.

van de Geer, S. (2000). Empirical Processes in M-Estimation. Cambridge Univ. Press, Cambridge.

van der Vaart, A. W. and Wellner, J. A. (1996). Weak Convergence and Empirical Processes: WithApplications to Statistics. Springer, New York.

Wainwright, M. J. (2019) High-Dimensional Statistics: A Non-Asymptotic Viewpoint. CambridgeUniv. Press, Cambridge.

Wellner, J. A. and Zhan, Y. (1996). Bootstrapping Z-estimators. Technical report. Department ofStatistics, University of Washington, Seattle.

Welsh, A. H. (1989). On M-processes and M-estimation. Ann. Statist. 15 337–361.

Zhao, L. C., Rao, C. R. and Chen, X. R. (1993). A note on the consistency of M-estimates in linearmodels. In Stochastic Processes: A Festschrift in Honour of Gopinath Kallianpur, 359–367.Springer, New York.

A Proofs of Propositions 1–4

A.1 Proof of Proposition 1

By (21), every ξβ∗ ∈ ∂Qn(β∗) satisfies ξβ∗ = ξ∗ := (1/n)∑n

i=1 xiI(εi ≤ 0) − τ with probability one.Hence, it suffices to bound ‖Σ−1/2ξ∗‖2 = sup‖u‖2=1〈u,Σ−1/2ξ∗〉. Via a standard covering argument,for any ε ∈ (0, 1), there exists an ε-net Nε of the unit sphere Sd−1 with |Nε | ≤ (1 + 2/ε)d such

28

Page 29: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

that ‖Σ−1/2ξ∗‖2 ≤ (1 − ε)−1 maxu∈Nε 〈u,Σ−1/2ξ∗〉. Along each direction u, define one-dimensionalmarginals

γu,i = 〈u,Σ−1/2xi〉I(εi ≤ 0) − τ, i = 1, . . . , n,

which satisfy E(γu,i) = 0 and var(γu,i) = τ(1 − τ) ≤ 1/4. By Condition 1, P(|〈u,Σ−1/2xi〉| ≥ υ0t) ≤2e−t2/2 for all t ≥ 0. Hence, for k = 1, 2, . . .,

Eγ2ku,i ≤ EI(εi ≤ 0) − τ2〈u,Σ−1/2xi〉

2k

=14υ2k

0 2k∫ ∞

0P(|〈u,Σ−1/2xi〉| ≥ υ0t

)t2k−1dt ≤ υ2k

0 2k−1k! ≤(2k)!2kk!

(a1υ0)2k

for some absolute constant a1 > 1. Following the proof of Theorem 2.6 in Wainwright (2019), itcan be shown that Eeλγu,i ≤ e(a1a2λυ0)2/2 for all λ ∈ R, where a2 > 1 is also an absolute constant.By the Hoeffding bound for sums of sub-Gaussian random variables (see, e.g., Proposition 2.5. inWainwright (2019)), for any y ≥ 0 we have

1n

n∑i=1

γu,i ≤ a1a2υ0

√2yn

with probability at least 1 − e−y. Taking the union bound over all vectors u ∈ Nε yields

‖Σ−1/2ξ∗‖2 ≤a1a2υ0

1 − ε

√2yn

with probability greater than 1 − elog(1+2/ε)d−y. Through a careful analysis, we select a1 = 1.09,a2 = 1.3 and ε = 0.314 so that all the requirements are satisfied. Finally, taking y = 2d + xcompletes the proof.

A.2 Proof of Proposition 2

By (20), every ξβ = (ξβ,1, . . . , ξβ,d)ᵀ ∈ ∂Qn(β) can be written as

ξβ, j = −τ

n

n∑i=1

xi j +1n

n∑i=1

xi jI(yi ≤ 〈xi,β〉) −1n

n∑i=1

xi jvi + (1 − τ)I(yi = 〈xi,β〉),

where vi ∈ [τ − 1, τ]. With δ = β − β∗, it follows that

〈ξβ − ξβ∗ ,β − β∗〉 ≥

1n

n∑i=1

〈xi, δ〉I(εi ≤ 〈xi, δ〉) − I(εi ≤ 0)

︸ ︷︷ ︸

:=Un(δ)

−1n

n∑i=1

|〈xi, δ〉|I(εi = 〈xi, δ〉) + I(εi = 0)

. (55)

Since the conditional distribution of ε given x is continuous, with probability one, there is no vectorδ ∈ Rd and 1 ≤ i ≤ n such that εi = 〈xi, δ〉. See Lemma A.1 of Ruppert and Carroll (1980). In otherwords, with probability one,

1n

n∑i=1

|〈xi, δ〉|I(εi = 〈xi, δ〉) + I(εi = 0)

= 0 for all δ ∈ Rd. (56)

29

Page 30: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

Turning to the first term on the right-hand side of (55), the main difficulty comes from thediscontinuity of Un(δ) as a function of δ. To construct a smooth version of Un, we introduce fourLipschitz continuous functions as follows. For any a, b > 0 and u ∈ R, define

ϕ+a (u) =

1 if u > 2a

−1 + ua if a < u ≤ 2a

0 otherwise

, ϕ−a (u) =

1 if u < −2a

−1 − ua if − 2a ≤ u < −a

0 otherwise

, (57)

and

ψ+b (u) =

1 if u ≤ b/2

2 − 2ub if b

2 < u ≤ b

0 otherwise

, ψ−b (u) =

1 if u ≥ −b/2

2 + 2ub if − b ≤ u < − b

2

0 otherwise

. (58)

Respectively, ϕ±a and ψ±b are (1/a)- and (2/b)-Lipschitz continuous; see Figure 2. Also, they satisfythe following properties: for a, b > 0 and u ∈ R,

I(u ≥ 2a) ≤ ϕ+a (u) ≤ I(u ≥ a), I(u ≤ −2a) ≤ ϕ−a (u) ≤ I(u ≤ −a), (59)

I(u ≤ b/2) ≤ ψ+b (u) ≤ I(u ≤ b), I(u ≥ −b/2) ≤ ψ−b (u) ≤ I(u ≥ −b), (60)

aϕ+a (u) ≤

12

maxu, 0, aϕ−a (u) ≤12

max−u, 0. (61)

-3.0 -2.5 -2.0 -1.5 -1.0 -0.5 0.0

0.0

0.2

0.4

0.6

0.8

1.0

1.2

u

Functionvalue

ϕ−a (u)

ψ−b (u)

(a) Plots of ϕ−a (u) and ψ−b (u).

0.0 0.5 1.0 1.5 2.0 2.5 3.0

0.0

0.2

0.4

0.6

0.8

1.0

1.2

u

Functionvalue

ϕ+a (u)

ψ+b (u)

(b) Plots of ϕ+a (u) and ψ+

b (u).

Figure 2: The Lipschitz continuous functions ϕ±a (u) and ψ±b (u) with a = 1 and b = 2.

Furthermore, for each εi, we define its positive and negative components as εi,+ = maxεi, 0

30

Page 31: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

and εi,− = max−εi, 0. For any r > 0, taking a = εi,± and b = 2r‖δ‖Σ yields

Un(δ) =1n

n∑i=1

〈xi, δ〉I(0 < εi ≤ 〈xi, δ〉) + 〈−xi, δ〉I(〈xi, δ〉 < εi ≤ 0)

1n

n∑i=1

εi,+ϕ

+εi,+

(〈xi, δ〉) + εi,−ϕ−εi,−

(〈xi, δ〉)

≥1n

n∑i=1

εi,+ϕ+εi,+

(〈xi, δ〉)I(〈xi, δ〉 ≤ 2r‖δ‖Σ) +1n

n∑i=1

εi,−ϕ−εi,−

(〈xi, δ〉)I(〈xi, δ〉 ≥ −2r‖δ‖Σ)

≥1n

n∑i=1

εi,+ϕ+εi,+

(〈xi, δ〉)ψ+2r‖δ‖Σ(〈xi, δ〉)︸ ︷︷ ︸

V+n (δ)

+1n

n∑i=1

εi,−ϕ−εi,−

(〈xi, δ〉)ψ−2r‖δ‖Σ(〈xi, δ〉)︸ ︷︷ ︸V−n (δ)

. (62)

To bound Vn(δ) = V+n (δ) + V−n (δ) from below, we follow a two-step procedure: in step one, we

derive a lower bound on the expectation EVn(δ), and in step two, we show concentration of Vn(δ)around EVn(δ) uniformly over δ ∈ Rd with high probability.

Step 1. Along each direction δ ∈ Rd \ 0, define the one-dimensional marginal ηδ = 〈x, δ〉/‖δ‖Σthat satisfies E(η2

δ) = 1. Using the lower bounds of ϕ±a and ψ±b given in (59) and (60) we obtain

EV+n (δ) ≥

1n

n∑i=1

Eεi,+I(2εi,+ ≤ 〈xi, δ〉 ≤ r‖δ‖Σ)

,

EV−n (δ) ≥1n

n∑i=1

Eεi,−I(−r‖δ‖Σ ≤ 〈xi, δ〉 ≤ −2εi,−)

.

Together, with Condition 2 and the law of total expectation, we have

EVn(δ) ≥1n

n∑i=1

E

∫ 〈xi,δ〉/2

0t fεi |xi(t) dt · I(0 ≤ 〈xi, δ〉 ≤ r‖δ‖Σ)

+1n

n∑i=1

E

∫ 0

〈xi,δ〉/2(−t) fεi |xi(t) dt · I(−r‖δ‖Σ ≤ 〈xi, δ〉 ≤ 0)

≥14‖δ‖2Σ · E

fε|x(0)η2

δI(|ηδ| ≤ r)−

L0

24‖δ‖3Σ · E

|ηδ|

3I(|ηδ| ≤ r)

( f4−

L0

24r‖δ‖Σ

)‖δ‖2Σ E

η2δI(|ηδ| ≤ r)

. (63)

Under Condition 1, P(|ηδ/υ0| ≥ t) ≤ 2e−t2/2 for all t ≥ 0 and δ ∈ Rd. Therefore,

Eη2δI(|ηδ| > r)

=

( ∫ r2

0+

∫ ∞

r2

)Pη2δI(|ηδ| > r) > t

dt

= 2υ20

∫ ∞

r/υ0

P(|ηδ/υ0| ≥ t)t dt + r2P(|ηδ/υ0| > r/υ0)

≤ 2r2e−(r/υ0)2/2 + 4υ20

∫ ∞

(r/υ0)2/2e−sds =

(2r2 + 4υ2

0)e−(r/υ0)2/2.

Taking r = 4υ20 with υ0 ≥ 1, it follows that Eη2

δI(|ηδ| ≤ r) ≥ 1 − supυ0≥1(32υ40 + 4υ2

0)e−8υ20 ≥

1 − 36e−8. Substituting this into (63) yields

EVn(δ) ≥ (2/9 − 8e−8) f ‖δ‖2Σ (64)

31

Page 32: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

for all δ ∈ Rd satisfying 0 ≤ ‖δ‖Σ ≤ f /(6L0υ20), where f and L0 are defined in Condition 2.

Step 2. We prove the concentration of Vn(δ) around EVn(δ) uniformly over δ via the peelingtechnique, which is widely used in empirical process theory (van de Geer, 2000). For some δ > 0to be specified, define Θ(δ) = δ ∈ Rd : ‖δ‖Σ ≥ δ = ∪∞`=1Θ`(δ) with

Θ`(δ) =δ ∈ Rd : 2(`−1)/2δ ≤ ‖δ‖Σ ≤ 2`/2δ

, ` = 1, 2 . . . .

For any R ≥ δ, define

∆n(R) = f (w1, . . . ,wn; R) := supδ≤‖δ‖Σ≤R

EVn(δ) − Vn(δ)

, (65)

where wi = (xi, εi) ∈ Rd × R. For δ ∈ Rd, write

E(δ; wi) = εi,+ϕ+εi,+

(〈xi, δ〉)ψ+2r‖δ‖Σ(〈xi, δ〉) + εi,−ϕ

−εi,−

(〈xi, δ〉)ψ−2r‖δ‖Σ(〈xi, δ〉).

Note that for any b > 0 and u ∈ R, at most one of εi,+ϕ+εi,+

(u)ψ+b (u) and εi,−ϕ

−εi,−

(u)ψ−b (u) can benon-zero. When 〈xi, δ〉 ≥ 0, by (58) and (61) we have

0 ≤ E(δ; wi) = εi,+ϕ+εi,+

(〈xi, δ〉)ψ+2r‖δ‖Σ(〈xi, δ〉)

≤〈xi, δ〉

1 if 〈xi, δ〉 ≤ r‖δ‖Σ2 − 〈xi,δ〉

r‖δ‖Σif r‖δ‖Σ < 〈xi, δ〉 ≤ 2r‖δ‖Σ

0 otherwise

≤r2‖δ‖Σ.

Following a similar argument, the same upper bound applies to E(δ; wi) when 〈xi, δ〉 < 0. Conse-quently, we have |E(δ; wi)| ≤ Rr/2, so that for any index i and an independent copy w′i = (x′i , ε

′i) of

wi = (xi, εi),

| f (w1, . . . ,wi, . . . ,wn; R) − f (w1, . . . ,w′i , . . . ,wn; R)| ≤Rrn.

Hence, applying McDiarmid’s inequality (McDiarmid, 1989), we obtain that for any t ≥ 0,

∆n(R) ≤ E∆n(R) + Rr

√t

2n(66)

with probability at least 1 − e−t. Next we evaluate E∆n(R). Again, using (61) it can be shown thatfor any a, b > 0, the functions u 7→ aϕ±a (u)ψ±b (u) are 1-Lipschitz continuous. Thus, for any samplewi = (xi, εi) ∈ Rd × R and parameters δ, δ′ ∈ Rd, we have

|E(δ; wi) − E(δ′; wi)| ≤ 2|〈xi, δ〉 − 〈xi, δ′〉|.

In other words, E(δ; wi) is a 2-Lipschitz continuous function in 〈xi, δ〉. Let e1, . . . , en be independentRademacher variables that are independent of the initial sample. By a classical symmetrization ar-gument and the Ledoux-Talagrand contraction inequality (see, e.g., (4.20) in Ledoux and Talagrand(1991)),

E∆n(R) ≤ 2E

supδ≤‖δ‖Σ≤R

1n

n∑i=1

eiE(δ; wi)≤ 4E

sup

δ≤‖δ‖Σ≤R

1n

n∑i=1

ei〈xi, δ〉

≤ 4RE∥∥∥∥∥1

n

n∑i=1

eiΣ−1/2xi

∥∥∥∥∥2≤ 4R

√dn. (67)

32

Page 33: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

Combining (66) and (67) yields that, with probability at least 1 − e−t,

∆n(R) ≤ Rr

√t

2n+ 4R

√dn. (68)

With the above preparations, we derive that for any t0 > 0,

P

∃ δ ∈ Θ(δ) s.t. − Vn(δ) + EVn(δ) ≥ ‖δ‖2Σrt0 + 4‖δ‖Σ

√2dn

(i)≤

∞∑`=1

P

∃ δ ∈ Θ`(δ) s.t. − Vn(δ) + EVn(δ) ≥

12

(2`/2δ)2rt0 + 4(2`/2δ)

√dn

(ii)≤

∞∑`=1

P

∆n(2`/2δ) ≥ (2`/2δ)r

√(2`/2t0δ)2

4+ 4(2`/2δ)

√dn

(iii)≤

∞∑`=1

e−(2`/2t0δ)2n/2 =

∞∑`=1

e−2`−1(t0δ)2n

(iv)≤

∞∑`=1

e−`(t0δ)2n =

e−(t0δ)2n

1 − e−(t0δ)2n:= P(n, t0, δ), (69)

where step (i) uses the union bound along with the decomposition Θ(δ) = ∪∞`=1Θ`(δ), step (ii)follows from the definition of ∆n(·) in (65), step (iii) uses the concentration inequality (68) withR = 2`/2δ for each ` ≥ 1, and step (iv) uses the elementary inequality that 2`−1 ≥ `.

Putting (55), (56), (62), (64) and (69) (with r = 4υ20) together, we conclude that with probability

at least 1 − P(n, t0, δ),

〈ξβ − ξβ∗ ,β − β∗〉 ≥

(2/9 − 8e−8) f − 4υ2

0t0‖δ‖2Σ − 4‖δ‖Σ

√2dn

(70)

uniformly over δ = β−β∗ satisfying δ ≤ ‖δ‖Σ ≤ f /(6L0υ20). In particular, we take t0 = (2/9−8e−8−

1/8) f /(4υ20) and recall that υ0 ≥ 1 from Condition 1, then the right-hand side of (70) is bounded

from below by18

f ‖δ‖2Σ − 4υ20‖δ‖Σ

√2dn.

By the convexity of Qn, 〈ξβ − ξβ∗ ,β − β∗〉 is always non-negative. Therefore, for any t ≥ 0, we mayassume

‖β − β∗‖Σ ≥ δ := (32υ20/ f )

√2(d + t)

n;

otherwise, (22) holds trivially. The above choices of (t0, δ) guarantee that P(n, t0, δ) ≤ e−t/2 in (69).Putting together the pieces, we conclude that with probability at least 1 − e−t/2,

〈ξβ − ξβ∗ ,β − β∗〉 ≥

18

f ‖β − β∗‖2Σ − 4υ20‖β − β

∗‖Σ

√2(d + t)

n

for all β satisfying 0 ≤ ‖β − β∗‖Σ ≤ f /(6L0υ20). This completes the proof.

33

Page 34: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

A.3 Proof of Proposition 3

By (21) and (23), every subgradient ξ[ = (ξ[1, . . . , ξ[d)ᵀ ∈ ∂Q[

n(β∗) coincides with (1/n)∑n

i=1 wiζixi

with probability one. Thus, without loss of generality, we assume ξ[ = (1/n)∑n

i=1 wiζixi. Note thatE∗ξ[ = (1/n)

∑ni=1 ζixi. Using a standard covering argument again, for any ε ∈ (0, 1), there exists an

ε-net Nε ⊆ Sd−1 with |Nε | ≤ (1 + 2/ε)d such that

‖Σ−1/2(ξ[ − E∗ξ[)‖2 ≤1

1 − εmaxu∈Nε

1n

n∑i=1

eiζi〈u, zi〉,

where ei are independent Rademacher random variables. For any u ∈ Nε and y ≥ 0, by Hoeffding’sinequality we have

P∗

1n

n∑i=1

eiζi〈u, zi〉 ≥

(2y

n∑i=1

ζ2i 〈u, zi〉

2)1/2 1

n

≤ e−y.

Moreover, note that ζi are bounded random variables that satisfy E(ζ2i |xi) = τ(1 − τ) ≤ 1/4,

E(ζ4i |xi) ≤ 1/12 and |ζi| ≤ 1. Following the calculations as in the proof of Proposition 1, for

every u ∈ Nε we have

E(ζ2i 〈u, zi〉

2)k ≤k!2

43υ4

0(2υ20)k−2, k = 2, 3, . . . .

Using Bernstein’s inequality, we obtain

P

(1n

n∑i=1

ζ2i 〈u, zi〉

2 ≥14

+ 2υ20

√2x3n

+ 2υ20

xn

)≤ e−x, valid for any x ≥ 0.

Finally, we set ε = 2/(e2 − 1) so that (1 + 2/ε)d = e2d. Taking the union bound twice over allu ∈ Nε with x = y = 2(d + t) yields

maxu∈Nε

2n

n∑i=1

ζ2i 〈u, zi〉

2 ≤12

+ 8υ20

√d + t3n

+ 8υ20

d + tn

(71)

with probability at least 1 − e−2t, and with P∗-probability at least 1 − e−2t conditioned on the eventthat (71) holds,

‖Σ−1/2(ξ[ − E∗ξ[)‖2 ≤ 2

√d + t

n

provided that n ≥ Cυ40(d + t) for some universal constant C > 0. Putting together the pieces

completes the proof of (24).

A.4 Proof of Proposition 4

We keep the notation used in the proof of Proposition 2, and follow a similar argument. To beginwith, note that every ξ[

β= (ξ[

β,1, . . . , ξ[β,d)ᵀ ∈ ∂Q[

n(β) can be written as

ξ[β, j =1n

n∑i=1

wixi jI(yi ≤ 〈xi,β〉) − τ −1n

n∑i=1

wixi jvi + (1 − τ)I(yi = 〈xi,β〉),

34

Page 35: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

where vi ∈ [τ − 1, τ]. As before, the bound 〈ξ[β− ξ[β∗,β − β∗〉 ≥ U[

n(δ) holds with probability one,where

U[n(δ) :=

1n

n∑i=1

wi〈xi, δ〉I(εi ≤ 〈xi, δ〉) − I(εi ≤ 0) for δ = β − β∗.

Again, introducing Lipschitz continuous functions ϕ±a (u) and ψ±b (u) as in (57) and (58), weobtain

U[n(δ) ≥

1n

n∑i=1

wiεi,+ϕ+εi,+

(〈xi, δ〉)ψ+2r‖δ‖Σ(〈xi, δ〉)

+1n

n∑i=1

wiεi,−ϕ−εi,−

(〈xi, δ〉)ψ−2r‖δ‖Σ(〈xi, δ〉) := Vn(δ) + V[n(δ), (72)

where Vn(δ) = V+n (δ) + V−n (δ) is defined in (62) and

V[n(δ) =

1n

n∑i=1

eiεi,+ϕ+εi,+

(〈xi, δ〉)ψ+2r‖δ‖Σ(〈xi, δ〉) +

1n

n∑i=1

eiεi,−ϕ−εi,−

(〈xi, δ〉)ψ−2r‖δ‖Σ(〈xi, δ〉).

Notice that E∗V[n(δ) = 0. For any R ≥ δ, define Γn(R) = f (e1, . . . , en; R) := supδ≤‖δ‖Σ≤R −V[

n(δ).For each index i and an independent copy ei of ei, we have

| f (e1, . . . , ei, . . . , en; R) − f (e1, . . . , ei, . . . , en; R)| ≤Rrn.

Applying McDiarmid’s inequality gives

Γn(R) ≤ E∗Γn(R) + Rr

√t

2n(73)

with P∗-probability at least 1 − e−t. Using the Lipschitz continuity of u 7→ εi,±ϕ+εi,±

(u)ψ±b (u), andTalagrand’s contraction principle, we obtain

E∗Γn(R) ≤ 2E∗(

supδ≤‖δ‖Σ≤R

1n

n∑i=1

ei〈xi, δ〉

)≤

2RnE∗

∥∥∥∥∥ n∑i=1

ei zi

∥∥∥∥∥2

≤2Rn

( n∑i=1

‖zi‖22

)1/2

= 2RMn,d

√dn, (74)

where zi are defined in (23) and M2n,d := (1/nd)

∑ni=1

∑dj=1 z2

i j. Together, (73) and (74) imply

Γn(R) ≤ 2RMn,d

√dn

+ Rr

√t

2n(75)

with P∗-probability at least 1 − e−t.Note that inequality (75) holds for every R ≥ δ. Again, via the slicing technique and taking

r = 4υ20, it can be shown that for any t1 > 0, with P∗-probability at least 1− e−(t1δ)

2n

1−e−(t1δ)2n

= 1−P(n, t1, δ),

V[n(δ) ≥ −2Mn,d‖δ‖Σ

√2dn− 4t1υ2

0‖δ‖2Σ

35

Page 36: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

uniformly over ‖δ‖Σ ≥ δ. For the data-dependent quantity Mn,d, note that M2n,d ≤ max1≤ j≤d(1/n)

∑ni=1 z2

i j.Under Condition 1, we have E(z2

i j) = 1, and for k = 2, 3, . . . ,

E(z2i j)

k = υ2k0 2k

∫ ∞

0P(|zi j| ≥ υ0x)x2k−1dx ≤ 2k+1υ2k

0 k! =k!2

16υ40(2υ2

0)k−2.

It then follows from Bernstein’s inequality that, for any 1 ≤ j ≤ d and x ≥ 0,

P

(1n

n∑i=1

z2i j ≥ 1 + 4υ2

0

√2xn

+ 2υ20

xn

)≤ e−x.

Taking x = log(2d) + t and applying the union bound, we obtain

M2n,d ≤ 1 + 4υ2

0

√2 log(2d) + 2t

n+ 2υ2

0log(2d) + t

n(76)

with probability at least 1 − e−t/2.Turning to Vn(δ) in (72), it follows from (69) and (70) that with probability at least 1−P(n, t0, δ),

Vn(δ) ≥(2/9 − 8e−8) f − 4υ2

0t0‖δ‖2Σ − 4‖δ‖Σ

√2dn

(77)

for all δ satisfying δ ≤ ‖δ‖Σ ≤ f /(6L0υ20). Let G(t, t0, δ) be the event that (76) and (77) hold. Then

PG(t, t0, δ) ≥ 1 − e−t/2 − P(n, t0, δ). Taking t0 = t1 = (2/9 − 8e−8 − 1/8) f /(8υ20) yields that with

P∗-probability at least 1 − P(n, t1, δ) conditioned on G(t, t0, δ),

〈ξ[β − ξ[β∗ ,β − β

∗〉 ≥18

f ‖δ‖2Σ − 8υ20‖δ‖Σ

√2dn

uniformly over δ ≤ ‖δ‖Σ ≤ f /(6L0υ20) as long as n ≥ Cυ4

0log(d) + t

for some universal constant

C > 0. For any t ≥ 0, we assume that

‖β − β∗‖Σ ≥ δ := (64υ20/ f )

√2(d + t)

n;

otherwise, (25) holds trivially, and the above choices of (t0, t1, δ) guarantee that P(n, t0, δ) = P(n, t1, δ) ≤e−t/2. This completes the proof.

B Additional Simulation Studies

This section presents additional numerical results under various combinations of the design anderror distributions.

36

Page 37: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

B.1 Confidence estimation

Independent Gaussian designCoverage probability Width

α pair pwy wild mb-per mb-norm pair pwy wild mb-per mb-norm0.05: 0.972 0.974 0.946 0.963 0.952 0.494 0.507 0.435 0.437 0.4340.1: 0.935 0.945 0.905 0.923 0.909 0.414 0.425 0.365 0.361 0.3640.2: 0.871 0.879 0.822 0.823 0.819 0.323 0.331 0.285 0.277 0.284

Weakly correlated Gaussian designCoverage probability Width

α pair pwy wild mb-per mb-norm pair pwy wild mb-per mb-norm0.05: 0.978 0.980 0.951 0.970 0.955 0.735 0.754 0.650 0.652 0.6460.1: 0.946 0.952 0.908 0.929 0.911 0.617 0.633 0.545 0.538 0.5420.2: 0.871 0.876 0.812 0.836 0.820 0.481 0.493 0.425 0.412 0.422

Equally correlated Gaussian designCoverage probability Width

α pair pwy wild mb-per mb-norm pair pwy wild mb-per mb-norm0.05: 0.980 0.982 0.950 0.974 0.958 0.784 0.804 0.693 0.694 0.6880.1: 0.947 0.954 0.914 0.934 0.912 0.658 0.675 0.581 0.573 0.5770.2: 0.871 0.881 0.821 0.837 0.820 0.512 0.526 0.453 0.438 0.450

Table 7: Average coverage probabilities and CI widths over all the coefficients under homoscedasticmodel (17) with t2 error.

Independent Gaussian designCoverage probability Width

α pair pwy wild mb-per mb-norm pair pwy wild mb-per mb-norm0.05: 0.981 0.980 0.958 0.965 0.966 0.435 0.447 0.379 0.389 0.3840.1: 0.958 0.959 0.919 0.920 0.926 0.365 0.375 0.318 0.320 0.3230.2: 0.879 0.891 0.814 0.824 0.826 0.285 0.292 0.248 0.243 0.251

Weakly correlated Gaussian designCoverage probability Width

α pair pwy wild mb-per mb-norm pair pwy wild mb-per mb-norm0.05: 0.980 0.984 0.960 0.970 0.957 0.667 0.685 0.586 0.593 0.5860.1: 0.952 0.957 0.922 0.926 0.920 0.560 0.575 0.492 0.488 0.4920.2: 0.877 0.889 0.843 0.835 0.834 0.436 0.448 0.383 0.372 0.383

Equally correlated Gaussian designCoverage probability Width

α pair pwy wild mb-per mb-norm pair pwy wild mb-per mb-norm0.05: 0.986 0.990 0.968 0.974 0.974 0.716 0.734 0.626 0.636 0.6290.1: 0.964 0.970 0.927 0.933 0.931 0.601 0.616 0.526 0.523 0.5280.2: 0.887 0.898 0.835 0.836 0.840 0.468 0.480 0.409 0.399 0.411

Table 8: Average coverage probabilities and CI widths over all the coefficients under heteroscedasticmodel (18) with t2 error.

37

Page 38: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

Independent Gaussian designCoverage probability Width

α pair pwy wild mb-per mb-norm pair pwy wild mb-per mb-norm0.05: 0.966 0.968 0.939 0.966 0.943 0.438 0.448 0.388 0.386 0.3830.1: 0.929 0.936 0.889 0.920 0.891 0.367 0.376 0.326 0.320 0.3220.2: 0.847 0.857 0.784 0.821 0.787 0.286 0.293 0.254 0.245 0.251

Weakly correlated Gaussian designCoverage probability Width

α pair pwy wild mb-per mb-norm pair pwy wild mb-per mb-norm0.05: 0.970 0.972 0.938 0.964 0.945 0.652 0.669 0.580 0.576 0.5720.1: 0.935 0.939 0.886 0.920 0.891 0.547 0.562 0.487 0.477 0.4800.2: 0.852 0.860 0.800 0.834 0.796 0.427 0.438 0.379 0.366 0.374

Equally correlated Gaussian designCoverage probability Width

α pair pwy wild mb-per mb-norm pair pwy wild mb-per mb-norm0.05: 0.962 0.967 0.937 0.966 0.941 0.695 0.713 0.618 0.614 0.6100.1: 0.932 0.939 0.891 0.924 0.891 0.583 0.599 0.519 0.509 0.5120.2: 0.850 0.859 0.803 0.825 0.801 0.454 0.466 0.404 0.390 0.399

Table 9: Average coverage probabilities and CI widths over all the coefficients under homoscedasticmodel (17) with type II mixture normal error.

Independent Gaussian designCoverage probability Width

α pair pwy wild mb-per mb-norm pair pwy wild mb-per mb-norm0.05: 0.977 0.980 0.956 0.966 0.961 0.386 0.397 0.339 0.344 0.3400.1: 0.949 0.956 0.910 0.924 0.914 0.324 0.333 0.284 0.283 0.2860.2: 0.870 0.879 0.816 0.821 0.818 0.253 0.259 0.222 0.216 0.223

Weakly correlated Gaussian designCoverage probability Width

α pair pwy wild mb-per mb-norm pair pwy wild mb-per mb-norm0.05: 0.976 0.980 0.951 0.963 0.950 0.591 0.607 0.522 0.525 0.5190.1: 0.941 0.946 0.901 0.922 0.908 0.496 0.509 0.438 0.433 0.4360.2: 0.869 0.883 0.815 0.825 0.819 0.387 0.397 0.341 0.331 0.340

Equally correlated Gaussian designCoverage probability Width

α pair pwy wild mb-per mb-norm pair pwy wild mb-per mb-norm0.05: 0.972 0.976 0.950 0.965 0.946 0.626 0.643 0.553 0.556 0.5510.1: 0.937 0.946 0.903 0.926 0.908 0.525 0.540 0.464 0.459 0.4620.2: 0.866 0.874 0.815 0.839 0.812 0.409 0.421 0.361 0.350 0.360

Table 10: Average coverage probabilities and CI widths over all the coefficients under heteroscedas-tic model (18) with type II mixture normal error.

38

Page 39: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

B.2 Goodness-of-fit testing

Independent Gaussian designType I error under null model Power under sparse model Power under dense model

α Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad0.01 0.330 0.005 0.000 0.005 0.945 0.490 0.535 0.600 0.685 0.095 0.075 0.0850.05 0.465 0.015 0.035 0.020 0.985 0.735 0.865 0.855 0.790 0.330 0.330 0.3200.1 0.550 0.070 0.080 0.055 0.990 0.845 0.945 0.930 0.840 0.445 0.505 0.440

Weakly correlated Gaussian designType I error under null model Power under sparse model Power under dense model

α Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad0.01 0.320 0.000 0.000 0.005 0.785 0.295 0.365 0.370 0.835 0.350 0.375 0.4050.05 0.475 0.040 0.020 0.015 0.900 0.595 0.645 0.595 0.940 0.650 0.710 0.6650.1 0.565 0.065 0.050 0.040 0.935 0.715 0.750 0.730 0.975 0.770 0.815 0.805

Equally correlated Gaussian designType I error under null model Power under sparse model Power under dense model

α Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad0.01 0.320 0.000 0.000 0.005 0.810 0.330 0.345 0.380 0.985 0.760 0.850 0.8450.05 0.475 0.040 0.020 0.015 0.890 0.570 0.630 0.595 0.995 0.935 0.965 0.9650.1 0.565 0.065 0.050 0.040 0.935 0.690 0.755 0.735 1.000 0.980 0.990 0.990

Table 11: Average type I error and power under homoscedastic model (17) with t2 error.

Independent Gaussian designType I error under null model Power under sparse model Power under dense model

α Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad0.01 0.325 0.005 0.000 0.000 0.935 0.650 0.770 0.810 0.745 0.205 0.160 0.2000.05 0.475 0.020 0.010 0.005 0.975 0.865 0.935 0.940 0.840 0.480 0.425 0.3950.1 0.520 0.055 0.045 0.020 0.990 0.940 0.965 0.960 0.890 0.620 0.630 0.555

Weakly correlated Gaussian designType I error under null model Power under sparse model Power under dense model

α Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad0.01 0.315 0.000 0.000 0.000 0.840 0.415 0.490 0.510 0.905 0.470 0.475 0.5350.05 0.465 0.050 0.010 0.010 0.910 0.685 0.705 0.685 0.955 0.750 0.785 0.7750.1 0.550 0.095 0.030 0.025 0.945 0.810 0.815 0.790 0.975 0.845 0.900 0.865

Equally correlated Gaussian designType I error under null model Power under sparse model Power under dense model

α Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad0.01 0.300 0.015 0.000 0.000 0.845 0.425 0.450 0.500 0.990 0.840 0.920 0.9300.05 0.445 0.040 0.015 0.005 0.925 0.675 0.750 0.745 0.995 0.970 0.985 0.9800.1 0.510 0.085 0.060 0.030 0.950 0.800 0.870 0.835 0.995 0.990 1.000 1.000

Table 12: Average type I error and power under heteroscedastic model (18) with t2 error.

39

Page 40: Multiplier Bootstrap for Quantile Regression: Non ...wez243/MBQR.pdf · estimator and its bootstrap counterpart. In Section2.3, we further disucss the use of multiplier boot-strap

Independent Gaussian designType I error under null model Power under sparse model Power under dense model

α Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad0.01 0.290 0.005 0.015 0.015 0.945 0.610 0.680 0.705 0.695 0.150 0.175 0.1800.05 0.435 0.040 0.035 0.030 0.990 0.800 0.895 0.900 0.860 0.395 0.495 0.4400.1 0.505 0.095 0.090 0.055 1.000 0.910 0.935 0.920 0.890 0.595 0.645 0.595

Weakly correlated Gaussian designType I error under null model Power under sparse model Power under dense model

α Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad0.01 0.345 0.005 0.005 0.005 0.860 0.420 0.530 0.565 0.905 0.455 0.530 0.5250.05 0.525 0.025 0.035 0.030 0.945 0.695 0.760 0.725 0.945 0.675 0.775 0.7250.1 0.600 0.075 0.110 0.085 0.965 0.815 0.875 0.855 0.960 0.810 0.860 0.830

Equally correlated Gaussian designType I error under null model Power under sparse model Power under dense model

α Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad0.01 0.345 0.005 0.005 0.005 0.875 0.385 0.485 0.525 0.980 0.890 0.950 0.9400.05 0.525 0.025 0.035 0.030 0.930 0.670 0.725 0.700 1.000 0.980 0.995 0.9900.1 0.600 0.075 0.110 0.085 0.960 0.810 0.845 0.790 1.000 0.990 0.995 0.995

Table 13: Average type I error and power under homoscedastic model (17) with type II mixturenormal error.

Independent Gaussian designType I error under null model Power under sparse model Power under dense model

α Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad0.01 0.320 0.010 0.005 0.005 0.960 0.760 0.840 0.840 0.815 0.300 0.250 0.2600.05 0.440 0.035 0.020 0.025 0.990 0.915 0.965 0.955 0.920 0.595 0.630 0.5650.1 0.495 0.085 0.060 0.045 0.995 0.970 0.975 0.970 0.950 0.755 0.755 0.725

Weakly correlated Gaussian designType I error under null model Power under sparse model Power under dense model

α Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad0.01 0.350 0.010 0.000 0.000 0.900 0.570 0.640 0.665 0.925 0.545 0.625 0.6550.05 0.485 0.040 0.025 0.020 0.950 0.790 0.850 0.835 0.975 0.790 0.870 0.8550.1 0.590 0.095 0.060 0.045 0.965 0.900 0.930 0.890 1.000 0.870 0.925 0.910

Equally correlated Gaussian designType I error under null model Power under sparse model Power under dense model

α Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad Wald rank mb-exp mb-Rad0.01 0.300 0.005 0.000 0.000 0.915 0.525 0.635 0.655 1.000 0.920 0.970 0.9700.05 0.415 0.040 0.015 0.015 0.955 0.805 0.855 0.835 1.000 0.980 0.995 0.9950.1 0.525 0.085 0.055 0.040 0.980 0.885 0.920 0.910 1.000 0.995 1.000 1.000

Table 14: Average type I error and power under heteroscedastic model (18) with type II mixturenormal error.

40