High Dimensional Inference in Partially Linear Models Ying Zhu∗, Zhuqing Yu†, and Guang Cheng‡

Abstract We propose two semiparametric versions of the debiased Lasso procedure for the model Yi = Xi β0 + g0 (Zi ) + εi , where the parameter vector of interest β0 is high dimensional but sparse (exactly or approximately) and g0 is an unknown nuisance function. Both versions are shown to have the same asymptotic normal distribution and do not require the minimal signal condition for statistical inference of any component in β0 . Our method also works when the vector of covariates Zi is high dimensional provided that the function classes E(Xij |Zi )s and E(Yi |Zi ) belong to exhibit certain sparsity features, e.g., a sparse additive decomposition structure. We further develop a simultaneous hypothesis testing procedure based on multiplier bootstrap. Our testing method automatically takes into account of the dependence structure within the debiased estimates, and allows the number of tested components to be exponentially high. Keywords: partial linear models, high dimensional inference, simultaneous testing, de-biased estimate, nonparametric projection This version: August 8, 2017

1

Introduction

Semiparametric regression is a longstanding statistical tool that leverages the flexibility of nonparametric models while avoiding the “curse of dimensionality” (Bickel et al, 1998). A leading example of semiparametric regression models is the partially linear regression Yi = Xi β0 + g0 (Zi ) + εi , i = 1, ..., n.

(1)

In (1), β0 ∈ Rp is an unknown vector and g0 is an unknown function; X := (Xi )ni=1 ∈ Rn×p and Z := (Zi )ni=1 ∈ Rn×d are observed covariates (Xi and Zi denote the ith row of X and Z, respectively), Y := (Yi )ni=1 ∈ Rn are the response variables, ε = (εi )ni=1 ∈ Rn is a noise vector with E(εi ) = 0 and E(ε2i ) = 1, and independent of (X, Z). Throughout the paper, we assume that the data {Xi , Zi , Yi }ni=1 are i.i.d.. The goal of this paper is to establish statistical inference results, e.g., confidence intervals and hypothesis testing, for the high dimensional component β0 in presence of the nuisance function g0 . In particular, we assume that p ≥ n and β0 exhibits sufficient sparsity (meaning that the ordered coefficients in β0 decay sufficiently fast). Our method also works when Zi is high dimensional (d ≥ n) provided that the function classes E(Xij |Zi )s and E(Yi |Zi ) belong to exhibit certain sparsity features, e.g., a sparse additive decomposition structure as defined in Raskutti, et al. (2012). ∗

Research Associate. Department of Economics. Michigan State University. Corresponding author Graduate Student. Department of Statistics. Purdue University. ‡ Professor of Statistics. Department of Statistics. Purdue University. Research Sponsored by NSF CAREER Award DMS-1151692, DMS-1418042, and Office of Naval Research. †

1

For statistical inference of β0 in (1), existing results mainly focus on the regime where p increases with n but smaller than n, for example, Li and Liang (2008), Xie and Huang (2009), and Cheng, et al. (2015). Sherwood and Wang (2016) allow p ≥ n but require the minimal signal condition. Such results therefore suffer the problems arising from the nonuniformity of limit theory. Recently, Javanmard and Montanari (2014), van de Geer, et al. (2014), and Zhang and Zhang (2014) have proposed the debiased Lasso for high dimensional linear models. These estimators are non-sparse, have a limiting normal distribution, and do not require the minimal signal condition. For the linear model Y = Xβ0 + ε, given an initial Lasso estimate βˆ of β0 , the debiased Lasso adds a correction ˆ to remove the bias introduced by regularization. In particular, term to βˆj (the jth component of β) the correction term takes the form of ˆ j 1 X T Y − X βˆ . Γ n 





(2)







In (2), n−1 X T Y − X βˆ is the sample analogue of the population score function E XiT (Yi − Xi β0 ) ; ˆ j denotes the jth row of Γ ˆ where Γ ˆ is an approximate inverse of n−1 X T X, whose population counΓ 



terpart is E XiT Xi . In our model (1), additional bias arises due to the presence of g0 ; consequently, the standard debiased Lasso cannot rid of the effect from g0 and thus will not have a limiting distribution centered around zero. Instead, we propose two modified versions of the debiased Lasso estimators for β0 . Both versions are shown to be asymptotically unbiased for β0 , have the same limiting (normal) distribution, and do not require the minimal signal condition. Our modified debiased Lasso estimators use a “nonparametric projection” strategy to remove the impact of g0 in (1). Such a strategy has been used in the semiparametric inference literature where p is assumed to be small relative to n (e.g., Robinson, 1988; Donald and Newey, 1994; Liang and Li, 2009). To be more specific, by taking the conditional expectations of the left side and the right side of (1) with respect to Zi , we obtain E (Yi |Zi ) = E (Xi |Zi ) β0 + g0 (Zi )

(3)

where we exploit the fact that E (εi |Zi ) = 0. Subtracting E (Yi |Zi ) from Yi and E (Xi |Zi ) + g0 (Zi ) from Xi + g0 (Zi ) in (1) yields ˜ i β0 + ε i Y˜i = X (4) 

˜ ij := Xij − E (Xij |Zi ) and X ˜ i := X ˜ ij where Y˜i := Yi − E (Yi |Zi ), X

p j=1

(which is a p−dimensional

row vector). Relating (4) to the linear model Yi = Xi β0 + εi , given nonparametric surrogates Yˆi := Yi − ˆ ˆ (Xij |Zi ) of X ˆ ij := Xij − E ˜ ij (j = 1, ..., p), we simply replace Yi with Yˆi , E (Yi |Zi ) of Y˜i and X  p ˆ j with the jth row (Θ ˆ j ) of an approximate inverse ˆ i := X ˆ ij Xi with the row vector X , and Γ j=1

ˆ of n−1 Pn X ˆT ˆ (denoted as Θ) i=1 i Xi in (2). This yields our first semiparametric version of the debiased procedure   ˆbj := βˆj + Θ ˆj 1X ˆ T Yˆ − X ˆ βˆ , (5) n   ˆ T Yˆ − X ˆ βˆ in (5) is simply where βˆ is an initial estimate of β0 . Alternatively, by noting that n−1 X 



˜ T εi , we arrive at our second debiased the sample analogue of the population score function E X i procedure   ˜bj := βˆj + Θ ˆj 1X ˆ T Y − X βˆ − gˆ , (6) n where gˆ is an estimate of g0 . 2

We provide theoretical implications on the impact of the estimation errors associated with the ˆ (Xij |Zi )s in our modified debiased Lasso procedures when each of p nonparametric surrogates E ˆ E (Xij |Zi )s concerns a large family of (regularized) nonparametric least squares estimators. These ˆ (Yi |Zi ) (which matters to (5)) and the surrogate implications also hold true for the surrogate E gˆ(Zi ) (which matters to (6)). After careful theoretical analysis, we find that if the error of the nonparametric regression per se (with respect to the prediction norm) is Op (rn ), it only contributes  Op rn2 in the asymptotic expansions of ˆbj −β0j and ˜bj −β0j for any j = 1, ..., p, where rn is related to the optimal rate for the nonparametric regression. This result implies that even with p much larger than n (and/or with the dimension d of Zi much larger than n), the limiting distribution of our modified debiased estimators for any individual component in β0 may behave as if the unknown conditional expectations E (Xij |Zi )s and E (Yi |Zi ) as well as the unknown function g0 (Zi ) were known. This theoretical finding motivates us to consider a multiplier-bootstrap-based simultaneous hypothesis testing procedure for any sub-vector of β0 . This extends the method developed by Zhang and Cheng (2017) from linear regressions to more flexible partially linear regressions. Our simultaneous testing procedure automatically takes into account of the dependence structure within our debiased estimators, and allows the number of tested components to be exponentially high. We illustrate the theoretical finding with four specific examples in terms of dim (Zi ) and the function class that E (Xij |Zi )s and E (Yi |Zi ) belong to. With regard to the specific forms of ˆ (Yi |Zi ), several modern techniques for the projection step are considered and ˆ (Xij |Zi )s and E E the rates achieved by these practical procedures are compared with the theoretical results. The techniques discussed in the paper include the smoothing splines estimator in Sobolev balls, the Lasso (Tibshirani, 1996) and Slope (Su and Candès, 2016) in sparse linear regression models, the (square-root) Lasso (Belloni, et al., 2014) in rearranged Sobolev balls, and the l1 −regularized kernel ridge regression (Raskutti et al., 2012) in sparse additive models. ˆ of n−1 Pn X ˆT ˆ When constructing an approximate inverse Θ i=1 i Xi in (5) and (6), we adopt the nodewise regression method proposed by van de Geer, et al. (2014). Since our analysis involves

ˆ

establishing Θ − Θ = o

j j p (1), as in van de Geer et al. (2014), we require a sparsity condition 1





˜TX ˜ i . This fact renders the method on the inverse Θ = Σ−1 of the population Hessian Σ := E X i ˆ inapplicable as their approach is only valid in Javanmard and Montanari (2014) for constructing Θ for fixed X, while our analysis accounts for the randomness in X and the estimation errors in ˆ (Xij |Zi )s. Furthermore, our analysis relaxes the exact sparsity of Θj (assumed in most literature E including van de Geer, et al., 2014) to accommodate for approximate sparsity which permits all the entries in Θj to be non-zero as long as they decay sufficiently fast. This extension provides a more realistic interpretation of most practical problems. The rest of the paper is organized as follows. Section 2 presents the detailed construction of the modified debiased estimators for β0 and the simultaneous testing procedure. Section 3 establishes the main theoretical results. All technical details are deferred to Section 4. Section 5 evaluates the performance of our methods with simulation experiments. Notation. The lq −norm of a p−dimensional vector ∆ is denoted by k∆k q for 1 ≤ q ≤ ∞. For a matrix H ∈ Rp1 ×p2 , write kHk ∞ := maxi,j |Hij | to be the elementwise l∞ −norm of H. Let Hj denote the jth row of H. The L2 (Pn )−norm of the vector ∆ := {∆(Xi )}ni=1 , denoted h P

i1

by k∆kn , is given by n1 ni=1 (∆(Xi ))2 2 . For functions f (n) and g(n), write f (n) % g(n) to mean that f (n) ≥ cg(n) for a universal constant c ∈ (0, ∞) and similarly, f (n) - g(n) to mean 0 0 that f (n) ≤ c g(n) for a universal constant c ∈ (0, ∞), and f (n)  g(n) when f (n) % g(n) and f (n) - g(n) hold simultaneously. Also denote max{a, b} by a ∨ b and min{a, b} by a ∧ b. As a 3

general rule for this paper, all the c ∈ (0, ∞) constants denote positive universal constants. The specific values of these constants may change from place to place.

2

Main methodology

In this section we discuss the construction of ˆbj and ˜bj in detail. Note that both (5) and (6) require estimators for the conditional expectations, an initial estimator βˆ for β0 (and an estimator gˆ for ˆ for n−1 Pn X ˆT ˆ g0 in ˜bj ), and also an approximate inverse Θ i=1 i Xi . We first discuss how to obtain these aforementioned quantities. Given ˆbj s and ˜bj s, we then present the simultaneous inference procedure. Estimators for the conditional expectations For either (5) or (6), we need to estimate the conditional expectations E (Xij |Zi )s (j = 1, ..., p). This step is easily paralleable as it involves solving p independent subproblems and each subproblem can be in general solved with an efficient algorithm. In contrast with (6), (5) does not require an estimate of g0 but an estimate of E (Yi |Zi ). Estimating conditional expectations is widely studied in the literature on nonparametric methods. For the purpose of this paper, global properties of ˆ (Xij |Zi )s and E ˆ (Yi |Zi ) are the key to our analysis of the debiased the nonparametric estimators E procedures and therefore, we focus on the following least squares estimators (

fˆj ∈ arg min

fj ∈Fj

n 1 X (wij − fj (zi ))2 , 2n i=1

)

(7)

ˆ (Yi |Zi ) and fˆj (Zi ) as E ˆ (Xij |Zi ). where wi0 = yi and wij = xij for j = 1, ..., p. Denote fˆ0 (Zi ) as E A nonparametric regression problem like (7) is a standard setup in many modern statistics books (e.g., van de Geer, 2000; Wainwright, 2015). Examples of (7) include linear regression, sparse linear regression, series projection, convex regression, Lipschitz Isotonic regression, and kernel ridge regression (KRR). In the case of KRR, we restrict Fj in (7) to be a compact subset of a reproducing kernel Hilbert space (RKHS) H, equipped with a norm k·kH ; (7)1 can then be reformulated in its Lagrangian form ( ) n X 1 2 2 (wij − fj (zi )) + µj kfj kH fˆj ∈ arg min (8) fj ∈H 2n i=1 where µj > 0 is a regularization parameter. In particular, smoothing spline estimators can be viewed as special cases of KRR. Initial estimators for β0 and g0 In a semiparametric regression model like (1), Zhu (2017) covers a wide spectrum of function classes that the nonparametric component g0 (·) may belong to and provides a general nonasymptotic theory for estimating β0 and g0 . The estimators βˆ and gˆ in Zhu (2017) can be used as initial estimators in (5)-(6). Given the way βˆ is obtained in Zhu (2017), the estimated conditional expectations, fˆj s, come in handy as byproducts (therefore, separate estimations for the conditional expectations are not needed in the construction of ˆbj or ˜bj ). 1

To be more specific, we let Fj be a ball of radius R in the norm k·kH and assume R ≤ 1 throughout the asymptotic analysis to avoid carrying “R”s around.

4

For special cases where Zi has a low dimension and g0 belongs to the mth order Sobolev ball other estimators for β0 and g0 are also available [Müller and van de Geer, 2015; Yu, et al. 2017] and take the following form

S m,

(

ˆ gˆ) ∈ arg (β,

min

β∈Rp ,g∈S m

n 1 X (Yi − Xi β − g(Zi ))2 + µJ 2 (g) + λ kβk 1 , 2n i=1

)

(9)

where µ, λ > 0 are regularization parameters, J 2 (g) := 01 (g m (z))2 dz and Z ∼ Uniform(0, 1). Due to the intractable limiting distribution of Lasso type estimators, these aforementioned papers do not provide any distributional results for their proposed estimators. In Section 3, we take the debiased versions, (5) and (6), of these aforementioned initial estimators and establish the asymptotic normality of individual components in the debiased estimators. R

Estimator for the inverse of the population Hessian ˆ of ˆ i of X ˜ i via (7), we obtain an approximate inverse Θ Given the estimates Yˆi of Y˜i and X Pn ˆ T ˆ −1 method proposed by van de Geer, et al. (2014). n i=1 Xi Xi using the nodewise regression

ˆ

Since our analysis involves establishing Θj − Θj = op (1), as in van de Geer, et al. (2014), we re1





˜TX ˜ i . Lack quire a sparsity condition on the inverse Θ = Σ−1 of the population Hessian Σ := E X i   ˆ j − Θj √1 X ˜ T ε in of sparsity in the off-diagonal elements of Θ will cause remainder terms like Θ n     √ √ the asymptotic expansions2 of n ˆbj − β0j or n ˜bj − β0j to diverge and the resulting limiting distribution may not be well-behaved for any practical purpose. This fact renders the method in ˆ inapplicable as their approach is only valid Javanmard and Montanari (2014) for constructing Θ for fixed X, while our analysis accounts for the randomness in X and the estimation errors in ˆ (Xij |Zi )s. E To apply the nodewise regression method in our context, for each 1 ≤ j ≤ p, let us define 

π ˆj ∈ arg min

π ˜j ∈Rp−1

2 1

ˆ

ˆ −j π ˜j + λj k˜ π j k1 ,

Xj − X 2 n 

ˆ −j denotes the submatrix of X ˆ without the jth column. Let where X   

ˆ :=  M  

n

0

1 −ˆ π1,2 · · · −ˆ π2,1 1 ··· .. .. .. . . . −ˆ πp,1 −ˆ πp,2 · · ·

−ˆ π1,p −ˆ π2,p .. . 1

   .  

o

Based on π ˆj := π ˆj,j 0 ; j 6= j , for each 1 ≤ j ≤ p, we compute τˆj2 := and write

2 1

ˆ

ˆ −j π ˆj + λj kˆ πj k1

Xj − X 2 n 



Tˆ2 := diag τˆ12 , ..., τˆp2 . ˆ := Tˆ−2 M ˆ. Finally, we define Θ 2

See (21)-(22) for more detail.

5

(10)

For later presentations of the theoretical results, we also introduce the population counter˜ ij on X ˜ i,−j = parts of the o above quantities: let πj be the population regression coefficients of X n 0 ˜ 0 ; j 6= j and X ij 

M

 

:=   

1 −π1,2 · · · −π2,1 1 ··· .. .. .. . . . −πp,1 −πp,2 · · · 

−π1,p −π2,p .. . 1

   ,  



T 2 := diag τ12 , ..., τp2 , such that τj2 := E



˜ ij − X ˜ i,−j πj X

2 

for j = 1, ..., p.

Simultaneous inference From a practical viewpoint, conducting simultaneous inference for a collection of parameters in highdimensional models may be of greater interest to researchers than inference of a single parameter. To be more specific, suppose we are interested in testing the hypothesis: H0,G : β0j = β˜j ∀j ∈ G ⊆ {1, 2, ..., p} against the alternative Ha,G : β0j 6= β˜j for some j ∈ G. In particular, we allow |G| ≥ n. Zhang and Cheng (2017) develop a bootstrap-assisted procedure to conduct simultaneous inference in sparse linear models. Here we propose similar test statistics TG = max

 √  n ˆbj − β˜j ,

or, TG = max

 √  n ˜bj − β˜j ,

j∈G

j∈G

and a multiplier bootstrap version n 1 X ˆ jX ˆ T i , Θ WG = max √ i j∈G n i=1

where (i )ni=1 are i.i.d. N (0, 1) random variables. The bootstrap critical value is then given by cG (α) = inf {t ∈ R : P (WG ≤ t|Y, X, Z) ≥ 1 − α} for any user-defined α ∈ (0, 1). In the case where the variance σε2 of εi in (1) is unknown, we can use n 1 X ˆ jX ˆ iT σ WG = max √ ˆε i , Θ j∈G n i=1 where σ ˆε2 =

Pn

2

(Yˆi −Xi βˆ) or σ ˆε2 = n−kβˆk1

i=1

Pn i=1

2

ˆ g (Zi )) (Yi −Xi β−ˆ is an estimator for σε2 . n−kβˆk1

6

3

Theoretical results

To make the key point of this paper, we first present the results for the case where β0 and πj are assumed to be exactly sparse (Theorem 1). Our second result (Theorem 2 in Section 3.4) relaxes the exact sparsity assumptions and allows β0 and πj to be approximately sparse. Note that the (exact or approximate) sparsity of πj along with the condition τ12 - 1 implies the sparsity of j

Θj =

h 

˜TX ˜i E X i

i−1 

.

j

We begin with the following definitions. Let s0 := |{j : β0j 6= 0}| , sj

n

0

0

o

:= j 6= j, 1 ≤ j ≤ p : πj,j 0 6= 0 .

To simplify our notations, we assume that kβ0 k1 - s0 and kπj k1 - sj in Theorem 1 and Corollary 1. Recall Fj in (7); for notation simplicity, we assume Fj = F from now on. Note that this restriction can be easily relaxed in our analysis. For any radius r˜n > 0, we define the conditional local complexity " # n 1 X ξi f (Zi ) | {Zi }ni=1 , Gn (˜ rn ; F) := Eξ sup (11) f ∈Ω(˜ rn ; F ) n i=1 where ξi s are i.i.d. zero-mean sub-Gaussian variables with parameter at most σ † and E (ξi |Zi ) = 0 for all i = 1, ..., n, and Ω(˜ rn ; F) :=

n

F¯ :=

n

o

f ∈ F¯ : kf kn ≤ r˜n , 0

f =f −f

00

0

00

o

: f ,f ∈F .

For any star-shaped class F¯ (that is, for any f ∈ F¯ and α ∈ [0, 1], αf ∈ F¯ ), Lemma A8 in Section 4 guarantees that the function t 7→ Gn (t;t F ) is non-increasing on the interval (0, ∞). Therefore, there exists some large enough r˜n > 0 that satisfies the critical inequality Gn (˜ rn ; F) ≤

r˜n2 ; 2

(12)

moreover, (12) has a smallest positive solution rn (which we will refer to as the critical radius). In practice, determining the exact value of this critical radius can be difficult; fortunately, reasonable upper bounds on rn are often available. Here we describe two common methods from existing literature. By a discretization argument and the Dudley’s entropy integral, we may bound (11) by c0

σ† √ n

!

Z r˜n q

log Nn (t; Ω(˜ rn ; F))dt +

0

r˜n2

for some universal constant c0 > 0, where Nn (t; Ω(˜ rn ; F)) is the t−covering number of the set Ω(˜ rn ; F). Let r˜n be a solution for σ† √ n

Z r˜n q 0

log Nn (t; Ω(˜ rn ; F))dt - r˜n2 .

(13)

The resulting r˜n is known to yield an upper bound on the critical radius rn for (12) (see Lemma A9 for a formal statement); moreover, such bounds achieve sharp scaling on rn for a wide variety of function classes (see e.g., Barlett and Mendelson, 2002; Koltchinski, 2006; Wainwright, 2015). 7

When F is a ball of radius R in the RKHS norm k·kH , we let n

o

Ω(˜ rn ; F) := f ∈ F¯ : kf kn ≤ r˜n , kf kH ≤ 1 . In this case, we can determine a good upper bound for rn using the result in Mendelson (2002) who shows that v u n u X †t 1 Gn (˜ rn ; F) - σ min {˜ rn2 , µ ˜i } n i=1 where µ ˜1 ≥ µ ˜2 ≥ ... ≥ µ ˜n ≥ 0 are the eigenvalues of the underlying kernel matrix for the KRR estimate. Consequently, we can solve for r˜nj via v u n u1 X σ†t min {˜ rn2 , µ ˜i } - r˜n2 . n i=1

This method above is known to yield r˜nj with sharp scaling for various choices of kernels.

3.1

The case of exact sparsity

We are now ready to establish our first main result (Theorem 1), which requires the following assumptions (in addition to those stated at the beginning of Section 1). 

˜ ij , Y˜i , εi Assumption 1 . For j = 1, ..., p, Xij , X rameters at most O(1). 

n i=1

are i.i.d. sub-Gaussian variables with pa-



˜TX ˜ i , the smallest eigenvalue Λ2 (Σ) is bounded away from Assumption 2 . (i) For Σ = E X min i 0, i.e.,

h

1 Λ2min (Σ)

i

- 1; moreover, Λ2max (Σ) - 1. (ii) For Σ1 = E E (Xi |Zi )T E (Xi |Zi ) (where

E (Xi |Zi ) := (E (Xi1 |Zi ) , ..., E (Xip |Zi )) is a p−dimensional row vector), Λ2max (Σ1 ) = O(1). Remark. Part (ii) of Assumption 2 is only used in the analysis for the second debiased estimator ˜b in (6). Assumption 3 . The n conditional expectations Eo(Yi |Zi ) and E (Xij |Zi ) (j = 1, ..., p) belong to 0 00 0 00 F. For any f ∈ F¯ = f˜ = f − f : f , f ∈ F and α ∈ [0, 1], αf ∈ F¯ (that is, F¯ is starshaped). Remark. This condition is needed to operationalize (5)-(6) which require estimators for the conditional expectations. In view of the following identity g0 (Zi ) = E (Yi |Zi ) − E (Xi |Zi ) β0 , note that imposing conditions on the function class E (Xij |Zi )s and E (Yi |Zi ) belong to automatically restricts the function class g0 belongs to.



Assumption 4 . (i) The initial estimator βˆ satisfies that βˆ − β0 = Op 2

 q

Op s0

log p n



.

 

(ii) E (X|Z) βˆ − β0

n

= Op 8

q

s0 log p n

q

s0 log p n







and βˆ − β0 = 1



and the estimator gˆ satisfies that

kˆ g − gk2n = Op



s0 rn2 ∨ 



s0 log p n



where rn is the critical radius.

Remark. Part (ii) of Assumption 4 is only used in the analysis for the second debiased estimator ˜b in (6). Under mild conditions, the rates in Assumption 4 are satisfied by the initial estimators based on Zhu (2017). For special cases where Zi has a low dimension and g0 belongs to the mth order Sobolev ball S m , the initial estimators based on Müller and van de Geer (2015) or Yu, et al. (2017) also satisfy the rate requirements in Assumption 4. q   2 log p Assumption 5 . kπj k1 ∨ 1 rn = O where rn is the critical radius. Moreover, n





nsj (s0 ∨ 1)

rn2

log p ∨ n



= o(1),

      s  √  2 log p log p log p  n sj (s0 ∨ 1) ∨ s3j (s0 ∨ 1) rn2 ∨ =  n n n    s     log p  log p 2  s2 ∨ 1  ∨ s3 ∨ 1 r ∨ = j

j

n

n

n

o(1),

o(1).

 √  ˜ T ε + REM Remark. With some algebraic manipulations, we show that n ˆbj − β0j = √1n Θj X   √ ˜ T ε + REM 0 , where the leading term √1 Θj X ˜ T ε has an asymptotic and n ˆbj − β0j = √1n Θj X n 0

normal distribution; see (21) and (22) for more details on the forms of REM and REM . Assumption 5 imposes requirements on the sparsity parameters s0 and sj s as well as the rates (reflected 0 by rn ) for the auxiliary estimators fˆj s so that REM = o(1) and REM = o(1). Assumption 6 . For all l 6= j, j = 1, ..., p, E

h

1 ˜T n Xl



˜j − X ˜ −j πj X

i

= O

q

log p n



as p → ∞

and n → ∞. Remark. For a general sub-Gaussian matrix, Assumption 6 is needed the scaling  in order to derive 

˜T ˜ j − πj X ˜ −j . Intuitively, X for λj , whose choice in (10) depends on an upper bound for n1 X

−j ∞ ˜ ˜ Assumption 6 hsays that as the number of terms ( X ) used to approximate X increases (that is, i,−j ij  i   log p 1 ˜T ˜ ˜ ˜ ˜ ˜ as p → ∞), E X Xj − πj X−j = o(1) provided = o(1). If E Xij |Xi,−j = Xi,−j πj (e.g., n

l

n

˜ i is a normal vector), then E 1 X ˜T X ˜ j − πj X ˜ −j when X n l considered in Theorem 2.2 in van de Geer, et al. (2014).

h



i

= 0 for all l 6= j. This special case is

Theorem 1 . Under Assumptions 1-6, if we choose λj 

q

log p n

 √ ˆ n bj − β0j

uniformly in j in (10), then

D

−→ N (0, 1) ,

σ ˆj  √ ˜ n bj − β0j

D

−→ N (0, 1) ,

σ ˆj ˆ j Xˆ T Xˆ Θ ˆ T , for each j = 1, ..., p. where σ ˆj2 = Θ j n

Based on Theorem 1, Corollary 1 justifies the use of multiplier bootstrap in testing H0,G even 9

when |G| diverges. Corollary 1 . Suppose Assumptions 1-4, 6 hold and Assumption 5 is satisfied with sj replaced q (log pn)7 log p by maxj∈G sj . Let λj  ≤ C1 n−c1 for some n uniformly in j in (10). Assume that n constants c1 , C1 > 0, and there exists a sequence of positive numbers αn → ∞ such that αpn = o(1) √ and αn (log p)2 maxj=1,...,p λj sj = o(1). Then under the null H0,G , for any G ⊆ {1, 2, ..., p}, we have sup |P (TG > cG (α)) − α| = o(1). (14) α∈(0,1)

With Corollary 1, the power analysis of TG then follows from Theorem 2.4 in Zhang and Cheng (2017). The above testing procedure can be easily adapted for constructing simultaneous confidence intervals and support recovery, as we will see in Sections 5.2 and 5.3.

3.2

Theoretical implication of Theorem 1

ˆ (Xij |Zi ) as ˆ ij = Xij − E The technique where we replace Xij s by the estimated partial residuals X in (5)-(6) is called “partialling out”. Note that this technique involves p nonparametric regressions where p ≥ n. Moreover, the estimation error from each nonparametric regression accumulates ˆ of n−1 Pn X ˆT ˆ in the approximate inverse Θ i=1 i Xi . Consequently, we first discuss what makes the “partialling out” strategy work in the statistical inference of β0 despite that p is high dimensional. Recall from our previous discussion that ˆbj − β0j and ˜bj − β0j can be decomposed into a ˜ T ε and several remainder terms as shown in (21)-(22). The rates of conleading term n1 Θj X vergence for the remainder terms thati are related to the nonparametric projection step depend on h 1 Pn ˜ ˆ maxj, j 0 n i=1 Xij fj 0 (Zi ) − fj 0 (Zi ) with fˆj defined in (7). In particular, we show that n 1 X h i ˜ ij fˆ 0 (Zi ) − f 0 (Zi ) ≤ ct2 max X n j j 0 n j, j

for any tn ≥ rn ,

(15)

i=1



0



00

0

00

with probability at least 1 − exp −c nt2n + c log p , for some constants c, c , c > 0. For many popular function classes, the critical radius rn defined earlier gives the optimal scaling for bounds



on fˆj 0 − fj 0 . In particular, for (7), one can show that n



00

ˆ

max

fj 0 − fj 0 ≤ c tn 0

for any tn ≥ rn ,

n

j

0



0

(16)



0

with probability at least 1 − c0 exp −c1 nt2n + c2 log p .   ˜ ij |Zi = 0 (for all j) introduced by our partialling out Note that the orthogonality condition E X strategy “reduces” the effects of the estimation errors from fˆj : The statistical error contributed by the projection step is rn2 instead of the optimal rate rn that one would expect from the nonparametric regression. Given this observation, for some function h(sj , s0 ) of sj and s0 only (where the exact form of h is detailed in Assumption 5), as long as √ 2 nrn h(sj , s0 ) = o(1),   √  √  the remainder terms related to (7) in the asymptotic expansions of n ˆbj − β0j and n ˜bj − β0j ˜ T ε, which has an asymptotic normal distribution. Note are dominated by the leading term √1 Θj X n

ˆ (Yi |Zi ) (which is used in (5)) that the above finding also holds true for the surrogate Yˆi = Yi − E and the surrogate gˆ(Zi ) (which is used in (6)). 10

3.3

Practical implication of Theorem 1

We illustrate the theoretical insight above with four specific examples in terms of dim (Zi ) and the function class F that fj s belong to. The initial estimators βˆ based on Zhu (2017) work for all four examples while βˆ based on Müller and van de Geer (2015) or Yu, et al. (2017) only works for the ˆ (Xij |Zi )s, several modern techniques for the first example. With regard to the specific forms of E projection step are considered. The rates achieved by these practical procedures are then compared with the theoretical results in Section 3.2. To facilitate the presentation, our following discussions ˆ (Xij |Zi )s; E (Yi |Zi ) and E ˆ (Yi |Zi ) can be argued in the same fashion. only concern E (Xij |Zi )s and E Example 1: Zi ∈ R and F ∈ S m (the mth order Sobolev ball). Estimating E (Xij |Zi )s via (7) or 2m

(8) can be reduced to the smoothing spline procedure, which achieves the sharp rate, n− 2m+1 , on 2m √ rn2 . In this case, we require nn− 2m+1 h(sj , s0 ) = o(1). Example 2: Zi ∈ R and F is the class of linear combinations of bounded basis functions ψl (·)s P 1 1 such that for f ∈ F, f (Zi ) = dl=1 θl ψl (Zi ) and θ := (θl )dl=1 belongs to the l0 −“ball” of “radius” k. Suppose d1 ≥ n and d1 ≥ 4k. Then the standard Lasso procedure would yield upper bounds with d1 scaling k log on the quantities in (15) using the fact that n n 1 X h i ˜ ij fˆ 0 (Zi ) − f 0 (Zi ) X j j n i=1

n 1 X



˜ ij ψl (Zi ) ≤ max X

θˆ − θ 1 l=1,...,d1 n i=1 s   s 

= Op  

= Op

log d1  Op k n

k log d1 n

log d1  n



k log

d1

d1 k where θˆ is the Lasso estimate. The scaling k log almost achieves the sharp rate, , on rn2 . In n n k log d 1 this case, we require √n h(sj , s0 ) = o(1). If we use the recently proposed Slope (Su and Candès, 2016) instead of the standard Lasso,

then the scaling

k log n

d1 k

d

can be attained. In this case, we require

k log k1 √ n

h(sj , s0 ) = o(1).

Example 3: Zi ∈ R and F ∈ RS m (the rearranged mth order Sobolev ball). Belloni, et al. (2014) show that, when used for estimating functions in an “enlarged” Sobolev space, i.e., RS m , the (square-root) Lasso achieves near oracle rates uniformly over the space. Hence, the square-root Lasso would be our estimation method in this example. The rearranged mth order Sobolev ball is P P∞ defined as the class of functions which take the form f (Zi ) = ∞ l=1 θl ψl (Zi ) such that l=1 |θl | < ∞ and (

θ := (θ1, θ2 , ...) ∈ ΘR (m, d1 , L) =

∃ permutation Υ : {1, ..., d1 } → {1, ..., d1 } P Pd1 2m ˜2 θ˜ ∈ `2 (N) : 2m θ ˜2 ≤ L θΥ(l) + ∞ l=d1 +1 l l l=1 l

Note that S m ⊆ RS m . For technical purposes, define an oracle model f 0 = 1 of a target parameter vector θ0 := (θ0l )dl=1 that solves min

˜ d1 θ∈R

1 n

n X i=1

 f (Zi ) −

d1 X l=1

11

2

θ˜l ψl (Zi ) +





σ 2 θ˜

0

n

,

Pd1

l=1 θ0l ψl (Zi )

)

.

in terms





where θ˜ denotes the number of non-zero components in θ˜ and σ 2 ( 1) is the noise variance. The 0 square-root Lasso θˆ then provides an estimator for θ0 and the resulting nonparametric estimator P for the true function f (Zi ) is given by fˆ = d1 θˆl ψl (Zi ); see Belloni, et al. (2014). l=1

We employ the existing theoretical results to show n 1 X h i ˜ ij fˆ 0 (Zi ) − f 0 (Zi ) X j j n



i=1



n n 1 X h h i i 1 X ˜ ij fˆ 0 (Zi ) − f 00 (Zi ) + ˜ ij f 00 (Zi ) − f 0 (Zi ) X X j j j j n n i=1 i=1 n n 1 X

1 X h i

0 ˜ ij ψl (Zi ) θˆ − θ0 + ˜ ij f 0 (Zi ) − f 0 (Zi ) max X X j j n 1 l=1,...,d1 n i=1 i=1      1 2m  2m log d 1

= Op n 2m+1 |

{z

|

}

T1



+ Op n− 2m+1 = Op n− 2m+1 log d1

n





1

where T1 follows from the bounds θˆ − θ0 = Op n 2m+1 1

Op

q

log d1 n







, T2 follows from the bound fj 0 − fj00

{z

}

T2

q

log d1 n





P n

and maxl n1 m

= Op n− 2m+1





˜

i=1 Xij ψl (Zi )

=

and (15). The bounds

n



ˆ

0 on θ − θ0 and fj 0 − fj 0 are established in Belloni, et al. (2014). In this case, we require 1 n  2m 2m √ 

n n− 2m+1 log d1 h(sj , s0 ) = o(1). Note that the scaling n− 2m+1 log d1 differs from the sharp 2m

rate, n− 2m+1 , on rn2 by a log d1 factor. Example 4: Zi ∈ Rd and F is the class of |S| := k sparse additive nonparametric functions P in the sense that any member f in F has the following decomposition form: f (Zi ) = dl=1 fl (Zil ) = P l∈S fl (Zil ); moreover, fl belongs to an RKHS of univariate functions. Suppose d ≥ n and d ≥ 4k. In practice, we may apply the method in Raskutti, et al. (2012) to estimate E (Xij |Zi )s. In particular, the estimators fˆj s defined in (7) can be written in the form fˆj (z1 , ..., zd ) =

d n X X

α ˆ ilj Kl (zl , zil )

i=1 l=1

where Kl denotes the kernel function for co-ordinate l such that the collection of empirical kernel  αlj ∈ Rn , l = 1, ..., d} are matrices K l ∈ Rn×n has entries Kiil 0 = Kl zil , zi0 l ; the optimal weights {ˆ any solution to the following convex program proposed by Raskutti, et al. (2012): min αlj ∈ Rn T K lα ≤ 1 αlj lj

  1  2n



2 d d r d q

 X X X 1

2 T K lα αlj kK l αlj k2 + µ2j ¯j − K l αlj + µ1j

wj − w lj

 n l=1

2

l=1

l=1

where w ¯j = n1 ni=1 wij and wj = (wij )ni=1 (recalling from Section 2 that wi0 = yi and wij = xij for each j = 1, ..., p).  2m √  If the underlying RKHS is S m , we would require nk n− 2m+1 ∨ logn d h(sj , s0 ) = o(1) in this P



case. Note that the scaling k n on

2m − 2m+1



log d n





almost achieves the sharp rate, k n

rn2 .

12

2m − 2m+1



log kd n



,

3.4

The case of approximate sparsity

With additional efforts, the exact sparsity assumptions of β0 and πj can be relaxed to accommodate for approximate sparsity, provided that the ordered coefficients decay sufficiently fast. To work with approximately sparse β0 and πj , we introduce two thresholded subsets: := {j ∈ {1, 2, ..., p} : |β0j | > τ } ,

Sτ Sτ j





:=

(17) o

n

l ∈ {1, 2, ..., p} \ j : |πjl | > τ j .

(18)



Let Sτ := s0 and Sτ j := sj . Note that the newly defined s0 and sj generalize the previous exact sparsity parameters; in Theorem 1, we simply take τ = 0 and τ j = 0. We also introduce four terms below. These terms are used in the proof for Theorem 2 in Section 4 and the first three also appear in Assumptions 4A and 5A (to be stated). The roles these assumptions play in the case of approximately sparse β0 and πj are similar to those Assumptions 4 and 5 play in the case of exact sparsity. Let s

D1 = s0 s

D2 =

s

s0 log p  + n

s

B1j

= sj s

B2j

=

 s

log p  + s0 n

s

sj log p  + n

2

1

2

log p

β0,Sτc  , n 1

 s

log p  + sj n

1



log p



β0,Sτc  + β0,Sτc , n 1 1



1 2





log p

πj,S c  + πj,S c ,



τ τ n j 1 j 1



1 2

log p

πj,S c  .

τ n j 1

p

Assumption 4A. With τ =

p c log n 2 Λ min (Σ)

in (17) for some universal constant c > 0, (i) the initial esti-



 



mator βˆ satisfies that βˆ − β0 = Op (D2 ) and βˆ − β0 = Op (D1 ); (ii) E (X|Z) βˆ − β0 = n 2 1   2 2 2

Op (D2 ) and the estimator gˆ satisfies that kˆ g − gkn = Op

kβ0 k1 rn ∨ D2 .

Remark. Under mild conditions, the rates in Assumption 4A are satisfied by the initial estimators based on Zhu (2017). As Assumption 4(ii), Assumption 4A(ii) is only used in the analysis for the second debiased estimator ˜b in (6).

13

Assumption 5A. kπj k1 ∨ 1 rn2 = O 



s

n max kπj k1 D22 , B1j





 

  

n max

n max

     

kπj k21 ∨ kπj k1

log p n



and

log p log p = o(1), , B1j D1 , kπj k1 (kβ0 k1 ∨ 1) rn2 ∨ n n 

 log p 



 



, kπj k31 ∨ kπj k1

n s

kπj k21 ∨ kπj k1

q



rn2 ∨

log p D1 , kπj k31 ∨ kπj k1 n 

s 

max B1j , kπj k21 ∨ kπj k1



log p  n



log p n



, kπj k31 ∨ kπj k1

rn2 ∨ 

s

 log p 

n 

= o(1),

 

log p = o(1), D1  n

rn2 ∨



 

log p = o(1). n 

In comparison with the case of exact sparsity, approximate sparsity permits all the entries in β0 (and  q

πj ) to be non-zero at the expense of incurring an extra approximation error s0  q



log p

β0,Sτc (respectively, sj n 1

log p n

 12

+

β0,Sτc 1

1



2

ˆ

πj,S c c ) in the upper bound for β + π − β

(respec0 j,S

τj τj 1 1

1

tively, kˆ πj − πj k1 ). Relative to Assumption 5, Assumption 5A imposes additional conditions on the







“small coefficients” via β0,Sτc and

πj,Sτc j so that the remainder terms in the asymptotic ex1  1    √ √ ˜ T ε (which pansions of n ˆbj − β0j and n ˜bj − β0j are dominated by the leading term √1n Θj X

  q

log p

c gives Theorem 2 below). Note that for the special case where πj,Sτ = O (sj ∨ 1) n j 1   q q q



(s0 ∨1) log p and β0,Sτc = O (s0 ∨ 1) logn p , we have D1 = (s0 ∨ 1) logn p , D2 = , B1j = n 1 q q (s ∨1) log p

j (sj ∨ 1) logn p , and B2j = in Assumptions 4A and 5A; these terms take the same forms n as those used in Assumptions 4 and 5.

Theorem 2 . Under Assumptions 1-3, 4A, 5A and 6, if we choose λj  p log p 0

in (10) and τ j = 1 still hold.

4

c n Λ2min (Σ)

q

log p n

uniformly in j

0

in (18) for some universal constant c > 0, then the claims in Theorem

Proofs

As a general rule for this appendix, all the c ∈ (0, ∞) constants denote positive universal constants. The specific values of these constants may change from place to place. For notational simplicity, we assume the regime of interest is p ≥ (n ∨ 2); the modification to allow p < (n ∨ 2) is trivial.

Proof of Theorem 1 and Corollary 1 Recall the two versions of the debiased estimators:   ˆbj := βˆj + 1 Θ ˆ jX ˆ T Yˆ − X ˆ βˆ , n   ˜bj := βˆj + 1 Θ ˆ jX ˆ T Y − X βˆ − gˆ(Z) ; n 14

(19) (20)

n

ˆ (Y |Z) := Yi − E ˆ (Yi |Zi ) gˆ(Z) := {ˆ g (Zi )}ni=1 , Yˆ = Y − E

on i=1

is an estimate for Y˜ = Y − E (Y |Z) := on

n

ˆ (Xj |Z) := Xij − E ˆ (Xij |Zi ) ˆ j = Xj −E {Yi − E (Yi |Zi )}ni=1 , and for j = 1, ..., p, X is an estimate i=1 ˜ j = Xj − E (Xj |Z) = {Xij − E (Xij |Zi )}n . We write Yˆ = Y˜ + Yˆ − Y˜ = Xβ ˜ 0 + ε + Yˆ − Y˜ for X i=1 ˆ =X ˜ +X ˆ − X, ˜ which are used in the following derivations. We show in the following that and X ˆbj and ˜bj have the same asymptotic distribution. For (19), we obtain ˆbj − β0j  1 ˆ ˆT ˆ ˆ βˆ = βˆj − β0j + Θ Y −X jX n h   i T 1ˆ ˜ 1 ˆ ˜T ˆ jX ˆ 0− X ˆ −X ˜ β0 + Yˆ − Y˜ − X ˆ βˆ ˆ ε + βˆj − β0j + 1 Θ ˆ T Xβ = Θj X ε − Θ j X −X n n n      T  1 1 1 1 ˆ ˆT ˆ ˆ T T ˆ ˆ ˜ ˜ ˆ ˜ = Θj − Θj X ε − Θj X − X ε + ej − Θj X X β − β0 Θj X ε + n n |n {z } |n {z } E0

|

E1

{z

}

E2

   1 ˆ ˆT  ˆ ˆ jX ˜ β0 + 1 Θ ˆ T Yˆ − Y˜ . X −X − Θ jX |n {z } |n {z } E3

(21)

E4

For (20), we obtain ˜bj − β0j  1 ˆ ˆT  = βˆj − β0j + Θ Y − X βˆ − gˆ(Z) jX n T 1 ˆ ˜T 1ˆ ˜ ˆ ε + βˆj − β0j = Θj X ε − Θ j X −X n n     i 1 ˆ ˆT h ˆ ˆ −X ˜ β0 + (X − X)β ˜ 0−X ˆ βˆ + X ˆ −X ˜ βˆ − (X − X) ˜ βˆ + g0 − gˆ + Θj X Xβ0 − X n    T 1 ˆ j − Θj X ˆj X ˜Tε + 1 Θ ˜Tε − 1 Θ ˜ −X ˆ ε Θj X = n |n {z } |n {z } E0

E1

    1 ˆ ˆT ˆ ˆ 1 ˆ ˆT  ˆ ˆ ˜ X X β − β Θ X X − X β − β + ej − Θ + j 0 j 0 n n {z } {z } | | 

0

E2

E3

  1  T   1 ˆ ˜T ˆ ˆ ˆ jX ˆ ˜ E (X|Z) βˆ − β0 − 1 Θ ˆ T (ˆ − Θ g − g0 ). j X E (X|Z) β − β0 − Θj X − X n n n | {z } | {z } | {z } 0

0

E4

E5

15

0

E6

(22)

By elementary inequalities, we have





1 ˜T

ˆ E0 ≤ Θ j − Θj X ε , 1 n ∞

 T

ˆ 1 ˜ ˆ ε , E1 ≤ Θj X −X

1 n



1 ˆ ˆT ˆ

ˆ E2 ≤

ej − n Θj X X β − β0 1 , ∞

 

1

ˆ T ˆ ˆ ˜

E3 ≤ Θj X X − X

kβ0 k1 , 1 n ∞

1  

ˆ ˆT ˆ E4 ≤ Θj X Y − Y˜

, 1 n ∞

and 0

E3 ≤ 0

E4 ≤ 0

E5 ≤ 0

E6 ≤



1 T 

ˆ − β0 ˆ X ˆ −X ˜

X β

,

1 1 n ∞

1



ˆ ˜T

Θj X E (X|Z)

βˆ − β0 ,

1 n 1 ∞



 

ˆ

ˆ ˜ j E (X|Z) βˆ − β0

,

Θj max Xj − X n n 1 j=1,...,p



1

ˆ ˆT g − g0 )

Θj X (ˆ

. 1 n



ˆ

Θj





ˆ

ˆ jX ˆTX ˆ We bound ej − n1 Θ

with (41) and Θ j − Θj with (38), which also implies that 1 ∞   s    

log p log p 

ˆ 2 + max s3j rnj ∨ .

Θj = Op max sj + Op max s2j 1

n

j

j

n

j

(23)

By Assumption 4, we have  s

ˆ

β − β0

1

= Op s0 s

 

E (X|Z) βˆ − β0

n

= Op 



log p  , n 

s0 log p  . n

(24)

By Assumption 1, standard tail bounds for sub-Exponential variables [e.g., Vershynin (2012)] yield

1 T ˜ ε

X

n

s

= Op 





1 T

˜ E (X|Z)

X

n

s

= Op 





log p  , n 

log p  , n

(25)

˜ ij is sub-Gaussian [implied by Assumption where we have used the fact that E (Xij |Zi ) = Xij − X 1 and that sub-Gaussianity is preserved by linear operations]. Note that (24)-(25) only matter to the second debiased estimator ˜bj in (20). 16

Note that we have

ˆ ˜j =

Xj − X n  T 1 ˜j − X ˆ j ε = X n   1 T ˆ X ˆ 0 −X ˜0 ≤ X n j j j

 1 T  ˆ Yˆ − Y˜ ≤ X n j 1 T ˆ g − g0 ) ≤ X n j (ˆ



ˆ (Xj |Z)

E (Xj |Z) − E

, n h iT 1 ˆ n E (Xj |Z) − E (Xj |Z) ε ,

   

ˆ X 0 |Z ˆ (Xj |Z)

E (Xj |Z) − E

E Xj 0 |Z − E j n n   i 1 T h  ˆ X 0 |Z − E X 0 |Z , ˜ E + X j j n j

i 1 T h

ˆ ˆ ˆ ˜

E (Xj |Z) − E (Xj |Z) E (Y |Z) − E (Y |Z) + Xj E (Y |Z) − E (Y |Z) , n n n

1

T ˆ ˜ g − g0 kn + Xj (ˆ g − g0 ) .

E (Xj |Z) − E (Xj |Z) kˆ n n

ˆ (Xj |Z) and fj := E (Xj |Z) for j = 1, ..., p, as well as fˆ0 := E ˆ (Y |Z) and f0 := We write fˆj := E E (Y |Z). Under Assumptions 1 and 3, by standard argument for nonparametric least squares estimators, for any tn ≥ rn ,

ˆ

(26)

fj − fj ≤ c1 tn n



0

with probability at least 1 − c exp −c

nt2n



. Moreover, under Assumption 4, 

√ kˆ g − g0 kn = Op  s0 rn ∨

s



s0 log p  . n

(27)

˜ For fixed elements f˜j ∈ F and g˜ ∈ F, respectively, we  can view fj − fj and  g˜ − g0 as functions of ˜ ij |Zi = 0 (by construction of Z only. Additionally, note that E (εi |Zi ) = 0 and E Y˜i |Zi = E X ˜ ij , i = 1, ..., n, j = 1, ..., p). The remaining argument uses results from empirical processes Y˜i and X theory and local function complexity. In particular, under the independent sampling assumption, Lemma A5 with (26)-(27) implies that h iT 1 ˆ max E (Xj |Z) − E (Xj |Z) ε j=1,...,p n   i 1 T h  ˆ ˜ max Xj E Xj 0 |Z − E Xj 0 |Z j,j 0 ∈{1,...,p} n i 1 T h ˆ (Y |Z) − E (Y |Z) ˜ E max X j j=1,...,p n 1 T ˜ j (ˆ max X g − g0 ) j=1,...,p n

provided that tn ≥ rn and

nt2n

 

= Op t2n ,  

= Op t2n ,  

= Op t2n , 



= Op s0 t2n + Op



s0 log p , n 



% log p. In our analysis, it the suffices to choose tn = rn ∨

17

q

log p n



.

Consequently,

 T

1 ˜ −X ˆ ε

X

n ∞



1 T  ˆ X ˆ −X ˜

X

n





1 T 

ˆ Yˆ − Y˜

X

n



1 T

ˆ (ˆ

X g − g0 )

n ∞

log p , n   log p 2 = Op rn ∨ , n   log p = Op rn2 ∨ , n   s0 log p 2 = Op s0 rn ∨ . n 



= Op rn2 ∨

(28) (29) (30)

Note that (29) only matters to the first debiased estimator ˆbj in (19) while (30) only matters to the second debiased estimator ˜bj in (20). √ ˆ n(bj −β0j ) D Putting all the pieces together, under Assumption 5, we apply the CLT and obtain −→ σj √ ˜   n(bj −β0j ) D ˜TX ˜ i ΘT . Now it remains to show that −→ N (0, 1) , where σj2 = Θj E X N (0, 1) and i j σj X T ˆ   ˆ ˆ X T T T ˆ − Θj E X ˜ X ˜ i Θ Θ Θj i j j n " #       ˆTX ˆ ˆ X ˆ T − Θj E X ˆ T + Θ ˆ jE X ˜TX ˜ i ΘT ˜TX ˜i Θ ˜TX ˜i Θ ≤ Θ −E X j i i j j i j n

X T ˆ     2  

ˆ X

ˆ ˆ T T ˜ T T ˜ ˆ ˜ ˜ iT X ˜i ˜ ≤ Θ − Θ E X X Θ −E X E X X

Θ

+ Θ j i j i j i j i

n

1 ∞   s    

= Op s2j rn2 ∨ 

+ Op s3j

log p n

+ s2j

log p log p  2 + s3j rnj ∨ n n

log p log p + s5j rn2 ∨ n n 

s

2

+ (sj ∨ 1)

log p + s2j ∨ 1 n 

(31) 

rn2 ∨



log p  n 

(32)

where (31) follows from (23) and (47)-(48); (32) follows from (40). Thus we have shown Theorem 1. To show Corollary 1, we adopt the following result (Lemma A1) from Lemma 1.1 of Zhang and Cheng (2016). Combining Lemma A1 with the facts established in Theorem 1, the claim in Corollary 1 follows. Lemma A1. Let ζj = (ζ1j  , ..., ζnj ), j = 1, ..., p, be 7centered i.i.d. sub-Gaussian random variables T ˜ ˜ with variance Θj E Xi Xi ΘTj . Assume that (lognpn) ≤ C1 n−c1 for some constants c1 , C1 > 0, and there exists a sequence of positive numbers αn → ∞ such that o(1). Then, for any G ⊆ {1, ..., p},

αn p

√ = o(1) and αn (log p)2 maxj=1,...,p λj sj =

    0 maxj∈G ζij maxj∈G ij 0 √ √ sup P ≤t −P ≤ t . n−c , c > 0, n n t∈R





˜TX ˜ i ΘT . where j = (j1 , ..., jn ) are i.i.d. normal random variables with mean 0 and variance Θj E X j i

18

Lemma A2. Suppose Assumptions 1, 2(i) regarding Λ2min (Σ), 3, and 6 hold. If kπj k1 ∨ 1 rn2 =   

O

q

log p n

, λj %

q

log p n ,

and 

rn2

max sj j

log p ≤ cΛ2min (Σ) ∨ n 

(33)

for some sufficiently small constant c > 0, then, max kˆ πj − πj k2 = Op j





sj λj ,

(34)

max kˆ πj − πj k1 = Op (sj λj ) .

(35)

j

˜j − X ˜ −j πj and Proof. First, write ηj := X ˜j X

ˆ (Xj |Z) − E (Xj |Z) ˆj + E = X h i ˆ (X−j |Z) − E (X−j |Z) πj + ηj , ˜ −j πj + ηj = X ˆ −j + E = X

thus we have ˆj X

i

h

i

h

ˆ (Xj |Z) − E (Xj |Z) + ηj ˆ (X−j |Z) − E (X−j |Z) πj − E ˆ −j πj + E = X ˆ −j πj + uj . = X

where

h

i

h

i

ˆ (X−j |Z) − E (X−j |Z) πj − E ˆ (Xj |Z) − E (Xj |Z) + ηj . uj := E

(36)

ˆT X ˆ By standard argument for the Lasso, applying Lemma A6 [which shows the n1 X −j −j satisfies a lower restricted eigenvalue (LRE) condition with probability at least q 1 −o(1)] and Lemma A7

1 ˆT

log p along with Assumption 6 [which implies that maxj X uj = Op ] yields (34) and (35). −j

n

n



Lemma A3. Suppose Assumptions in Lemma A2 hold. Let λj  every j = 1, ..., p, we have

q

log p n

uniformly in j. Then for

  s    log p log p 2 , + max s2j ∨ 1 rn2 ∨ τˆj − τj2 = Op max (sj ∨ 1)

n

j

where τj2 := E



˜ ij − X ˜ i,−j πj X

2 

; moreover, 



ˆ

Θj − Θj

1

s

= Op max s2j j



ˆ

Θj − Θj

2

  ˆ ˆ T − Θj,j ˜TX ˜i Θ Θj E X i j

3 2



s

= Op max sj j

= Op



+

+ Op max (sj ∨ 1)



max s5j j

s j

s

= Op max s2j j





rn2

log p ∨ n





(39)

2 !  

  log p log p  + max s2j ∨ 1 rn2 ∨ , (40) j n n 



log p log p  + max s3j rn2 ∨ . j n n

19

(38)

5 log p log p  + max sj2 rn2 ∨ , j n n

log p max s3j j n





log p log p  + max s3j rn2 ∨ , j n n





X

T ˆ

ˆ ˆ X

− ej

Θj

n

(37)

n

j





(41)

1 n

Proof. Note that we have τˆj2 :=

2

ˆ ˆ −j π πj k1 and ˆj + λj kˆ

Xj − X 2

n  1 X 2 ˆ ij − X ˆ i,−j π X ˆj − τj2 n i=1 n h n h n i 1 X i2 2 X 1X 2 2 ˆ i,−j (πj − π ˆ i,−j (πj − π ≤ uij X ˆj ) + uij − τj X ˆj ) + n n n i=1 i=1 i=1

n h n 

2 X T   i2 2X

2 ˜ i,−j (πj − π ˆ i,−j − X ˜ i,−j ˆ i,−j − X ˜ i,−j ≤ X ˆj ) + kπj − π ˆj k1 X X

n

n i=1 i=1 ∞

n  n  1 X  1 X 

2 T

2 2 ˆ

ˆj k1 + + u2ij − ηij ηij − τj2 . +

n X−j uj kπj − π n n ∞ i=1

i=1

Under the condition Λ2max (Σ) = O(1) [Assumption 2], applying (34) in Lemma A2 and Lemma A4 yields   n h i2 2X maxj sj log p ˜ . Xi,−j (πj − π ˆj ) = Op n i=1 n 

By choosing tn = rn ∨

q

log p n



kπj −

as in the proof for Theorem 1, (26) and (35) imply that



2 π ˆj k21

n

n  X

= Op

T 



ˆ i,−j − X ˜ i,−j X

i=1

log p max s2j j n



ˆ i,−j − X ˜ i,−j X







Op

rn2

log p ∨ . n 

By (35) and Lemma A7 along with Assumption 6, we have

2 T

ˆ uj

X

n −j



P n

For the term n1

i=1



s

kπj − π ˆj k1 = Op 





log p  Op max sj j n

s



log p  . n



2 , it suffices to show u2ij − ηij

log p ∨ n



i 1 T hˆ log p ηj E (X−j |Z) − E (X−j |Z) πj = Op max sj rn2 ∨ j n n   i 1 T hˆ log p η E (Xj |Z) − E (Xj |Z) = Op rn2 ∨ . n j n



h i 2

ˆ

E (X−j |Z) − E (X−j |Z) πj

n

2

ˆ

E (Xj |Z) − E (Xj |Z)

n



= Op 

max s2j j

= Op rn2 ∨



rn2

log p n



(42)



(43)





In the above, (42) and (43) follow from (26) (where we choose tn = rn ∨

q

log p n

(44) (45) 

). In terms of (44)

and (45), for fixed elements f˜j 0 , fj 0 ∈ F, we can view f˜j 0 − fj 0 as functions of Z only. Meanwhile,   ˜ only, so E X ˜ ij |Zi = 0 (i = 1, ..., n, j = 1, ..., p) implies that ηj is a function of X h

i

h

i

h

i

˜ ij |Zi − f˜ 0 (Zi )E X ˜ i,−j |Zi πj = 0. E ηij f˜j 0 (Zi )|Zi = f˜j 0 (Zi )E X j 20

(46)

h

i

Furthermore, E ηij f˜j 0 (Zi ) = 0. As for (28), we apply Lemma A5 with (26) and obtain the (44) and (45). P   2 − τ 2 , we note that by Assumption 1 and the definition of η , for In terms of n1 ni=1 ηij ij j j = 1, ..., p, an application of standard tail bounds for sub-Exponential variables yields s  n  1 X  log p  2 . ηij − τj2 = Op  n n i=1

Moreover, by (35) and the choice λj  

q

s

λj kˆ πj k1 = Op max sj j

log p n ,

for j = 1, ..., p, we have s



log p  + Op  n





log p   O max sj j n

s



log p  n

Putting the pieces together, we have (37). Next we show (38) and (39). Note that

ˆ

Θj − Θj = 1



M kˆ πj − πj k1 Mj

ˆj + kπj k1

2 − 2 ≤

τˆj τj τˆj2

1 1 − 2 2 τˆj τj

1

 q

where the first term is Op sj

log p n



!

1 τj2

by (35) and the fact that τˆj2 = τj2 + op (1) while

- 1 [by

Assumption 2]. For the second term, we have kπj k1 = O(sj ) and 

 1 1 max (sj ∨ 1) − = O p  j τˆj2 τj2

s



  log p  log p ∨ max s2j ∨ 1 rn2 ∨ j n n 



 

.



Therefore, we have (38). Similarly, we can obtain (39) by exploiting

kˆ πj − πj k2

ˆ

+ kπj k2

Θj − Θj ≤ 2 2

τˆj

1 1 − 2 2 τˆj τj

!

.

We now show (41). Note that



X T ˆ

ˆ ˆ X

− ej

Θj

n





X T ˆ

ˆ

ˆ X

≤ Θj − Θj

1 n





X T ˆ  

ˆ X T ˜ ˜ + kΘj k1 − E Xi Xi

n

.



By (28), we have

X T ˆ ˜TX ˜ X

ˆ X



n n









X

(X

(X T ˆ − X) ˜ ˜ TX ˜ ˜ T (X ˆ − X) ˜

˜ (X

ˆ − X)

ˆ − X)

+

+





n n n ∞ ∞ ∞  

= Op rn2 ∨

log p . n

(47)

Moreover, by standard tail bounds for sub-Exponential variables, we have

X T ˜  

˜ X T ˜ ˜ − E Xi Xi

n



21

s

= Op 



log p  . n

(48)

Consequently, we obtain (41).     ˆ j,j = ˜TX ˜ i = eT , Θj E X ˜TX ˜ i ΘT = Θj,j , Θ Finally we show (40). Using the facts that Θj E X i j i j 1 , τˆj2

and Θj,j = 

1 , τj2

we have 

ˆ T − Θj,j ˆ jE X ˜ iT X ˜i Θ Θ j 











ˆ j − Θj )T + Θj E X ˆ j − Θj )T + 2Θj E X ˆ j − Θj )E X ˜ iT X ˜ i ΘTj − Θj,j ˜ iT X ˜ i (Θ ˜ iT X ˜ i (Θ =(Θ   ˆ j − Θj )T + 2 − 2 . ˆ j − Θj )E X ˜ iT X ˜ i (Θ =(Θ τˆj2 τj2

Note that

2





ˆ ˆ j − Θj )T ≤ Λ2 ˆ j − Θj )E X ˜ iT X ˜ i (Θ (Θ max Θj − Θj . 2

Applying (39) yields the claim.  Lemma A4. Let U ∈ Rn×p1 be a sub-Gaussian matrix with parameter σU and each row be independently sampled. Then kU ∆k 22 n kU ∆k 22 n

≤ ≥

3¯ α 0 log p1 k∆k 22 + α k∆k 21 , 2 n α 0 log p1 k∆k 22 − α k∆k 21 , 2 n

for all ∆ ∈ Rp1 for all ∆ ∈ Rp1

0

with probability at least 1 − c1 exp(−bn), where α, α ¯ , α , and b are positive constants that only depend on σU and the eigenvalues of ΣU = E(UiT Ui ) for i = 1, ..., n. Remark. This lemma is Lemma 13 in Loh and Wainwright (2012). Lemma A5. Suppose {vi }ni=1 are i.i.d. sub-Gaussian variables with parameter at most σv and E (vi |Zi ) = 0. Let F¯ be a star-shaped function class. Then, there are universal positive constants (c, c1 , c2 ) such that for any tn ≥ rn (where rn is the critical radius), n 1 X vi f (Zi ) ≤ ct2n sup n f ∈F¯ , kf kn ≤tn i=1

(49)

2 with probability at least 1 − c1 exp −c2 n σtn2 . If F¯ concerns a ball of a RKHS H equipped with a v norm k·kH , then for any f ∈ F¯ and tn ≥ rn , we have





n 1 X 0 00 vi f (Zi ) ≤ c t2n kf kH + c tn kf kn n

(50)

i=1

0



0

2



with probability at least 1 − c1 exp −c2 n σtn2 . v

Proof. The proof for bound (49) follows the proof for Lemma 13.2 in Wainwright (2015). To show (49), we let (

A(u) :=

n 1 X ¯ ∃f ∈ F ∩ {kf kn ≥ u} : vi f (Zi ) ≥ 2 kf kn u n i=1



22



)

where u ≥ rn . Suppose that there exists some f ∈ F¯ with kf kn ≥ u such that n 1 X vi f (Zi ) ≥ 2 kf kn u. n i=1





(51)



Let the function f˜ := kfuk f and note that f˜ = u. Since f ∈ F¯ and u ≤ kf kn by construction, n n f˜ ∈ F¯ under the star-shaped assumption. Consequently, whenever the event A(u) is true so that there exists a function f satisfying inequality (51), then there exists a function f˜ ∈ F¯ with kf˜ kn = u such that n n u 1 X 1 X ˜ vi f (Zi ) = vi f (Zi ) ≥ 2u2 . n i=1 kf kn n i=1

To summarize, we have established i

h

P [A(u) | {Zi }ni=1 ] ≤ P An (u) ≥ 2u2 | {Zi }ni=1 , where n 1 X An (u) := sup vi f˜(Zi ) f˜∈Ω(u; F ) n i=1



n



o

with Ω(u; F) = f ∈ F¯ : kf kn ≤ u . Conditioning on {Zi }ni=1 , note that under our assumptions, P for each fixed f˜, the variable n1 ni=1 vi f˜(Zi ) is zero-mean sub-Gaussian. Then by standard tail bounds, we have P [An (u) ≥ Ev [An (u) |

{Zi }ni=1 ]

+ b|

{Zi }ni=1 ]

{Zi }ni=1 ]

2

{Zi }ni=1

nb2 ≤ c0 exp −c 2 2 u σv

!

.

Consequently, for any b = u2 , h

P An (u) ≥ Ev [An (u) |

+u |

i

nu2 ≤ c0 exp −c 2 σv

!

.

(52)

Finally, note that we have Ev [An (u) | {Zi }ni=1 ] = Gn (u; F). Since u ≥ rn and the function t 7→ Gn (t; F ) is non-increasing (by Lemma A8), we obtain t Gn (u; F) Gn (rn ; F) rn ≤ ≤ , u rn 2 where the last inequality uses the definition of rn . Putting everything together, we have established that Ev [An (u) | {Zi }ni=1 ] ≤ ur2n . Combined with the bound (52), we have h

P An (u) ≥ 2u2 | {Zi }ni=1

i

h

≤ P An (u) ≥ urn + u2 | {Zi }ni=1 u2 ≤ c1 exp −c2 n 2 σv

i

!

where the inequality uses the fact that u2 ≥ urn . This proves (49). The second bound from Lemma 1 in Raskutti et al. (2012). P (50) follows  0 If 2kfkH - 1 and 0 1 n 2 kf kn - tn , then n i=1 vi f (Zi ) ≤ c3 tn with probability at least 1 − c1 exp −c2 n σtn2 .  v

23

Lemma A6 (LRE condition). Suppose Assumptions 1, 3, and (33) hold. Then, for any

n





o

∆j ∈ C(J(πj )) := ∆ ∈ Rp−1 : ∆J(πj )c ≤ 3 ∆J(πj ) 1

(53)

1

[J(πj ) is the support of πj ] and every j = 1, ..., p, we have ∆Tj 

ˆT X ˆ X −j −j ∆j ≥ κ1 k∆j k22 n 0

(54)



00

0

with probability at least 1 − c exp −c nt2n + c log p for any tn ≥ rn , where κ1 = c0 Λ2min (Σ) for 0

some universal constant c0 > 0. Proof. By elementary inequalities, we have ˆT ˆ T X−j X−j ∆j ≥ ∆j n

≥ ≥ ≥

! ˜T ˜ ˜T ˜ ˆT ˆ T X−j X−j T X−j X−j − X−j X−j ∆j − ∆j ∆j ∆j n n

X ˜T ˜ ˜T ˜ ˆT ˆ T X−j X−j −j X−j − X−j X−j 2 ∆j − ∆j

k∆j k1

n n ∞



! T T



(X ˜ ˆ ˜ ˜ ˜ ˜ −j )T X ˆ −j

X−j (X−j − X−j ) T X−j X−j

ˆ −j − X

∆j − k∆j k21 ∆j

+



n n n ∞ ∞

T T



˜ ˜ ˜ ˆ ˜ T X−j X−j X−j (X−j − X−j ) 2 ∆j −

k∆j k1 ∆j

n n ∞



(X

T ˆ ˜ ˜ ˆ ˜ −j )T (X ˆ −j − X ˜ −j )

−j − X−j ) X−j

(X−j − X

2 2 −

k∆j k1 −

k∆j k1 .



n n ∞

T

X˜ −j (Xˆ −j −X˜ −j )

. We first bound

n ˜ −j ) ˜ T (X ˆ −j −X X −j , n



0

Note that in terms of the (l, l )th element of the matrix



for any tn ≥ rn , we have n h i 1 T 1 X ˜ ˆ ˜ ˜ il fˆ0 (Zi ) − f 0 (Zi ) . X X n l (Xl0 − X l0 ) = n l l i=1

Lemma A5 and (26) imply that, for any tn ≥ rn , 1 T ˜ ˆ ˜ max Xl (Xl0 − X l0 ) ≤ c0 t2n 0 n

(55)

l, l

with probability at least 1−c1 exp we have

−c2 nt2n



(Xˆ −j −X˜ −j )T (Xˆ −j −X˜ −j )

,

+ c3 log p . To bound the term

n 





(X

2 ˜ −j )T (X ˆ −j − X ˜ −j ) 0 2

ˆ

ˆ −j − X

0 (Z) − f 0 (Z) ≤ c t f

≤ max

0 n l l 0

n n l ∞  0  0 0

(56)

with probability at least 1 − c1 exp −c2 nt2n + c3 log p .

T X ˜ −j T X˜ −j from below, we apply the second result in Lemma A4; since this To bound ∆j ∆ j n

result holds for all ∆j ∈ Rp−1 , we can specialize it to any ∆j in the restricted sets specified in (53). 24

Putting everything above together and choosing tn = rn ∨ ∆Tj

q

log p n

yields

  ˆT X ˆ X log p −j −j ∆j ≥ c1 Λ2min (Σ) k∆j k22 − c2 ∨ rn2 k∆j k21 n n 

0

(57)



00

with probability at least 1 − c exp −c nt2n + c log p . As a result, the cone condition kˆ πj − πj k1 ≤ √ πj − πj k2 implied by (53) along with condition (33) yields (54). 4 sj kˆ



ˆT

2 Lemma A7 (Upper bound on maxj n1 X −j uj ). Suppose kπj k1 ∨ 1 rn = O 

q

log p n



, Assump-



h i q 

ˆT

1 ˜T ˜j − X ˜ −j πj + log p max E X X tions 1 and 3 hold. Then, we have maxj=1,...,p n1 X u

j l,j −j n l n ∞  0   0  0 0 0

with probability at least 1 − c1 exp −c2 log p − c3 exp −c4 nt2n + c5 log p . Proof. Recall the definition of uj in (36) and ˆ (X−j |Z) + E (X−j |Z) ˆ −j = X ˜ −j − E X



˜ −j = X−j − E (X−j |Z). For any l 6= j, after expanding 1 X ˆT where X n l uj , we note that in order

ˆ

ˆT

1 ˜T ˆ

2 ˜ to bound maxj n1 X −j uj , we need to bound maxl0 fl0 (Z) − fl0 (Z) , maxl, l0 n Xl (Xl0 − X l0 )





0





n









ˆl − X ˜ l ) , and maxl=1,...,p, l6=j 1 X ˜T (l l = 1, ..., p, l 6= j), maxl=1,...,p, l6=j n1 ηjT (X n l ηj . In particular, the first two terms are bounded in (56) and (55); the fourth term can be bounded by a standard sub-Exponential concentration inequality; for the third term, we evoke (46) and the argument that is used to bound (55), which yields 1 T 00 ˆ ˜ max ηj (Xl − X l ) ≤ c0 t2n l6=j n 00



00

00



with probability at least 1 − c1 exp −c2 nt2n + c3 log p . Putting the pieces together and choosing tn = rn ∨

q

log p n ,

the claim in Lemma A7 follows. 

Lemma A8. Let the shifted function class F¯ be star-shaped. Then the function t 7→ Gn (t;t F ) is non-increasing on the interval (0, ∞); as a result, the critical inequality has a smallest positive solution (the critical radius). Remark. This is Lemma 13.1 from Wainwright (2015). Lemma A9: Let Nn (t; Ω(˜ rn ; F)) denote the t−covering number of the set n

Ω(˜ rn ; F) = f ∈ F¯ : |f |n ≤ r˜n

o

in the L2 (Pn ) norm. Then the smallest positive solution (the critical radius) to the critical inequality is bounded above by any r˜n ∈ (0, σ † ] such that 0

c √

n

Z r˜n r ˜2 √n 4 2σ †

q

log Nn (t; Ω(˜ rn ; F))dt ≤

r˜n2 . 4σ †

Remark. This result is established by van der Vaart and Wellner (1996), van de Geer (2000), Barlett and Mendelson (2002), Koltchinski (2006), Wainwright (2015), etc. 25

Lemma A10 (Approximately sparse ). Suppose Assumptions 1, 2(i) regarding Λ2min (Σ), 3, q πj q  2 log p log p , λ % and 6 hold. If kπj k1 ∨ 1 rn = O j n n , and 

kπj k1

s

log p ∨ rn2 ≤ c n 

log p 2 Λ (Σ) n min

(58)

!



λj

πj,Sτc j ,

(59)

for some sufficiently small constant c > 0, then, s



max kˆ πj − πj k2 = Op λj sj + j

s

max kˆ πj − πj k1 = Op λj sj + j

1



!





λj sj πj,Sτc + πj,Sτc

. j j 1

(60)

1





ˆj = π ˆ

Proof. Let ∆ ˆj − πj . The basic inequality and the choice of λj yield that

∆j,Sτc j ≤





ˆ

3 ∆j,Sτ + 4 πj,Sτc

. Consequently, j j 1 1







ˆ ˆ j,S + 4 πj,S c ∆

∆j ≤ 4

τ τ 1

j

1

1





ˆ j

≤ 4√sj

πj,S c . ∆ + 4



τj 2 j 1

(61)

1

Moreover, we have X l=1,...,p, l6=j

X

|πjl | ≥

l∈Sτ

|πjl | ≥ τ j sj j

and therefore sj ≤ τ −1 j kπj k1 . Putting the pieces together yields

q





ˆ

ˆ −1

∆j ≤ 4 τ j kπj k1 ∆j + 4 πj,Sτc

. 1 2 j

(62)

1

ˆ j satisfying (62), applying (57) yields Therefore, for any vector ∆

2    

2  ˆT X ˆ

X log p log p

ˆ −j −j ˆ −1 2 2 2

∆j ≥ ∆j c1 Λmin (Σ) − c2 τ j kπj k1 ∨ rn − c3 πj,Sτc ∨ rn . 2 n n n j 1 (63) With the choice of

s  − 1

log p 2 2

Λmin (Σ) ∨ rn2  δ ∗ ,

πj,Sτc j n 1

ˆT ∆ j

under condition (58), we have

2 ˆT ˆ ˆ T X−j X−j ˆ

ˆ ∆j ≥ c4 Λ2min (Σ) ∆ ∆j j 2 n



ˆ j such that ˆ j for any ∆



≥ δ ∗ . We can then apply Theorem 1 in Negahban, et. al (2010) to 2 obtain (59). By (61), we also obtain (60). 

26

Proof of Theorem 2 For Theorem 2, we apply (59) and (60) when proving Lemma A3 and obtain

1 1 − 2 2 τˆj τj







2 = Op max B2j + Op max kπj k1 ∨ 1 j

j







+ Op max kπj k21 ∨ 1 j



ˆ

Θj − Θj

1

s

= Op max B1j + Op max kπj k21 ∨ kπj k1



j

j







rn2 ∨

log p n



log p  n



,





s

= Op max B2j + Op max kπj k1 kπj k2 ∨ kπj k2 j







j



= Op

2 max B2j j





+ Op max



j

+ Op max kπj k41 kπj k22 ∨ kπj k22 j

 

s

max kπj k1 ∨ 1



 j 

rn2 ∨

log p n

kπj k21 kπj k22



+ Op



rn2 ∨



s



+ Op max j

j



kπj k31

∨ kπj k1



rn2

log p ∨ n



 log p 

n

2 !

log p + max kπj k21 ∨ 1 j n 

j

,









log p  n



kπj k22

log p n

= Op max B1j + Op max kπj k21 ∨ kπj k1





j

+ Op max kπj k21 kπj k2 ∨ kπj k2



X T ˆ

ˆ ˆ X − ej

Θj

n

, s

j

  ˆ ˆ T − Θj,j ˜TX ˜i Θ Θj E X i j





+ Op max kπj k31 ∨ kπj k1

2

log p n

rn2 ∨



log p  n











ˆ

Θj − Θj





rn2 ∨

 

log p , n 



log p  n



.

Now, we adopt the same argument as in the proof for Theorem 1 with the minor dif following  0 0 1 √ as well as that ferences: in showing the remainder terms E0 − E4 and E3 − E6 are op n   ˆ Xˆ T Xˆ ˆ T ˜TX ˜ i ΘT = op (1), we apply the above rates and Assumptions 4A-5A (which Θ − Θj E X Θj n

j

i

j

replace Assumptions 4-5). Putting these pieces together gives the claims in Theorem 2. 

5

Simulations

In this section, we evaluate the performance of our methods with simulation experiments. To generate the full covariates X, we first generate X0 from the p−dimensional normal distribution with mean 0 and variance ΣX0 = (ΣX0 ,ij )pi,j=1 , which takes three different forms: (S1) Independent: ΣX0 = Ip ; (S2) AR(1): ΣX0 ,ij = 0.5|i−j| ; (S3) Exchangeable/Compound Symmetric: ΣX0 ,ii = 1 and ΣX0 ,ij = 0.5 if i 6= j. 27

The covariates {Zi }ni=1 are i.i.d. from U [0, 2]. To incorporate the dependence between X and Z, we set Xi1 = X0,i1 + 3Zi , Xi2 = X0,i2 + 3Zi2 , Xi3 = X0,i3 − 3Zi and Xij = X0,ij , 1 ≤ i ≤ n, 4 ≤ j ≤ p. The set of nonzero coefficients in β0 is from a fixed realization of s0 = 3 i.i.d. U [0, 3]. The active set is set to be S0 = {1, 2, 3}. We consider two different non-linear functions g0 : (G1) g0 (z) = 1.5 sin(2πz); (G2) g0 (z) = z 10 (1 − z)4 /B(11, 5) + 4z 4 (1 − z)10 /B(5, 11) where B(· , ·) denotes the beta distribution. The error terms are generated from a standard normal ˆ gˆ) ˆ by regressing X on Z with the smoothing spline procedure and (β, distribution. We estimate X are obtained from (9). As a result, the debiased estimator in our simulations concerns (6). We do not test the performance of (5) with simulation experiments but expect it to behave similarly as (6). Similar to Zhang and Cheng (2017), the estimated variance σ ˆε2 is calculated as follows: Pn

σ ˆε2 =

i=1



2

Yi − Xi βˆ − gˆ(Zi )



n − βˆ

.

1

We set the tuning parameter µ = n−2/5 /10 (Müller and van de Geer, 2015) and let λ and λj (1 ≤ j ≤ p) be calculated from the 10-fold cross validation (van de Geer, et al., 2014). Across all the simulations, we set the sample size n = 100 and the number of variables p = 500. Results in sections 5.1 and 5.2 are based on 100 replications, while those in section 5.3 are based on 500 replications.

5.1

Component-Wise Confidence Interval

Average coverage and average length of the intervals for individual coefficients corresponding to variables in either S0 or S0c are considered. Denote CIj as a two-sided confidence interval for βj0 . In Table 1, we report the empirical versions of Avgcov S0 = s−1 ∈ CIj ); j∈S0 P(β 0 P 0j c −1 Avgcov S0 = (p − s0 ) j∈S0c P(0 ∈ CIj ); −1 P Avglength S0 = s0 j ); j∈S0 length(CI P Avglength S0c = (p − s0 )−1 j∈S0c length(CIj ). P

The results in Table 1 agree with our theoretical predictions. The average coverage probabilities of confidence intervals for S0c are close to the nominal 95% level, while those for S0 are slightly lower than 95%. The confidence intervals for S0c are comparably narrower than those for S0 . We also notice that as the columns in X0 become more correlated (so the inverse Θ of the Hessian becomes less sparse), the coverage performance becomes worse. This finding confirms our earlier comment (in Section 2) that the sparsity condition on the off diagonal elements of Θ plays a crucial role in the makes remainder terms like   effectiveness of the debiased approach as this condition  √ ˜ 1 T ˆ ˜ Θj − Θj √n X ε of order op (1) in the asymptotic expansion of n bj − β0j .

5.2

Simultaneous Confidence Intervals

In Table 2, we present the coverage probabilities and interval widths for the simultaneous confidence intervals for β0j , 1 ≤ j ≤ p. For each simulation run, we record whether the simultaneous confidence interval contains β0j for 1 ≤ j ≤ p and the corresponding interval width. Again, it is not surprising 28

Table 1: Average coverage probabilities and lengths of confidence intervals at the 95% nominal level with 100 iterations; n = 100, p = 500 Setup Measure Avgcov S0 Avglength S0 Avgcov S0c Avglength S0c

Active set S0 = {1, 2, 3}; Error ε ∼ N (0, 1) S2, G1 S3, G1 S1, G2 S2, G2 0.857 0.693 0.887 0.823 0.812 0.807 0.798 0.789 0.955 0.963 0.953 0.955 0.510 0.547 0.480 0.500

S1 ,G1 0.896 0.802 0.953 0.476

S3, G2 0.683 0.827 0.963 0.559

Table 2: Coverage probabilities and interval widths for the simultaneous confidence intervals based on the non-studentized (NST) and studentized (ST) test statistics with 100 iterations; n = 100, p = 500 Setup Measure NST coverage NST width ST coverage ST width

S1 ,G1 0.95 1.05 0.88 0.88

Active set S0 = {1, 2, 3}; Error ε ∼ N (0, 1) S2, G1 S3, G1 S1, G2 S2, G2 0.74 0.72 0.96 0.86 1.09 1.12 1.06 1.06 0.94 0.87 0.82 0.91 0.96 1.04 0.87 0.93

S3, G2 0.74 1.12 0.83 1.04

that the coverage probability is affected by the amount of correlations between the columns in X0 . Overall, both studentized and non-studentized method provide satisfactory coverage probability. When ΣX0 is the identity matrix, non-studentized method has better coverage; while when ΣX0 takes the form of S2 or S3, the performance of the studentized method is better.

5.3

Support Recovery

The major goal of this section is to identify signal locations of β0 in a pre-specified set G = {1, 2, . . . , p}, i.e. support recovery. Similarly as the procedure in Zhang and Cheng (2017), we take the signal set ˜ : |˜bj | > λ∗j }, Sˆ0 = {j ∈ G q

ˆ j Xˆ T Xˆ Θ ˆ T . Note that similar arguments as Proposition where λ∗j = 2ˆ ωjj log(p)/n and ω ˆ jj = σ ˆε2 Θ j n 3.1 of Zhang and Cheng (2017) implies this support recovery procedure is consistent. To assess the performance, we consider the following similarity measure |Sˆ0 ∩ S0 | d(Sˆ0 , S0 ) = q . |Sˆ0 | · |S0 | ˆ Table 3 summarizes the mean and standard deviation q of d(S0 , S0 ) as well as the number of false positives (FP) and false negatives (FN) normalized by |Sˆ0 | · |S0 |. When the amount of correlations between the columns in X0 increases (as in S3), the false positive rates are comparably higher.

References Bartlett, P. and Mendelson, S. (2002). Gaussian and Rademacher complexities: Risk bounds and structural results. J. Mach. Learn. Res., 3: 463-482.

29

Table 3: The mean and standard deviation (SD) of d(Sˆ0 , S0 ), and the numbers of false positives (FP) and false negatives (FN) with 500 iterations; n = 100, p = 500 Setup Measure Mean SD FP FN

S1 ,G1 0.96 0.07 0.09 0.00

Active set S0 = {1, 2, 3}; Error ε ∼ N (0, 1) S2, G1 S3, G1 S1, G2 S2, G2 0.97 0.94 0.97 0.97 0.06 0.08 0.07 0.06 0.06 0.13 0.09 0.07 0.00 0.00 0.00 0.00

S3, G2 0.94 0.08 0.12 0.00

Belloni, A., Chernozhukov V. and Wang L. (2014). Pivotal estimation via square-root Lasso in nonparametric regression. Ann. Statist., 42: 757-788. Bickel, P. J., Klaassen, C. A. J., Ritov, Y. and Wellner, J. A. (1998). Efficient and Adaptive Estimation for Semiparametric Models. Springer-Verlag, New York. Cheng G., Zhang H. and Shang Z. (2015). Sparse and efficient estimation for partial spline models with increasing dimension. Annals of the Institute of Statistical Mathematics, 67: 93-127. Javanmard, A. and Montanari, A. (2014). Condence intervals and hypothesis testing for high-dimensional regression. J. Mach. Learn. Res., 15: 2869-2909. Koltchinskii, V. (2006). Local Rademacher complexities and oracle inequalities in risk minimization. Ann. Statist., 34: 2593-2656. Li, R., and Liang, H. (2008). Variable selection in semiparametric regression modeling. Ann. Statist., 36: 261-286. Liang, H., and Li R. (2009). Variable selection for partially linear models with measurement errors. Journal of the American Statistical Association. 104: 234-248. Loh, P., and Wainwright, M. (2012). High-dimensional regression with noisy and missing data: Provable guarantees with non-convexity. Ann. Statist. 40: 1637-1664. Mendelson, S. (2002). Geometric parameters of kernel machines. In Proceedings of COLT. 29-43. Müller, P. and van de Geer, S. (2015). The partial linear model in high dimensions. Scandinavian Journal of Statistics, 42: 580-608. Negahban, S., P. Ravikumar, M. J. Wainwright, and Yu, B. (2012). A unified framework for highdimensional analysis of M-estimators with decomposable regularizers. Statistical Science, 27: 538-557. 2010 version: arXiv:1010.2731v1. Donald, S. G., and Newey, W. (1994). Series estimation of semilinear models. Journal of Multivariate Analysis. 50: 30-40. Raskutti, G., Wainwright, J. M., and Yu, B. (2012). Minimax-optimal rates for sparse additive models over kernel classes via convex programming. J. Mach. Learn. Res., 13: 389-427. Robinson, P. M. (1988). Root-n-consistent semiparametric regression. Econometrica. 56: 932-954. Sherwood, B. and Wang L. (2016). Partially linear additive quantile regression in ultra-high dimension. Ann. Statist., 44: 288-317. Su, W. and Candès E. (2016). Slope is adaptive to unknown sparsity and asymptotically minimax. Ann. Statist., 44: 1038-1068. Tibshirani, R. (1996). Regression shrinkage and selection via the Lasso. Journal of the Royal Statistical Society Series B. 58: 267-288. van de Geer, S. (2000). Empirical Processes in M-Estimation. Cambridge University Press. van de Geer, S., Bühlmann, P., Ritov, Y. and Dezeure, R. (2014). On asymptotically optimal confidence regions and tests for high-dimensional models. Ann. Statist., 42: 1166-1202. van der Vaart, A. W. and J. Wellner (1996). Weak Convergence and Empirical Processes. Springer-

30

Verlag, New York, NY. Vershynin, R. (2012). Introduction to the non-asymptotic analysis of random matrices, in Eldar, Y. and G. Kutyniok, Eds, Compressed Sensing: Theory and Applications. 210-268, Cambridge. Wainwright, J. M. (2015). High-Dimensional Statistics: A Non-Asymptotic Viewpoint. In preparation. University of California, Berkeley. Xie, H. L. and Huang J. (2009). Scad-penalized regression in high-dimensional partially linear models. Ann. Statist., 37: 673-696. Yu, Z., Levine, M. and Cheng, G. (2017). Minimax optimal estimation in high dimensional semiparametric models. arXiv preprint. Zhang, C. H. and Zhang S. (2014). Confidence intervals for low dimensional parameters in high dimensional linear models. J. R. Stat. Soc. Ser. B. Stat. Methodol., 76: 217-242. Zhang, X. and Cheng, G. (2017). Simultaneous inference for high-dimensional linear models. Forthcoming in Journal of the American Statistical Association - Theory & Methods. Zhu, Y. (2017). Nonasymptotic analysis of semiparametric regression models with high-dimensional parametric coefficients. Forthcoming in Ann. Statist.

31

High Dimensional Inference in Partially Linear Models

Aug 8, 2017 - belong to exhibit certain sparsity features, e.g., a sparse additive ...... s2 j ∨ 1. ) √ log p n.. ∨. [( s3 j ∨ 1. ) ( r2 n ∨ log p n. )] = o(1). 8 ...

525KB Sizes 3 Downloads 338 Views

Recommend Documents

Inference in partially identified models with many moment
Apr 25, 2016 - ‡Department of Economics and Business, Aarhus University, ..... later, ˆµL(θ) in Eq. (3.2) is closely linked to the soft-thresholded least squares.

Sparsity and statistical inference in high-dimensional ...
(Donoho, 2005, Candes/Tao, 2005; Donoho & Tanner, 2006). – standard Gaussian ensemble: X ∈ R n×p with Xij ∼ N(0,1) i.i.d.. – signal β∗ is k-sparse, with k ...

PDF Partially Linear Models (Contributions to Statistics)
... statistical problems, including least squares regression, asymptotically efficient estimation, bootstrap resampling, and censores data analysis. Full description.

Variable Clustering in High-Dimensional Linear ... - The R Journal
Reducing the dimensionality is a challenge that most statistical methodologies ... The core of the package is a C++ program interfaced with R using the ... illustrating the good predictive performances of CLERE, with noticeable parsimony ...

Variable Clustering in High-Dimensional Linear ... - The R Journal
The CLusterwise Effect REgression (CLERE) methodology (Yengo et al., 2014), was recently ...... 50. ∑ j=43. 15 × xij,1.. . All algorithms were run using 10 different random ... illustration of the relative predictive performance of our model.

Estimation and Inference for Linear Models with Two ...
Estimation and Inference for Linear Models with Two-Way. Fixed Effects and Sparsely Matched Data. Appendix and Supplementary Material. Valentin Verdier∗. April 21, 2017. ∗Assistant Professor, Department of Economics, University of North Carolina,

Penalised Additive Least Squares Models for High Dimensional ...
sumption in high dimensional regression models is to assume that the function ... have a few samples, using a simpler model to fit our data may give us a better ...

Inference in Incomplete Models
Program for Economic Research at Columbia University and from the Conseil Général des Mines is grate- ... Correspondence addresses: Department of Economics, Harvard Uni- versity ..... Models with top-censoring or positive censor-.

Power in High-Dimensional Testing Problems - ULB
In the sequel we call tests ... As a consequence, in many high-dimensional testing problems the power .... for all possible rates d(n) at which the dimension of the parameter space can increase ..... that the total variation distance between Nn(θn,n

Reporting Neighbors in High-Dimensional Euclidean Space
(c) For each cell τ with |Pτ | ≥ 2, go over all pairs of points of Pτ and report those pairs at ...... Geometry, Second Edition, CRC Press LLC, Boca Raton, FL, 2004.

Power in High-Dimensional Testing Problems - ULB
In regression models power properties of F-tests when the ...... But a(n) → ∞ hence implies (via the triangle inequality, together with dw-continuity of (µ, σ2) ↦→.

Sparse Linear Models and l1−Regularized 2SLS with High ...
High-Dimensional Endogenous Regressors and Instruments. Ying Zhu ... (2) for all j. Our primary interest concerns the regime where p ≥ (n ∨ 2), β∗ and π∗ ..... quantity erra accounts for the remaining error from π∗ j,Sc τj ..... (2013).

Linear and Linear-Nonlinear Models in DYNARE
Apr 11, 2008 - iss = 1 β while in the steady-state, all the adjustment should cancel out so that xss = yss yflex,ss. = 1 (no deviations from potential/flexible output level, yflex,ss). The log-linearization assumption behind the Phillips curve is th

Context model inference for large or partially ...
are also extensible to the partially observable case (Ve- ness et al., 2009; Ross .... paper, we shall only look at methods for approximat- ing the values of nodes ...

bayesian inference in dynamic econometric models pdf
bayesian inference in dynamic econometric models pdf. bayesian inference in dynamic econometric models pdf. Open. Extract. Open with. Sign In. Main menu.

Optimal Inference in Regression Models with Nearly ...
ymptotic power envelopes are obtained for a class of testing procedures that ..... As a consequence, our model provides an illustration of the point that “it is de-.

Inference in Panel Data Models under Attrition Caused ...
ter in a panel data'model under nonignorable sample attrition. Attrition can depend .... (y&,x&,v), viz. the population distribution of the second period values.

Inference in Second-Order Identified Models
Jan 9, 2017 - where fs(X, θ) is the s-th element of f(X, θ). The following assumption defines the identification configuration maintained throughout our analysis. ...... The relative performance of the LM statistic is less clear. Theorem 2 indicate

Inference in models with adaptive learning
Feb 13, 2010 - Application of this method to a typical new Keynesian sticky-price model with perpetual ...... Princeton, NJ: Princeton University Press. Hodges ...

Simultaneous Inference in General Parametric Models
Simultaneous inference is a common problem in many areas of application. If multiple null hypotheses are tested ...... Stefanski, L. A. and Boos, D. D. (2002).

inference in models with multiple equilibria
May 19, 2008 - When the set of observable outcomes is infinite, the problem remains infinite ...... For standard definitions in graph theory, we refer the reader to ...

Inference in Panel Data Models under Attrition Caused ...
j+% ) 6E -'(y%,y&,x%,x&,β) g (я$ (z%,z&,v)) 6S φ 1,x%j,x&j.*& . To estimate. 6. E F ...... problem in that it satisfies the conditions S3'S6 of the consistency and ...