Banded and tapered estimates for autocovariance matrices and the linear process bootstrap Dimitris N. Politis

DePaul University

University of California, San Diego June 20, 2010





Timothy L. McMurry

[email protected]

1

Banded and tapered estimates for autocovariance matrices and the linear process bootstrap

Abstract We address the problem of estimating the autocovariance matrix of a stationary process. Under short range dependence assumptions, convergence rates are established for a gradually tapered version of the sample autocovariance matrix and for its inverse. The proposed estimator is formed by leaving the main diagonals of the sample autocovariance matrix intact while gradually down-weighting off-diagonal entries towards zero. In addition we show the same convergence rates hold for a positive definite version of the estimator, and we introduce a new approach for selecting the banding parameter. The new matrix estimator is shown to perform well theoretically and in simulation studies. As an application we introduce a new resampling scheme for stationary processes termed the linear process bootstrap (LPB). The LPB is shown to be asymptotically valid for the sample mean and related statistics. The effectiveness of the proposed methods are demonstrated in a simulation study.

Key words: autocovariance matrix; stationary process; bootstrap; block bootstrap; sieve bootstrap

1

Introduction

Let X1 , . . . , Xn be a realization of a mean zero, stationary process {Xt }t∈Z , and let γk = cov [X0 , Xk ] be its autocovariance function. The goal of the present work is to estimate the n × n autocovariance matrix  n Σn = γ|i−j| i,j=1 . The lag-k autocovariance γk has a natural estimate given by the sample autocovariance γˆk = n

−1

n−k X

Xi Xi+k

i=1

However, plugging in γˆk instead of γk in Σn does not work because   ˆ n = γˆ|i−j| n Σ i,j=1 2

ˆ n does not converge is not a consistent estimator of Σn in the sense that the operator norm of Σn − Σ to zero. To achieve consistency, Wu and Pourahmadi (2009) proposed a banded estimator of the sample covariance matrix. In the present work, we propose a more general estimator of Σn which leaves the 2l + 1 main ˆ n intact, and then gradually down-weighs increasingly distant off-diagonal entries diagonals of Σ instead of setting them to zero as in the banded matrix case. We establish rates of convergence and demonstrate the efficacy of the proposed method. In addition, by analogy with the related problem of spectral density estimation, we introduce a natural estimate for the banding parameter, l, that is useful even for the Wu and Pourahmadi (2009) estimator. The remainder of the paper is structured as follows: Section 2 presents our main results; Section 3 addresses a correction to positive definiteness; Section 4 presents a method to choose the banding parameter; Section 5 introduces as an application the linear process bootstrap, a new bootstrap for stationary processes; Section 6 provides a small simulation study; and Section 7 contains all technical proofs.

2

A tapered covariance matrix estimator

In the present section, we establish convergence rates for the tapered sample covariance matrix to Σn in the operator norm, defined by ρ(A) =

max

x∈Rn :|x|=1

|Ax|,

(1)

where | · | denotes the usual Euclidean norm on Rn . It is worth noting that for a real valued matrix p A, ρ(A) = λmax (At A), where λmax (M ) is the largest eigenvalue of the matrix M ; see Horn and Johnson (1990), p. 296.   ˆ κ,l := w|i−j| γˆ|i−j| n , where w|i−j| is a weight We propose estimating Σn by the matrix Σ i,j=1 function which down-weighs the values of γˆ|i−j| when |i − j| is large; this is desirable because estimated covariances with large values of |i − j| are known to be less reliable (see, for example, Brockwell and Davis, 1991). The motivation for our approach lies in the relationship between this problem and that of spectral density estimation. The spectral density is defined as f (ω) =

∞ 1 X γj e−iωj , 2π j=−∞

3

and is nonparametrically estimated by l 1 X ˆ wj γˆj e−iωj , f (ω) = 2π j=−l

where the wj are weights that play a role analogous to those used in the present problem. In the context of spectral density estimation, the weighting scheme we propose here has shown to provide optimal convergence rates (Politis and Romano, 1995) and to allow for a straightforward method of banding parameter selection (Politis, 2003a); we show that these advantages carry over to the present setting. With this motivation in mind, we denote our weight function by κ(·) and define it as follows. Definition 1. The tapered weight function κ is given by    1 if |x| ≤ 1   κ(x) = g(|x|) if 1 < |x| ≤ cκ     0 if |x| > c ,

(2)

κ

where g(·) is a function satisfying |g(x)| < 1, and cκ is a constant satisfying cκ ≥ 1. The l-scaled version of κ(·) will be denoted by κl (x) = κ(x/l). With this notation, our tapered estimator of Σn is given by  n ˆ κ,l = κl (i − j)ˆ Σ γ|i−j| i,j=1 .

(3)

A simple example of a weight function satisfying Definition 1 is the trapezoid proposed by Politis and Romano (1995), i.e.,

κ(x) =

        

if |x| ≤ 1

1

2 − |x| if 1 < x ≤ 2

(4)

if |x| > 2

0

but other choices are possible. For example, McMurry and Politis (2004) consider an infinitely differentiable weight function, and Politis (2007) considers several smooth tapers. Remark 1. The function g(x) will typically also be decreasing in |x| in such a way that κ(x) is continuous; these restrictions do not impact asymptotic convergence rates but they tend to improve finite sample results. The banded estimator of Wu and Pourahmadi (2009) can be put in 4

the framework of our general tapered estimator (3) with the choice cκ = 1 and no function g, i.e., a rectangular window κ(x). However, the rectangular window does not perform well for spectral estimation, and similarly here the use of a non-rectangular window is recommended. ˆ κ,l to Σn , we need to impose some short range In order to establish convergence rates of Σ dependence assumptions on the time series. We follow Wu and Pourahmadi (2009) in adopting the physical dependence measure of Wu (2005). Let i , i ∈ Z be a sequence of i.i.d. random variables. Moreover, assume that Xi is a causal process of the form Xi = f (. . . , i−1 , i ),   where f is a measurable function such that Xi is well defined and E Xi2 < ∞. In order to quantify the dependence, let 0i be an independent copy of i , i ∈ Z. Let ξi = (. . . , i−1 , i ), ξi0 = (. . . , −1 , 00 , 1 , . . . , i ), and Xi0 = g(ξi0 ). For α > 0, we define the physical dependence measure δα (i) := ||Xi − Xi0 ||α , where ||Y ||α := E [|Y |α ]1/α . Note that the difference between Xi and Xi0 is due only to the difference between 0 and 00 , and therefore δα (i) measures the dependence of Xi on an event i units of time in the past. To measure the cumulative dependence across all time, the quantity ∆α :=

∞ X

δα (i)

i=1

is helpful. We will say that {Xi } is short-range dependent with moment α if ∆α < ∞. This notion of short-range dependence provides conditions strong enough to establish our main ˆ κ,l to Σn . result, which gives an upper bound for the rate of convergence of Σ Theorem 1. Let 1 < q ≤ 2. Assume ||X1 ||2q < ∞, ∆2q < ∞, and 0 ≤ cκ l < n − 1. Then bcκ lc n X 2 X −(q−1)/q ˆ ||ρ(Σκ,l − Σn )||q ≤ dq (bcκ lc + 1)n + i|γi | + 2 |γj |, n i=1

(5)

i=l+1

where dq is a constant depending on ||X1 ||2q , ∆2q , and q, and cκ is as given in (2). Theorem 1 generalizes Theorem 2 of Wu and Pourahmadi (2009) who consider an estimate Σn of the same form as that given in equation (3) but with the weights restricted to being only 1 or 0, i.e., a rectangular window. The additional generality is achieved without changing the overall rate of convergence. 5

Remark 2. Theorem 1 is stated for mean zero data, but the result applies equally well to the ¯ . . . , Xn − X. ¯ centered data X1 − X, The inequality (5) suggests approximately optimal rates for l depending on the rate at which |γi | → 0 as i → ∞. Corollary 1. The convergence rate for the bound in inequality (5) can be optimized by minimizing the bound (5) as a function of l. The optimal bounds are found to be: i. If |γi | = O(i−d ) for some d > 1, then the rate for the bound in Theorem 1 is optimized by   choosing l ∝ n(q−1)/(dq) , and the bound (5) becomes of order O n−(d−1)(q−1)/(dq) . ii. If |γi | = O(θi ) for some θ with |θ| < 1 and if l = ba log nc for a large enough, then the  bound (5) becomes of order O n−(q−1)/q log n . iii. If there exists B such that γi = 0 for all i > B, then if l = B, the bound (5) becomes of order  O n(−q−1)/q . In all three cases above, the second term of the bound (5) is dominated by the other two terms.

3

Positive definite autocovariance matrix estimation

ˆ κ,l is asymptotically invertible and Under some additional conditions, Theorem 1 implies that Σ ˆ −1 to Σ−1 . provides a bound for the convergence rate of Σ n κ,l Theorem 2. Assume l grows fast enough to ensure the convergence (5) and that l = o(n(q−1)/q ). Also assume that the spectral density ∞ X

f (ω) = (2π)−1

γk e−iωk ,

k=−∞

satisfies 0 < c1 ≤ f (ω) ≤ c2 < ∞ for some positive constants c1 and c2 . Then, under the conditions ˆ κ,l is positive definite with probability tending to 1, and of Theorem 1, Σ ∞   X ˆ −1 − Σ−1 = Op (rn ), where rn = ln−(q−1)/q + ρ Σ |γj |. n κ,l i=l

ˆ κ,l is not guaranteed to be positive definite for finite samples. If positive definiteness However, Σ ˆ κ,l is desired, a modified estimator achieves this goal without compromising accuracy. In of Σ

6

ˆ κ,l = Tn DT t where Tn is an orthogonal matrix, particular, consider the spectral decomposition Σ n ˆ κ,l . Now let and D = diag(d1 , . . . , dn ), a diagonal matrix containing the eigenvalues of Σ ˆ  := Tn D Tnt , Σ κ,l where D = diag(d1 , . . . , dn ) and di = max(di , ˆ γ0 /nβ ); here β and  are user-defined positive ˆ constants to be discussed below. The presence of the term γˆ0 in the definition of di makes Σ κ,l scale-equivariant. ˆ  is positive definite by construction. The following is the analog of It is obvious that Σ κ,l ˆ . Theorem 1 for the modified estimator Σ κ,l Theorem 3. Let 1 < q ≤ 2. Assume ||X1 ||2q < ∞, ∆2q < ∞, and 0 ≤ cκ l < n − 1. Then bcκ lc n X 4 X −(q−1)/q  ˆ + ||ρ(Σκ,l − Σn )||q ≤ 2dq (bcκ lc + 1)n i|γi | + 4 |γj | n i=1

β

+γ0 /n + O(n

1/q−1−β

i=l+1

),

(6)

where dq , ||X1 ||2q , ∆2q , and q, and cκ are as in Theorem 1. The two last terms on the right hand side of (6) are dominated by the first term when β > ˆ  maintains the same 1/2. The following corollary ensues showing that the modified estimator Σ κ,l ˆ κ,l . asymptotic rate of convergence as Σ Corollary 2. Assume the conditions of Theorem 3 and that β > 1/2. Then, ˆ κ,l − Σn )||q ). ˆ  − Σn )||q = O(||ρ(Σ ||ρ(Σ κ,l For practical use, it is advisable not to take β close to the threshold 1/2. In simulation, we found β = 1 in conjunction with  = 1 worked well. Taking  = 0 will result into an estimator that is non-negative definite but not necessarily positive definite. An immediate corollary of the two preceding theorems is that the inverse positive definite version of the estimator also achieves the same convergence rates as given in Theorem 2. Corollary 3. Under the conditions of Theorems 2 and 3,   ˆ  )−1 − Σ−1 = Op (rn ). ρ (Σ n κ,l

7

4

Banding parameter selection

In this section we recall the rule introduced in Politis (2003a) for estimating the bandwidth in spectral density estimation using flat-top kernels.

Empirical rule for picking l. (Politis, 2003a) Let %(k) = γk /γ0 and %ˆ(k) = γˆk /ˆ γ0 . Let ˆl be the p smallest positive integer such that |ˆ %(ˆl + k)| < c log n/n for k = 1, . . . , Kn where c > 0 is a fixed constant, and Kn is a positive, nondecreasing sequence that satisfies Kn = o(log n). The rates of increase of ˆl chosen by the above rule vary according to how quickly the autocorrelation function of the process decays; they are summarized in the following theorem. Theorem 4. (Politis, 2003a) Assume conditions strong enough to ensure that for all finite m √ max |ˆ %(s + i) − %(s + i)| = OP (1/ n)

i=1,...,m

uniformly in s, and max

i=0,1,...,n−1

|ˆ %(i) − %(i)| = OP

p  log n/n .

Also assume there exists a positive i0 such that |γi | > 0 for all i < i0 . i. Assume that γi ∼ Ci−d , and for some C > 0, and a positive integer d. Then, A1 n1/2d P ˆl ∼ (log n)1/2d ii. Assume γi ∼ Cθi , where C > 0, and |θ| < 1 are some constants. Then P ˆl ∼ A2 log n

where A2 = −1/ log |θ|. iii. If γi = 0 for all k > B ≡ i0 , but γB 6= 0, then ˆl = B + op (1). Note ˆl automatically adapts to the underlying correlation structure by switching its rate of increase without any decision from the practitioner. Remark 3. Theorem 4 remains valid for all c > 0 and 1 ≤ Kn ≤ n, although different choices of c and Kn lead to very different finite sample performances. Nonetheless, there are some guidelines 8

for practically useful choices of these tuning parameters. The factor



log n varies slowly, so it has

little influence. For example, if log is taken to denote base 10 logarithm, then for sample sizes √ between 100 and 1000, as is quite typical, log n varies between 1.41 and 1.73. Thus, if c is chosen p to be around 2 and Kn about 5, the event %(ˆl + k) < c log n/n for k = 1, . . . , Kn corresponds by the Bonferroni inequality to an approximate 95% hypothesis test for %(ˆl + 1), . . . , %(ˆl + Kn ) simultaneously equal to 0. We have found values in this range work well in practice. An immediate consequence of Theorem 4 is that, in the case where q = 2, the above rule proves close to optimal in the present setting of estimating Σn . As a matter of fact, except for the slowly varying factor (log n)1/2d in case i, the rates of increase of ˆl are the same as the optimal rates for q = 2 given in Corollary 1. We thus have the following Corollary that gives credence to the applicability of ˆl for use in estimating the autocovariance matrix. Corollary 4. Assume ||X1 ||4 < ∞, ∆4 < ∞, 0 ≤ cκ l < n − 1, and let ˆl be picked by the above empirical rule. Then i. If γi ∼ Ci−d for some C > 0 and positive integer d, then,   ˆ κ,l − Σn )||2 = OP (n/ log n)(−1/2)(1−d−1 ) . ||ρ(Σ ii. If γk ∼ Cθi for some C > 0 and |θ| < 1, then   ˆ κ,l − Σn )||2 = OP n−1/2 log n . ||ρ(Σ iii. If γi = 0 for all k > B ≡ i0 , but γB 6= 0, then   ˆ κ,l − Σn )||2 = OP n−1/2 . ||ρ(Σ

5

Linear Process Bootstrap

There are several bootstraps for time series data; see, for example, B¨ uhlmann (2002), Lahiri (2003), or Politis (2003b) for reviews. The most popular methods in the literature are the block bootstrap and the AR sieve. The block bootstrap of K¨ unsch (1989) and Liu and Singh (1992) create bootstrap pseudo-data by resampling from blocks of b consecutive observations. If b, which is assumed to grow with n, is sufficiently large, the pseudo-data will have a dependence structure which closely mimics that of the original process. The AR sieve bootstrap of Kreiss (1992), Paparoditis and

9

Streitberg (1992), and B¨ uhlmann (1997) fits an AR(p) model to the original data, and then uses the fitted model in conjunction with a residual bootstrap to simulate pseudo-data. Letting p grow with n allows the sieve bootstrap to asymptotically capture the covariance structure of the original time series. A natural extension of the AR sieve would be an MA sieve, which models the observed time series by fitting increasingly high order MA(q) processes to the data; this has not been done because of the relative difficulty of fitting MA models. MA models are either fit by numerical optimization, which is not feasible for large values of q, or by algorithms such as the innovations algorithm presented in Theorem 8.3.1 of Brockwell and Davis (1991). Unfortunately, the innovations algorithm requires estimating MA coefficients of orders much greater than q in order to assess the stability of the first q fitted parameters; see the discussion following Theorem 8.3.1 in Brockwell and Davis (1991). Below, we propose a new bootstrap, termed the linear process bootstrap (LPB) which is an ˆ  makes it possible to generate an alternative to an MA sieve; it works because knowledge of Σ κ,l MA process without knowing the MA coefficients. The LPB is also more general because one could, in principle, use a taper κ(·) that is not identically zero after a point, but just tends to zero, (see, for example Politis, 2007); in that case the LPB generates linear MA(∞) rather than MA(q) processes. We prove the validity of the LPB for the mean, and we conjecture its validity for all statistics whose asymptotic distribution depends only on the mean and covariance of the data. The LPB algorithm is as follows. ¯ for i = 1, . . . , n, and let Y = (Y1 , . . . , Yn )t . 1. Let Yi = Xi − X ˆ  )−1/2 Y . 2. Let W = (Σ κ,l ¯ )/ˆ ¯ = n−1 Pn Wi 3. Let Z be the standardized version of W , with Zi = (Wi − W σW , where W i=1 Pn 2 −1 2 ¯ and σ ˆW = n i=1 (Wi − W ) . 4. Generate Z1∗ , . . . Zn∗ by an i.i.d. bootstrap of Z1 , . . . , Zn . ˆ  )1/2 is taken to be the lower triangular matrix L in ˆ  )1/2 Z ∗ , where (Σ 5. Compute Y ∗ = (Σ κ,l κ,l ˆ  = LLt . the Cholesky decomposition Σ κ,l ˆ  )−1/2 in step 2 can be any matrix square root that converges Remark 4. The matrix square root (Σ κ,l −1/2

to Σn

ˆ  converges to Σn , such as those obtained by the Cholesky or spectral at the same rate as Σ κ,l

decompositions (see, for example, Horn and Johnson, 1990, p. 411). We conjecture that the same is true of the square root used in step 5, but our proof of Theorem 5 (below) is specific to the 10

Cholesky decomposition. For reasons of symmetry, it seems preferable to use the same square root in step 2 as in step 5. Under assumptions of the preceding theorems, the algorithm above can be used to produce confidence intervals for the mean which are justified by the following theorem. Theorem 5. Let E [Xi ] = µ. Then under the conditions of Theorems 1, 2, and 3, with q = 2, h i i h ¯ − µ) ≤ x − P∗ n1/2 Y¯ ∗ ≤ x →P 0, sup P n1/2 (X

(7)

x

and h i var∗ n1/2 Y¯ ∗ →P σ 2 , where σ 2 = (γ0 + 2

P∞

k=1 γk )

  ¯ . = limn→∞ var n1/2 X

Remark 5. Surprisingly, the assumptions of Theorem 5 do not include linearity of the original time series, i.e., an MA(∞) model. The LPB manages to generate a linear process with covariance structure approaching that of the original time series while requiring only the weak dependence conditions used in estimating Σn . Thus, the range of applicability of the LPB is quite wide and includes a vast class of nonlinear processes. The name ‘Linear Process Bootstrap’ therefore only alludes to the fact that a linear process is being generated in the bootstrap world. Furthermore, we conjecture that the LPB is applicable to all statistics which depend only on the first two moments of the process.

6 6.1

Simulations Covariance matrix estimation

We conducted several simulations to assess the performance of our estimator and to draw a direct comparison with the method of Wu and Pourahmadi (2009); they use a subsampling rule to estimate the banding parameter l, whereas we employ the empirical rule of Section 4. We also use three different weight functions: rectangular, defined by κ(x) = 1{|x| < 1}; trapezoidal as defined in equation (4); and infinitely differentiable, as defined in McMurry and Politis (2004) and shown in Figure 1. Simulations were performed in R (R Development Core Team, 2009). Results are based on N = 100 replications, and the parameters in the bandwidth choice rule were chosen to be c = 2, and Kn = 5. Each model we considered was tested with sample sizes n = 250, 500, and 750.

11

0.8 0.4 0.0

κ(x)

-2

-1

0

1

2

x

Figure 1: An infinitely differentiable weight function

We assessed the the performance by computing a matrix norm of the difference between the ˆ κ,l −Σn ), and then averaged these numbers over all 100 repliestimated matrix and the true, i.e. ρ(Σ cations. We used two matrix norms, the operator norm ρ(·) and, to allow for a direct comparison with Wu and Pourahmadi (2009), the matrix infinity norm |||A|||∞ :=

max i∈{1,...,n}

n X

|aij |.

j=1

The averaged errors in the two norms will be referred to respectively as operator norm losses and infinity norm losses. We also tested the adjustment to positive definiteness given in Theorem 3 using the trapezoid weight function. While negative eigenvalues were occasionally observed, they were so close to zero that the results were numerically identical in operator norm loss to the unadjusted trapezoid estimator. For this reason, these results will not be further discussed. The rectangular weight function is expected to produce many more nonpositive matrices, as the negative sidelobes of its Fourier transform are more pronounced than those of the trapezoid’s. In what follows we will qualitatively describe the experiments we ran and highlight the notable results. Full numeric results can be found in Tables 1–3 of McMurry and Politis (2010). While most results were specific to individual simulations, across simulations the trapezoid weight function was consistently the best performing of the three weight functions. The infinitely differentiable weight function was often close but never better, and both were substantially better than the rectangular weight function.

12

6.1.1

MA(1)

In the first simulation, the data were generated by the moving average process Xt = t + 0.5t−1 , with t and iid sequence of N (0, 1) random variables. For all three weight functions, the infinity norm losses given in Wu and Pourahmadi (2009) are more than 10 times our losses, and our method comes close to achieving the oracle bound given in their Table 1. 6.1.2

AR(1)

In the second experiment, data were simulated from the AR(1) process Xt = φXt−1 + t , where the t were iid N (0, 1 − φ2 ), for φ = 0.1, 0.5, and 0.9; the error variance was chosen to make var [Xt ] = 1 for all simulations. Wu and Pourahmadi (2009) do not provide numeric results for this case. In all simulations the performance of our estimator was substantially worse for φ = 0.9 than for the two smaller values; this is to be expected, as off diagonal terms of Σn are substantially larger for large φ. Losses in this case were also somewhat larger than losses for the MA model; this is also unsurprising as the MA covariance matrix can be estimated without tapering or truncating any non-zero entries, whereas tapering non-zero entries cannot be avoided for an AR process. Finally, as expected, results consistently improve with increasing sample size. 6.1.3

Absolute value AR(1)

For the final simulation, data were simulated from the model Xt = φ|Xt−1 | + t , for φ = 0.1, 0.5, and 0.9, and where t were i.i.d. N (0, 1). The autocovariance function for the absolute AR model does not have a simple closed form, so we approximate the true Σn with its empirical version using a very large amount of simulated data. While this provides (crude) estimates of the loss, it induces some additional difficult to quantify uncertainty. Nonetheless, our results again show significant improvement over those presented in Wu and Pourahmadi (2009), particularly for smaller values of φ. With n = 250, our infinity norm losses for φ = 0.1 were always less than 25% of theirs, for φ = 0.5 our losses were always less than 50% of theirs, and for φ = 0.9, our loss was about 90%; our method’s advantage improved with increasing n.

13

6.2

Linear Process Bootstrap

Finally, we ran several simulation experiments to assess the performance of the linear process bootstrap. For comparison we also tested each simulated data set using two other popular resampling schemes. First, we considered the block bootstrap as implemented in Canty and Ripley (2009), using the block length selection of Politis and White (2004) (see also Patton, Politis, and White, 2009). Second, we also used the AR sieve bootstrap with AIC for model selection. Reviews of the block and AR-sieve bootstraps can be found in B¨ uhlmann (2002) and Politis (2003b). Each experiment was repeated 1000 times using 1000 bootstrap replications. Full numeric results can be found in Table 4 of McMurry and Politis (2010). 6.2.1

MA(1)

The first model we considered was the moving average process Xt = t + φt−1 for φ = 0.1, 0.5, and 0.9, and i iid N (0, 1). We used sample sizes n = 250, 500, and 750. All three bootstraps performed well for most combinations of φ and n, with typical coverage of 93–94% for nominal 95% confidence intervals. The LPB seemed to be slightly better than its two competitors for the two larger values of φ, and comparable or very slightly worse for φ = 0.1. It is expected that the LPB should excel for MA processes, as its banded covariance matrix structure can produce data that exactly possesses an MA covariance structure. The AR sieve should have the most difficulty in this setting because an MA(1) process can only be well represented as a very high order AR process; this was evident when n = 250 and φ = 0.9, where the AR sieve only produced coverage of 90%. In all other situations the AR-sieve’s performance was very similar to the other bootstrap estimators. 6.2.2

AR(1)

The second model we considered was the AR(1) process considered in Section 6.1.2, with the same values of φ and n. The LPB performed slightly worse than the block bootstrap for the two smaller values of φ and noticeably better when φ = 0.9. The AR sieve consistently outperformed the other two methods; this was expected because an AR model is exactly correct in this setting. Perhaps the most surprising feature of the simulations was that the sieve bootstrap performed somewhat poorly when the AR coefficient was 0.9. In this case, for nominal 95% confidence intervals, we observed coverages of 88% with n = 250, 91% with

14

n = 500, and 94% with n = 750. Although the AR sieve bootstrap is expected to break down when φ is close to 1, the bad behavior for φ = 0.9 was unexpected. 6.2.3

Absolute value AR(1)

Finally we considered the absolute value AR(1) process described in Section 6.1.3. In this case, the LPB’s coverage was identical to the block bootstrap’s when φ = 0.1, noticeably worse when φ = 0.5, and somewhat better when φ = 0.9. The AR sieve again consistently outperformed the other two estimators in this setting. Surprisingly, in this situation, the AR sieve did as well as or better than it did for the AR(1) processes, for which it should be ideal. Here, the AR sieve’s coverage was never more than 3% different from the LPB’s.

7

Technical proofs

The proof of Theorem 1 rests on the following lemma. Lemma 1. (Wu and Pourahmadi, 2009) Assume that {Xi } satisfies ∆2q < ∞ with 1 < q ≤ 2. Then for any j ∈ Z, ||

n X

Xi Xi+|j| − nγj ||q ≤ 2Bq n1/q ||X1 ||2q ∆2q

i=1

where Bq =

 

18q 3/2 (q−1)1/2

if q 6= 2



1

if q = 2.

ˆ κ,l − Σn is Proof of Theorem 1: By Problem 21, p. 313 in Horn and Johnson (1990), and since Σ symmetric, ˆ κ,l − Σn ) ≤ ρ(Σ ≤

n X

max

1≤j≤n n−1 X

|ˆ γi−j κl (|i − j|) − γi−j |

i=1

|ˆ γi κl (i) − γi |

i=1−n

≤ 2

l X

bcκ lc

X

|ˆ γi − γi | + 2

i=0

i=l+1

= T1 + T2 + T3 .

15

|ˆ γi κl (i) − γi | + 2

n X i=bcκ lc+1

|γi |

We first examine T1 . By Lemma 1, there exists a constant d0q depending on ||X1 ||2q and ∆2q , but not l or n, such that i |γi | n i|γi | + . n

||ˆ γi − γi ||q ≤ ||ˆ γi − E [ˆ γi ] ||q + d0q (n − i)1/q n

≤ Therefore

l

2X i|γi |, n

||T1 ||q ≤ dq (l + 1)n−(q−1)/q +

i=1

where dq = d0q /2. The second term, T2 , can be studied in a similar fashion. bcκ lc

T2 ≤ 2

X

bcκ lc

κl (i)|ˆ γi − γi | + 2

i=l+1

X

|κl (i) − 1||γi |.

i=l+1

Therefore, ||T2 ||q ≤ dq (bcκ lc − l)n−(q−1)/q +

bcκ lc bcκ lc X 2 X |γi |. i|γi | + 2 n i=l+1

(8)

i=l+1

The final term of (8) can now be combined with T3 . Since cκ ≥ 1, bcκ lc

2

X

|γi | + T3 = 2

i=l+1

n X

|γi |,

i=l+1

the last term in (5). Proof of Theorem 2: All the eigenvalues of Σn lie in the interval [2πc1 , 2πc2 ] (see Grenander and   ˆ κ,l − Σn = Op (rn ). Since rn tends to zero, the Szeg¨o, 1958, Section 5.2). By Theorem 1, ρ Σ ˆ κ,l is positive definite tends to 1. probability Σ −1/2

Let An = Σn

ˆ κ,l An . Then and Γn = An Σ   ˆ κ,l − Σn ρ (Γn − In ) ≤ ρ(An )2 ρ Σ = Op (rn ).

Similarly, ρ Γ−1 n − In



≤ ρ(Γ−1 n ) ρ (Γn − In ) = Op (rn ).

ˆ −1 − Σ−1 = An (Γ−1 − In )An , the result follows. Since Σ n n κ,l 16

Proof of Theorem 3: By the triangle inequality, ˆ  − Σn ) ≤ ρ(Σ ˆ κ,l − Σn ) + ρ(Σ ˆ − Σ ˆ κ,l ). ρ(Σ κ,l κ,l

(9)

ˆ κ,l = Tn DTnt , where without loss of generality, we assume that the eigenvalues of Σ ˆ κ,l Recall that Σ have been ordered so that D = diag(d1 , . . . , dn ), where d1 ≥ d2 ≥ . . . ≥ dn . Let λmax (A) and λmin (A) respectively denote the largest and smallest eigenvalues of a symmetric matrix A. Then, ˆ κ,l ) dn = λmin (Σ ˆ κ,l ) = −λmax (−Σ ˆ κ,l ) ≥ −λmax (Σn − Σ ˆ κ,l ), ≥ −ρ(Σn − Σ

(10)

where the first inequality follows because Σn is non-negative definite (see Corollary 4.3.3, p.182 in Horn and Johnson, 1990). We now focus on the second term of (9). ˆ κ,l = Tn D− Tnt ˆ − Σ Σ κ,l  where D− = diag max(d1 , ˆ γ0 /nβ ) − d1 , . . . , max(dn , ˆ γ0 /nβ ) − dn . By the above spectral decomposition and inequality (10),   β ˆ − Σ ˆ κ,l ) = max 0, ˆ ρ(Σ γ /n − d 0 n κ,l   ˆ κ,l ) ≤ max 0, ˆ γ0 /nβ + ρ(Σn − Σ ˆ κ,l ). ≤ ˆ γ0 /nβ + ρ(Σn − Σ The result now follows from Theorem 1 and Lemma 1. Proof of Theorem 5: By Theorem 3 in Wu (2005), D

¯ − µ) −→ N (0, σ 2 ). n1/2 (X We establish (7) by showing n1/2 Y¯ ∗ has the same limiting normal distribution. For clarity of exposition, the proof proceeds through a sequence of lemmas. Lemma 2. Define Z˜ ∗ to be the equivalent bootstrap resample to Z ∗ , except the resample is drawn −1/2

from the standardized values of Σn

ˆ  )−1/2 Y . Let 1 Y rather than its data driven counterpart (Σ κ,l 17

be the n-vector of 1’s. Under the conditions of Theorem 5, ˆ  )1/2 Z ∗ n1/2 Y¯ ∗ = n−1/2 1t (Σ κ,l h i −1/2 t ˜∗ ˆ  )1/2 − Σn1/2 Z ∗ = n−1/2 1t (Σ1/2 1 (Σn1/2 )(Z ∗ − Z˜ ∗ ) + n−1/2 1t (Σ n )Z + n κ,l ˜∗ = n−1/2 1t (Σ1/2 n )Z + R1 + R2

(11)

˜∗ = n−1/2 1t (Σ1/2 n )Z + oP (1). Proof of Lemma 2. We first consider R2 . It has bootstrap mean 0 and variance h   i ˆ  )1/2 − Σ1/2 Z ∗ var∗ n−1/2 1t (Σ n κ,l h h i h i i ˆ  )1/2 − Σ1/2 Z ∗ (Z ∗ )t (Σ ˆ  )1/2 − Σ1/2 1 = E ∗ n−1 1t (Σ n

κ,l

κ,l

n

= oP (1),     ˆ  )1/2 − Σ1/2 →P 0. where the final equality follows because E ∗ Z ∗ (Z ∗ )t = I and ρ (Σ n κ,l −1 ˆ −1/2 Y where 1n is the n × n matrix of ones, M ∗ (I − n−1 1n )Σ For R1 , we can write, Z ∗ = σ ˆW κ,l

and M ∗ is a random n × n matrix, where each row is independently and uniformly selected from −1/2 −1 ∗ −1 the standard basis vectors e1 , . . . , en . With this notation, Z˜ ∗ = σ ˆW Y . Since ˜ M (I − n 1n )Σn 2 −σ 2 | = o (1), and both σ 2 and σ 2 are bounded away from 0 and from above with probability |ˆ σW ˆW ˆW ˆW P ˜ ˜

tending to 1, i h −1 −1/2 −1 ˆ −1/2 ∗ ˜ ∗ ) = n−1/2 1t (Σ1/2 )M ∗ (I − n−1 1n ) σ Σ − σ ˆ Σ Y n−1/2 1t (Σ1/2 )(Z − Z ˆ n n n ˜ W κ,l W i h −1 −1/2 −1 ˆ −1/2 ∗ −1 Y + oP (1) −σ ˆW = n−1/2 1t (Σ1/2 ˆW n )M (I − n 1n ) σ ˜ Σn ˜ Σκ,l = R3 + oP (1). It is clear by construction that E ∗ [R3 ] = 0. Its bootstrap variance is var∗ [R3 ] h i −2 ∗ ∗ −1 ˆ −1/2 − Σ−1/2 )Y Y t (Σ ˆ −1/2 − Σ−1/2 )(I − n−1 1n )(M ∗ )t Σ1/2 1 = σ ˆW n−1 1t Σ1/2 n M (I − n 1n )(Σκ,l n n n ˜ E κ,l h i −2 ∗ −1 t 1/2 ∗ ∗ t 1/2 = σ ˆW E n 1 Σ V (V ) Σ 1 n n ˜

ˆ −1/2 − Σ−1/2 where V ∗ is an n-vector of bootstrap resamples of the elements of (I − n−1 1n )(Σ )Y . n κ,l   Since the sample is i.i.d., E ∗ V ∗ (V ∗ )t = σV2 I, where

σV2

ˆ −1/2 − Σn−1/2 )(I − n−1 1n )2 (Σ ˆ −1/2 − Σn−1/2 )Y = n−1 Y t (Σ κ,l κ,l −1/2

ˆ = n−1 Y t (Σ κ,l

−1/2

ˆ − Σn−1/2 )(Σ κ,l 18

− Σn−1/2 )Y (1 + OP (1)).

−1/2

ˆ Since Y t Y = OP (n) (see Wu, 2005) and ρ(Σ κ,l

−1/2

− Σn

) = OP (rn ), we have σV2 = oP (1) which

implies var∗ [R3 ] = oP (1) and therefore R3 = oP (1). Lemma 3.

 h i ˜ ∗ = n−1 var∗ n−1/2 1t (Σ1/2 n )Z



n−1 X

(n − |i|)γi 

i=−(n−1)

Proof of Lemma 3. i h i h ˜ ∗ = E ∗ n−1 1t (Σ1/2 ˜ ∗ ˜ ∗t −1/2 1 var∗ n−1/2 1t (Σ1/2 n )Z n )Z Z Σn   n−1 X = n−1 (n − |i|)γi  , i=−(n−1)

i h where the final equality follows because E ∗ Z˜ ∗ Z˜ ∗t = I. Lemma 4. Let An and Bn be sequences of n × n symmetric matrices bounded in operator norm and satisfying ρ(An − Bn ) → 0. Then n−1/2 1t An Z˜ ∗ = n−1/2 1t Bn Z˜ ∗ + oP (1). Proof of Lemma 4. Since Z˜ ∗ has bootstrap mean 0, it is sufficient to show the variance converges to 0 in probability. h i var∗ n−1/2 1t (An − Bn )Z˜ ∗ = σZ2˜ ∗ n−1 1t (An − Bn )2 1 ≤ ρ(An − Bn )(1 + op (1)) = op (1)

h i Lemma 5. Under the assumptions of Theorem 5, E Z˜i4 uniformly bounded in i. −1/2

Proof of Lemma 5. Denote the entries of Σn

P = [aij ]ni,j=1 . With this notation, Z˜i = nj=1 aij Yj .

Following Theorem 2 in Wu (2005), define the projection operator Pk Y = E [Y |ξk ] − E [Y |ξk−1 ], P P and define Mk,n = nj=1 aij Pj−k Yj . Then, Z˜i = ∞ k=0 Mk,n . By Proposition 4 in Dedecker and Doukhan (2003),  ||Mk,n ||4 ≤ ||P0 Yk ||4 8

n X

1/2 a2ij 

j=1

Pn 2 = eti Σ−1 n ei , where ei is the i’th standard basis vector, for large enough n, j=1 aij is P bounded from above and away from 0 uniformly in i. By Theorem 1 in Wu (2005), ∞ k=0 ||P0 Yk ||4 < Since

Pn

2 j=1 aij

∞. Therefore, ||Z˜i ||4 is uniformly bounded in i. 19

Lemma 6. (Horn and Johnson, 1990, p. 411) Let A and B be symmetric and positive definite. Let A1/2 and B 1/2 denote the lower triangular square roots obtained through the Cholesky decomposition of A and B. Then ρ(A1/2 − B 1/2 ) ≤ ρ(A−1/2 )ρ(A − B). 1/2

We are now in a position to complete the proof of Theorem 5. We do so by approximating Σn  n in (11) in the following manner. Let Σn,k = γ|i−j| 1|i−j|≤k i,j=1 be the k-banded version of Σn . By Horn and Johnson (1990) p. 313, ρ(Σn,k − Σn ) ≤ 2

∞ X

|γi |.

i=k+1 1/2

Therefore ρ(Σn,k − Σn ) → 0 for any sequence k → ∞. Let Ln,k and Σn

be the lower-triangular

matrices associated with the Cholesky decompositions of Σn,k and Σn respectively. By Lemma 6, 1/2

1/2

ρ(Ln,k − Σn ) → 0, so by Lemma 4 we can approximate Σn

in (11) by Ln,k .

Matrix multiplication shows that Ln,k is nonzero only on the main diagonal and the first k 1/2

diagonals below the main, and that the entries of Ln,k are bounded in absolute value by γ0 . P Letting c1,n , . . . , cn,n denote the column sums of Ln,k , we immediately see ni=1 c4i,n = O(k 4 n) = O(n(log n)4 ), if we choose k ∝ log n. We can now establish the main result. 1/2

In order to proceed with the proof of Theorem 5, we use Ln,k to approximate Σn

in the first

term of (11). n−1/2 1t (Ln,k )Z˜ ∗ = n−1/2 (c1,n Z˜1∗ + . . . + cn,n Z˜n∗ ). We establish the desired result via the central limit theorem for triangular arrays (Resnick, 1999, p. 321) which is implied by the Liapunov condition n i h X 1 ∗ ˜ ∗ |2+δ → 0, E |c Z i,n i (c21,n + . . . + c2n,n )1+δ/2 i=1

(12)

for some δ > 0. Convergence (12) will be shown to hold for δ = 2. We first examine the numerator. ! ! n n n h i X X X E ∗ |ci,n Z˜i∗ |2+δ = c4i,n n−1 Zi4 = OP (k 4 n) i=1

i=1

i=1

by Lemma 5 and the preceding calculation. We now turn our attention to the denominator of (12). c21,n + . . . + c2n,n = 1t Ln,k L0n,k 1 = 1t Σn,k 1.

20

By Problem 21, p. 313 in Horn and Johnson (1990), ρ(Σn,k − Σn ) ≤ 2

Pn

i=k+1 |γi |.

Therefore

ρ(Σn,k − Σn ) → 0 for any sequence k → ∞. Since the eigenvalues of Σn lie in the interval [2πc1 , 2πc2 ] for c1 and c2 as in Theorem 2, for any  > 0 there exists k large enough such that (2πc1 − )n < 1t Σn,k 1 < (2πc2 + )n. This establishes that the denominator of (12) is O(n2 ) when δ = 2. The only requirement on k is that k → ∞, so we now choose k = log n, and it immediately follows that the left side of (12) converges to 0 in probability. With Lemma 3, this completes the proof.

Acknowledgements Many thanks to Mohsen Pourahmadi, Wei Biao Wu, and an anonymous referee for their comments, critiques, and helpful suggestions.

References P. J. Brockwell and R. A. Davis. Time Series: Theory and Methods. Springer, New York, 1991. P. B¨ uhlmann. Sieve bootstrap for time series. Bernoulli, 3(2):123–148, 1997. P. B¨ uhlmann. Bootstraps for time series. Statist. Sci., 17(1):52–72, 2002. A. Canty and B. D. Ripley. boot: Bootstrap R (S-Plus) Functions, 2009. R package version 1.2-41. J. Dedecker and P. Doukhan. A new covariance inequality and applications. Stochastic Process. Appl, 106(1):63–80, 2003. U. Grenander and G. Szeg¨ o. Toeplitz Forms and Their Applications. University of California Press, Los Angeles, 1958. R. A. Horn and C. R. Johnson. Matrix Analysis. Cambridge University Press, New York, 1990. J.-P. Kreiss. Bootstrap procedures for AR(∞)-processes. In Bootstrapping and related techniques (Trier, 1990), volume 376 of Lecture Notes in Econom. and Math. Systems, pages 107–113. Springer, Berlin, 1992. H. R. K¨ unsch. The jackknife and the bootstrap for general stationary observations. Ann. Statist, 17:1217–1241, 1989. 21

S. N. Lahiri. Resampling Methods for Dependent Data. Springer Series in Statistics. Springer-Verlag, New York, 2003. R. Y. Liu and K. Singh. Moving blocks jackknife and bootstrap capture weak dependence. In R. LePage and L. Billard, editors, Exploring the Limits of Bootstrap, pages 225–248. Wiley, New York, 1992. T. McMurry and D. N. Politis. Nonparametric regression with infinite order flat-top kernels. Journal of Nonparametric Statistics, 16(3-4):549–562, 2004. T. McMurry and D. N. Politis. Banded and tapered estimates for autocovariance matrices and the linear process bootstrap. http://www.math.ucsd.edu/∼politis/PAPER/covmat3-2010.pdf, 2010. E. Paparoditis and B. Streitberg. Order identification statistics in stationary autoregressive movingaverage models: vector autocorrelations and the bootstrap. J. Time Ser. Anal., 13(5):415–434, 1992. A. Patton, D. N. Politis, and H. White. CORRECTION TO “Automatic block-length selection for the dependent bootstrap”. Econometric Reviews, 28(4):372–375, 2009. D. N. Politis. Adaptive bandwidth choice. Journal of Nonparametric Statistics, 25(4-5):517–533, 2003a. D. N. Politis. The impact of bootstrap methods on time series analysis. Statist. Sci., 18(2):219–230, 2003b. D. N. Politis. Higher-order accurate, positive semi-definite estimation of large-sample covariance and spectral density matrices. UCSD Deptartment of Economics, Discussion Paper 2005-03, 2007. D. N. Politis and J. P. Romano. Bias-corrected nonparametric spectral estimation. Journal of Time Series Analysis, 16(1):67–103, 1995. D. N. Politis and H. White. Automatic block-length selection for the dependent bootstrap. Econometric Reviews, 23(1):53–70, 2004. R Development Core Team. R: A Language and Environment for Statistical Computing. R Foundation for Statistical Computing, Vienna, Austria, 2009. URL http://www.R-project.org. ISBN 3-900051-07-0. 22

S. I. Resnick. A Probability Path. Birkhauser, Boston, 1999. W. B. Wu. Nonlinear system theory: Another look at dependence. Proceedings of the National Academy of Sciences, 102:14150–14154, 2005. W. B. Wu and M. Pourahmadi. Banding sample autocovariance matrices of stationary processes. Statistica Sinica, 19(4):1755–1768, 2009.

23

Banded and tapered estimates for autocovariance ... - TL McMurry, LLC

Jun 20, 2010 - A natural extension of the AR sieve would be an MA sieve, which models the ... in that case the LPB generates linear MA(∞) rather than MA(q).

274KB Sizes 2 Downloads 186 Views

Recommend Documents

Nonparametric Regression with Infinite Order Flat ... - TL McMurry, LLC
Apr 24, 2003 - All programming was done in Mathematica. Before results are presented, a few words should be said about the choice of the parameters c and h. The simulations were performed using the kernel defined by equation (3) with c = 0.05 and b =

EIGENVALUE ESTIMATES FOR STABLE MINIMAL ...
KEOMKYO SEO. (Communicated by H. Martini). Abstract. In this article ... negative constants, we can extend Theorem 1.1 as follows. THEOREM 1.3. Let M be an ...

Northwest corner and banded matrix ... - Wiley Online Library
Abstract: In this paper, we consider approximations to discrete time Markov chains with countably infinite state spaces. We provide a simple, direct proof for the convergence of certain probabilistic quantities when one uses a northwest corner or a b

Sensitivity Estimates for Compound Sums
driven models of asset prices and exact sampling of a stochastic volatility ... For each λ in some parameter domain Λ, (1) determines the distribution of X(λ) ...... Then, from standard properties of Poisson processes, it is straightforward to che

Estimates For Appliance Repair Scottsdale.pdf
Page 1 of 5. https://sites.google.com/site/appliancerepairinscottsdale/. How​ ​Hard​ ​Is​ ​It​ ​to​ ​Find​ ​a​ ​Quality​ ​Dishwasher​ ​Repair​ ...

The wild tapered block bootstrap
Sep 24, 2014 - ∗I acknowledge support from CREATES - Center for Research in Econometric Analysis of Time Series (DNRF78), funded by the Danish National Research ...... and call {. µ. ∗(TBB). Nt. } the resampled version of {µNt}.

Selberg's zeta function and Dolgopyat's estimates for ...
(see for example [1, 24, 13, 14, 18] ) on the asymptotic distribution of closed orbits for various flows of hyperbolic nature and counting problems on higher rank symmetric spaces. Some recent applications of Dolgopyat's techniques to number theory a

Distributional National Accounts - Methods and Estimates for the ...
Distributional National Accounts - Methods and Estimates for the United States.pdf. Distributional National Accounts - Methods and Estimates for the United ...

Comparing and Combining Effort and Catch Estimates ...
the confluence of its east and west branches at. Hancock, New ... our team. The assigned usage levels for the access points remained fixed within each of the seven time blocks. ... during each calendar week; sampling of low-usage sites was ...

Kinderstart.com LLC v. Google, Inc. - Balough Law Offices, LLC
Arizona Corporation Commission, 720. F.2d 578, 581 (9th Cir .... an ostensibly private association's members were public schools which “largely provided for the.

USA - WTS LLC
Jan 3, 2013 - Taxpayer Relief Act of 2012 (“ATRA”). The ATRA was ... ANY TAXPAYER OR (ii) PROMOTING, MARKETING ... Email: [email protected].

Kinderstart.com LLC v. Google, Inc. - Balough Law Offices, LLC
“PageRank is not a mere statement of opinion of the innate value or human ...... In that case, the court held that America Online (“AOL”), an Internet .... The AdSense agreement contains no express promises that serving ads will increase.

ml harper, llc
Sep 20, 2016 - Emergency Contact. Contact Name: _Day Phone: Night Phone: Cellular Phone: Alternate Contact: Phone: ... Policy Number: I hereby authorize ...

USA - WTS LLC
Jan 3, 2013 - spending cuts. The ATRA is effective on January 1, ... ANY TAXPAYER OR (ii) PROMOTING, MARKETING ... Email: [email protected].

Comparing and Combining Effort and Catch Estimates ...
surveys produced similar estimates of boat angler effort and little evidence of bias, but shore anglers were undercounted in the aerial ... Published online August 28, 2006. 727 ...... From a statistical point of view, the price to pay for employing

ENTROPY ESTIMATES FOR A FAMILY OF EXPANDING ... - CiteSeerX
where Js denotes the Bessel function of the first kind and order s. We recall ..... by evaluation of the Lyapunov exponent on a grid of stepsize 10−3 and use of the.

TL-100.pdf
Sign in. Page. 1. /. 1. Loading… Page 1 of 1. Page 1 of 1. Main menu. Displaying TL-100.pdf. Page 1 of 1.

Pre-launch estimates for GLAST sensitivity to dark ...
Jul 17, 2008 - 3 Physics Department, Stockholm University, Albanova, SE-10691 ..... Figure 3. γ-ray spectrum of the inner Galaxy (300° < l < 30°, |b| < 5) ...

Integral estimates for the family of B-Operators ...
Post Graduate Department of Mathematics,. Govt.College,Baramulla,Kashmir-193101 India. e-mail: [email protected]. ABSTRACT. Let P(z) be a polynomial of degree n.In this paper, we discuss inequalities for the family of operators B = B(λ0,λ1,λ2,.)

(19) United States W Tl. l
of access points include Apple's Airport and 3COM's Air connect wireless ... Connection speeds range from 1.6 Mbps with OpenAir tech nology to 11 Mbps with ...