SuppPub
Self-convolved Bootstrap for M-regression under Complex Temporal Dynamics
Abstract
The paper considers simultaneous nonparametric inference for a wide class of M-regression models with time-varying coefficients. The covariates and errors of the regression model are tackled as a general class of nonstationary time series and are allowed to be cross-dependent. A novel and easy-to-implement self-convolved bootstrap procedure is proposed. With only one tuning parameter, the bootstrap facilitates a -consistent inference of the cumulative regression function for the M-estimators under complex temporal dynamics, even under the possible presence of breakpoints in time series. Our methodology leads to a unified framework to conduct general classes of Exact Function Tests, Lack-of-fit Tests, and Qualitative Tests for the time-varying coefficients. These tests enable one to, among many others, conduct variable selection, check for constancy and linearity, as well as verify shape assumptions, including monotonicity and convexity. As applications, our method is utilized to study the time-varying properties of global climate data and Microsoft stock return, respectively.
Keywords: Bahadur representation, cumulative regression function, M estimator, nonstationary time series, quantile regression, shape test.
1 Introduction
There has been an increasing necessity to account for temporal non-stationarity in finance and economics. For instance, financial returns are frequently non-stationary (Francq and Sucarrat, 2023). In climate economics, climate change has become progressively unpredictable with extreme weather happening globally with greater intensity (Ebi et al., 2021). See also Dette et al. (2011); Vogt (2012); Nason (2013); Kreiss and Paparoditis (2014); Dahlhaus et al. (2019) among others for the recent literature on non-stationary time series analysis. As deviations from temporal homogeneity become more pronounced, static models may fail to capture the dynamic features of the data. In the context of regression analysis, one successful generalization is to use the varying-coefficient model (Hastie and Tibshirani, 1993), which is capable of describing the non-constant relationship between the predictors and the response (Hoover et al., 1998; Fan and Zhang, 2000; Karmakar et al., 2022). In the present work, we shall further consider the following time-varying M-regression (Huber, 1964, 1973) model for non-stationary time series:
(1) |
where is a -dimensional predictor process, is the error process satisfying almost surely with being the left derivative of a convex function , and is the coefficient evaluated at time . We study model (1) in a general framework where and belong to a general class of non-stationary time series models with both smoothly and abruptly time-varying dynamics. The detailed assumptions of the model are deferred to Section 7. Observe that (1) includes a wide range of regression models with flexible choices of . Prominent examples include the least square regression with , the quantile regression with , , , the regression with , , the expectile regression with , , and Huber’s robust regression with , .
This article aims to provide a unified framework for nonparametric inference of model (1). Due to the complex temporal dynamics and the broad choice of loss functions, it is difficult to directly estimate the limiting distributions of the M-estimators, in which case resampling methods such as the bootstrap is typically adopted to circumvent the difficulty. In the context of nonparametric M-regression, however, so far there exist no valid bootstrap methods for its simultaneous inference when the regressors and errors are non-stationary time series with possible jumps in the underlying data generating mechanism. The most related work can be found in Wu and Zhou (2017), who explored time-varying coefficient quantile regression with smoothly time-varying dynamics. We also refer to Friedrich and Lin (2022) who proposed an autoregressive sieve bootstrap framework, valid under regression with strictly stationary -mixing predictors. Extending to our current setup, the primary challenge in designing bootstrap methods originates from the inconsistency in estimating various key quantities around the jump points. Additionally, bootstrap inference for M-regression involves estimating a density-like function, which requires the delicate choice of a tuning parameter (Koenker, 2005). For instance, in quantile regression where , the asymptotic behavior of local estimates of is tied to a density-type quantity (Wu and Zhou, 2017). In moderate samples, the quality of the bootstrap inference depends on the estimation accuracy of , which substantially entails the nontrivial task of selecting the tuning parameter. Lastly, the existing literature on bootstrap inference of nonparametric M-regression, such as Friedrich and Lin (2022) and Wu and Zhou (2017), often constructs the bootstrap based on residuals. In moderate samples, the performance of such residual-based bootstrap can be significantly hampered by the estimation error of the residuals as estimators of the true regression errors.
The methodological innovation of this paper lies in proposing a novel self-convolved bootstrap for the simultaneous nonparametric inference of Model (1), which shares a simple and unified construction among a broad class of M-estimators and hypothesis tests, and is consistent under very general forms of non-stationary temporal dependence and predictor-error dependence. As its name implies, the self-convolved bootstrap only necessitates the convolution of the M-estimators with i.i.d. auxiliary standard normal variables, eliminating the need for additional computation. Particularly, it avoids estimating the aforementioned density-like quantities or residuals. The construction principle applies uniformly to a wide class of M-estimators and therefore saves lots of effort when conducting inference based on various estimators simultaneously. The validity of the proposed bootstrap arises from two main observations. Firstly, by taking the second-order difference of the local M-estimators as shown in Section 3, their progressive sum can be well approximated by a weighted block sum of the data through a uniform Bahadur representation. Secondly, the weighted block sum convolved by i.i.d. auxiliary standard normal variables takes the form of a robust multiplier bootstrap (Zhou, 2013), which can consistently mimic the probabilistic behavior of their partial sum processes. We will show that the easy-to-implement bootstrap simulates the probabilistic behavior of the M-estimators consistently under complex temporal dynamics. Compared to the bootstrap methods developed in Karmakar et al. (2022), Wu and Zhou (2017) and Friedrich and Lin (2022), the self-convolved bootstrap enjoys broader applicability yet maintains a simple implementation.
Facilitating the self-convolved bootstrap, our framework utilizes the cumulative regression function (CRF)
(2) |
instead of for inference, driven by two main reasons. First, as is suggested in Theorem 1, the CRF can be estimated at a parametric rate of . Such merits of getting convergence via integration have been demonstrated by Hall and Marron (1987), Bickel and Ritov (1988), Huang and Fan (1999) and many others. Recently, Mies (2023) employed the method to estimate parameters of locally stationary time series, leading to a functional central limit theorem. Cai and Juhl (2023) also obtained a similar result regarding rolling regression estimators, where the error process is assumed to be a martingale difference. Although the results in the existing literature have gradually adapted from the i.i.d. case to a broader non-stationary time series setup, they generally rely on a closed-form statistic, e.g., the least square estimator, and therefore do not apply to the M-regression setting. To our knowledge, no comparable research has been conducted on uniform nonparametric inference of the CRF under the time-varying M-regression framework. The second reason, possibly more salient, pertains to the complexity of the temporal dynamics considered. Specifically, it is difficult to consistently estimate the (asymptotic) distributional behavior of the M-estimators around the time points where and/or experience abrupt changes. As a result, directly inferring the M-estimators of simultaneously over time becomes challenging. Nevertheless, we discover that aggregated M-estimators of the CRF converge to a Gaussian process uniformly over time under mild conditions, which can be effectively approximated by the self-convolved bootstrap mentioned earlier. Consequently, it is simpler and more accurate to make simultaneous inferences of than under complex temporal dynamics. As far as we know, no previous literature has used the method of integration or aggregation to alleviate the effect of complex dependence and heteroscedasticity in regression analysis.
The self-convolved bootstrap facilitated by the CRF addresses different types of tests using a unified asymptotic result, which ventures into a new avenue compared to the conventional methods. To elaborate, we consider three types of hypothesis tests: Exact Function Tests, Lack-of-fit Tests, and Qualitative Tests. The first type, Exact Function Test (EFT), takes the form of for some specific , which, for example, enables one to check the significance of the variables by letting . Secondly, the Lack-of-fit Test (LOFT), also referred to as diagnostic test (Stute, 1997), tackles the circumstance where the null hypothesis is a parametric family while the alternative remains nonparametric (He and Zhu, 2003). This includes, but is not limited to, the tests of constancy and linearity of the function versus a general smooth alternative. An example in the realm of econometrics is (Chen and Hong, 2012), where the LOFT is used to tackle the long-standing problem of detecting structural changes in economic relationships. Apart from the aforementioned two types of tests, one may also want to check qualitative beliefs on functions where the null hypothesis is a relatively general nonparametric class of functions (Komarova and Hidalgo, 2020). For example, many economic models use certain shape restrictions (e.g., monotonicity, convexity) as plausible restrictions (Chetverikov, 2019; Fang and Seo, 2021). While those prespecified qualitative shape assumptions facilitate the inference and enhance performance (Friedman and Tibshirani, 1984; Matzkin, 1994), wrong conclusions can be drawn when these assumptions are not satisfied. The necessity to conduct Qualitative Tests (QT), therefore, becomes axiomatic.
Historically in the context of regression, much literature has focused on the Lack-of-fit Test on the regression mean function, serving as a key component in model checking. One popular approach is to estimate the parametric and nonparametric curves separately and compute their discrepancies as a natural test statistic, as seen in Hardle and Mammen (1993) and Hong and White (1995). Similarly, Chen and Hong (2012) proposed a generalized Hausman test for checking parameter stability in time series models, applicable to stationary data under a framework. Alternatively, another distinguished direction is to study the empirical process based on residuals; see Stute et al. (1998); Koul and Stute (1999). Recently, Mies (2023) utilized the integration technique for change-point detection of local parameters. In comparison, our current work is based on a new self-convolved bootstrap technique and applies to a broader range of tests and a more general M-regression setting. Different from the LOFT where the null hypothesis forms a parametric family, qualitative hypotheses often include shape constraints that are more complicated to deal with. To test the monotonicity of a regression curve, Bowman et al. (1998) exploited Silverman (1981)’s idea of critical bandwidth. Gijbels et al. (2000) and Ghosal et al. (2000) formulated monotonicity of the regression curve as the concordance of and . As a distinctive approach, Durot (2003) transformed the test of monotonicity into testing whether the integral coincides with its least concave majorant. A localized version of the test was later elaborated by Akakpo et al. (2014). Other contributions in nonparametric qualitative tests include Hall and Heckman (2000); I. Gijbels and Verhasselt (2017); Komarova and Hidalgo (2020); Fang and Seo (2021), among others.
The rest of the paper is structured as follows. Section 2 introduces the piece-wise locally stationary time series, dependence measures, and local linear M-estimators. The estimator of the CRF is proposed, with its asymptotic results stated in Theorem 1. Section 3 presents the self-convolved bootstrap. In Section 4, three types of hypothesis tests are discussed, where detailed algorithms and their asymptotic behaviors are provided. The finite-sample performance of our method is demonstrated in Section 5, followed by empirical illustrations using global climate data and Microsoft stock return in Section 6. Regularity conditions of the model and auxiliary theoretical results about Bahadur representation are given in Section 7. Detailed proofs are provided in the supplementary material.
2 Preliminary
We start by introducing some notations. For an -dimensional (random) vector , let be the Euclidean norm and . Let be the indicator function and we denote the weak convergence by .
2.1 Non-stationary time series models
The following piece-wise locally stationary (PLS) time series models (Zhou, 2013) are used to describe the complex temporal dynamics of the covariates and errors. For a sequence of i.i.d. random variables , let be an i.i.d. copy of . Define , and .
Definition 1.
We say that is a -dimensional piece-wise locally stationary time series with break points and filtration if
(3) |
where , and , are possible nonlinear filters.
The breakpoints are assumed to be fixed but unknown, and the number of breaks is assumed to be bounded. The PLS process can capture a broad class non-stationary behavior in practice because it allows the underlying data-generating mechanism to evolve smoothly between breakpoints (provided that the filters are smooth in ) and permits the latter mechanism to change abruptly from to at the breakpoint . Let be the index such that . Then (3) is equivalent to . For and , we define the dependence measures
(4) |
where is the norm. Write . Let if . Intuitively, measures uniformly the changes in a dynamic system’s output when the innovations of the system -steps ahead are replaced with i.i.d. copies. Therefore the decay speed of as a function of measures the strength of the system’s temporal dependence. We refer the readers to (Zhou, 2013) for more discussions and examples of the PLS time series models and the associated dependence measures.
Here we consider filtrations and , where and are independent. Assume piece-wise locally stationary time series , . Observe that and can be dependent as the innovations which generate are also involved in ’s generating process. On the other hand, the factors influencing but are unrelated to are captured in . Without loss of generality, assume and share the same breakpoints .
2.2 Aggregated local estimation of the CRF
Throughout this paper, we assume that the regression function . For any , since in a small neighborhood of , we define the preliminary local linear M-estimates of and by
(7) |
where , is a bandwidth satisfying , and is a convex loss function with left derivative . In this paper, we also assume robustness of the loss functions in the sense that for and some positive constants . is a kernel function satisfying and is collection of symmetric and density functions with support .
Recall the CRF defined in (2), then the regression function can be retrieved via
where and represent the left and right derivative of , respectively.
As is discussed in Section 7, the M-estimator obtained in (7) involves a bias term regarding . In order to proceed without estimating , we shall employ a Jackknife bias-corrected estimator denoted as , where
(8) |
The implementation of the bias-correction procedure is asymptotically equivalent to using the second-order kernel . In this way, the bias of is asymptotically negligible under mild conditions. Based on the Jackknife bias-corrected estimator, we propose to estimate via
(9) |
and . For any , let be the linear interpolation of the sequence . We omit the subscript of hereafter, whenever no confusion caused.
Let be a fixed , full rank matrix, we are interested in testing the dynamic pattern of , which includes any linear combination of . As a result, for the CRF , its estimator can be obtained via . Based on the Bahadur representation presented in Section 7, the following theorem establishes a -consistent Gaussian weak convergence result for the proposed aggregated M-estimator:
3 The Self-convolved Bootstrap
The Gaussian process expressed in Theorem 1 provides a theoretical foundation for the simultaneous inference of , but the implementation is still unaccomplished due to its complex covariance structure. To circumvent the problem, we propose a self-convolved bootstrap method to mimic the behavior of the Gaussian process .
Let , be i.i.d. standard normal random variables independent of . For a given bandwidth , define the bootstrap process as the linear interpolation of , where
The self-convolved bootstrap only requires the convolution of local M-regression estimates with i.i.d. standard normal random variables . The heuristics stem from two fundamental observations. Firstly, the Bahadur representation of in Section 7 suggests that can be expressed as a weighted block sum of , where we take the second-order difference as a bias-correction technique. Secondly, Zhou (2013) demonstrated that for a broad range of non-stationary time series, progressive convolutions of their block sums with i.i.d. standard normal random variables can consistently mimic the joint probabilistic behavior of their partial sum processes. Observe that represents the limiting behavior of a weighted partial sum process of . By leveraging these observations, the self-convolved bootstrap achieves an accurate simulation of the limiting Gaussian process of the CRF and retains a concise form for a broad range of M-estimators, utilizing just a single tuning parameter.
In contrast, it is difficult to make inference of directly based on , as the density-like quantity involved in the distributional behavior of is hard to estimate in practice due to possible discontinuities and the no-trivial task of choosing smoothing parameters. As a result, the CRF considered in this paper, combined with the proposed self-convolved bootstrap, enables us to conduct inference more efficiently and accurately.
To quantify the consistency of , we derive a direct comparison of the distribution between and in the following theorem. First, we define the -dimensional vector with
and similarly define the -dimensional vector with
Recall is defined in Theorem 1. Furthermore, we define
as a measure of the difference in covariance structure between the bootstrapped process and the limiting Gaussian process.
Theorem 2.
Suppose assumptions (A1)-(A7) in Section 7 hold, , and , then we have . Define the sequence of events where is a sequence diverging at an arbitrarily slow rate. Then . On the event , for any , we have
(10) | ||||
where with some finite constant that does not depend on .
Remark 1.
Theorem 2 guarantees that the conditional behavior of the bootstrap process consistently mimics the distribution of the limiting process in norm. The restricted range is to make the Gaussian approximation valid under the circumstance that the variances of have no lower bounds. Note that and the upper bound in (10) converge to as . Accordingly, for a given and when is sufficiently large, will be dominated by the -th percentile of , thereby ensuring that the results in Theorem 2 asymptotically validate the bootstrap procedure. Furthermore, under the conditions of Theorem 2, converges to , hence the consistency of the bootstrap demonstrated on is asymptotically uniform on .
4 Applications to Hypothesis Testing
In principle, hypotheses regarding the regression function can be expressed equivalently in terms of the CRF . In this Section, we shall explore the application of the CRF and the self-convolved bootstrap to three general classes of hypothesis tests of the M-regression: Exact Function Test (EFT), Lack-of-fit Test (LOFT), and Qualitative Test (QT).
4.1 Exact Function Test
The Exact Function Test amounts to testing , where is a known function. It can be reformulated as . Define
(11) |
For any prespecified level , our goal is to find the critical value , such that asymptotically. The detailed procedures for conducting EFT are given in Algorithm 1. Propositions 1 and 2 in Section 4.2 validate the latter algorithm theoretically.
4.2 Lack-of-fit Test
The Lack-of-fit Test investigates whether belongs to a given parametric family; i.e., for a given family of functions with unknown . Under , oftentimes the unknown can be expressed (solved) in terms of for some given points . See Section 4.2.1 for a detailed example of testing whether is a polynomial of a given degree. In this case, the null hypothesis is written as
(12) |
where is a finite integer. Assumptions on are discussed in Theorem 3. Consider
(13) |
then the critical value can be obtained as shown in Algorithm 2.
4.2.1 Polynomial Test
The polynomial test aims to verify whether the coefficients are -th order polynomials for a fixed . For instance, allows one to test if for some unknown . Therefore under , and by setting . Hence can be written equivalently as . This special case of is closely related to CUSUM tests in structural change detection and has been widely discussed in the statistics literature recently; see for instance Mies (2023) and the references therein.
In general, testing whether is a -th order polynomial is equivalent to testing . Choosing and , we solve via the system of linear equations
By plugging in the solutions to ’s, is of the form (12), and the LOFT can be applied.
Theorem 3 lays a theoretical foundation for the Lack-of-fit Test and we summarise by listing the procedures in Algorithm 2.
Theorem 3.
Under the conditions of Theorem 1 and 2, assume is continuously differentiable with regard to . Denote the partial derivative of with regard to as , we assume are Lipschitz continuous for uniformly over , and that . Let be a Gaussian process defined in Theorem 2, , , . Let and , then
(14) |
Additionally, let , then on a sequence of events with , we have
(15) |
where as defined in Theorem 2.
It’s straightforward to see that the EFT addressed in Section 4.1 is a special case of the LOFT in Section 4.2. Therefore, we state the asymptotic rejection rate of the proposed algorithm of LOFT in Proposition 1 and 2, which also applies to EFT.
Proposition 1.
Suppose the conditions of Theorem 3 hold, under the null hypothesis, for a given significance level , the rejection rate of the proposed LOFT satisfies
Proposition 2.
Suppose the conditions of Theorem 3 hold, under the alternative hypothesis of LOFT, if , then
Propositions 1 and 2 demonstrate the accuracy and power of our proposed testing framework. In particular, both EFT and LOFT are asymptotically accurate under the null hypothesis, while attaining asymptotic power of under local alternatives deviating from with a distance much greater than . These results highlight the asymptotic reliability and robustness of our methodology.
4.3 Qualitative Test
In this subsection, we are interested in testing qualitative hypotheses on M-regression coefficients in the sense that the null hypothesis is nonparametric and can be written as
(16) |
where is a non-empty subset of . Based on the cumulative M-estimator , consider the optimization problem:
(17) | |||
(18) |
Suppose that there exists a projection under the norm on for any . Then the solution to (17), denoted by , is the projection of onto . Define
(19) |
we observe that becomes a natural indicator of the distance between and , as guaranteed by the uniform consistency of . Therefore, in principle, should be small under and large under the alternative. The critical value of the test can be easily obtained by the self-convolved bootstrap, similar to the implementation of the EFT in Algorithm 1. Detailed steps are given in Algorithm 3.
Proposition 3 stated below shows the capability of our approach to asymptotically control the type I error under the specified significance level. Meanwhile, it attains an asymptotic power of when the CRF deviates from the qualitative hypothesis with a distance dominating , as illustrated in Proposition 4.
Proposition 3.
Proposition 4.
4.3.1 Shape Test
One instance of such Qualitative Tests involves testing the shape of . For example, an attempt is to check the monotonicity of , which is equivalent to testing
As a straightforward extension, we may consider the null hypothesis . Particularly, corresponds to nonnegativity, corresponds to the null hypothesis that is a monotonically increasing function, and refers to the convexity of . Clearly, such differential-type shape constraints belong to the realm of QT, by letting the feasible region .
In practice, to tackle the optimization problem in the finite sample, the functional optimization problem is reduced to that regarding discrete vectors. Let for a slight gain of simplicity, we denote the -th differential matrix as and let , then (17) can be formulated as follows:
(20) | |||
(21) |
Further, as is discussed in Knight (2017), the projection under norm in (20) can be expressed as the solution to a linear program, which can be obtained easily.
5 Simulation Studies
This section utilizes Monte Carlo simulations to demonstrate the finite sample performance of the proposed method. All three hypotheses stated in Section 4 are considered.
Let , , . For , define , , where , and , , where , , and , , and are i.i.d standard normals. Let , , and be the -th quantile of . Consider the following models:
-
I
. Covariates and errors are independent.
-
II
when using quantile loss, and when using quadratic loss. In this model, covariates and errors are correlated.
Case I and II share the same but have different errors. By investigating both cases, we can illustrate the applicability of our method under error-covariate independence and dependence. Throughout this section, we use the Epanechnikov kernel. The number of replications is , and the bootstrap sample size is .
5.1 Bandwidth Selection
The framework involves smoothing parameters and , where is used for estimating , while is used for constructing the bootstrapped process .
To select the appropriate bandwidth , we recommend using Leave-One-Out Cross-Validation (LOOCV) to select from a predetermined grid . Denote as the estimator of using bandwidth and the training set , then we select via the minimum average error:
(22) |
Ideally by the proof of Theorem 4 and a similar argument from Wu and Zhou (2017) regarding quantile regression. Based on Chen and Hong (2012) where was chosen as a rule-of-thumb approach, a feasible grid of can be selected around it, say equispaced points within the interval . The proposed procedure works reasonably well in our simulation studies.
Regarding the bandwidth , we can use as an easy implementation. For refinements, we recommend the extended Minimum Volatility (MV) method suggested by Zhou and Wu (2010), which is an extension of the minimal volatility method proposed in Politis et al. (1999). Suppose we are to get bootstrap samples of , where is a matrix. Let the candidates of possible bandwidths be . For , define and . We further define over interval by letting if and if . Based on this, we then compute , respectively. For a positive integral , say , define the integrated standard error (ise) as
where represents the entry of the matrix . Further, define
then is chosen as if .
5.2 Exact Function Test
We generate data under the setting of Case I and Case II, focusing on 2 null hypotheses and for both cases. Sample sizes and are considered, and 4 different losses are applied: quadratic loss, quantile losses with , and .
Table 1 and 2 demonstrate the type I error of EFT with nominal levels . Using the bandwidth selected via the procedures described in Section 5.1, a reasonable type I error can be observed for both cases. We also present results using and for sensitivity analysis. The results show a good approximation of the nominal level, as long as the bandwidths are not very far from the ones chosen by the proposed procedures.
To demonstrate the power of EFT, we generate data under the mechanism , while and remain unchanged from the original setting. Let be a positive constant, then the power of the test should increase with . For comparison, we also replicate the SCB test proposed by Wu and Zhou (2017) under the same setting. For both Case I and Case II, we present the power when using quantile loss with and sample size . The nominal level used here is . Figure 2 and Figure 2 show that, in both cases, the power of both methods goes to as increases from to . As expected, the power of EFT converges faster to than the other method, which is guaranteed by the convergence rate theoretically.


5.3 Lack-of-fit Test
Case I Case II Case I Case II Case I Case II Case I Case II
Case I Case II Case I Case II Case I Case II Case I Case II
For illustration of the Lack-of-fit Test, we focus on in Case I and II. Under both settings, satisfies constancy and linearity, thus the two types of Lack-of-fit Tests can be conducted to demonstrate the accuracy of our method. Following the procedures in Section 4.2, we obtain the simulated type I errors for the 4 losses and sample size . The optimal bandwidths are the same as those selected in Section 5.2 in that we are considering the same settings. Table 3 and Table 4 show that the accuracy is quite satisfactory and is not very sensitive to the selection of the bandwidths.
Under the setting of Case II, we also conducted the power analysis under 2 scenarios: (1) ; (2) , where the first scenario allows us to investigate constancy while the other enables the testing of linearity. The sample size is chosen to be and the nominal level is . For each test, power curves regarding the 4 different losses are simulated and compared in Figure 4 and Figure 4. The quadratic loss is the most powerful and the quantile loss with comes second. As for the extreme quantiles and , the convergence is relatively slower.


We then compare our method with Chen and Hong (2012) to test the constancy of the coefficients. Since their method assumes stationary covariates and errors, we consider the following stationary model for a more fair comparison:
-
III
, where , , and are i.i.d standard normal random variables independent of each other.

Note that when , the model corresponds to constant coefficients , and for the joint constancy test of , we expect to see an increasing rejection rate as increases. We generate the data under and use loss to implement our method. The nominal level is chosen as . Figure 5 indicates that our method has comparable accuracy and power with Chen and Hong (2012)’s method, where ‘H-homo’ and ‘H-het’ represent the homoscedastic and heteroscedasticity-robust version of the generalized Hausman test proposed in Chen and Hong (2012). Considering that their method is designed for the constancy test of -regression models and applies to stationary variables, our Lack-of-fit Test enjoys more flexibility and broader applicability, with a faster convergence rate theoretically.
5.4 Qualitative Test
For the Qualitative Test, exemplarily, we mainly focus on the hypothesis of monotonicity. Under the setting of Case II with , we generate the data using the following mechanism: for , and we expect to see an increasing power curve as increases.
Applying a similar projection idea to the SCB test in Wu and Zhou (2017), we also manage to obtain the power curve regarding the monotonicity test for quantile loss with for their method. Figure 7 provides the results with nominal level , demonstrating the performance of our method regarding 4 losses and the comparison with the SCB test. The power curves of quadratic loss and quantile loss with converge faster than the extreme quantiles, as we expect. Figure 7 also shows that, under the same quantile loss with , our qualitative test is more powerful than the SCB test with projection. This is due to the fact that our CRF-based test can detect local alternatives of the order while the SCB in Wu and Zhou (2017) are sensitive to those of the order with bandwidth .


6 Real Data Illustrations
6.1 Global Temperature Anomalies
In this section, we delve into monthly global temperature anomalies from 1882.1 to 2005.12 (available at HadCRUT5 dataset), investigating the anthropogenic warming trend and the time-varying relationship between the anomalies and potential factors. The initial candidate factors include lags of multivariate ENSO index (MEI), total solar irradiance (TSI), aerosol optical depth (AOD) and Atlantic multidecadal oscillation (AMO). These factors are typical regressors in existing studies, where multivariate linear regression is commonly employed to filter out the fluctuations caused by the factors and to reveal the underlying anthropogenic warming (Zhou and Tung, 2013). The authors believe that the warming trend has been remarkably steady since the mid-twentieth century. Foster and Rahmstorf (2011) also applied a similar approach to the global temperature anomalies from 1979 to 2010, where the deduced warming rate was also found to be steady over the investigated time interval. All previous analyses assume a stationary structure of the errors and constant coefficients over time, in addition to which a linear trend of anthropogenic warming is also hypothesized. Nevertheless, by utilizing the change-point detection method in (Zhou, 2013), nonstationarity of the residuals has been detected after fitting the multivariate linear model, suggesting the necessity to use a more general model that allows for the nonstationarity of the time series. Furthermore, no rigorous proof has been provided in the existing literature to draw convincing conclusions about the shape of the warming trend.
We hereby apply our method to the aforementioned time series. Following the steps stated in Section 4, a full model consisting of all the factors is first constructed and analyzed, then a backward variable selection is implemented based on our EFT. Most covariates are therefore removed due to insignificance or multicollinearity. The final model includes the intercept, MEI, and the 6-month lag of MEI under loss, quantile loss with and , while the quantile regression model is composed of the intercept, 2-month and 6-month lags of MEI, and 5-month lag of AOD. All models involve the intercept and corresponding lags of MEI, enabling us to conduct tests regarding the anthropogenic warming trend and the coefficient of the ENSO effect. The outcome agrees with the fact that ENSO has conventionally been recognized as a leading contributor to global temperature fluctuations (Trenberth et al., 2002; Foster and Rahmstorf, 2011). The 6-month lag of MEI chosen in each model also aligns with the belief that El Niño warms up the global temperature with a lag of 6 months (Trenberth et al., 2002). As a distinctive predictor, the 5-month lag of AOD is selected under the quantile, emphasizing the influence of volcano activities on extreme high temperatures. The model selection result coincides with vast literature in climatology, where ENSO and volcano events are identified as significant sources of variance in global temperature, compared to solar activities and other factors (Lean and Rind, 2008; Foster and Rahmstorf, 2011). On the other hand, our conclusion is drawn under general non-stationary assumptions which could be more reliable.
Test Statistic p-val Test Statistic p-val Test Statistic p-val Test Statistic p-val 0.132 1.355 1 0.853 0.69 0.007 0.217 0.702 0.986 0.221 0.499 0.842 0.157 1.02 1 0.886 0.647 0.002 0.401 0.711 0.613 0.4 0.528 0.244 0.143 1.464 1 0.877 0.752 0.013 0.333 0.729 0.804 0.216 0.643 0.978 0.293 1.495 0.999 0.954 0.733 0.001 0.322 0.527 0.468 0.431 0.496 0.121


We are interested in the hypotheses on the shape of the anthropogenic trend and the coefficients regarding MEI and AOD. It is commonly believed that the anthropogenic trend is increasing, and we can further check whether it is steady over time, which is equivalent to testing whether the intercept is a linear function. As for the coefficients of MEI and AOD, we examine whether they are time-invariant, which amounts to testing the constancy. The test results are summarized in Table 5, which demonstrates similar results under different losses. For the shape of the intercept, the positive monotonicity is not rejected, while the linearity is rejected at significance level. This strengthens the belief of discernible human influences on global warming and also suggests that the warming rate is not steady from 1882 to 2005. The non-uniform rate is consistent with the conclusion from IPCC (2007), where global warming is observed to accelerate until 2005 after a cooling period in the 1960s and 1970s. This can also be illustrated by Figure 9, with a relatively flat trend from 1960 to 1970, and a steep increase afterward. While a similar pattern can be detected by solely analyzing the temperature anomalies, our approach rigorously assesses the anthropogenic fluctuations by separating them from the natural ones, making the conclusion more reliable and statistically verifiable. As for the coefficients of MEI and its lags, we fail to reject the null hypothesis of constancy at significance level. This gives us the insight that the overall ENSO effect on global temperature is likely to be steady from 1882 to 2005.
To represent the temperature change induced by volcano activities, we demonstrate the influence of the 5-month lag of AOD in Figure 9, which is computed by multiplying the estimated coefficient and the factor itself. Unlike the traditional multivariate linear regression where only the overall negative effect of AOD can be statistically detected, Figure 9 reveals the effect of AOD as a combination of heating and long-term cooling. This finding complies with the mechanism of volcano activities, where volcano eruptions can cool the surface due to an increased aerosol loading in the stratosphere that scatters solar radiation back to space. Simultaneously, these eruptions can lead to regional heating, particularly evident during winters in the Northern Hemisphere (Robock, 2000; Christiansen, 2008). Such unstable and complicated impact is also confirmed using our Lack-of-fit Test, where the -values of the test of constancy and linearity turn out to be 0.029 and 0.002, respectively.
6.2 Microsoft Stock Return
As another application, we incorporate the time-varying M-regression framework with the Fama-French 5-factor (FF5) model to study the monthly return of Microsoft stock from 1986.5 to 2023.10. Initially proposed by Fama and French (2015), the FF5 model aims to capture the size, value, profitability, and investment patterns in average stock returns. There have been subsequent discussions on the time-stability of the regression coefficients of FF5. Racicot et al. (2019) accounted for the time-varying nature of the model using a GMM approach, presenting evidence against a static market systematic risk parameter . Similarly, the time-varying parameters were examined by Horváth and Wang (2021), especially during selected events such as the dot-com bubble around 2000, the 2008 financial crisis, and the COVID-19 outbreak. Noda (2022) further argued that the parameters change over time, exhibiting distinct patterns of change across different countries and regions.
In the meantime, a quantile version of the FF5 model has been discussed in order to analyze the extremes of the stock returns (Allen and Powell, 2011). Emerging as an alternative to least squares estimation, quantile regression not only looks beyond the conditional mean, but also alleviates some of the statistical problems which plague model including errors-in-variables, omitted variables bias, sensitivity to outliers, and non-normal error distributions (Barnes and Hughes, 2002).
While the aforementioned alternatives have been primarily studied in a parallel way in the literature, both directions can be accommodated under our framework. Denoting the monthly return of Microsoft stock as , the model is written as
(23) |
where represents the risk-free rate, is the monthly return on the value-weight (VW) market portfolio, and are size factor, value factor, profitability factor, and investment factor, respectively. The data are accessible from Professor French’s website, and we refer to Fama and French (2015) for a detailed explanation of the variables.
We estimate the regression coefficients under 4 losses: loss, quantile losses with , , and . Using the test proposed in Zhou (2013), second-order stationarity of the residuals is rejected under all losses with -value and , suggesting the existence of time non-stationarity in the underlying data. Additionally, following the procedures of LOFT as presented in Section 4.2, a joint constancy test is applied to the regression coefficients, which yields a -value of , , , under the 4 losses, respectively. These results confirm the necessity of utilizing a time-varying coefficient model, and the original FF5 model that features static parameters might be misspecified.


We further focus on the behavior of the estimators under loss, as the intercept and the market coefficient have been widely recognized as important indicators for investment. Jensen (1968) proposed the intercept as a measure of performance, where a positive value suggests that the stock tends to outperform the average market. Meanwhile, is regarded as the measure of systematic risk (Sharpe, 1964). As is indicated in Figure 11, the estimated intercept is greater than most of the time, which aligns with the result of non-negativity test with a -value equal to . When it comes to the systematic market risk , fluctuations and extreme values have been observed from Figure 11, especially during the dot-cum bubble around 2000. Additionally, is smaller than most of the time, indicating a relatively lower risk of Microsoft stock compared to the average market.
Intercept SMB HML RMW CMA 0 0 0 0 0.482 0.028 0 0.029 0 0 0.976 0 0 0 0 0 0.368 0.375 0 0 0.68 0.002 0.075 0.428
While it remains unclear how to interpret the quantile coefficients in an economic context due to scarce literature on the quantile FF5 model, we gain some insights into the predictors by conducting significance tests on the regression coefficients. Presented in Table 6 are -values of the significance tests. The intercept and the market factor serve as significant predictors under all losses. Similarly, the value factor HML continues to be significant as well. The size factor SMB plays its role under loss and loss with and , but becomes insignificant under the higher quantile . This suggests that the higher quantile of Microsoft stock return is no longer attributable to the size factor. An opposite pattern can be observed for the profitability factor RMW, where it remains insignificant while starting to make a difference under higher quantile . Finally, the investment factor CMA might contribute to the stock return under the loss and lower quantile , but is considered insignificant otherwise with or .
7 Assumptions and Auxiliary Theoretical Results
Let , and define
(26) |
where . Compare this with equation (7), simple calculations yield that
(27) |
Let be a vector, . Write , . We assume .
Remark 2.
The assumption controls the magnitude of , which is used to establish the Bahadur representation. If is continuous, this assumption is automatically satisfied as . While in the case where is discontinuous, the solution may not always exist. A key example of such discontinuous arises in quantile regression (Wu, 2007, Remark 4) and the proof of it for quantile regression is provided in Lemma LABEL:lem:4 of the supplementary.
7.1 Model Assumptions
For , define for ,
-
(A1)
For , assume that
(28) (29) where are measurable r.v. with finite fourth moments. We also require that for , is stochastically Lipschitz continuous for , that is, , s.t. , ,
(30) -
(A2)
Assume for the covariate process, , s.t. . In addition, for some constant . For all , for some constant . Recall the definition of in Section 2.1.
-
(A3)
For , and any integer , define
We assume that , s.t. for some constant .
-
(A4)
Define , we require for all and any -dimensional vector that
and a.s. for for some . Let be the smallest eigenvalue of , assume that . Let be the smallest eigenvalue of if . We require that i) for some constant , and a.s. for some constant ; ii) for some positive constant ; iii) for some constant , and a.s. for some constant , where .
-
(A5)
Assume for some , where if , . Also assume for the error process that for all and . Further assume for all , for some constant .
-
(A6)
Define the long-run covariance function of
Let . Assume that the minimum eigenvalue of is bounded away from 0 on .
-
(A7)
Define . Assume that, for sufficiently large , there exists , such that holds for .
Here are some insights on the above regularity conditions. Assumption (A1) guarantees the smoothness of the conditional loss functions. (28) holds if , which is satisfied by least square regression, and can be achieved by quantile regression under some mild constraints. A sufficient condition for (28) is provided in Wu and Zhou (2018) by virtue of the robustness of loss functions. (A2) requires that the covariate process is stochastic Lipschitz continuous with geometrically decaying dependence measures. The moment requirement of in (A2) can also be significantly relaxed when is bounded or light-tailed. (A3) controls the dependence measures of the derivatives of the loss functions’ conditional expectations, and is easy to verify for a large class of non-stationary processes, see Wu and Zhou (2018) for further details. (A4) plays a key role in the consistency of by endowing smoothness on to some extent. These conditions altogether enable Bahadur representation of the estimators. Condition (A6) means the time-varying long-run covariance matrices of are non-degenerate on , which is mild in most applications. Condition (A7) assumes Lipschitz continuity of . Observe that is the integration of and hence the assumption is mild even though both and may experience jumps. (A7) is utilized to guarantee the tightness of the estimators, see Lemma LABEL:lem:7 in the supplementary material for more details.
7.2 Uniform Bahadur Representation
This section establishes uniform Bahadur representations for the local linear M-estimators and the estimated CRF, by which the limiting distribution of the estimators can be derived in conjunction with some Gaussian approximation results.
Define a matrix , where is defined in (A4) and for some positive integer .
Theorem 4.
Suppose (A1)-(A4), , and for some positive constant . Let , and , then
(31) |
where .
The Bahadur representation in (31) involves a bias term involving . Recall the Jackknife bias-corrected M-estimator defined in (8) and (9). We extend the above Bahadur representation result to the estimated CRF as follows.
Corollary 1.
Suppose the conditions of Theorem 4 hold, then for any , full rank matrix , ,
SUPPLEMENTARY MATERIAL
- Theoretical proof:
-
This file contains detailed proof of all theorems. (.pdf file)
References
- Akakpo et al. (2014) Akakpo, N., F. Balabdaoui, and C. Durot (2014). Testing monotonicity via local least concave majorants. Bernoulli 20(2), 514 – 544.
- Allen and Powell (2011) Allen, D. E. and S. R. Powell (2011). Asset Pricing, the Fama—French Factor Model and the Implications of Quantile-Regression Analysis, pp. 176–193. London: Palgrave Macmillan UK.
- Barnes and Hughes (2002) Barnes, M. L. and A. T. W. Hughes (2002, November). A quantile regression analysis of the cross section of stock market returns. (458522).
- Bickel and Ritov (1988) Bickel, P. J. and Y. Ritov (1988). Estimating integrated squared density derivatives: sharp best order of convergence estimates. Sankhya A (1961-2002) 50(3), 381–393.
- Bowman et al. (1998) Bowman, A. W., M. C. Jones, and I. Gijbels (1998). Testing monotonicity of regression. J. Comput. Graph. Stat. 7(4), 489–500.
- Cai and Juhl (2023) Cai, Z. and T. Juhl (2023). The distribution of rolling regression estimators. Journal of Econometrics 235(2), 1447–1463.
- Chen and Hong (2012) Chen, B. and Y. Hong (2012). Testing for smooth structural changes in time series models via nonparametric regression. Econometrica 80(3), 1157–1183.
- Chetverikov (2019) Chetverikov, D. (2019). Testing regression monotonicity in econometric models. Econometric Theory 35(4), 729–776.
- Christiansen (2008) Christiansen, B. (2008). Volcanic eruptions, large-scale modes in the northern hemisphere, and the el niño–southern oscillation. Journal of Climate 21(5), 910–922.
- Dahlhaus et al. (2019) Dahlhaus, R., S. Richter, and W. B. Wu (2019). Towards a general theory for nonlinear locally stationary processes. Bernoulli 25(2), 1013 – 1044.
- Dette et al. (2011) Dette, H., P. Preuß, and M. Vetter (2011). A measure of stationarity in locally stationary processes with applications to testing. J. Am. Statist. Ass. 106(495), 1113–1124.
- Durot (2003) Durot, C. (2003). A Kolmogorov-type test for monotonicity of regression. Statist. Probab. Lett. 63(4), 425–433.
- Ebi et al. (2021) Ebi, K. L., J. Vanos, J. W. Baldwin, J. E. Bell, D. M. Hondula, N. A. Errett, K. Hayes, C. E. Reid, S. Saha, J. Spector, and P. Berry (2021). Extreme weather and climate change: population health and health system implications. Annu. Rev. Public Health 42(1), 293–315.
- Fama and French (2015) Fama, E. F. and K. R. French (2015). A five-factor asset pricing model. Journal of Financial Economics 116(1), 1–22.
- Fan and Zhang (2000) Fan, J. and W. Zhang (2000). Simultaneous confidence bands and hypothesis testing in varying-coefficient models. Scandinavian Journal of Statistics 27(4), 715–731.
- Fang and Seo (2021) Fang, Z. and J. Seo (2021). A projection framework for testing shape restrictions that form convex cones. Econometrica 89(5), 2439–2458.
- Foster and Rahmstorf (2011) Foster, G. and S. Rahmstorf (2011, dec). Global temperature evolution 1979–2010. Environ. Res. Lett. 6(4), 044022.
- Francq and Sucarrat (2023) Francq, C. and G. Sucarrat (2023). Volatility estimation when the zero-process is nonstationary. J. Bus. Econ. Statist. 41(1), 53–66.
- Friedman and Tibshirani (1984) Friedman, J. and R. Tibshirani (1984). The monotone smoothing of scatterplots. Technometrics 26(3), 243–250.
- Friedrich and Lin (2022) Friedrich, M. and Y. Lin (2022). Sieve bootstrap inference for linear time-varying coefficient models. Journal of Econometrics, 105345.
- Ghosal et al. (2000) Ghosal, S., A. Sen, and A. W. van der Vaart (2000). Testing monotonicity of regression. Ann. Statist. 28(4), 1054–1082.
- Gijbels et al. (2000) Gijbels, I., P. Hall, M. C. Jones, and I. Koch (2000). Tests for monotonicity of a regression mean with guaranteed level. Biometrika 87(3), 663–673.
- Hall and Heckman (2000) Hall, P. and N. E. Heckman (2000). Testing for monotonicity of a regression mean by calibrating for linear functions. Ann. Statist. 28(1), 20–39.
- Hall and Marron (1987) Hall, P. and J. Marron (1987). Estimation of integrated squared density derivatives. Statist. Probab. Lett. 6(2), 109–115.
- Hardle and Mammen (1993) Hardle, W. and E. Mammen (1993). Comparing nonparametric versus parametric regression fits. Ann. Statist. 21(4), 1926 – 1947.
- Hastie and Tibshirani (1993) Hastie, T. and R. Tibshirani (1993). Varying-coefficient models. J. R. Statist. Soc. B 55(4), 757–796.
- He and Zhu (2003) He, X. and L.-X. Zhu (2003). A lack-of-fit test for quantile regression. J. Am. Statist. Ass. 98(464), 1013–1022.
- Hong and White (1995) Hong, Y. and H. White (1995). Consistent specification testing via nonparametric series regression. Econometrica 63(5), 1133–1159.
- Hoover et al. (1998) Hoover, D. R., J. A. Rice, C. O. Wu, and L.-P. Yang (1998). Nonparametric smoothing estimates of time-varying coefficient models with longitudinal data. Biometrika 85(4), 809–822.
- Horváth and Wang (2021) Horváth, D. and Y.-L. Wang (2021). The examination of fama-french model during the covid-19. Finance Research Letters 41, 101848.
- Huang and Fan (1999) Huang, L.-S. and J. Fan (1999). Nonparametric estimation of quadratic regression functionals. Bernoulli 5(5), 927 – 949.
- Huber (1964) Huber, P. J. (1964). Robust estimation of a location parameter. Ann. Math. Stat. 35(1), 73 – 101.
- Huber (1973) Huber, P. J. (1973). Robust regression: asymptotics, conjectures and Monte Carlo. Ann. Statist. 1(5), 799 – 821.
- I. Gijbels and Verhasselt (2017) I. Gijbels, M. A. I. and A. Verhasselt (2017). Shape testing in quantile varying coefficient models with heteroscedastic error. J. Nonparametr. Stat. 29(2), 391–406.
- IPCC (2007) IPCC (2007). Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change [Solomon, S., D. Qin, M. Manning, Z. Chen, M. Marquis, K.B. Averyt, M. Tignor and H.L. Miller (eds.)]. Cambridge: Cambridge university press.
- Jensen (1968) Jensen, M. C. (1968). The performance of mutual funds in the period 1945-1964. The Journal of Finance 23(2), 389–416.
- Karmakar et al. (2022) Karmakar, S., S. Richter, and W. B. Wu (2022). Simultaneous inference for time-varying models. Journal of Econometrics 227(2), 408–428.
- Knight (2017) Knight, K. (2017). On the asymptotic distribution of the estimator in linear regression.
- Koenker (2005) Koenker, R. (2005). Quantile Regression. Econometric Society Monographs. Cambridge University Press.
- Komarova and Hidalgo (2020) Komarova, T. and J. Hidalgo (2020, June). Testing nonparametric shape restrictions. (arXiv:1909.01675). arXiv:1909.01675 [econ, math, stat].
- Koul and Stute (1999) Koul, H. L. and W. Stute (1999). Nonparametric model checks for time series. Ann. Statist. 27(1), 204 – 236.
- Kreiss and Paparoditis (2014) Kreiss, J.-P. and E. Paparoditis (2014, 04). Bootstrapping Locally Stationary Processes. J. R. Statist. Soc. B 77(1), 267–290.
- Lean and Rind (2008) Lean, J. L. and D. H. Rind (2008). How natural and anthropogenic influences alter global and regional surface temperatures: 1889 to 2006. Geophys. Res. Lett. 35(18).
- Matzkin (1994) Matzkin, R. L. (1994). Chapter 42 restrictions of economic theory in nonparametric methods. Volume 4 of Handbook of Econometrics, pp. 2523–2558. Elsevier.
- Mies (2023) Mies, F. (2023). Functional estimation and change detection for nonstationary time series. J. Am. Statist. Ass. 118(542), 1011–1022.
- Nason (2013) Nason, G. (2013, 03). A Test for Second-Order Stationarity and Approximate Confidence Intervals for Localized Autocovariances for Locally Stationary Time Series. J. R. Statist. Soc. B 75(5), 879–904.
- Noda (2022) Noda, A. (2022, August). Estimating the time-varying structures of the fama-french multi-factor models. (arXiv:2208.01270). arXiv:2208.01270 [econ, q-fin].
- Politis et al. (1999) Politis, D. N., J. P. Romano, and M. Wolf (1999). Choice of the Block Size, pp. 188–212. New York, NY: Springer New York.
- Racicot et al. (2019) Racicot, François-Éric., W. F. Rentz, D. Tessier, and R. Théoret (2019, 09). The conditional fama-french model and endogenous illiquidity: A robust instrumental variables test. PLOS ONE 14(9), 1–26.
- Robock (2000) Robock, A. (2000). Volcanic eruptions and climate. Reviews of Geophysics 38(2), 191–219.
- Sharpe (1964) Sharpe, W. F. (1964). Capital asset prices: A theory of market equilibrium under conditions of risk*. The Journal of Finance 19(3), 425–442.
- Silverman (1981) Silverman, B. W. (1981). Using kernel density estimates to investigate multimodality. J. R. Statist. Soc. B 43(1), 97–99.
- Stute (1997) Stute, W. (1997). Nonparametric model checks for regression. Ann. Statist. 25(2), 613 – 641.
- Stute et al. (1998) Stute, W., S. Thies, and L.-X. Zhu (1998). Model checks for regression: an innovation process approach. Ann. Statist. 26(5), 1916 – 1934.
- Trenberth et al. (2002) Trenberth, K. E., J. M. Caron, D. P. Stepaniak, and S. Worley (2002). Evolution of El Niño–Southern Oscillation and global atmospheric surface temperatures. Journal of Geophysical Research: Atmospheres 107(D8), AAC 5–1–AAC 5–17.
- Vogt (2012) Vogt, M. (2012). Nonparametric regression for locally stationary time series. Ann. Statist. 40(5), 2601–2633.
- Wu and Zhou (2017) Wu, W. and Z. Zhou (2017). Nonparametric inference for time-varying coefficient quantile regression. J. Bus. Econ. Statist. 35(1), 98–109.
- Wu and Zhou (2018) Wu, W. and Z. Zhou (2018). Gradient-based structural change detection for nonstationary time series M-estimation. Ann. Statist. 46(3), 1197–1224.
- Wu (2007) Wu, W. B. (2007). M-estimation of linear models with dependent errors. Ann. Statist. 35(2), 495 – 521.
- Zhou and Tung (2013) Zhou, J. and K.-K. Tung (2013). Deducing multidecadal anthropogenic global warming trends using multiple regression analysis. Journal of the Atmospheric Sciences 70(1), 3 – 8.
- Zhou (2013) Zhou, Z. (2013). Heteroscedasticity and autocorrelation robust structural change detection. J. Am. Statist. Ass. 108(502), 726–740.
- Zhou and Wu (2010) Zhou, Z. and W. B. Wu (2010). Simultaneous inference of linear models with time varying coefficients. J. R. Statist. Soc. B 72(4), 513–531.