Skip to main content
Log in

Joint tests for dynamic and spatial effects in short panels with fixed effects and heteroskedasticity

  • Published:
Empirical Economics Aims and scope Submit manuscript

Abstract

Simple and reliable tests are proposed for testing the existence of dynamic and/or spatial effects in fixed-effects panel data models with small T and possibly heteroskedastic errors. The tests are constructed based on the adjusted quasi scores (AQS), which correct the conditional quasi scores given the initial differences to account for the effect of initial values. To improve the finite sample performance, standardized AQS tests are also derived, which are shown to have much improved finite sample properties. All the proposed tests are robust against nonnormality, but some are not robust against cross-sectional heteroskedasticity (CH). A different type of adjustments is made on the AQS functions, leading to a set of tests that are fully robust against unknown CH. Monte Carlo results show excellent finite sample performance of the standardized versions of the AQS tests.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. Other works on short SDPD models include Elhorst (2010), Su and Yang (2015), Qu et al. (2016), and Kuersteiner and Prucha (2018). However, most of the research on SDPD models focuses on long panels (with large n and large T), see, e.g., Yang et al. (2006), Mutl (2006), Yu et al. (2008), Yu and Lee (2010), Lee and Yu (2010b, (2012, (2014); Bai and Li (2015), and Shi and Lee (2017).

  2. Interestingly, this method finds root in Neyman and Scott (1948) on modified likelihood equations. Chudik and Pesaran (2017) use similar ideas to give a bias-corrected method of moments estimation.

  3. This typically occurs to the estimator of the spatial error parameter; see Lee (2004), Liu and Yang (2015), Su and Yang (2015), and Yang (2018a). However, this feature is not explicitly reflected in the subsequent developments as the implementations of the tests do not require \(\iota \).

  4. The concentrated AQS function for \(\rho \) contained in (2.16) clearly shows that the M-estimator is not only consistent when T is fixed but also eliminates the bias of order \(O(T^{-1})\). In contrast, the estimator based on the unadjusted score is inconsistent when T is fixed and has a bias of order \(O(T^{-1})\) when T grows with n. See Hahn and Kuersteiner (2002), and Yang (2018a, (2018b) for more discussions.

  5. As \({{\mathbf {M}}}^{*}=\Omega ^{-1}-\Omega ^{-1}\Delta X(\Delta X^{\prime }\Omega ^{-1}\Delta X)^{-1}\Delta X^{\prime }\Omega ^{-1}\), calculations of \(\Omega ^{\frac{1}{2}}\) and \(\Omega ^{-\frac{1}{2}}\) are avoided.

  6. The Rook and Queen schemes are standard. For Group-I, we first generate \(k=\sqrt{n}\) groups of sizes \(n_g \sim U(.5{\bar{n}}, 1.5{\bar{n}})\), \(g=1, \cdots , k\) and \({\bar{n}}=n/k\), and then adjust \(n_g\) so that \(\sum _{g=1}^k n_g = n\). For Group-II, we first generate 6 groups of fixed sizes (3, 5, 7, 9, 11, 15), and replicate these groups r times to give \(n=r\times 50\). See Lin and Lee (2010) and Yang (2018a) for details in generating these spatial layouts.

  7. In both (ii) and (iii), the generated errors are standardized to have mean zero and variance \(\sigma _v^2\).

  8. See Lee and Yu (2016) for a detailed discussion on parameter identification of the SDPD model.

References

  • Anselin L, Bera AK, Florax R, Yoon MJ. Simple diagnostic tests for spatial dependence. Reg Sci Urban Econ. 1996;26:77–104.

    Google Scholar 

  • Anselin L. Rao’s score test in spatial econometrics. J Stat Plan Inference. 2001;97:113–39.

    Google Scholar 

  • Anselin L, Bera AK. Spatial dependence in linear regression models with an introduction to spatial econometrics. In: Ullah A, Giles DEA, editors. Handbook of applied economic statistics. New York: Marcel Dekker; 1998.

    Google Scholar 

  • Bai J, Li K. Dynamic spatial panel data models with common shocks. Working paper, Columbia University, New York; 2015

  • Baltagi BH, Song SH, Koh W. Testing panel data regression models with spatial error correlation. J Econom. 2003;117:123–50.

    Google Scholar 

  • Baltagi BH, Song SH, Jung Koh W. Testing for serial correlation, spatial autocorrelation and random effects using panel data. J Econom. 2007;140:5–51.

    Google Scholar 

  • Baltagi B, Yang ZL. Standardized LM tests for spatial error dependence in linear or panel regressions. Econom J. 2013a;16:103–34.

    Google Scholar 

  • Baltagi B, Yang ZL. Heteroskedasticity and non-normality robust LM tests of spatial dependence. Reg Sci Urban Econ. 2013b;43:725–39.

    Google Scholar 

  • Bera AK, Dogan O, Taspinar S, Leiluo Y. Robust LM tests for spatial dynamic panel data models. Reg Sci Urban Econ. 2019;76:47–66.

    Google Scholar 

  • Born B, Breitung J. Simple regression based tests for spatial dependence. Econom J. 2011;14:330–42.

    Google Scholar 

  • Chudik A, Pesaran MH. A bias-corrected method of moments approach to estimation of dynamic short-\(T\) panels. CESifo working paper series 2017; No. 6688.

  • Debarsy N, Ertur C. Testing for spatial autocorrelation in a fixed effects panel data model. Reg Sci Urban Econ. 2010;40:453–70.

    Google Scholar 

  • Elhorst JP. Dynamic panels with endogenous interaction effects when \(T\) is small. Reg Sci Urban Econ. 2010;40:272–82.

    Google Scholar 

  • Hahn J, Kuersteiner G. Asymptotically unbiased inference for a dynamic panel model with fixed effects when both \(n\) and \(T\) are large. Econometrica. 2002;70:1639–57.

    Google Scholar 

  • Jin F, Lee LF. On the bootstrap for Moran’s I test for spatial dependence. J Econom. 2015;184:295–314.

    Google Scholar 

  • Jin F, Lee LF. Outer-product-of-gradients tests for spatial autoregressive models. Reg Sci Urban Econ. 2018;72:35–57.

    Google Scholar 

  • Kelejian HH, Prucha IR. A generalized moments estimator for the autoregressive parameter in a spatial model. Int Econ Rev. 1999;40:509–33.

    Google Scholar 

  • Kelejian HH, Prucha IR. On the asymptotic distribution of the Moran \(I\) test statistic with applications. J Econom. 2001;104:219–57.

    Google Scholar 

  • Kuersteiner GM, Prucha IR. Dynamic panel data models: networks, common shocks, and sequential exogeneity. Working Paper, University of Maryland, College Park; 2018

  • Lee LF. Consistency and efficiency of least squares estimation for mixed regressive spatial autoregressive models. Econom Theory. 2002;18:252–77.

    Google Scholar 

  • Lee LF. Asymptotic distributions of quasi-maximum likelihood estimators for spatial autoregressive models. Econometrica. 2004;72:1899–925.

    Google Scholar 

  • Lee LF, Yu J. Estimation of spatial autoregressive panel data models with fixed effects. J Econom. 2010a;154:165–85.

    Google Scholar 

  • Lee LF, Yu J. A spatial dynamic panel data model with both time and individual fixed effects. Econom Theory. 2010b;26:564–97.

    Google Scholar 

  • Lee LF, Yu J. Spatial panels: random components vs. fixed effects. Int Econ Rev. 2012;53:1369–412.

    Google Scholar 

  • Lee LF, Yu J. Efficient GMM estimation of spatial dynamic panel data models with fixed effects. J Econom. 2014;180:174–97.

    Google Scholar 

  • Lee LF, Yu J. Identification of spatial Durbin panel models. J Appl Econom. 2016;31:133–62.

    Google Scholar 

  • Li LY, Yang ZL. Spatial dynamic panel data models with correlated random effects. J Econom. 2020a (forthcoming).

  • Li LY, Yang ZL. Estimation of fixed effects spatial dynamic panel data models with small \(T\) and unknown heteroskedasticity. Reg Sci Urban Econ. 2020b (forthcoming).

  • Lin X, Lee LF. GMM estimation of spatial autoregressive models with unknown heteroskedasticity. J Econom. 2010;157:34–52.

    Google Scholar 

  • Liu SF, Yang ZL. Asymptotic distribution and finite sample bias correction of QML estimators for spatial error dependence model. Econometrics. 2015;3:376–411.

    Google Scholar 

  • Liu X, Prucha IR. A robust test for network generated dependence. J Econom. 2018;207:92–113.

    Google Scholar 

  • Mutl J, Dynamic panel data models with spatially correlated disturbances. Ph.D. Thesis, University of Maryland, College Park; 2006

  • Neyman J, Scott EL. Consistent estimates based on partially consistent observations. Econometrica. 1948;16:1–32.

    Google Scholar 

  • Qu X, Wang X, Lee LF. Instrumental variable estimation of a spatial dynamic panel model with endogenous spatial weights when \(T\) is small. Econom J. 2016;19:261–90.

    Google Scholar 

  • Robinson PM, Rossi F. Improved Lagrange multiplier tests in spatial autoregressions. Econom J. 2014;17:139–64.

    Google Scholar 

  • Robinson PM, Rossi F. Refined tests for spatial correlation. Econom Theory. 2015a;31:1249–80.

    Google Scholar 

  • Robinson PM, Rossi F. Refinements in maximum likelihood inference on spatial autocorrelation in panel data. J Econom. 2015b;189:447–56.

    Google Scholar 

  • Shi W, Lee LF. Spatial dynamic panel data models with interactive fixed effects. J Econom. 2017;197:323–47.

    Google Scholar 

  • Su LJ, Yang ZL. QML estimation of dynamic panel data models with spatial errors. J Econom. 2015;185:230–58.

    Google Scholar 

  • Taspinar S, Dogan O, Bera AK. GMM gradient tests for spatial dynamic panel data models. Reg Sci Urban Econ. 2017;65:65–88.

    Google Scholar 

  • van der Vaart AW. Asymptotic statistics. Cambridge: Cambridge University Press; 1998.

    Google Scholar 

  • Xu YH, Yang ZL. Specification tests for temporal heterogeneity in spatial panel data models with fixed effects. Reg Sci Urban Econ. 2019 (forthcoming).

  • Yang ZL. A robust LM test for spatial error components. Reg Sci Urban Econ. 2010;40:299–310.

    Google Scholar 

  • Yang ZL. LM tests of spatial dependence based on bootstrap critical values. J Econom. 2015;185:33–59.

    Google Scholar 

  • Yang ZL. Unified \(M\)-estimation of fixed-effects spatial dynamic models with short panels. J Econom. 2018a;205:423–47.

    Google Scholar 

  • Yang ZL. Supplement to Unified \(M\)-estimation of fixed-effects spatial dynamic models with short panels. J Econom. 2018b;205:423–47.

    Google Scholar 

  • Yang ZL. Bootstrap LM tests for higher order spatial effects in spatial linear regression models. Empir Econ. 2018c;55:35–68.

    Google Scholar 

  • Yang ZL, Li C, Tse YK. Functional form and spatial dependence in dynamic panels. Econ Lett. 2006;91:138–45.

    Google Scholar 

  • Yu J, de Jong R, Lee LF. Quasi-maximum likelihood estimators for spatial dynamic panel data with fixed effects when both \(n\) and \(T\) are large. J Econom. 2008;146:118–34.

    Google Scholar 

  • Yu J, Lee LF. Estimation of unit root spatial dynamic panel data models. Econom Theory. 2010;26:1332–62.

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Zhenlin Yang.

Ethics declarations

Conflict of Interest

The author declares that he has no conflict of interest.

Ethical approval

This article does not contain any studies with human participants or animals performed by the author.

Additional information

Publisher's Note

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

An early version was circulated under the title: Joint Tests for Dynamic and Spatial Effects in Short Panels with Fixed Effects. I would like to thank Peter Robinson, Fei Jin, Lung-Fei Lee, James LeSage, the participants of the 15th International Workshop on Spatial Econometrics and Statistics, Orleans, 27–28 May 2016 and the 2nd Econometrics Workshop at the Chinese University of Hong Kong, 29 April 2017, Guest Editors Qi Li and Vasilis Sarafidis, and two anonymous referees, for their helpful comments. I am grateful to Singapore Management University for financial support under Grant C244/MSS14E002.

Appendices

Appendix A: Some useful lemmas

The development and the proofs of theoretical results reported in this paper depend critically on the following lemmas.

Lemma A.1

(Kelejian and Prucha 1999; Lee 2002). Let \(\{A_n\}\) and \(\{B_n\}\) be two sequences of \(n\times n\) matrices that are uniformly bounded in both row and column sums. Let \(C_n\) be a sequence of conformable matrices whose elements are uniformly \(O(\iota _{n}^{-1})\). Then,

  1. (i)

    the sequence \(\{A_n B_n\}\) is uniformly bounded in both row and column sums,

  2. (ii)

    the elements of \(A_n\) are uniformly bounded and tr\((A_n) = O(n)\), and

  3. (iii)

    the elements of \(A_n C_n\) and \(C_n A_n\) are uniformly \(O(\iota _{n}^{-1})\).

Lemma A.2

(Lee 2004, p.1918). For \(W_r\) and \(B_r\), \(r=1,3\), defined in Model (1.1), if \(\Vert W_r\Vert \) and \(\Vert B_{r0}^{-1}\Vert \) are uniformly bounded, where \(\Vert \cdot \Vert \) is a matrix norm, then \(\Vert B_{r}^{-1}\Vert \) is uniformly bounded in a neighborhood of \(\lambda _{r0}\).

Lemma A.3

(Lee 2004, p.1918). Let \(X_n\) be an \(n\times p\) matrix. If the elements \(X_n\) are uniformly bounded and \(\lim _{n\rightarrow \infty }\frac{1}{n}X^{\prime }_n X_n\) exists and is nonsingular, then \(P_n = X_n(X^{\prime }_n X_n)^{-1}X^{\prime }_n\) and \(M_n = I_n - P_n\) are uniformly bounded in both row and column sums.

Lemma A.4

(Li and Yang 2020b) Let \(\{A_n\}\) be a sequence of \(n \times n\) matrices that are uniformly bounded in either row or column sums. Suppose that the elements \(a_{n,ij}\) of \(A_n\) are \(O(\iota ^{-1})\) uniformly in all i and j. Let \(v_n\) be a random n-vector of inid elements satisfying Assumption B, and \(b_n\) a constant n-vector of elements of uniform order \(O(\iota ^{-{1}/{2}})\). Then,

$$\begin{aligned} \begin{array}{ll} (i) \; \mathrm{E}(v^{\prime }_n A_n v_n) = O(\frac{n}{\iota _n}), &{} (ii) \mathrm{Var}(v^{\prime }_n A_n v_n) = O(\frac{n}{\iota _n}), \\ (iii) \mathrm{Var}(v^{\prime }_n A_n v_n + b_n^{\prime } v_n) = O(\frac{n}{\iota _n}), &{} (iv) v^{\prime }_n A_n v_n = O_p(\frac{n}{\iota _n}), \\ (v) \; v^{\prime }_n A_n v_n -\mathrm{E}(v^{\prime }_n A_n v_n) = O_p((\frac{n}{\iota _n})^{\frac{1}{2}}), &{} (vi) v^{\prime }_n A_n b_n = O_p((\frac{n}{\iota _n})^{\frac{1}{2}}), \end{array} \end{aligned}$$

and (vii), the results (iii) and (vi) remain valid if \(b_n\) is a random n-vector independent of \(v_{n}\) such that \(\{\mathrm{E}(b_{ni}^{2})\}\) are of uniform order \(O(\iota _{n}^{-1})\).

Lemma A.5

(Li and Yang 2020b): Let \(\{\Phi _{n}\}\) be a sequence of \(n\times n\) matrices with row and column sums being uniformly bounded, and elements of uniform order \(O(\iota _{n}^{-1})\). Let \(v_{n}\) be a random n-vector satisfying Assumption B. Let \(b_{n}=\{b_{ni}\}\) be an \(n\times 1\) random vector, independent of \(v_{n}\), such that \((i)\ \{\mathrm{E}(b_{ni}^2)\}\) are of uniform order \(O(\iota _{n}^{-1})\), (ii) sup\(_{i}E|b_{ni}|^{2+\epsilon _{0}} < \infty \), (iii) \(\frac{\iota _n}{n}\sum _{i=1}^{n}[\phi _{n,ii}(b_{ni}-\mathrm{E}b_{ni})]=o_p(1)\) where \(\{\phi _{n,ii}\}\) are the diagonal elements of \(\Phi _{n}\), and (iv) \(\frac{\iota _n}{n}\sum _{i=1}^{n}[b_{ni}^{2}-\mathrm{E}(b_{ni}^{2})]=o_p(1)\). Let \({{\mathcal {H}}}_n \)= \(\mathrm{diag}(h_{n1},\ldots , h_{nn})\). Define the bilinear-quadratic form:

$$\begin{aligned} Q_{n} = b_{n}'v_{n} + v_{n}^{\prime }\Phi _{n}v_{n} - \sigma _{v}^{2}tr(\Phi _{n} {{\mathcal {H}}}_n), \end{aligned}$$

and let \(\sigma _{Q_{n}}^{2}\) be the variance of \(Q_n\). If \(\lim _{n\rightarrow \infty }\iota _{n}^{1+2/\epsilon _{0}}/n=0\) and \(\{\frac{\iota _n}{n}\sigma _{Q_{n}}^{2}\}\) are bounded away from zero, then \(Q_{n}/\sigma _{Q_{n}}\overset{d}{\longrightarrow }N(0,1)\).

The following lemma extends the formulations in Sect. 2.2 to allow for unknown CH. Its proof follows the results of Theorems 3.2 and 3.3 of Li and Yang (2020b). Recall: \(A^{u}\), \(A^{l}\) and \(A^{d}\) denote the upper-triangular, lower-triangular, and diagonal matrix of a square matrix A; \(\Pi _{t}\), \(\Phi _{ts}\) and \(\Psi _{ts}\) the submatrices of \(\Pi \), \(\Phi \), and \(\Psi \) partitioned according to \(t, s = 2, \ldots , T\); \(\Psi _{t+}=\sum _{s=2}^{T}\Psi _{ts}\), \(\Theta =\Psi _{2+}(B_{30}B_{10})^{-1}\), \(\Delta y_{1}^{\circ }=B_{30}B_{10}\Delta y_{1}\), and \(\Delta y_{1t}^{*}=\Psi _{t+}\Delta y_{1}\).

Lemma A.6

Suppose Assumptions A, B\(^{*}\), C-E hold for Model (2.1). Consider the linear, quadratic, and bilinear forms, \(Q(\psi _{0})=\{(\Pi \Delta ^{\prime }v)^{\prime }, \Delta v'\Phi \Delta v, (\Delta v'\Psi \Delta {{\mathbf {y}}}_1)^{\prime }\}^{\prime }\), associated with the model. Assume the elements of \(\Pi \) (\(N\times 1\)) are uniformly bounded, and the matrices \(\Phi \) and \(\Psi \) (\(N\times N\)) are uniformly bounded in both row and column sums. Define

$$\begin{aligned}\begin{array}{lll} g_{1i} &{}=&{} \sum _{t=2}^{T} \Pi '_{it} \Delta v_{it}, \\ g_{2i} &{}=&{} \sum _{t=2}^{T}(\Delta v_{it}\Delta \xi _{it} + \Delta v_{it}\Delta v_{it}^{*}-\sigma _{v0}^{2}d_{it}),\\ g_{3i} &{}=&{} \Delta v_{2i}\Delta \zeta _{i} + \Theta _{ii}(\Delta v_{2i}\Delta y_{1i}^{\circ }+\sigma _{v0}^{2}h_{ni}) + \sum _{t=3}^{T}\Delta v_{it}\Delta y_{1i t}^{*}, \end{array} \end{aligned}$$

where \(\xi _{t} = \sum _{s=2}^{T}(\Phi _{st}^{u \prime } + \Phi _{ts}^{l}) \Delta v_{s}\), \(\Delta v_{t}^{*} = \sum _{s=2}^{T}\Phi _{ts}^{d}\Delta v_{s}\), and \(\{d_{it}\}\) are the diagonal elements of \(\Phi (C\otimes {{\mathcal {H}}}_n)\), \(\{\Delta \zeta _i\} = \Delta \zeta = (\Theta ^{u}+\Theta ^{l})\Delta y_{1}^{\circ }\), and \(\mathrm{diag}\{\Theta _{ii}\}=\Theta ^{d}\). Then, we have,

\((i)\;\;\; Q(\psi _{0})-\mathrm{E}[Q(\psi _{0})] = \sum _{i=1}^{n}{} \mathbf{g}_{i}\), where \(\mathbf{g}_{i} = (g_{1i}, g_{2i}, g_{3i})'\),

\((ii)\;\; \frac{1}{N}[Q(\psi _{0})-\mathrm{E}(Q(\psi _{0}))] \overset{D}{\longrightarrow } N(0, \lim _{n\rightarrow \infty }\frac{1}{N}\Gamma )\), where \(\Gamma =\mathrm{Var}(Q(\psi _{0}))\).

\((iii)\ \mathrm{Var}[Q(\psi _{0})] = \sum _{i=1}^{n}\mathrm{E}(\mathbf{g}_{i}{} \mathbf{g}'_{i})\), and \((iv)\; \frac{1}{N} \sum _{i=1}^{n}[\mathbf{g}_{i}{} \mathbf{g}'_{i}-\mathrm{E}(\mathbf{g}_{i}\mathbf{g}'_{i})] = o_p(1)\).

Appendix B: Some technical details

We sketch the proofs of the theorems, corollaries, and lemmas. Details are given in Supplementary Appendix available at http://www.mysmu.edu/faculty/zlyang/, including the detailed expressions for the derivative matrices of the three AQS functions, which is referred to loosely as the Hessian matrix in this paper.

Proof Theorem 2.1

The proof of Theorem 2.1 follows closely the proofs of Theorems 3.2 and 3.3 of Yang (2018a), and is typically simpler as under \(H_{0}^\mathtt{M}\) the model becomes simpler. The Hessian matrix \(\frac{\partial }{\partial \psi ^{\prime }}S^{*}(\psi )\) used to estimate \(\Sigma ^{*}(\psi _{0})\) can be easily derived based on the expression of \(S^{*}(\psi )\) given in (2.2). It can found in Yang (2018a, Proof of Theorem 3.2), and also in the Supplementary Appendix to this paper containing additional ‘asymmetric components’ that did not appear in Yang (2018a).

Proof of Corollary 2.1

The quantities needed for evaluating the AQS function defined in (2.4) become: \(\Pi _{1}=\frac{1}{\sigma _{v0}^2}{{\mathbf {C}}}^{-1}\Delta X\), \(\Pi _{2}=\frac{1}{\sigma _{v0}^2}{{\mathbf {C}}}^{-1}{{\mathbb {B}}}_{-1}\Delta X\beta \), \(\Pi _{3}=\frac{1}{\sigma _{v0}^2}{{\mathbf {C}}}^{-1}{{\mathbf {W}}}_{1}\Delta X\beta \), \(\Pi _{4}=\frac{1}{\sigma _{v0}^2}{{\mathbf {C}}}^{-1}{{\mathbf {W}}}_{2}{{\mathbb {B}}}_{-1}\Delta X\beta \), \(\Phi _{1}= \frac{1}{2\sigma _{v0}^4}{{\mathbf {C}}}^{-1}\), \(\Phi _{2}\)=\(\frac{1}{\sigma _{v0}^2}{{\mathbf {C}}}^{-1}{{\mathbb {B}}}_{-1}\), \(\Phi _{3}=\frac{1}{\sigma _{v0}^2}{{\mathbf {C}}}^{-1}{{\mathbf {W}}}_{1}\), \(\Phi _{4}=\frac{1}{\sigma _{v0}^2}{{\mathbf {C}}}^{-1}{{\mathbf {W}}}_{2}{{\mathbb {B}}}_{-1}\), \(\Phi _{5}=\frac{1}{2\sigma _{v0}^2}[C^{-1}\otimes (W'_{3}+W_{3})]\), \(\Psi _{1}=\frac{1}{\sigma _{v0}^2}{{\mathbf {C}}}^{-1}{{\mathbb {R}}}_{-1}\), \(\Psi _{2}=0\), \(\Psi _{3}=\frac{1}{\sigma _{v0}^2}{{\mathbf {C}}}^{-1}{{\mathbf {W}}}_{2}{{\mathbb {R}}}_{-1}\), \({{\mathbb {R}}}_{-1}=\mathrm{blkdiag}(I_n, 0,\ldots , 0)\), \({{\mathbb {B}}}_{-1} = I_{T-1}^{*} \otimes I_n\), and \(I_{T-1}^{*}\) is a \((T-1)\times (T-1)\) matrix with elements 1 on the positions immediately below the diagonal elements, and zero elsewhere. Further, \({{\mathcal {B}}}_0 = 0_n\), and hence \({{\mathbf {D}}}_0=-C\otimes I_n\) and \({{\mathbf {D}}}_{-10}=-C_{-1}\otimes I_n\), where

$$\begin{aligned} C_{-1} = \left( \begin{array}{ccccccc} -1&{}\quad 0&{}\quad 0&{}\quad \cdots &{}\quad 0&{}\quad 0&{}\quad 0\\ 2&{}\quad -1&{}\quad 0&{}\quad \cdots &{}\quad 0&{}\quad 0&{}\quad 0\\ -1&{}\quad 2&{}\quad -1&{}\quad \cdots &{}\quad 0&{}\quad 0&{}\quad 0\\ \vdots &{}\quad \vdots &{}\quad \vdots &{}\quad \ddots &{}\quad \vdots &{}\quad \vdots &{}\quad \vdots \\ 0&{}\quad 0&{}\quad 0&{}\quad \cdots &{}\quad 0&{}\quad 2&{}\quad -1\\ \end{array} \right) _{(T-1)\times (T-1)}. \end{aligned}$$

These show that with \(\psi _0=(\beta '_{0},\sigma _{v0},0,0,0,0)'\), all the \(\Phi \) and \(\Psi \) matrices are either of the form \(A\otimes I_n\) or \(A\otimes W\) for some \((T-1)\times (T-1)\) matrix A and a spatial weight matrix W satisfying Assumption D. Thus, \(\mathrm{E}[S^{*}(\psi _0)]=0\) even when the errors are heteroskedastic. Hence by Lemma A.5, we have \(\frac{1}{\sqrt{N}}S^*(\psi _{0}) \overset{D}{\longrightarrow } N[0, \lim _{n\rightarrow \infty }\frac{1}{N}\Gamma ^{*}(\psi _{0})]\).

By the mean value theorem (MVT), one easily shows that \(\frac{1}{\sqrt{N}}[S_{\delta }^*({\tilde{\psi }})-S_{\delta }^*(\psi _{0})]=o_p(1)\), where \({\tilde{\psi }}=({\tilde{\beta }}', {\tilde{\sigma }}_{v0}^{2},0,0,0,0)'\) and we note that the OLS estimators \({\tilde{\beta }}\) and \({\tilde{\sigma }}_{v0}^{2}\) are robust against unknown heteroskedasticity \(\{h_{ni}\}\). Now, since by (2.12) \(S^*(\psi _{0})=\sum _{i=1}^{n} \mathbf{g}_{i}\), where \(\{\mathbf{g}_{i},{{\mathcal {F}}}_{n,i}\}\) form a vector MD sequence, we have \(\frac{1}{N}\sum _{i=1}^{n}[\mathbf{g}_{i}{} \mathbf{g}'_{i}-\mathrm{E}(\mathbf{g}_{i}{} \mathbf{g}'_{i})]=o_p(1)\) by Lemma A.6. By MVT and the consistency of \({\tilde{\beta }}\) and \({\tilde{\sigma }}_{v0}^{2}\), one shows that \(\frac{1}{N}\sum _{i=1}^{n}(\tilde{\mathbf{g}}_{i}\tilde{\mathbf{g}}'_{i}-\mathbf{g}_{i}{} \mathbf{g}'_{i})=o_p(1)\) under heteroskedasticity. Finally, it is easy to show that \(\mathrm{plim}_{n\rightarrow \infty }(\widetilde{\Lambda }-{\Lambda })=0\), using the simplified expression of \(H^{*}(\psi )\) and MVT. \(\square \)

Proof of Lemma 2.1

Consider the AQS vector \(S^{*}(\beta ,\sigma _{v}^{2},\rho )\) for the DPD model, and the concentrated AQS function which defines \({\tilde{\rho }}\) under \(H_{0}^\mathtt{DPD}\):

$$\begin{aligned} \textstyle S_\mathtt{DPD}^{*c}(\rho )=\frac{1}{{\tilde{\sigma }}_v^{2}(\rho )}\Delta {\tilde{v}}'(\rho ){{\mathbf {C}}}^{-1}\Delta Y_{-1} + n\big (\frac{1}{1-\rho }-\frac{1-\rho ^T}{T(1-\rho )^2}\big ), \end{aligned}$$

where \(\Delta {\tilde{v}}(\rho )=\Delta Y-\rho \Delta Y_{-1}-\Delta X{\tilde{\beta }}(\rho )\), \({\tilde{\beta }}(\rho ) = (\Delta {X}^{\prime }{{\mathbf {C}}}^{-1}\Delta {X})^{-1}\Delta {X}^{\prime }{{\mathbf {C}}}^{-1}(\Delta Y-\rho \Delta Y_{-1})\) and \({\tilde{\sigma }}_v^2(\rho ) = \textstyle \frac{1}{N}\Delta {\tilde{v}}^{\prime }(\rho ){{\mathbf {C}}}^{-1}\Delta {\tilde{v}}(\rho )\).

Define \({\bar{S}}^{*}(\beta ,\sigma _{v}^{2},\rho )=\mathrm{E}[S^{*}(\beta ,\sigma _{v}^{2},\rho )]\). Given \(\rho \), \({\bar{S}}^{*}(\beta ,\sigma _{v}^{2},\rho ) = 0\) is partially solved at \({\bar{\beta }}(\rho )=(\Delta {X}^{\prime }{{\mathbf {C}}}^{-1}\Delta {X})^{-1}\Delta {X}^{\prime }{{\mathbf {C}}}^{-1}(\mathrm{E}\Delta Y-\rho \mathrm{E}\Delta Y_{-1})\) and \({\bar{\sigma }}_{v}^{2}(\rho )=\frac{1}{N}\mathrm{E}[\Delta {\bar{v}}(\rho )'{{\mathbf {C}}}^{-1}\Delta {\bar{v}}(\rho )]\), where \(\Delta {\bar{v}}(\rho )=\Delta Y-\rho \Delta Y_{-1}-\Delta X{{\bar{\beta }}}(\rho )\). Substituting \({\bar{\beta }}(\rho )\) and \({\bar{\sigma }}_{v}^{2}(\rho )\) back into \({\bar{S}}^{*}(\beta ,\sigma _{v}^{2},\rho )\) gives the population counterpart of \(S_\mathtt{DPD}^{*c}(\rho )\) as

$$\begin{aligned} \textstyle {\bar{S}}_\mathtt{DPD}^{*c}(\rho )=\frac{1}{{{\bar{\sigma }}}_v^{2}(\rho )}\mathrm{E}[\Delta {\bar{v}}'(\rho ){{\mathbf {C}}}^{-1}\Delta Y_{-1}] + n\big (\frac{1}{1-\rho }-\frac{1-\rho ^T}{T(1-\rho )^2}\big ). \end{aligned}$$

By Theorem 5.9 of van der Vaart (1998), \({\tilde{\rho }}\) will be consistent if \((i) \inf _{\rho : |\rho -\rho _0|\ge \epsilon }|{\bar{S}}_\mathtt{DPD}^{*c}(\rho )| > 0\) for every \(\epsilon > 0\), and \((ii) \sup _{\rho \in \Upsilon }\frac{1}{\sqrt{N}}|S_\mathtt{DPD}^{*c}(\rho )-{\bar{S}}_\mathtt{DPD}^{*c}(\rho )| \overset{p}{\longrightarrow } 0\), which are straightforward. The asymptotic normality can be proved using Lemma A.5. \(\square \)

Proof

First, with \(\psi _{0}=(\beta _{0},\sigma _{v0}^{2},\rho _{0},0'_{3})'\) it is easy to show that \(\mathrm{E}[S^{*}(\psi _0)]=0\) under the general heteroskedasticity \(\{h_{ni}\}\). By Lemma A.5, one shows that \(\frac{1}{\sqrt{N}}S^{*}(\psi _{0}) \overset{D}{\longrightarrow } N(0, \Gamma ^{*}(\psi _0))\). By Lemma A.6, one shows that \(\frac{1}{N}\sum _{i=1}^{n}[\mathbf{g}_{n,i}{} \mathbf{g}'_{n,i}-\mathrm{E}(\mathbf{g}_{i}{} \mathbf{g}'_{i})] \overset{p}{\longrightarrow } 0\). By the mean value theorem, and \(\sqrt{N}\) consistency and robustness of \({\tilde{\beta }}\), \({\tilde{\sigma }}_{v}^{2}\) and \({\tilde{\rho }}\) against unknown heteroskedasticity \(\{h_{ni}\}\) as shown in Lemma 2.1, we have \(\frac{1}{\sqrt{N}}[S_{\lambda }^{*}({\tilde{\psi }})-S_{\lambda }^{*}(\psi _{0})] \overset{p}{\longrightarrow } 0\) where \({\tilde{\psi }}=({\tilde{\beta }}', {\tilde{\sigma }}_{v}^{2},{\tilde{\rho }},0_3)'\), and \(\frac{1}{N}\sum _{i=1}^{n}(\tilde{\mathbf{g}}_{n,i}\tilde{\mathbf{g}}'_{n,i}-\mathbf{g}_{i}{} \mathbf{g}'_{i}) \overset{p}{\longrightarrow } 0\). Finally, using the simplified expression of \(H^{*}(\psi )\) and MVT, we show \(\mathrm{plim}_{n\rightarrow \infty }(\widetilde{\Lambda }-{\Lambda })=0\).

Proof of Theorem 2.2

The proof is similar to that of Theorem 2.1. The partial derivatives of \(S^{\diamond }(\delta )\) required to estimate the components of \(\Sigma _{\varphi \pi }^{\diamond }(\delta _{0})\) and \(\Sigma _{\varphi \varphi }^{\diamond }(\delta _{0})\) can be easily obtained from the expression \(S^{\diamond }(\delta )\) given in (2.19). The full expression of \(\frac{\partial }{\partial \delta ^{\prime }}S^{\diamond }(\delta )\) is given in the Supplementary Appendix to this paper. \(\square \)

Proof of Theorem 3.1

The proof of Theorem 3.1 follows closely the proofs of Theorems (3.2) and (3.3) of Li and Yang (2020b). The Hessian matrix \(\frac{\partial }{\partial \psi ^{\prime }}S_\mathtt{H}^{*}(\psi )\) used to estimate \(\Sigma _\mathtt{H}^{*}(\psi _{0})\) is given in Li and Yang (2020b, Proof of Theorem 3.2), and can also be found in the Supplementary Appendix to this paper, where the ‘asymmetric components’ that did not appear in Li and Yang (2020b) are also given. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yang, Z. Joint tests for dynamic and spatial effects in short panels with fixed effects and heteroskedasticity. Empir Econ 60, 51–92 (2021). https://doi.org/10.1007/s00181-020-01935-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00181-020-01935-y

Keywords

JEL Classification

Navigation