Log in

Skewness-based test diagnosis of technical inefficiency in spatial autoregressive stochastic frontier models

  • Published:
Journal of Productivity Analysis Aims and scope Submit manuscript

Abstract

In the Spatial Autoregressive (SAR) Stochastic Frontier (SF) model, the inefficiency term, which distinguishes it from the SAR model, can capture the effects of technical inefficiency. To determine whether inefficiency significantly exists in the cross-sectional SARSF model, this paper proposes a skewness-based test. This test does not rely on the normality assumption for the disturbances and allows inefficiency to follow an unknown one-sided distribution. We establish the asymptotic theory of the test statistic under spatial near-epoch dependent properties. Furthermore, we extend this test to the panel SARSF data model, accounting for both individual and time fixed-effects. Additionally, Monte Carlo simulations demonstrate the robustness of our test against non-normal disturbances and its satisfactory performance with different one-sided distributions for inefficiency. Finally, we provide an empirical application using data from 137 dairy farms in Northern Spain to illustrate the presence of technical inefficiency in production according to our test.

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

Access this article

Subscribe and save

Springer+ Basic
EUR 32.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or Ebook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

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

Instant access to the full article PDF.

Fig. 1

Similar content being viewed by others

Notes

  1. This means that at least one component of Un is strictly positive and the rest are nonnegative.

  2. Following Smith (2008), introducing dependence between vni and uni in the stochastic frontier models has become a focus of interest. Griffiths and Hajargasht (2016) and Amsler et al. (2016, 2017) exemplify papers that address the endogeneity problem while allowing the correlation of vni and uni.

  3. Here, β1 and β−1 are, respectively, the first element and the remaining sub-vector of β, i.e., \(\beta ={[{\beta }_{1},{\beta }_{-1}^{{\prime} }]}^{{\prime} }\). In addition, we can see from (2.1) and (2.1’) that \({X}_{n}^{{\dagger} }\) must contain an intercept, no matter whether Xn contains or not.

  4. Under Assumption 1, no matter what distribution \({\epsilon }_{ni}^{{\dagger} }\)’s follow, QMLE treats the \({\epsilon }_{ni}^{{\dagger} }\)’s as normal.

  5. The first element of \({\tilde{\beta }}_{n}^{{\dagger} }\) is consistent for β1 − E[uni] but β1 as model (2.1) includes an intercept, and converges to − E[uni] as model (2.1) do not include an intercept.

  6. Lp-NED also implies Lq-NED for any q < p.

  7. When λ = 0, In − λWn reduces to the identity matrix In such that \({M}_{n}^{{\dagger} }({I}_{n}-\lambda {W}_{n})^{-1}{X}_{n}^{{\dagger} }={M}_{n}^{{\dagger} }{X}_{n}^{{\dagger} }={{{{\bf{0}}}}}_{n\times k}\).

  8. See Papadopoulos and Parmeter (2023) for a survey on wrong skewness and Badunenko and Henderson (2023) for an alternative SF model, which aims to avoid the wrong skewness problem by introducing an asymmetric two-sided disturbance.

  9. As in the footnote 5, \({\beta }_{0}^{{\dagger} }\) is equal to β0 + E[uit] but β0.

  10. This means that at least one component of Unt is strictly positive and the rest are nonnegative.

  11. Available on https://github.com/Dr-Man-go/Cubic-CLT.

  12. This is generated from the algebra equality that \({{{\rm{vec}}}}(ABC)=({C}'\otimes A){{{\rm{vec}}}}(B)\), where A, B and C are compatible matrices, and vec( ⋅ ) denotes the vectorization operator.

References

  • Aigner D, Lovell CAK, Schmidt P (1977) Formulation and estimation of stochastic frontier production function models. J Econom 6:21–37

    Article  MathSciNet  Google Scholar 

  • Amsler C, Prokhorov A, Schmidt P (2016) Endogenous stochastic frontier models. J Econom 190:280–288

    Article  Google Scholar 

  • Amsler C, Prokhorov A, Schmidt P (2017) Endogenous environmental variables in stochastic frontier models. J Econom 199:131–140

    Article  MathSciNet  Google Scholar 

  • Amsler C, Schmidt P, Tsay WJ (2019) Evaluating the cdf of the distribution of stochastic frontier composed error. J Prod Anal 52:29–35

    Article  Google Scholar 

  • Ayouba K (2023) Spatial dependence in production frontier models. J Prod Anal 60:21–36

    Article  Google Scholar 

  • Badunenko O, Henderson DJ (2023) Production analysis with asymmetric noise. J Prod Anal. https://doi.org/10.1007/s11123-023-00680-5

  • Baptista R (2000) Do innovations diffuse faster within geographical clusters? Int J Ind Organ 18:515–535

    Article  Google Scholar 

  • Berndt EK, Hall HB, Hall RE, Hausman J (1974) Estimation and inference in non-linear structural model. Anal Econ Soc Measur 4:653–666

    Google Scholar 

  • Billingsley P (2012) Probability and Measure, Anniversary edn. John Wiley & Sons, Inc., Hoboken, New Jersey

  • Bonanno G, Giovanni DD, Domma F (2017) The ‘wrong skewness’ problem: a re-specification of stochastic frontiers. J Prod Anal. 47:49–64

    Article  Google Scholar 

  • Chen Y-T, Wang H-J (2012) Centered-residuals-based moment estimator and test for stochas tic frontier models. Economet Rev 31:625–653

    Article  CAS  Google Scholar 

  • Cliff A, Ord J (1973) Spatial autocorrelation. Pion, London

  • Cliff A, Ord J (1981) Spatial process: models and applications. Pion, London

  • Coelli TJ (1995) Estimators and hypothesis tests for a stochastic frontier function: a Monte Carlo analysis. J Prod Anal 6:247–268

    Article  Google Scholar 

  • Davidson R, MacKinnon J G (2004) Econometric theory and methods. Oxford University Press, New York

  • Deng M-Y, Wang M (2022) Artificial regression test diagnostics for impact measures in spatial models. Econom Lett 217, 110689

  • Fried HO, Lovell CK, Schmidt SS (2008) The Measurement of Productive Efficiency and Productivity Growth. Oxford University Press, New York

  • Gallant AR, White H (1988) A Unified Theory of Estimation and Inference for Nonlinear Dynamic Models. Basil Blackwell, New York, 1988

  • Galli F (2023) A spatial stochastic frontier model including both frontier and error-based spatial cross-sectional dependence. Spatial Economic Anal 18:239–258

    Article  Google Scholar 

  • Glass A, Kenjegalieva K, Paez-Farrell J (2013) Productivity growth decomposition using a spatial autoregressive frontier model. Econom Lett 119:291–295

    Article  Google Scholar 

  • Glass A, Kenjegalieva K, Sickles RC (2014) Estimating efficiency spillovers with state level evidence for manufacturing in the US. Econom Lett 123:154–159

    Article  Google Scholar 

  • Glass AJ, Kenjegalieva K (2019) A spatial productivity index in the presence of efficiency spillovers: evidence for US banks, 1992–2015. Eur J Oper Res 273:1165–1179

    Article  MathSciNet  Google Scholar 

  • Glass AJ, Kenjegalieva K, Sickles RC (2016) A spatial autoregressive stochastic frontier model for panel data with asymmetric efficiency spillovers. J Econom 190:289–300

    Article  MathSciNet  Google Scholar 

  • Greene WH (1990) A Gamma-distributed stochastic frontier model. J Econom 46:141–163

    Article  MathSciNet  Google Scholar 

  • Greene WH (2005) Fixed and random effects in stochastic frontier models. J Prod Anal 23:7–32

    Article  Google Scholar 

  • Greene WH (2005) Reconsidering heterogeneity in panel data estimators of the stochastic frontier model. J Econom 126:269–303

    Article  MathSciNet  Google Scholar 

  • Griffiths WE, Hajargasht G (2016) Some models for stochastic frontiers with endogeneity. J Econom 190:341–348

    Article  MathSciNet  Google Scholar 

  • Gude A, Álvarez I, Orea L (2018) Heterogeneous spillovers among Spanish provinces: a generalized spatial stochastic frontier model. J Prod Anal 50:155–173

    Article  Google Scholar 

  • Hou Z, Zhao S, Kumbhakar SC (2023) The GMM estimation of semiparametric spatial stochastic frontier models. Eur J Oper Res 305:1450–1464

    Article  MathSciNet  Google Scholar 

  • Jenish N, Prucha IR (2009) Central limit theorems and uniform laws of large numbers for arrays of random fields. J Econom 150:86–98

    Article  MathSciNet  PubMed  PubMed Central  Google Scholar 

  • Jenish N, Prucha IR (2012) On spatial processes and asymptotic inference under near-epoch dependence. J Econom 170:178–190

    Article  MathSciNet  PubMed  PubMed Central  Google Scholar 

  • ** F, Lee L-F (2020) Asymptotic properties of a spatial autoregressive stochastic frontier model. J Spat Econom 1, 2. https://doi.org/10.1007/s43071-020-00002-z

  • ** F, Wang Y (2023) Consistent non-Gaussian pseudo maximum likelihood estimations of spatial autoregressive models. Economet Theor 0:1–39

    Google Scholar 

  • Kelejian H, Piras G (2017) Spatial econometrics. Academic Press

  • Kelejian H, Prucha I (1998) A generalized spatial two-stage least squares procedure for estimating a spatial autoregressive model with autoregressive disturbances. J Real Estate Finance Econ 17:99–121

    Article  Google Scholar 

  • Kopp RJ, Mullahy J (1990) Moment-based estimation and testing of stochastic frontier models. J Econom 46:165–183

    Article  Google Scholar 

  • Kumbhakar SC, Lovell CAK (2000) Stochastic Frontier Analysis. Cambridge University Press, Cambridge

  • Kutlu L (2018) Estimating efficiency in a spatial autoregressive stochastic frontier model. Econom Lett 163:155–157

    Article  Google Scholar 

  • Kutlu L (2022) Spatial stochastic frontier model with endogenous weighting matrix. Empir Econ 63:1947–1968

    Article  Google Scholar 

  • Kutlu L (2023) Calculating efficiency for spatial autoregressive stochastic frontier model. Econom Lett 225, 111043

  • Kutlu L, Tran KC, Tsionas MG (2019) A time-varying true individual effects model with endogenous regressors. J Econom 211:539–559

    Article  MathSciNet  Google Scholar 

  • Kutlu L, Tran KC, Tsionas MG (2020) A spatial stochastic frontier model with endogenous frontier and environmental variables. Eur J Oper Res 286:389–399

    Article  MathSciNet  Google Scholar 

  • Lai H-p, Tran KC (2022) Persistent and transient inefficiency in a spatial autoregressive panel stochastic frontier model. J Prod Anal 58:1–13

    Article  Google Scholar 

  • Lee L-F (2004) Asymptotic distributions of quasi-maximum likelihood estimators for spatial autoregressive models. Econometrica 72:1899–1925

    Article  MathSciNet  Google Scholar 

  • Lee L-F (2007) GMM and 2SLS estimation of mixed regressive, spatial autoregressive models. J Econom 137:489–514

    Article  MathSciNet  Google Scholar 

  • Lee L-F, Yu J (2010) Estimation of spatial autoregressive panel data models with fixed effects. J Econom 154:165–185

    Article  MathSciNet  Google Scholar 

  • Liu T, Lee L-F (2019) A likelihood ratio test for spatial model selection. J Econom 213:434–458

    Article  MathSciNet  Google Scholar 

  • Meeusen W, van den Broeck J (1977) Efficiency estimation from Cobb-Douglas production functions with composed error. Internat Econom Rev 18:435–444

    Article  Google Scholar 

  • Meintanis SG, Papadimitriou CK (2022) Goodness–of–fit tests for stochastic frontier models based on the characteristic function. J Prod Anal 57:285–296

    Article  Google Scholar 

  • Neyman J, Scott E (1948) Consistent estimation from partially consistent observations. Econometrica 16:1–32

    Article  MathSciNet  Google Scholar 

  • Papadopoulos A, Parmeter CF (2023) The wrong skewness problem in stochastic frontier analysis: a review. J Prod Anal. https://doi.org/10.1007/s11123-023-00708-w

  • Parmeter CF, Sickles RC (2021) Advances in Efficiency and Productivity Analysis. Springer Proceedings in Business and Economics. Springer, Cham, Switzerland

  • Schmidt P, Lin T-F (1984) Simple tests of alternative specifications in stochastic frontier models. J Econom 24:349–361

    Article  Google Scholar 

  • Sickles RC, Zelenyuk V (2019) Measurement of Productivity and Efficiency. Cambridge Books. University Press, Cambridge, UK

  • Smith MD (2008) Stochastic frontier models with dependent error components. Economet J 11:172–192

    Article  Google Scholar 

  • Stevenson R (1980) Likelihood functions for generalized stochastic frontier estimation. J Econom 13:57–66

    Article  Google Scholar 

  • Wang HJ, Ho CW (2010) Estimating fixed-effect panel stochastic frontier models by model transformation. J Econom 157:286–296

    Article  MathSciNet  Google Scholar 

  • Wang WS, Amsler C, Schmidt P (2011) Goodness of fit tests in stochastic frontier models. J Prod Anal 35:95–118

    Article  CAS  Google Scholar 

  • Xu X, Lee L-F (2015) Maximum likelihood estimation of a spatial autoregressive Tobit model. J Econom 188:164–280

    Article  MathSciNet  Google Scholar 

  • Xu X, Lee L-F (2018) Sieve maximum likelihood estimation of the spatial autoregressive Tobit model. J Econom 203:96–112

    Article  MathSciNet  Google Scholar 

  • Yu J, Jong RD, Lee L-F (2008) Quasi-maximum likelihood estimators for spatial dynamic panel data with fixed effects when both n and T are large. J Econom 146:118–134

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

This work is supported by the National Natural Science Foundation of China (No. 71971062), and the Excellent Young Scholar Foundation of UIBE (No. 21JQ01).

We are very grateful to the editor, Prof. Christopher Parmeter, and two anonymous reviewers for their valuable comments and suggestions, which have important guiding significance to our work.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Levent Kutlu.

Appendices

Appendix A. Useful definition and lemma

This appendix lists some useful definitions and lemmas.

Definition 1

(Lp-norm and Lp-NED). For any random variable t with a finite pth absolute moment, where p≥1, denote its Lp-norm by \(\parallel t{\parallel }_{p}={\left[{{{\rm{E}}}}| t{| }^{p}\right]}^{1/p}\). Let \(g=\left\{{g}_{ni},i\in {D}_{n},n\ge 1\right\}\) and \(v=\left\{{v}_{ni},i\in {D}_{n},n\ge 1\right\}\) be two random fields, where Dn satisfies Assumption 3. Assume that \(\mathop{\sup }\nolimits_{i,n}{\left\Vert {g}_{ni}\right\Vert }_{p} < \infty\). The random field g is said to be Lp-NED on v if

$${\left\Vert {g}_{ni}-{{{\rm{E}}}}\left({g}_{ni}| {{{{\mathcal{F}}}}}_{ni}(s)\right)\right\Vert }_{p}\le {d}_{ni}\psi (s)$$

for some arrays of finite positive constants \(\left\{{d}_{ni},i\in {D}_{n},n\ge 1\right\}\) and for some sequence ψ(s) ≥ 0 such that \(\mathop{\lim }\nolimits_{s\to \infty }\psi (s)=0\), where \({{{{\mathcal{F}}}}}_{ni}(s)\) is the σ-field generated by the random variables vnj ’s with units j ’s located within the ball Bi(s) that is centered at i with radius s, and ψ(s) is the NED coefficient. If we further have \(\mathop{\sup }\nolimits_{n}\mathop{\sup }\nolimits_{i\in {D}_{n}}{d}_{ni}\, < \,\infty\), then g is said to be uniformly Lp-NED on v.

Lemma 1

(Theorem 1 in Jenish and Prucha 2012). Let \(\left\{{D}_{n}\right\}\) be a sequence of arbitrary finite subsets of D such that \(\left\vert {D}_{n}\right\vert \to \infty\) as n → , where \(D\subset {{\mathbb{R}}}^{d},d\ge 1\) is as in Assumption 3, and let Tn be a sequence of subsets of D such that Dn ⊆ Tn. Suppose further that \(Z=\left\{{Z}_{i,n},i\in {D}_{n},n\ge 1\right\}\) is L1-NED on \(\varepsilon =\left\{{\varepsilon }_{i,n},i\in {T}_{n},n\ge 1\right\}\) with the scaling factors di,n. Suppose that Z and ε satisfy: (a) there exist nonrandom positive constants \(\left\{{c}_{i,n},i\in {D}_{n},n\ge 1\right\}\) such that Zi,n/ci,n is uniformly Lp-bounded for some p > 1, i.e., \(\mathop{\sup }\nolimits_{n}\mathop{\sup }\nolimits_{i\in {D}_{n}}E{\left\vert {Z}_{i,n}/{c}_{i,n}\right\vert }^{p}\, < \,\infty\), and (b) the α-mixing coefficients of the input field ε satisfy \(\bar{\alpha }(u,r,v)\le \varphi (u,v)\hat{\alpha }(r)\) for some function φ(u, v) which is nondecreasing in each argument, and some \(\widehat{\alpha }(r)\) such that \(\mathop{\sum }\nolimits_{r = 1}^{\infty }{r}^{d-1}\widehat{\alpha }(r) < \infty\). Then

$$\frac{1}{{M}_{n}\left\vert {D}_{n}\right\vert }\mathop{\sum}\limits_{i\in {D}_{n}}\left({Z}_{i,n}-{{{\rm{E}}}}{Z}_{i,n}\right)\mathop{\to }\limits^{{L}_{1}}0,$$

where \({M}_{n}=\mathop{\max }\limits_{i\in {D}_{n}}\max \left({c}_{i,n},{d}_{i,n}\right)\).

Lemma 2

(Proposition 1 in Xu and Lee 2015). Suppose the parameter space Θ of θ is compact. Consider the SAR model Yn = λWnYn + Xnβ + Vn.

  1. (1)

    Under Assumption 2, if \(\mathop{\sup }\nolimits_{1\leqslant\, j\,\leqslant k,i,n}{{{\rm{E}}}}{\left\vert {x}_{ij,n}\right\vert }^{p}\, < \,\infty\) and \(\mathop{\sup }\nolimits_{i,n}{{{\rm{E}}}}{\left\vert {\epsilon }_{ni}\right\vert }^{p}\, < \,\infty\) for some p ⩾ 1, then \({\left\{{y}_{ni}\right\}}_{i = 1}^{n}\) and \({\left\{{{{{\bf{w}}}}}_{i,n}{Y}_{n}\right\}}_{i = 1}^{n}\) are uniformly Lp bounded.

  2. (2)

    Under Assumptions 3-4(a) and 5, \({\left\{{y}_{ni}\right\}}_{i = 1}^{n}\) is geometrically L2-NED on \({\left\{{{{{\bf{x}}}}}_{i,n},{v}_{ni}\right\}}_{i = 1}^{n}:{\left\Vert {y}_{ni}-{{{\rm{E}}}}\left[{y}_{ni}| {{{{\mathcal{F}}}}}_{ni}(s)\right]\right\Vert }_{2}\leqslant C({\zeta }^{1/{\bar{d}}_{0}})^{s}\) for some C > 0 that does not depend on i and n. The same conclusion also holds for \({\left\{{{{{\bf{w}}}}}_{i,,n}{Y}_{n}\right\}}_{i = 1}^{n}\).

  3. (3)

    Under Assumptions 3-4(b) and 5, \({\left\{{y}_{ni}\right\}}_{i = 1}^{n}\) is L2-NED on \({\left\{{{{{\bf{x}}}}}_{i,n},{v}_{ni}\right\}}_{i = 1}^{n}:{\left\Vert {y}_{ni}-{{{\rm{E}}}}\left[{y}_{ni}| {{{{\mathcal{F}}}}}_{ni}(s)\right]\right\Vert }_{2}\leqslant C/{s}^{\alpha -d}\) for some C > 0 that does not depend on i and n. The same conclusion also holds for \({\left\{{{{{\bf{w}}}}}_{i,n}{Y}_{n}\right\}}_{i = 1}^{n}\).

Lemma 3

(Generalization of Corollary 4.3(b) in Gallant and White 1988). If \(\left\{{Y}_{i,n}\right\}\) and \(\left\{{Z}_{i,n}\right\}\) are both uniformly L2r bounded for some r > 2, and uniformly and geometrically L2-NED, then \(\left\{{Y}_{i,n}{Z}_{i,n}\right\}\) is uniformly and geometrically L2-NED.

Appendix B. Technical details in Section 2 and 3

This appendix provides some expressions in Section 2 and main proofs for the results in Section 3.

  • Π in Assumption 2 by the best GMME, or by the QMLE under H0

    For the best GMME \({\tilde{\theta }}_{n,bgmme}^{{\dagger} }\),

    $${\Pi }^{{\dagger} -1}=\mathop{\lim }\limits_{n\to \infty }\frac{1}{n}{{{\rm{E}}}}\left[\begin{array}{rr}{{{\rm{tr}}}}({G}_{n1}^{ds}{G}_{n})+\frac{1}{{\sigma }^{{\dagger} 2}}{({G}_{n}{X}_{n}^{{\dagger} }{\beta }^{{\dagger} })}^{{\prime} }({G}_{n}{X}_{n}^{{\dagger} }{\beta }^{{\dagger} })&\frac{1}{{\sigma }^{{\dagger} 2}}{({G}_{n}{X}_{n}^{{\dagger} }{\beta }^{{\dagger} })}^{{\prime} }{X}_{n}^{{\dagger} }\\ \frac{1}{{\sigma }^{{\dagger} 2}}{X}_{n}^{{\dagger} {\prime} }({G}_{n}{X}_{n}^{{\dagger} }{\beta }^{{\dagger} })&\frac{1}{{\sigma }^{{\dagger} 2}}{X}_{n}^{{\dagger} {\prime} }{X}_{n}^{{\dagger} }\end{array}\right],$$
    (B.1)

    where \({G}_{n1}^{ds}=[{G}_{n}-{{{\rm{Diag}}}}({G}_{n})]+{[{G}_{n}-{{{\rm{Diag}}}}({G}_{n})]}^{{\prime} }\), and Diag(Gn) is the diagonal matrix generated by the diagonal elements of Gn. Under H0, for the QMLE \({\tilde{\theta }}_{n,mle}^{{\dagger} }\),

    $${\Pi }^{{\dagger} -1}=\mathop{\lim }\limits_{n\to \infty }\frac{1}{n}{{{\rm{E}}}}\left[\begin{array}{rr}{{{\rm{tr}}}}({G}_{n2}^{ds}{G}_{n})+\frac{1}{{\sigma }^{{\dagger} 2}}{({G}_{n}{X}_{n}^{{\dagger} }{\beta }^{{\dagger} })}^{{\prime} }({G}_{n}{X}_{n}^{{\dagger} }{\beta }^{{\dagger} })&\frac{1}{{\sigma }^{{\dagger} 2}}{({G}_{n}{X}_{n}^{{\dagger} }{\beta }^{{\dagger} })}^{{\prime} }{X}_{n}^{{\dagger} }\\ \frac{1}{{\sigma }^{{\dagger} 2}}{X}_{n}^{{\dagger} {\prime} }({G}_{n}{X}_{n}^{{\dagger} }{\beta }^{{\dagger} })&\frac{1}{{\sigma }^{{\dagger} 2}}{X}_{n}^{{\dagger} {\prime} }{X}_{n}^{{\dagger} }\end{array}\right],$$
    (B.2)

    where \({G}_{n2}^{ds}=[{G}_{n}-({{{\rm{tr}}}}({G}_{n})/n){I}_{n}]+[{G}_{n}-({{{\rm{tr}}}}({G}_{n})/n){I}_{n}]^{{\prime} }\). No matter in (B.1) or (B.2), the following fact holds.

    Fact: the second column of Π†−1 is \((1/\sigma^{\dagger 2}){\mathbf{h}}_2^{\dagger \prime} = (1/\sigma^{\dagger 2})\lim_{n\to\infty} n^{-1} {\mathrm{E}} [{\mathbf{1}}_n^\prime G_n X_n^{\dagger} \beta^{\dagger}, {\mathbf{1}}_n^\prime X_n^{\dagger}]^\prime\).

    The above fact is useful in Corollary 1, and more details can be found in the proof of the corollary.

  • Elements of \({\Sigma }_{n}^{{\dagger} }\) and Σn in Proposition 2

    \({\Sigma }_{n}^{{\dagger} }\) is a symmetric 3 × 3 matrix, whose (i, j)th element is denoted as \({\delta }_{ij,n}^{{\dagger} }\). Recall that \({\sigma }^{{\dagger} 2}={{{\rm{E}}}}[{\epsilon }_{ni}^{{\dagger} 2}],{\mu }_{s}^{{\dagger} }={{{\rm{E}}}}[{\epsilon }_{ni}^{{\dagger} s}]\) for s = 3, 4, 5, 6, and \({{{{\bf{h}}}}}_{1}^{{\dagger} }\) and \({{{{\bf{h}}}}}_{2}^{{\dagger} }\) are defined in (3.2). For the diagonal elements of Σn,

    $$\begin{array}{ll}{\delta }_{11,n}^{{\dagger}}={\sigma }^{{\dagger} 2}+{{{{\bf{h}}}}}_{2}^{{\dagger} }{\Pi }^{{\dagger} }{{{{\bf{h}}}}}_{2}^{{\dagger} {\prime} }-2{{{{\bf{h}}}}}_{2}^{{\dagger} }{C}_{\epsilon ,1n},\\{\delta }_{22,n}^{{\dagger} }={\mu }_{4}^{{\dagger} }-{\sigma }^{{\dagger} 4}+4{\sigma }^{{\dagger} 4}{{{{\bf{h}}}}}_{1}^{{\dagger} }{\Pi }^{{\dagger} }{{{{\bf{h}}}}}_{1}^{{\dagger} {\prime} }-4{\sigma }^{{\dagger} 2}{{{{\bf{h}}}}}_{1}^{{\dagger} }{C}_{\epsilon ,2n},\\ {\delta }_{33,n}^{{\dagger} }={\mu }_{6}^{{\dagger} }-{\mu }_{3}^{{\dagger} 2}-6({\sigma }^{{\dagger} 2}{{{{\bf{h}}}}}_{2}^{{\dagger} }+{\mu }_{3}^{{\dagger} }{{{{\bf{h}}}}}_{1}^{{\dagger} }){{{{\bf{h}}}}}_{2}^{{\dagger} }{C}_{\epsilon ,3n}\\\qquad+9({\sigma }^{{\dagger} 2}{{{{\bf{h}}}}}_{2}^{{\dagger} }+{\mu }_{3}^{{\dagger} }{{{{\bf{h}}}}}_{1}^{{\dagger} }){\Pi }^{{\dagger} }{({\sigma }^{{\dagger} 2}{{{{\bf{h}}}}}_{2}^{{\dagger} }+{\mu }_{3}^{{\dagger} }{{{{\bf{h}}}}}_{1}^{{\dagger} })}^{{\prime} }.\end{array}$$
    (B.3)

    For the off-diagonal elements,

    $$\begin{array}{l}{\delta }_{21,n}^{{\dagger} }={\delta }_{12,n}^{{\dagger} }={\mu }_{3}^{{\dagger} }-{{{{\bf{h}}}}}_{2}^{{\dagger} }{C}_{\epsilon ,2n}-2{\sigma }^{{\dagger} 2}{{{{\bf{h}}}}}_{1}^{{\dagger} }{C}_{\epsilon ,1n}+2{\sigma }^{{\dagger} 2}{{{{\bf{h}}}}}_{1}^{{\dagger} }{\Pi }^{{\dagger} }{{{{\bf{h}}}}}_{2}^{{\dagger} {\prime} },\\ {\delta }_{31,n}^{{\dagger} }={\delta }_{13,n}^{{\dagger} }={\mu }_{4}^{{\dagger} }+3({\sigma }^{{\dagger} 2}{{{{\bf{h}}}}}_{2}^{{\dagger} }+{\mu }_{3}^{{\dagger} }{{{{\bf{h}}}}}_{1}^{{\dagger} }){\Pi }^{{\dagger} }{{{{\bf{h}}}}}_{2}^{{\dagger} {\prime} }-{{{{\bf{h}}}}}_{2}^{{\dagger} }{C}_{\epsilon ,3n}-3({\sigma }^{{\dagger} 2}{{{{\bf{h}}}}}_{2}^{{\dagger} }+{\mu }_{3}^{{\dagger} }{{{{\bf{h}}}}}_{1}^{{\dagger} }){C}_{\epsilon ,1n},\\ {\delta }_{32,n}^{{\dagger} }={\delta }_{23,n}^{{\dagger} }={\mu }_{5}^{{\dagger} }-{\mu }_{3}^{{\dagger} }{\sigma }^{{\dagger} 2}-2{\sigma }^{{\dagger} 2}{{{{\bf{h}}}}}_{1}^{{\dagger} }{C}_{\epsilon ,3n}-3({\sigma }^{{\dagger} 2}{{{{\bf{h}}}}}_{2}^{{\dagger} }+{\mu }_{3}^{{\dagger} }{{{{\bf{h}}}}}_{1}^{{\dagger} }){C}_{\epsilon ,2n}+6{\sigma }^{{\dagger} 2}{{{{\bf{h}}}}}_{1}^{{\dagger} }{\Pi }^{{\dagger} }({\sigma }^{{\dagger} 2}{{{{\bf{h}}}}}_{2}^{{\dagger} }+{\mu }_{3}^{{\dagger} }{{{{\bf{h}}}}}_{1}^{{\dagger} })^{{\prime} }.\end{array}$$
    (B.4)

    Here, \({C}_{\epsilon ,jn}={{{\rm{E}}}}[{n}^{-1/2}\mathop{\sum }\nolimits_{i = 1}^{n}{\epsilon }_{ni}^{{\dagger} j}\cdot {n}^{1/2}({\tilde{\theta }}_{n}^{{\dagger} }-{\theta }^{{\dagger} })]\) as the (k + 1)-dimensional expectation vector for any j ∈ {1, 2, 3}. Under \({H}_{0},{\mu }_{3}^{{\dagger} }={\mu }_{5}^{{\dagger} }=0\), then \({\Sigma }_{n}^{{\dagger} }\) is reducible to Σn.

Proof of Proposition 1.

The probability limit of \({n}^{-1}\mathop{\sum }\nolimits_{i = 1}^{n}{\tilde{\epsilon }}_{ni}^{{\dagger} }\) is shown in p. 5–6. We now only analyze that of \({n}^{-1}\mathop{\sum }\nolimits_{i = 1}^{n}{\tilde{\epsilon }}_{ni}^{{\dagger} 3}\), and \({n}^{-1}\mathop{\sum }\nolimits_{i = 1}^{n}{\tilde{\epsilon }}_{ni}^{{\dagger} 2}-{\sigma }^{{\dagger} 2}={o}_{p}(1)\) can be proved by a similar argument. By the expansion, it follows that

$$\begin{array}{rl}&\frac{1}{n}\mathop{\sum }\limits_{i=1}^{n}{\tilde{\epsilon }}_{ni}^{{\dagger} 3}-\frac{1}{n}\mathop{\sum }\limits_{i=1}^{n}{\epsilon }_{ni}^{{\dagger} 3}\\=&\frac{3}{n}\mathop{\sum }\limits_{i=1}^{n}{\epsilon }_{ni}^{{\dagger} 2}{{{{\bf{z}}}}}_{i,n}^{{\prime} }({\theta }^{{\dagger} }-{\tilde{\theta }}_{n}^{{\dagger} })+\frac{3}{n}\mathop{\sum }\limits_{i=1}^{n}{\epsilon }_{ni}^{{\dagger} }[{{{{\bf{z}}}}}_{i,n}^{{\prime} }({\theta }^{{\dagger} }-{\tilde{\theta }}_{n}^{{\dagger} })]^{2}\\&+\frac{1}{n}\mathop{\sum }\limits_{i=1}^{n}[{{{{\bf{z}}}}}_{i,n}^{{\prime} }({\theta }^{{\dagger} }-{\tilde{\theta }}_{n}^{{\dagger} })]^{3}.\end{array}$$
(B.5)

Under Assumptions 1–5, {zi,n} is uniformly Lr-norm bounded for some r = 4 + ι > 4 and is uniformly L2-NED on \({\{{{{{\bf{x}}}}}_{i,n}^{{\dagger} },{\epsilon }_{ni}^{{\dagger} }\}}_{i = 1}^{n}\), as the argument below (3.1). Assumption 1 tells us that E∣vni6+ι <  for some ι > 0, so {ϵni} must be uniformly Lr-norm bounded for some r > 4 under H0. Hence, according to Lemma 3 (in Appendix A), \(\{{\epsilon }_{ni}^{{\dagger} }{{{{\bf{z}}}}}_{i,n}\}\) is uniformly L2-NED on \({\{{{{{\bf{x}}}}}_{i,n}^{{\dagger} },{\epsilon }_{ni}^{{\dagger} }\}}_{i = 1}^{n}\) under H0. Similarly, under H0, all of \({\epsilon }_{ni}^{{\dagger} 2}{{{{\bf{z}}}}}_{i,n},{\epsilon }_{ni}^{{\dagger} }{{{{\bf{z}}}}}_{i,n}{{{{\bf{z}}}}}_{i,n}^{{\prime} }\) and \(\{{z}_{ij,n}{{{{\bf{z}}}}}_{i,n}{{{{\bf{z}}}}}_{i,n}^{{\prime} }\}\) are also uniformly L2-NED on \({\{{{{{\bf{x}}}}}_{i,n}^{{\dagger} },{\epsilon }_{ni}^{{\dagger} }\}}_{i = 1}^{n}\). Besides, let hn,ij be either 1 or any element of \({{{{\bf{z}}}}}_{i,n}^{{\prime} }\) for j = 1, 2, 3. According to the generalized Hölder’s inequality, under Assumptions 1–5,

$$\begin{array}{l}\quad\mathop{\sup }\limits_{i,n}{{{\rm{E}}}}| {h}_{n,i1}{h}_{n,i2}{h}_{n,i3}| \\\le \mathop{\sup }\limits_{i,n}\left\{{({{{\rm{E}}}}| {h}_{n,i1}{| }^{3})}^{1/3}{({{{\rm{E}}}}| {h}_{n,i2}{| }^{3})}^{1/3}{({{{\rm{E}}}}| {h}_{n,i3}{| }^{3})}^{1/3}\right\}\, < \,\infty\end{array}$$

and

$$\begin{array}{l}\quad\mathop{\sup }\limits_{i,n}{{{\rm{E}}}}| {v}_{ni}^{k}{h}_{n,i1}{h}_{n,i2}| \\\le \mathop{\sup }\limits_{i,n}\left\{{({{{\rm{E}}}}| {v}_{ni}{| }^{3k})}^{1/3}{({{{\rm{E}}}}| {h}_{n,i1}{| }^{3})}^{1/3}{({{{\rm{E}}}}| {h}_{n,i2}{| }^{3})}^{1/3}\right\}\, < \,\infty\end{array}$$

hold for k = 0, 1, 2. By making use of the LLN in Jenish and Prucha (2012) again, \({n}^{-1}\mathop{\sum }\nolimits_{i = 1}^{n}{\epsilon }_{ni}^{{\dagger} 2}{{{{\bf{z}}}}}_{i,n}^{{\prime} }={{{\rm{E}}}}[{n}^{-1}\mathop{\sum }\nolimits_{i = 1}^{n}{\epsilon }_{ni}^{{\dagger} 2}{{{{\bf{z}}}}}_{i,n}^{{\prime} }]+{o}_{p}(1)={O}_{p}(1)\). Similarly, \({n}^{-1}\mathop{\sum }\nolimits_{i = 1}^{n}{\epsilon }_{ni}^{{\dagger} }{{{{\bf{z}}}}}_{i,n}{{{{\bf{z}}}}}_{i,n}^{{\prime} }={n}^{-1}\mathop{\sum }\nolimits_{i = 1}^{n}{{{{\bf{z}}}}}_{i,n}{{{{\bf{z}}}}}_{i,n}^{{\prime} }{z}_{n,ij}={O}_{p}(1)\). Working with \({\theta }^{{\dagger} }-{\tilde{\theta }}_{n}^{{\dagger} }={o}_{p}(1)\), it yields that \({n}^{-1}\mathop{\sum }\nolimits_{i = 1}^{n}{\epsilon }_{ni}^{{\dagger} 2}{{{{\bf{z}}}}}_{i,n}^{{\prime} }({\theta }^{{\dagger} }-{\tilde{\theta }}_{n}^{{\dagger} })={o}_{p}(1),{n}^{-1}\mathop{\sum }\nolimits_{i = 1}^{n}{\epsilon }_{ni}^{{\dagger} }{[{{{{\bf{z}}}}}_{i,n}^{{\prime} }({\theta }^{{\dagger} }-{\tilde{\theta }}_{n}^{{\dagger} })]}^{2}={o}_{p}(1)\), and \({n}^{-1}\mathop{\sum }\nolimits_{i = 1}^{n}{[{{{{\bf{z}}}}}_{i,n}^{{\prime} }({\theta }^{{\dagger} }-{\tilde{\theta }}_{n}^{{\dagger} })]}^{3}={o}_{p}(1)\). Therefore, \(\frac{1}{n}\mathop{\sum }\nolimits_{i = 1}^{n}{\tilde{\epsilon }}_{ni}^{{\dagger} 3}-\frac{1}{n}\mathop{\sum }\nolimits_{i = 1}^{n}{\epsilon }_{ni}^{{\dagger} 3}=\frac{1}{n}\mathop{\sum }\nolimits_{i = 1}^{n}{\tilde{\epsilon }}_{ni}^{{\dagger} 3}-{\mu }_{3}^{{\dagger} }={o}_{p}(1)\).

Under H0, by the Lindeberg-Lévy CLT, \({n}^{-1}\mathop{\sum }\nolimits_{i = 1}^{n}{\epsilon }_{ni}^{{\dagger} 3}={O}_{p}({n}^{-1/2})\) since \({{{\rm{E}}}}[{\epsilon }_{ni}^{{\dagger} 3}]={{{\rm{E}}}}[{v}_{ni}^{3}]=0\). The proof is accomplished. □

Proof of Proposition 2.

To check the correctness of (3.3), we take the term regarding \({n}^{-1}{\sum \nolimits_{i = 1}^{n}}{\tilde{\epsilon}}_{ni}^{{\dagger}3}\) as the example. Applying the mean-value theorem to \({n}^{-1}{\sum \nolimits_{i = 1}^{n}}{\tilde{\epsilon }}_{ni}^{{\dagger} 3}={n}^{-1}{\sum \nolimits_{i = 1}^{n}}({y}_{ni}-{\bf{z}}_{i,n}^{\prime}{\tilde{\theta}}_{n}^{{\dagger} })^{3}\triangleq \,{n}^{-1}{\sum \nolimits_{i = 1}^{n}}{\epsilon }_{ni}^{{\dagger} 3}({\tilde{\theta }}_{n}^{{\dagger}})\) at θ, it yields that

$$\begin{array}{ll}\frac{1}{n}\mathop{\sum }\limits_{i=1}^{n}{\tilde{\epsilon }}_{ni}^{{\dagger} 3}=\frac{1}{n}\mathop{\sum }\limits_{i=1}^{n}{\epsilon }_{ni}^{{\dagger} 3}-\frac{3}{{n}^{3/2}}\mathop{\sum }\limits_{i=1}^{n}{\epsilon }_{ni}^{{\dagger} 2}({\bar{\theta }}_{n}^{{\dagger} })[{{{{\bf{w}}}}}_{i,n}{Y}_{n},{{{{\bf{x}}}}}_{i,n}^{{\dagger} {\prime} }]\cdot \sqrt{n}({\tilde{\theta }}_{n}-{\theta }^{{\dagger} })\\ =\frac{1}{n}\mathop{\sum }\limits_{i=1}^{n}{\epsilon }_{ni}^{{\dagger} 3}-\frac{3}{{n}^{1/2}}\left[\frac{1}{n}\mathop{\sum }\limits_{i=1}^{n}{\epsilon }_{ni}^{{\dagger} 2}({\bar{\theta }}_{n}^{{\dagger} }){{{{\bf{w}}}}}_{i,n}{Y}_{n},\frac{1}{n}\mathop{\sum }\limits_{i=1}^{n}{\epsilon }_{ni}^{{\dagger} 2}({\bar{\theta }}_{n}^{{\dagger} }){{{{\bf{x}}}}}_{i,n}^{{\dagger} {\prime} }\right]\cdot \sqrt{n}({\tilde{\theta }}_{n}-{\theta }^{{\dagger} }),\end{array}$$

where \({\bar{\theta }}_{n}^{{\dagger} }\) satisfies \(\parallel {\bar{\theta }}_{n}^{{\dagger} }-{\theta }^{{\dagger} }\parallel \le \parallel {\tilde{\theta }}_{n}^{{\dagger} }-{\theta }^{{\dagger} }\parallel\) and therefore, \({\bar{\theta }}_{n}^{{\dagger} }-{\theta }^{{\dagger} }={o}_{p}(1)\). By expansion we can obtain

$$\frac{1}{n}\mathop{\sum }\limits_{i=1}^{n}{\epsilon }_{ni}^{{\dagger} 2}({\bar{\theta }}_{n}^{{\dagger} }){{{{\bf{w}}}}}_{i,n}{Y}_{n}=\frac{1}{n}\mathop{\sum }\limits_{i=1}^{n}{\epsilon }_{ni}^{{\dagger} 2}{{{{\bf{w}}}}}_{i,n}{Y}_{n}+\frac{1}{n}\mathop{\sum }\limits_{i=1}^{n}[{{{{\bf{z}}}}}_{i,n}^{{\prime} }({\theta }^{{\dagger} }-{\bar{\theta }}_{n}^{{\dagger} })]^{2}{{{{\bf{w}}}}}_{i,n}{Y}_{n}+\frac{2}{n}\mathop{\sum }\limits_{i=1}^{n}{\epsilon }_{ni}^{{\dagger} }{{{{\bf{z}}}}}_{i,n}^{{\prime} }({\theta }^{{\dagger} }-{\bar{\theta }}_{n}^{{\dagger} }){{{{\bf{w}}}}}_{i,n}{Y}_{n}.$$

As mentioned above, any element of \(\{{{{{\bf{z}}}}}_{i,n}:{{{{\bf{z}}}}}_{i,n}^{{\prime} }=[{{{{\bf{w}}}}}_{i,n}{Y}_{n},{{{{\bf{x}}}}}_{i,n}^{{\dagger} {\prime} }]\}\) are uniformly L2r-norm bounded for some r > 2, and uniformly L2-NED on \(\{{{{{\bf{x}}}}}_{i,n}^{{\dagger} {\prime} },{\epsilon }_{ni}^{{\dagger} }\}\). Based on this, we can further conclude that \(\{{\epsilon }_{ni}^{{\dagger} 2}{{{{\bf{w}}}}}_{i,n}{Y}_{n}\}\) is uniformly L4-norm bounded, and uniformly L2-NED on \(\{{{{{\bf{x}}}}}_{i,n}^{{\dagger} {\prime} },{\epsilon }_{ni}^{{\dagger} }\}\), thereby \({n}^{-1}\mathop{\sum }\nolimits_{i = 1}^{n}{\epsilon }_{ni}^{{\dagger} 2}{{{{\bf{w}}}}}_{i,n}{Y}_{n}-{\sigma }^{{\dagger} 2}{{{\rm{E}}}}[{n}^{-1}{{{{\bf{1}}}}}_{n}^{{\prime} }{G}_{n}{X}_{n}^{{\dagger} }{\beta }^{{\dagger} }]-{\mu }_{3}^{{\dagger} }{n}^{-1}{{{\rm{tr}}}}({G}_{n})={o}_{p}(1)\) by the LLN in Jenish and Prucha (2012) and the fact that \({{{\rm{E}}}}[{n}^{-1}\mathop{\sum }\nolimits_{i = 1}^{n}{\epsilon }_{ni}^{{\dagger} 2}{{{{\bf{w}}}}}_{i,n}{Y}_{n}]={{{\rm{E}}}}[{n}^{-1}\mathop{\sum }\nolimits_{i = 1}^{n}{\epsilon }_{ni}^{{\dagger} 2}{{{{\bf{w}}}}}_{i,n}{S}^{-1}{X}_{n}^{{\dagger} }{\beta }^{{\dagger} }]+{{{\rm{E}}}}[{n}^{-1}\mathop{\sum }\nolimits_{i = 1}^{n}{\epsilon }_{ni}^{{\dagger} 2}{{{{\bf{w}}}}}_{i,n}{S}^{-1}{\epsilon }_{n}^{{\dagger} }]={\sigma }^{{\dagger} 2}{{{\rm{E}}}}[{n}^{-1}{{{{\bf{1}}}}}_{n}^{{\prime} }{G}_{n}{X}_{n}^{{\dagger} }{\beta }^{{\dagger} }]+{\mu }_{3}^{{\dagger} }{n}^{-1}{{{\rm{tr}}}}({G}_{n})\). Ditto for \({n}^{-1}\mathop{\sum }\nolimits_{i = 1}^{n}{z}_{ij,n}^{{\prime} }{{{{\bf{w}}}}}_{i,n}{Y}_{n}\) and \({n}^{-1}\mathop{\sum }\nolimits_{i = 1}^{n}{\epsilon }_{ni}^{{\dagger} }{z}_{ij,n}{{{{\bf{w}}}}}_{i,n}{Y}_{n}\). Taking with \({\theta }^{{\dagger} }-{\bar{\theta }}_{n}^{{\dagger} }={o}_{p}(1)\) together, it follows that \({n}^{-1}\mathop{\sum }\nolimits_{i = 1}^{n}{\left[{{{{\bf{z}}}}}_{i,n}^{{\prime} }({\theta }^{{\dagger} }-{\bar{\theta }}_{n}^{{\dagger} })\right]}^{2}{{{{\bf{w}}}}}_{i,n}{Y}_{n}\) and \({n}^{-1}\mathop{\sum }\nolimits_{i = 1}^{n}{\epsilon }_{ni}^{{\dagger} }\left[{{{{\bf{z}}}}}_{i,n}^{{\prime} }({\theta }^{\dagger}{-}{\bar{\theta }}_{n}^{{\dagger} })\right]{{{{\bf{w}}}}}_{i,n}{Y}_{n}\) are op(1). Therefore, \({n}^{-1}\mathop{\sum }\nolimits_{i = 1}^{n}{\epsilon }_{ni}^{{\dagger} 2}({\bar{\theta }}_{n}^{{\dagger} }){{{{\bf{w}}}}}_{i,n}{Y}_{n}{-}{\sigma }^{{\dagger} 2}{{{\rm{E}}}}[{n}^{-1}{{{{\bf{1}}}}}_{n}^{{\prime} }{G}_{n}{X}_{n}^{{\dagger} }{\beta }^{{\dagger} }]={o}_{p}(1)\). By a similar argument for

$$\begin{array}{*{20}{l}}\frac{1}{n}\mathop{\sum }\limits_{i=1}^{n}\epsilon_{ni}^{\dagger 2}(\bar{\theta}_n^{\dagger}){\mathbf{x}}_{i,n}^{\dagger \prime} = \frac{1}{n}\mathop{\sum }\limits_{i = 1}^{n}\epsilon_{ni}^{\dagger 2}{\mathbf{x}}_{i,n}^{\dagger \prime} + \frac{1}{n}\mathop{\sum }\limits_{i = 1}^{n}[{\mathbf{z}}_{i,n}^{\prime}(\theta^{\dagger} - \bar{\theta}_n^{\dagger})]^{2}\\ \times\, {\mathbf{x}}_{i,n}^{\dagger \prime} + \frac{2}{n}\mathop{\sum }\limits_{i = 1}^{n}\epsilon_{ni}^{\dagger}{\mathbf{z}}_{i,n}^{\prime}(\theta^{\dagger} - \bar{\theta}_n^{\dagger}){\mathbf{x}}_{i,n}^{\dagger \prime},\end{array}$$

the last two terms on the right-hand side satisfy \({n}^{-1}\mathop{\sum }\nolimits_{i = 1}^{n}[{{{{\bf{z}}}}}_{i,n}^{{\prime} }({\theta }^{{\dagger} }-{\bar{\theta }}_{n}^{{\dagger} })]^{2}{{{{\bf{x}}}}}_{i,n}^{{\dagger} {\prime} }={o}_{p}(1)\) and \((2/n)\mathop{\sum }\nolimits_{i = 1}^{n}{\epsilon }_{ni}^{{\dagger} }[{{{{\bf{z}}}}}_{i,n}^{{\prime} }({\theta }^{{\dagger} }-{\bar{\theta }}_{n}^{{\dagger} })]{{{{\bf{x}}}}}_{i,n}^{{\dagger} {\prime} }={o}_{p}(1)\). Working with \({n}^{-1}\mathop{\sum }\nolimits_{i = 1}^{n}{\epsilon }_{ni}^{{\dagger} 2}{{{{\bf{x}}}}}_{i,n}^{{\dagger} {\prime} }-{{{\rm{E}}}}[{n}^{-1}{\sigma }^{{\dagger} 2}{{{{\bf{1}}}}}_{n}^{{\prime} }{X}_{n}^{{\dagger} }]={o}_{p}(1)\) by the Chebyshev LLN, it yields that \({n}^{-1}\mathop{\sum }\nolimits_{i = 1}^{n}{\epsilon }_{ni}^{{\dagger} 2}({\bar{\theta }}_{n}^{{\dagger} }){{{{\bf{x}}}}}_{i,n}^{{\dagger} }={{{\rm{E}}}}[{n}^{-1}{\sigma }^{{\dagger} 2}{{{{\bf{1}}}}}_{n}^{{\prime} }{X}_{n}^{{\dagger} }]+{o}_{p}(1)\). Therefore, we obtain

$$\frac{1}{n}\mathop{\sum }\limits_{i=1}^{n}{\tilde{\epsilon }}_{ni}^{{\dagger} 3}=\frac{1}{n}\mathop{\sum }\limits_{i=1}^{n}{\epsilon }_{ni}^{{\dagger} 3}-\frac{3}{\sqrt{n}}({\sigma }^{{\dagger} 2}{{{{\bf{h}}}}}_{2}^{{\dagger} }+{\mu }_{3}^{{\dagger} }{{{{\bf{h}}}}}_{1}^{{\dagger} })\sqrt{n}({\tilde{\theta }}_{n}^{{\dagger} }-{\theta }^{{\dagger} })+{o}_{p}({n}^{-1/2}).$$

Similarly, the other two equalities in (3.3) hold.

To derive the asymptotic distribution of (3.4), it is equivalent to deriving that of

$$\begin{array}{ll}{F}_{n}\triangleq \frac{{c}_{3}}{n}\mathop{\sum }\limits_{i=1}^{n}({\epsilon }_{ni}^{{\dagger} 3}-{\mu }_{3}^{{\dagger} })+\frac{{c}_{2}}{n}\mathop{\sum }\limits_{i=1}^{n}{\epsilon }_{ni}^{{\dagger} 2}+\frac{{c}_{1}}{n}\mathop{\sum }\limits_{i=1}^{n}{\epsilon }_{ni}^{{\dagger} }\\\qquad\,\,\,\,+\,\frac{1}{\sqrt{n}}{{{\bf{f}}}}({\bf{c}})\sqrt{n}({\tilde{\theta }}_{n}^{{\dagger} }-{\theta }^{{\dagger} })\end{array}$$
(B.6)

with f(c) being defined in (3.5), which depends on the limiting distribution of \(\sqrt{n}({\tilde{\theta }}_{n}^{{\dagger} }-{\theta }^{{\dagger} })\). By Assumption 2, we know \({n}^{1/2}({\tilde{\theta }}_{n}^{{\dagger} }-{\theta }^{{\dagger} })={g}_{n}({\theta }^{{\dagger} })+{o}_{p}(1)\), where gn(θ) is quadratic forms in \({\epsilon }_{ni}^{{\dagger} }\). Define \({A}_{n}^{d}={A}_{n}-[{n}^{-1}{{{\rm{tr}}}}({A}_{n})]\cdot {I}_{n}\). Therefore, without lose of generality, we can assume \({g}_{n}({\theta }^{{\dagger} })={n}^{-1/2}D{[{\epsilon }_{n}^{{\dagger} {\prime} }{P}_{1}^{d}{\epsilon }_{n}^{{\dagger} },\cdots ,{\epsilon }_{n}^{{\dagger} {\prime} }{P}_{s}^{d}{\epsilon }_{n}^{{\dagger} },{\epsilon }_{n}^{{\dagger} {\prime} }Q]}^{{\prime} }\in {{\mathbb{R}}}^{k+1}\), where \(D=[{d}_{1},\cdots \,,{d}_{s},{{{{\bf{d}}}}}_{q}]\in {{\mathbb{R}}}^{(k+1)\times (s+q)},P\in {{\mathbb{R}}}^{n\times n}\), and \(Q\in {{\mathbb{R}}}^{n\times q}\) are constant IV matrix with q and s being arbitrary constants. Accordingly, (B.6) becomes

$$\begin{array}{lll}{F}_{n}=\frac{{c}_{3}}{n}\mathop{\sum }\limits_{i=1}^{n}({\epsilon }_{ni}^{{\dagger} 3}-{\mu }_{3}^{{\dagger} })+\frac{{\epsilon }_{n}^{{\dagger} {\prime} }}{n}[{c}_{2}{I}_{n}+{{{\bf{f}}}}({\bf{c}})\mathop{\sum }\limits_{j=1}^{s}{d}_{j}{P}_{j}^{d}]{\epsilon }_{n}^{{\dagger} }\\\qquad\;\;+\,\frac{1}{n}\left[{c}_{1}{\bf{1}}_{n}^{{\prime} }+{{{\bf{f}}}}({\bf{c}}){{{{\bf{d}}}}}_{q}{Q}^{{\prime} }\right]{\epsilon }_{n}^{{\dagger} }\\\quad\;\; \triangleq \frac{{c}_{3}}{n}\mathop{\sum }\limits_{i=1}^{n}({\epsilon }_{ni}^{{\dagger} 3}-{\mu }_{3}^{{\dagger} })+\frac{1}{n}{\epsilon }_{n}^{{\dagger} {\prime} }{A}_{n}{\epsilon }_{n}^{{\dagger} }+\frac{1}{n}{{{{\bf{b}}}}}_{n}^{{\prime} }{\epsilon }_{n}^{{\dagger} }.\end{array}$$

The cubic CLTFootnote 11 tells us that

$$\sqrt{n}{F}_{n} \sim AN\left(0,\mathop{\lim }\limits_{n\to \infty }{{{{\bf{c}}}}}^{{\prime} }{\Sigma }_{n}^{{\dagger} }{{{\bf{c}}}}\right),$$
(B.7)

where \({\Sigma }_{n}^{{\dagger} }\) is defined as in Appendix B. Based on (B.7), taking (3.4), (B.6) and (B.7) together, (3.6) follows from the Slutsky theorem. Under \({H}_{0},{\mu }_{3}^{{\dagger} }={\mu }_{5}^{{\dagger} }=0\) and \({\Sigma }_{n}^{{\dagger} }\) reduces to Σn. The proof is accomplished. □

Proof of Proposition 3.

Employing the delta method for (3.6), we can obtain

$$\begin{array}{ll}&{n}^{1/2}[{S}_{n}({\tilde{\epsilon }}_{n}^{{\dagger} })-{S}^{{\dagger} }]\\=&{n}^{1/2}\left[\psi (\frac{1}{n}\mathop{\sum }\limits_{i=1}^{n}{\tilde{\epsilon }}_{ni}^{{\dagger} },\frac{1}{n}\mathop{\sum }\limits_{i=1}^{n}{\tilde{\epsilon }}_{ni}^{{\dagger} 2},\frac{1}{n}\mathop{\sum }\limits_{i=1}^{n}{\tilde{\epsilon }}_{ni}^{{\dagger} 3})-\psi(0,{\sigma }^{{\dagger} 2},{\mu }_{3}^{{\dagger} })\right]\\ \sim & AN\left(0,\frac{\partial \psi (0,{\sigma }^{{\dagger} 2},{\mu }_{3}^{{\dagger} })}{\partial [a,b,c]}\mathop{\lim }\limits_{n\to \infty }{\Sigma }_{n}^{{\dagger} }\frac{\partial \psi (0,{\sigma }^{{\dagger} 2},{\mu }_{3}^{{\dagger} })}{\partial {[a,b,c]}^{{\prime} }}\right),\end{array}$$

where \((\partial \psi (0,{\sigma }^{{\dagger} 2},{\mu }_{3}^{{\dagger} })/\partial [a,b,c])=[-3/{\sigma }^{{\dagger} },-3{\mu }_{3}^{{\dagger} }/(2{\sigma }^{{\dagger} 5}),1/{\sigma }^{{\dagger} 3}]={{{{\bf{r}}}}}_{1}^{{\prime} }.\)

Further under H0, we have \({S}^{{\dagger} }={\mu }_{3}^{{\dagger} }=0\), and \(\partial \psi (0,{\sigma }^{{\dagger} 2},0)/\partial [a,b,c]=[-3/{\sigma }^{{\dagger} },0,1/{\sigma }^{{\dagger} 3}]={{{{\bf{r}}}}}_{2}^{{\prime} }\). So \({n}^{1/2}{S}_{n}({\tilde{\epsilon }}_{n}^{{\dagger} }) \sim AN\left(0,\mathop{\lim }\nolimits_{n\to \infty }{{{{\bf{r}}}}}_{2}^{{\prime} }{\Sigma }_{n}{{{{\bf{r}}}}}_{2}\right)\). The proof is accomplished. □

Proof of Corollary 1.

It follows from Proposition 3 that under Assumptions 1–5, \({n}^{1/2}[{S}_{n}({\tilde{\epsilon}}_{n}^{\dagger})-{S}^{\dagger}]\cdot ({\bf{r}}_{1}^{\prime} {\Sigma}_{n}^{\dagger}{\bf{r}}_{1})^{-1/2} \,{\sim}\, {AN}\left(0,1\right)\). Then, since \({\tilde{\sigma }}^{{\dagger} }-{\sigma }^{{\dagger} }={o}_{p}(1)\) and \({\tilde{\Sigma}}_{\ast}-{\Sigma }_{n}^{{\dagger}}={o}_{p}(1),{\sqrt{n}}[{S}_{n}({\tilde{\epsilon }}_{n}^{\dagger})-{S}^{\dagger}]\cdot ({{\tilde {\bf{r}}}}_{1}^{\prime}{\tilde{\Sigma}}_{\ast }{{\tilde {\bf{r}}}}_{1})^{-1/2} \,{\sim}\, {AN}(0,1)\) holds by the Slutsky theorem. Under \({H}_{0},{\tilde{{{{\bf{r}}}}}}_{2}={\tilde{{{{\bf{r}}}}}}_{1}\). (3.9) follows from two facts that S = 0, and

$${n}^{1/2}{S}_{n}({\tilde{\epsilon }}_{n}^{{\dagger} })-\frac{n{\sum \nolimits_{i = 1}^{n}}{\tilde{\epsilon }}_{ni}^{{\dagger} 3}}{({\sum \nolimits_{i = 1}^{n}}{\tilde{\epsilon }}_{ni}^{{\dagger} 2})^{3/2}}={o}_{p}(1).$$

We now prove (3.10). As \({\tilde{\epsilon }}_{n}^{{\dagger} }\) is generated from either MLE (Lee 2004, Theorem 3.2) or the best GMM estimation (Lee 2007, Proposition 3) under \({H}_{0},{{{{\bf{h}}}}}_{2}^{{\dagger} }\) is the second column of the matrix σ†2Π†−1 (i.e., the fact below (B.2)), since \({X}_{n}^{{\dagger} }\) must have an intercept (as said in the footnote 3). Then, \((1/{\sigma }^{{\dagger} 2}){\Pi }^{{\dagger} }{{{{\bf{h}}}}}_{2}^{{\dagger} {\prime} }={{{{\bf{e}}}}}_{2,k+1}\), where e2,k+1 is the second column of the identity matrix Ik+1. By direct calculation, we obtain

$${{{{\bf{h}}}}}_{2}^{{\dagger} }{\Pi }^{{\dagger} }{{{{\bf{h}}}}}_{2}^{{\dagger} {\prime} }={\sigma }^{{\dagger} 2}\,\,{{\mbox{and}}}\,\,{{{{\bf{h}}}}}_{1}^{{\dagger} }{\Pi }^{{\dagger} }{{{{\bf{h}}}}}_{2}^{{\dagger} {\prime} }={{{{\bf{h}}}}}_{2}^{{\dagger} }{\Pi }^{{\dagger} }{{{{\bf{h}}}}}_{1}^{{\dagger} {\prime} }=0,$$

since the second element of \({{{{\bf{h}}}}}_{2}^{{\dagger} }\) is 1 and that of \({{{{\bf{h}}}}}_{1}^{{\dagger} }\) is 0. Recall that \({C}_{\epsilon ,jn}={{{\rm{E}}}}[{n}^{-1/2}\mathop{\sum }\nolimits_{i = 1}^{n}{\epsilon }_{ni}^{{\dagger} j}\cdot {n}^{1/2}({\tilde{\theta }}_{n}^{{\dagger} }-{\theta }^{{\dagger} })]\) for j = 1, 2, 3. When vni’s are normal, it is from straight calculation that

$${C}_{\epsilon ,1n}={\Pi }^{{\dagger} }{{{{\bf{h}}}}}_{2}^{{\dagger} {\prime} }={\sigma }_{v}^{2}{{{{\bf{e}}}}}_{2,k+1}\,\,{{\mbox{and}}}\,\,{C}_{\epsilon ,3n}=3{\sigma }_{v}^{2}{\Pi }^{{\dagger} }{{{{\bf{h}}}}}_{2}^{{\dagger} {\prime} }=3{\sigma }_{v}^{4}{{{{\bf{e}}}}}_{2,k+1}.$$

For the elements in (B.3) and (B.4), \({\delta }_{11,n}^{{\dagger} }={\delta }_{13,n}^{{\dagger} }=0\), \({\delta }_{33,n}^{{\dagger} }={\mu }_{6}^{{\dagger} }-9{\sigma }_{v}^{6}=6{\sigma }_{v}^{6}\), whereby we can derive

$${{{{\bf{r}}}}}_{2}^{{\prime} }{\Sigma }_{n}{{{{\bf{r}}}}}_{2}={{{{\bf{r}}}}}_{2}^{{\prime} }{\Pi }_{n}^{{\dagger} }{{{{\bf{r}}}}}_{2}={\delta }_{11,n}^{{\dagger} }{r}_{1}^{2}+{\delta }_{33,n}^{{\dagger} }{r}_{3}^{2}+2{\delta }_{13,n}^{{\dagger} }{r}_{1}{r}_{3}=6{\sigma }_{v}^{6}(1/{\sigma }_{v}^{3})^{2}=6.$$

where \({{{{\bf{r}}}}}_{2}\triangleq [{r}_{1},0,{r}_{3}]^{{\prime} }\). The proof is accomplished. □

Appendix C. Technical details in Section 4

We first show how to transform the panel SARSF model (4.1’) for eliminating individual and time fixed-effects.

Let \([{F}_{T,T-1},\frac{1}{\sqrt{T}}{{{{\bf{1}}}}}_{T}]\) be the orthogonal matrix consisting of the eigenvectors of JT, where \({F}_{T,T-1}\in {{\mathbb{R}}}^{T\times (T-1)}\) corresponds to the eigenvalues of one and \(\frac{1}{\sqrt{T}}{{{{\bf{1}}}}}_{T}\in {{\mathbb{R}}}^{T}\) corresponds to the eigenvalue of zero. So does the orthogonal matrix \([{F}_{n,n-1},\frac{1}{\sqrt{n}}{{{{\bf{1}}}}}_{n}]\) for Jn. For any n × T matrix [Zn1, ⋯ , ZnT], define the n × (T − 1) transformed matrix \([{Z}_{n1}^{* },\cdots \,,{Z}_{n,T-1}^{* }]=[{Z}_{n1},\cdots \,,{Z}_{nT}]{F}_{T,T-1}\). Similarly, we can obtain \({X}_{nt}^{* },{\alpha }_{to}^{* }\) and \({\epsilon }_{nt}^{{\dagger} * }\). Then, (4.1’) implies \({Y}_{nt}^{* }=\lambda {W}_{n}{Y}_{nt}^{* }+{X}_{nt}^{* }\beta +{\alpha }_{t}^{* }{{{{\bf{1}}}}}_{n}+{\epsilon }_{nt}^{{\dagger} * }\) for t = 1, 2, ⋯ , T − 1, where cn is eliminated since \({{{{\bf{1}}}}}_{T}^{{\prime} }{F}_{T,T-1}={{{{\bf{0}}}}}_{T-1}^{{\prime} }\). After multiplying \({F}_{n,n-1}^{{\prime} }\) to the left on both sides of the transformed model, model (4.1’) can further be simplified to

$${Y}_{nt}^{* * }=\lambda {W}_{n}^{* }{Y}_{nt}^{* * }+{X}_{nt}^{* * }\beta +{\epsilon }_{nt}^{{\dagger} * * },t=1,2,\cdots \,,T-1,$$
(C.1)

where \({Y}_{nt}^{* * }={F}_{n,n-1}^{{\prime} }{Y}_{nt}^{* }\in {{\mathbb{R}}}^{n-1},{W}_{n}^{* }={F}_{n,n-1}^{{\prime} }{W}_{n}{F}_{n,n-1}\in {{\mathbb{R}}}^{(n-1)\times (n-1)},{X}_{nt}^{* * }={F}_{n,n-1}^{{\prime} }{X}_{nt}^{* }\in {{\mathbb{R}}}^{(n-1)\times k}\), and \({\epsilon }_{nt}^{{\dagger} * * }={F}_{n,n-1}^{{\prime} }{\epsilon }_{nt}^{{\dagger} * }\in {{\mathbb{R}}}^{n-1}\).

Remark 4

In the transformed panel SARSF model (C.1), the transformed time-effect term \({\alpha }_{t}^{* }{{{{\bf{1}}}}}_{n}\) is eliminated, because \({F}_{n,n-1}^{{\prime} }{{{{\bf{1}}}}}_{n}={{{{\bf{0}}}}}_{n-1}\). Thus, there does not exist an intercept term in model (C.1).

Recall \({{{{\boldsymbol{\epsilon }}}}}_{nT}^{{\dagger} }=[{\epsilon }_{n1}^{{\dagger} {\prime} },{\epsilon }_{n2}^{{\dagger} {\prime} },\cdots \,,{\epsilon }_{n,T}^{{\dagger} {\prime} }]\) and define \({{{{\boldsymbol{\epsilon }}}}}_{n-1,T-1}^{{\dagger} * * }=[{\epsilon }_{n1}^{{\dagger} * * {\prime} },{\epsilon }_{n2}^{{\dagger} * * {\prime} },\cdots \,,{\epsilon }_{n,T-1}^{{\dagger} * * {\prime} }]\). Both vectors satisfy the following condition:Footnote 12

$$\begin{array}{*{20}{l}}{{{{\boldsymbol{\epsilon }}}}}_{n-1,T-1}^{{\dagger} * * }=({I}_{T-1}\otimes {F}_{n,n-1}^{{\prime} })[{\epsilon }_{n1}^{{\dagger} * {\prime} },{\epsilon }_{n2}^{{\dagger} * {\prime} },\cdots \,,{\epsilon }_{n,T-1}^{{\dagger} * {\prime} }]\\\qquad\quad\;\;\;=({F}_{T,T-1}^{{\prime} }\otimes {F}_{n,n-1}^{{\prime} }){{{{\boldsymbol{\epsilon }}}}}_{nT}^{{\dagger} }.\end{array}$$
(C.2)

It follows that

$$\begin{array}{*{20}{l}}&{{{\rm{Var}}}}[{{{{\boldsymbol{\epsilon }}}}}_{n-1,T-1}^{{\dagger} * * }]\\=&({F}_{T,T-1}^{{\prime} }\otimes {F}_{n,n-1}^{{\prime} }){{{\rm{Var}}}}[{{{{\boldsymbol{\epsilon }}}}}_{nT}^{{\dagger} }]{({F}_{T,T-1}^{{\prime} }\otimes {F}_{n,n-1}^{{\prime} })}^{{\prime} }\\ =&{\sigma }^{{\dagger} 2}({F}_{T,T-1}^{{\prime} }\otimes {F}_{n,n-1}^{{\prime} }){({F}_{T,T-1}^{{\prime} }\otimes {F}_{n,n-1}^{{\prime} })}^{{\prime} }\\=&{\sigma }^{{\dagger} 2}({I}_{T-1}\otimes {I}_{n-1}).\end{array}$$
(C.3)

This indicates even if the original panel SARSF model (4.1’) is transformed into (C.1), their centered composite errors do not only have the same variance σ†2, but also are uncorrelated mutually. Next, for estimating the transformed model in (C.1), we consider QMLE as example in what follows. Given that QMLE treats \({{{{\boldsymbol{\epsilon }}}}}_{n-1,T-1}^{{\dagger} * * }\) as normal, the concentrated log-likelihood function of (C.1) with respect to λ is

$$\begin{array}{*{20}{l}}\ln L_{n,T}(\lambda) = -(1/2)(n-1)(T-1)\cdot\ln(2\pi + 1) \\\qquad\qquad\quad\;\;+\, (T-1)\cdot\ln \left|(1 - \lambda)^{-1} (I_n - \lambda W_n)\right|\\\qquad\qquad\quad\;\; -\,(1/2)(n-1)(T-1)\cdot\sigma_{nT}^{\dagger 2}(\lambda),\end{array}$$
(C.4)

whose derivation will be introduced below, where

$$\begin{array}{*{20}{l}}\quad\beta(\lambda) = \left[\mathop{\sum}\limits_{t=1}^{T} \tilde{X}_{nt}^\prime J_n \tilde{X}_{nt}\right]^{-1}\mathop{\sum}\limits_{t=1}^{T}\tilde{X}_{nt}^\prime J_n (I_n - \lambda W_n) \tilde{Y}_{nt}, \\\,\sigma^{\dagger 2}(\lambda) = \frac{1}{(n-1)(T-1)} \mathop{\sum}\limits_{t=1}^{T} \left[(I_n - \lambda W_n)\tilde{Y}_{nt} - \tilde{X}_{nt}\beta(\lambda)\right]^\prime J_n\\\qquad\qquad\; \times \left[(I_n - \lambda W_n)\tilde{Y}_{nt} - \tilde{X}_{nt}\beta(\lambda)\right]\end{array}$$
(C.5)

with \({\tilde{X}}_{nt}={X}_{nt}-{T}^{-1}\mathop{\sum }\nolimits_{t = 1}^{T}{X}_{nt}\) and \({\tilde{Y}}_{nt}={Y}_{nt}-{T}^{-1}\mathop{\sum }\nolimits_{t = 1}^{T}{Y}_{nt}\). The QMLE \({\tilde{\lambda }}_{nT}\) can be obtained by maximizing \(\ln {L}_{n,T}(\lambda )\), as well as the plug-in estimator for β and σ†2 are defined by \({\tilde{\beta }}_{nT}=\beta ({\tilde{\lambda }}_{nT})\) and \({\tilde{\sigma }}_{nT}^{{\dagger} 2}={\sigma }^{{\dagger} 2}({\tilde{\lambda }}_{nT})\), respectively.

We now discuss how to derive (C.4). First the log-likelihood function of the transformed panel SARSF model in (C.1) is

$$\begin{array}{ll}\ln L_{n,T}(\theta) = -(1/2)(n-1)(T-1)\cdot\ln(2\pi + \sigma^{\dagger 2})\\\qquad\qquad\quad\;\;+\, (T-1)\cdot\ln |I_{n-1} - \lambda W_n^*| \\\qquad\qquad\quad\;\,\,- \left[1/(2\sigma^{\dagger 2})\right]\cdot\sum\nolimits_{t=1}^{T-1}\epsilon_{nt}^{\dagger ** \prime}(\theta)\epsilon_{nt}^{\dagger **}(\theta), \end{array}$$
(C.6)

where \(\theta ={[\lambda ,{\beta }^{{\prime} }]}^{{\prime} }\) and \({\epsilon }_{nt}^{{\dagger} * * }(\theta )=({I}_{n-1}-\lambda {W}_{n}^{* }){Y}_{nt}^{* * }-{X}_{nt}^{* * }\beta\). On the one hand, notice that

$$\begin{array}{*{20}{l}}\left| I_{n-1} - \lambda W_n^{\ast} \right| = \left| I_{n-1} - \lambda F_{n,n-1}^{\prime} W_n F_{n,n-1} \right| \\\qquad\qquad\quad\;\;\;\;= (1-\lambda)^{-1}\left| I_n - \lambda W_n \right|\end{array}$$
(C.7)

holds for the row-normalized Wn, which is assumed in Section 4. On the other hand, by (C.2), we obtain

$$\begin{array}{ll}&\mathop{\sum }\limits_{t=1}^{T-1}{\epsilon }_{nt}^{{\dagger} ** {\prime} }(\theta ){\epsilon }_{nt}^{{\dagger} ** }(\theta )\,\\=&\,{{{{\boldsymbol{\epsilon }}}}}_{nT}^{{\dagger} {\prime} }(\theta )({F}_{T,T-1}\otimes {F}_{n,n-1})({F}_{T,T-1}^{{\prime} }\otimes {F}_{n,n-1}^{{\prime} }){{{{\boldsymbol{\epsilon }}}}}_{nT}^{{\dagger} }(\theta )\\ =&\,{{{{\boldsymbol{\epsilon }}}}}_{nT}^{{\dagger} {\prime} }(\theta )({J}_{T}\otimes {J}_{n}){{{{\boldsymbol{\epsilon }}}}}_{nT}^{{\dagger} }(\theta )=\mathop{\sum }\limits_{t=1}^{T}{\tilde{\epsilon }}_{nt}^{{\dagger} {\prime} }(\theta ){J}_{n}{\tilde{\epsilon }}_{nt}^{{\dagger} }(\theta ),\end{array}$$
(C.8)

where \({\tilde{\epsilon }}_{nt}^{{\dagger} }(\theta )=({I}_{n}-\lambda {W}_{n}){\tilde{Y}}_{nt}-{\tilde{X}}_{nt}\beta\). Substituting (C.7) and (C.8) into (C.6) yields that

$$\begin{array}{ll}\ln L_{n,T}(\theta) = -(1/2)(n-1)(T-1)\cdot\ln(2\pi + \sigma^{\dagger 2}) \\\qquad\qquad\quad\;\;+\, (T-1) \cdot \ln \left|(1 - \lambda)^{-1}(I_{n} - \lambda W_{n})\right| \\\qquad\qquad\quad\;\;-\, [1/(2\sigma^{\dagger 2})]\cdot\sum\nolimits_{t=1}^{T}\tilde{\epsilon}_{nt}^{\dagger}(\theta) J_{n} \tilde{\epsilon}_{nt}^{\dagger}(\theta).\end{array}$$
(C.9)

Taking derivatives for \(\ln {L}_{n,T}(\theta )\) with respect to σ†2 and β, and setting such derivatives to zero, we have

$$\begin{array}{ll}&\frac{(n-1)(T-1)}{{\sigma }^{{\dagger} 2}}-\frac{1}{{\sigma }^{{\dagger} 4}}\mathop{\sum }\limits_{t=1}^{T}{\tilde{\epsilon }}_{nt}^{{\dagger} }(\theta ){J}_{n}{\tilde{\epsilon }}_{nt}^{{\dagger} }(\theta )=0\,\,{{\mbox{and}}}\,\\&\,\mathop{\sum }\limits_{t=1}^{T}{\tilde{X}}_{nt}^{{\prime} }{J}_{n}{\tilde{X}}_{nt}\beta -\mathop{\sum }\limits_{t=1}^{T}{\tilde{X}}_{nt}^{{\prime} }{J}_{n}({I}_{n}-\lambda {W}_{n}){\tilde{Y}}_{nt}=0,\end{array}$$

both which indicate (C.5) holds. Finally, the concentrated log-likelihood function in (C.4) follows from taking (C.5) instead of the counterpart in (C.9).

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Deng, MY., Kutlu, L. & Wang, M. Skewness-based test diagnosis of technical inefficiency in spatial autoregressive stochastic frontier models. J Prod Anal (2024). https://doi.org/10.1007/s11123-024-00721-7

Download citation

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11123-024-00721-7

JEL classification

Keywords

Navigation