Skip to main content

Advertisement

Log in

Imposing Regularity Conditions to Measure Banks’ Productivity Changes in Taiwan Using a Stochastic Approach

  • Original Research
  • Published:
Asia-Pacific Financial Markets Aims and scope Submit manuscript

Abstract

This paper develops a stochastic approach to impose regularity properties on a directional output distance function (DODF) and an output distance function, which can be estimated by maximum likelihood. We use the resulting parameter estimates to evaluate efficiency and total factor productivity (TFP) growth for Taiwan’s commercial banks over the period 2002–2015 and claim that the failure of considering the regularity restrictions and the exclusion of undesirables lead to miscalculated efficiency measures and productivity gains. The outcomes from the regularity-constrained DODF reveal that almost all data-points satisfy the regularity properties, that the managerial abilities of the banks improve after the subprime crisis of 2007, and that the sample banks’ TFP grow at an average rate of 1.93% per annum, whereby technical change is the driving force. However, our estimates show downward trends in the growth rate of TFP and technical change.

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.

Fig. 1

Similar content being viewed by others

Notes

  1. We do not show the relevant results for brevity.

  2. We implicitly impose the assumption of null-joint production and weak disposability of good and bad outputs, and strong disposability of the good output y. Please see, e.g., Färe and Grosskopf (2005), for the related definitions.

  3. Färe et al. (2005) estimate the stochastic DODF without imposing monotonicity, which violates the monotonicity conditions 57 out of 209 times in 1993 and 20 out of 209 times in 1997.

  4. It is noteworthy that if different distributions are assumed, then the resulting parameter estimates and the number of constraint violations may change somewhat.

  5. It can be shown that \(\left| {F_{2} } \right| = \left| {\begin{array}{*{20}c} {\alpha_{22} } & {\alpha_{23} } \\ {\alpha_{23} } & {\alpha_{33} } \\ \end{array} } \right| = \alpha_{22} \alpha_{33} - \alpha_{23}^{2} = \left| {F_{3} } \right| = \left| {\begin{array}{*{20}c} {\alpha_{11} } & {\alpha_{13} } \\ {\alpha_{13} } & {\alpha_{33} } \\ \end{array} } \right| = \left| {F_{1} } \right| \ge 0\).

  6. Since our likelihood function may have multiple local maxima, a change in initial conditions tends to result in distinct parameter estimates.

  7. Note that our simultaneous estimation procedure with respect to (2) and (7)–(12) can still raise the efficiency of the estimators somewhat due to the imposition of cross-equation restrictions on the coefficients under study.

  8. Although the use of ODF allows one to decompose the Malmquist TFP change index into three items, i.e., TCS, ECS, and SC (a scale component), DODF is unable to address the term of SC. Since this paper mainly focuses on the use of DODF, we follow Feng and Serletis (2014) who ignore this scale component.

  9.  A financial holding company (FHC) is a financial institution engaged in banking-related activities, offering customers a wide range of financial services, such as purchasing insurance products and investment in securities. An FHB is a subsidiary of an FHC, and its scale is usually larger than Non-FHBs, which do not belong to any financial group. Here, a foreign bank is obligated to follow the regulations of both the home and host countries, and its scale is close to Non-FHBs in Taiwan.

  10. Four out of the five foreign banks started their businesses in Taiwan after 2007. Hence, there are no observations for the four banks prior to 2007.

References

  • Aigner, D., Knox Lovell, C. A., & Schmidt, P. (1977). Formulation and estimation of stochastic frontier production function models. Journal of Econometrics, 6, 21–37.

    Google Scholar 

  • Amsler, C., Prokhorov, A., & Schmidt, P. (2014). Using copulas to model time dependence in stochastic frontier models. Econometric Reviews, 33(5–6), 497–522.

    Google Scholar 

  • Battese, G. E., & Coelli, T. J. (1988). Prediction of firm-level technical efficiencies with a generalized frontier production function and panel data. Journal of Econometrics, 38, 387–399.

    Google Scholar 

  • Berger, A. N., Demsetz, R. S., & Strahan, P. E. (1999). The consolidation of the financial services industry: Causes, consequences, and implications for the future. Journal of Banking and Finance, 23, 135–194.

    Google Scholar 

  • Berger, A. N., Hunter, W. C., & Timme, S. G. (1993). The efficiency of financial institutions: A review and preview of research past, present, and future. Journal of Banking & Finance, 17, 221–249.

    Google Scholar 

  • Chambers, R. G. (2002). Exact nonradial input, output, and productivity measurement. Econometric Theory, 20, 751–765.

    Google Scholar 

  • Chang, T. P., Hu, J. L., Chou, R. Y., & Sun, L. (2012). The sources of bank productivity growth in China during 2002–2009: A disaggregation view. Journal of Banking and Finance, 36, 1997–2006.

    Google Scholar 

  • Chen, K. H. (2012). Incorporating risk input into the analysis of bank productivity: Application to the Taiwanese banking industry. Journal of Banking & Finance, 36, 1911–1927.

    Google Scholar 

  • Chen, I. I., Schmidt, P., & Wang, H. J. (2014). Consistent estimation of the fixed effects stochastic frontier model. Journal of Econometrics, 181, 65–76.

    Google Scholar 

  • Chung, Y. H., Färe, R., & Grosskopf, S. (1997). Productivity and undesirable outputs: a directional distance function approach. Journal of EnvironmentalManagement, 51, 229–240.

    Google Scholar 

  • Diewert, W. E., & Fox, K. J. (2008). On the estimation of returns to scale, technical progress and monopolistic markups. Journal of Econometrics, 145, 174–193.

    Google Scholar 

  • Diewert, W. E., & Wales, T. J. (1987). Flexible functional forms and global curvature conditions. Econometrica, 55, 43–68.

    Google Scholar 

  • Du, P., Parmeter, C. F., & Racine, J. S. (2013). Nonparametric kernel regression with multiple predictors and multiple shape constraints. Statistica Sinica, 23(3), 1347–1371.

    Google Scholar 

  • Färe, R., & Grosskopf, S. (1994). Cost and Revenue Constrained Production. New York: Springer.

    Google Scholar 

  • Färe, R., & Grosskopf, S. (2005). New directions: Efficiency and productivity. Boston: Kluwer Academic Publishers.

    Google Scholar 

  • Färe, R., Grosskopf, S., Noh, D. W., & Weber, W. (2005). Characteristics of a polluting technology: Theory and practice. Journal of Econometrics, 126, 469–492.

    Google Scholar 

  • Färe, R., Grosskopf, S., Norris, M., & Zhang, Z. (1994). Productivity growth, technical progress, and efficiency changes in industrialized countries. American Economic Review, 84, 66–83.

    Google Scholar 

  • Feng, G. H., & Serletis, A. (2010). Efficiency, technical change, and returns to scale in large US banks: Panel data evidence from an output distance function satisfying theoretical regularity. Journal of Banking & Finance, 34, 127–138.

    Google Scholar 

  • Feng, G., & Serletis, A. (2014). Undesirable outputs and a primal Divisia productivity index based on the directional output distance function. Journal of Econometrics, 183(1), 135–146.

    Google Scholar 

  • Feng, G. H., & Serletis, A. (2015). Imposing theoretical regularity on flexible functional forms. Econometric Reviews, 34(1–2), 198–227.

    Google Scholar 

  • Feng, G. H., & Zhang, X. H. (2012). Productivity and efficiency at large and community banks in the US: A Bayesian true random effects stochastic distance frontier analysis. Journal of Banking & Finance, 36, 1883–1895.

    Google Scholar 

  • Geweke, J. (1986). Exact inference in the inequality constrained normal linear regression model. Journal of Applied Econometrics, 1, 127–141.

    Google Scholar 

  • Griffiths, W. E., O’Donnell, C. J., & Tan Cruz, A. (2000). Imposing regularity conditions on a system of cost and cost-share equations: A Bayesian approach. Australian Journal of Agricultural and Resource Economics, 44(1), 107–127.

    Google Scholar 

  • Hsiao, H. C., Chang, H., Cianci, A. M., & Huang, L. H. (2010). First financial restructuring and operating efficiency: Evidence from Taiwanese commercial banks. Journal of Banking & Finance, 34, 1461–1471.

    Google Scholar 

  • Huang, T. H., Chiang, D. L., & Tsai, C. M. (2015). Applying the new metafrontier directional distance function to compare banking efficiencies in Central and Eastern European countries. Economic Modelling, 44, 188–199.

    Google Scholar 

  • Huang, T. H., & Chung, M. T. (2017). Do undesirables matter on the examination of banking efficiency using stochastic directional distance functions. Quarterly Review of Economics and Finance, 65, 194–211.

    Google Scholar 

  • Huang, T. H., Liu, N. H., & Kumbhakar, S. C. (2018). Joint estimation of the Lerner index and cost efficiency using copula methods. Empirical Economics, 54(2), 799–822.

    Google Scholar 

  • Kauko, K. (2009). Managers and efficiency in banking. Journal of Banking & Finance, 33, 546–556.

    Google Scholar 

  • Kleit, A. N., & Terrell, D. (2001). Measuring potential efficiency gains from deregulation of electricity generation: A Bayesian approach. Review of Economics and Statistics, 83(3), 523–530.

    Google Scholar 

  • Koutsomanoli-Filippaki, A., Margaritis, D., & Staikouras, C. (2009). Efficiency and productivity growth in the banking industry of Central and Eastern Europe. Journal of Banking & Finance, 33(3), 557–567.

    Google Scholar 

  • Kuosmanen, T. (2008). Representation theorem for convex nonparametric least squares. The Econometrics Journal, 11(2), 308–325.

    Google Scholar 

  • Kuosmanen, T., & Kortelainen, M. (2012). Stochastic non-smooth envelopment of data: Semi-parametric frontier estimation subject to shape constraints. Journal of Productivity Analysis, 38(1), 11–28.

    Google Scholar 

  • Lai, H. P., & Huang, C. J. (2013). Maximum likelihood estimation of seemingly unrelated stochastic frontier regressions. Journal of Productivity Analysis, 40(1), 1–14.

    Google Scholar 

  • Li, Z., Liu, G., & Li, Q. (2017). Nonparametric knn estimation with monotone constraints. Econometric Reviews, 36(6–9), 988–1006.

    Google Scholar 

  • Meeusen, W., & van den Broeck, J. (1977). Efficiency estimation from Cobb–Douglas production functions with composed error. International Economic Review, 18(2), 435–444.

    Google Scholar 

  • Mekaroonreung, M., & Johnson, A. L. (2012). Estimating the shadow prices of SO2 and NOx for US coal power plants: A convex nonparametric least squares approach. Energy Economics, 34(3), 723–732.

    Google Scholar 

  • Morey, E. R. (1986). An introduction to checking, testing, and imposing curvature properties: the true function and the estimated function. Canadian Journalof Economics, 19, 207–235.

    Google Scholar 

  • O’Donnell, C. J., & Coelli, T. J. (2005). A Bayesian approach to imposing curvature on distance functions. Journal of Econometrics, 126(2), 493–523.

    Google Scholar 

  • Olley, S., & Pakes, A. (1996). The dynamics of productivity in the telecommunications equipment industry. Econometrica, 64(6), 1263–1298.

    Google Scholar 

  • Orea, L. (2002). Parametric decomposition of a generalized Malmquist productivity index. Journal of Productivity Analysis, 18, 5–12.

    Google Scholar 

  • Poirier, D. (1995). Intermediate statistics and econometrics: A comparative approach. Cambridge, MA: The MIT Press.

    Google Scholar 

  • Robert, C. (2007). The Bayesian Choice: From Decision-Theoretic Foundations to Computational Implementation. NEW York: Springer.

    Google Scholar 

  • Ryan, D. L., & Wales, T. J. (2000). Imposing local concavity in the translog and generalized Leontief cost functions. Economics Letters, 67(3), 253–260.

    Google Scholar 

  • Sealey, C., & Lindley, J. (1977). Inputs, outputs, and a theory of production and cost at depository financial institutions. Journal of Finance, 32, 1251–1266.

    Google Scholar 

  • Seijo, E., & Sen, B. (2011). Nonparametric least squares estimation of a multivariate convex regression function. The Annals of Statistics, 39(3), 1633–1657.

    Google Scholar 

  • Terrell, D. (1996). Incorporating monotonicity and concavity conditions in flexible functional forms. Journal of Applied Econometrics, 11, 179–194.

    Google Scholar 

  • Trivedi, P. K. (1981). Some discrete approximations to Divisia integral indices. International Economic Review, 22, 71–77.

    Google Scholar 

  • Wang, H. J., & Ho, C. W. (2010). Estimating fixed-effect panel stochastic frontier models by model transformation. Journal of Econometrics, 157, 286–296.

    Google Scholar 

  • Wiley, D. E., Schmidt, W. H., & Bramble, W. J. (1973). Studies of a class of covariance structure models. Journal of the American Statistical Association, 68(342), 317–323.

    Google Scholar 

  • Yagi, D., Chen, Y., Johnson, A. L., & Kuosmanen, T. (2018). Shape constrained kernel-weighted least squares: Application to production function estimation for Chilean manufacturing industries. The Journal of Business and Economic Statistics (accepted).

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Tai-Hsin Huang.

Additional information

Publisher's Note

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

Appendix

Appendix

1.1 Imposition of Monotonicity and Curvature on an ODF

We re-define the vector of y as \(y = \left( {y_{1} ,y_{2} } \right)^{\prime } \in R_{ + }^{2}\), because undesirables are excluded from ODF. The vector x is intact. We define the output set \(S(x)\) as:

$$S(x) = \left\{ {y|y{\text{ can be produced by }}x} \right\}.$$

The output distance function \(D_{o} \left( {y,x,t} \right)\) can be written as:

$$D_{o} \left( {y,x,t} \right) = min\left\{ {\delta |0 < \delta \le 1, y/\delta \in S\left( x \right)} \right\}.$$
(24)

A value of \(D_{o} \left( {y,x,t} \right)\) equaling unity means that the bank is already producing on the efficient frontier, while a value of \(D_{o} \left( {y,x,t} \right)\) less than unity reveals that the bank is technically inefficient due to managerial inability. Following Orea (2002), O’Donnell and Coelli (2005), and Feng and Serletis (2010, 2014), we specify the translog functional form for ODF:

$$\begin{aligned} \ln D_{o} (y,x,t) & = a_{0} + \sum\limits_{i = 1}^{2} {a_{i} \ln y_{i} } + \frac{1}{2}\sum\limits_{i = 1}^{2} {\sum\limits_{k = 1}^{2} {a_{ik} \ln y_{i} \ln y_{k} + } } \sum\limits_{n = 1}^{3} {b_{n} } \ln x_{n} + \frac{1}{2}\sum\limits_{n = 1}^{3} {\sum\limits_{j = 1}^{3} {b_{nj} \ln x_{n} \ln x_{j} } } \\ & \quad + \delta_{\tau } t + \frac{1}{2}\delta_{\tau \tau } t^{2} + \sum\limits_{n = 1}^{3} {\sum\limits_{i = 1}^{2} {g_{ni} \ln x_{n} \ln y_{i} } } + \sum\limits_{i = 1}^{2} {\delta_{\tau i} t\ln y_{i} } + \sum\limits_{n = 1}^{3} {\delta_{\tau n} t\ln x_{n} } \, \\ \end{aligned}$$
(25)

After imposing the homogeneity and symmetrical properties and appending a statistical noise term, we transform ODF into a regression equation:

$$\begin{aligned} - \ln y_{1} & = a_{0} + a_{2} \ln \left( {\frac{{y_{2} }}{{y_{1} }}} \right) + \frac{1}{2}a_{22} \left[ {\ln \left( {\frac{{y_{2} }}{{y_{1} }}} \right)} \right]^{2} + b_{1} \ln x_{1} + b_{2} \ln x_{2} + b_{3} \ln x_{3} + \frac{1}{2}b_{11} \left( {\ln x_{1} } \right)^{2} \\ & \quad + \frac{1}{2}b_{22} \left( {\ln x_{2} } \right)^{2} + \frac{1}{2}b_{33} \left( {\ln x_{3} } \right)^{2} + b_{12} \ln x_{1} \ln x_{2} + b_{13} \ln x_{1} \ln x_{3} + b_{23} \ln x_{2} \ln x_{3} \\ & \quad + \delta_{\tau } t + \frac{1}{2}\delta_{\tau \tau } t^{2} + g_{12} \ln x_{1} \ln \left( {\frac{{y_{2} }}{{y_{1} }}} \right) + g_{22} \ln x_{2} \ln \left( {\frac{{y_{2} }}{{y_{1} }}} \right) + g_{32} \ln x_{3} \ln \left( {\frac{{y_{2} }}{{y_{1} }}} \right) \\ & \quad + \delta_{\tau 2} t\ln \left( {\frac{{y_{2} }}{{y_{1} }}} \right) + \delta_{\tau 1} t\ln x_{1} + \delta_{\tau 2} t\ln x_{2} + \delta_{\tau 3} t\ln x_{3} { + }V_{1} { + }U_{1} \\ \end{aligned}$$
(26)

where \(U_{1} = - \ln D_{o} (y,x,t)\sim\left| {N\left( {0,\sigma_{U1}^{2} } \right)} \right|\) is a one-sided error signifying technical inefficiency, \(V_{1} \sim N\left( {0,\sigma_{V1}^{2} } \right)\) is the random disturbance, and \(U_{1}\) and \(V_{1}\) are assumed to be statistically independent.

1.2 Monotonicity and Curvature Constraints

Monotonicity requires that \(D_{o} \left( {y,x,t} \right)\) be non-decreasing in outputs and non-increasing in inputs—that is:

$$\frac{{\partial D_{o} }}{{\partial y_{2} }} = \frac{{\partial \ln D_{o} }}{{\partial \ln y_{2} }}\frac{{D_{o} }}{{y_{2} }} \, \ge \, 0\;{\text{and}}\;\frac{{\partial D_{o} }}{{\partial x_{n} }} = \frac{{\partial \ln D_{o} }}{{\partial \ln x_{n} }}\frac{{D_{o} }}{{x_{n} }} \, \le \, 0,\quad n = 1,2,3,$$
(27)

or, equivalently:

$$\frac{{\partial \ln D_{o} }}{{\partial \ln y_{2} }} \ge 0\;{\text{and}}\;\frac{{\partial \ln D_{o} }}{{\partial \ln x_{n} }} \le 0,\quad n = 1,2,3,$$
(28)

since \(D_{o} /y_{2}\) and \(D_{o} /x_{n}\) are positive. The property \(\partial D_{o} /\partial y_{2} \ge 0\) implies that a bank’s technical efficiency does not diminish when it can produce more of an output, say \(y_{2}\), after employing a given input mix. The property \(\partial D_{o} /\partial x_{n} \le 0\) implies that a bank’s technical efficiency does not rise when it hires more of an input, say \(x_{n}\), to produce the same output levels.

We re-write (28) for output \(y_{2}\) as:

$$\frac{{\partial \ln D_{o} }}{{\partial \ln y_{2} }} = a_{2} + a_{22} \ln \left( {\frac{{y_{2} }}{{y_{1} }}} \right) + g_{12} \ln x_{1} + g_{22} \ln x_{2} + g_{32} \ln x_{3} + \delta_{\tau 2} t \, \ge 0.$$

Since the ODF must be homogeneous in \(y_{1}\) and \(y_{2}\), we have:

$$\frac{{\partial \ln D_{o} }}{{\partial \ln y_{1} }} = 1 - \frac{{\partial \ln D_{o} }}{{\partial \ln y_{2} }} \ge 0\;{\text{and}}\;\frac{{\partial \ln D_{o} }}{{\partial \ln y_{2} }} \le 1.$$

Translating the above inequalities into equalities, we obtain:

$$a_{2} + a_{22} \ln \left( {\frac{{y_{2} }}{{y_{1} }}} \right) + g_{12} \ln x_{1} + g_{22} \ln x_{2} + g_{32} \ln x_{3} + \delta_{\tau 2} t \, = V_{2} + U_{2}$$
(29)
$$a_{2} - 1 + a_{22} \ln \left( {\frac{{y_{2} }}{{y_{1} }}} \right) + g_{12} \ln x_{1} + g_{22} \ln x_{2} + g_{32} \ln x_{3} + \delta_{\tau 2} t \, = V_{3} - U_{3} .$$
(30)

Similarly, the monotonicity conditions for the three inputs can be expressed as:

$$b_{1} + b_{11} \ln x_{1} + b_{12} \ln x_{2} + b_{13} \ln x_{3} + g_{12} \ln \left( {\frac{{y_{2} }}{{y_{1} }}} \right) + \delta_{\tau 1} t = V_{4} - U_{4} ,$$
(31)
$$b_{2} + b_{22} \ln x_{2} + b_{12} \ln x_{1} + b_{23} \ln x_{3} + g_{22} \ln \left( {\frac{{y_{2} }}{{y_{1} }}} \right) + \delta_{\tau 2} t = V_{5} - U_{5} ,$$
(32)
$$b_{3} + b_{33} \ln x_{3} + b_{13} \ln x_{1} + b_{23} \ln x_{2} + g_{32} \ln \left( {\frac{{y_{2} }}{{y_{1} }}} \right) + \delta_{\tau 3} t = V_{6} - U_{6} .$$
(33)

Here, \(U_{n} \sim \left| {N\left( {0,\sigma_{Un}^{2} } \right)} \right|\) and \(V_{n} \sim N\left( {0,\sigma_{Vn}^{2} } \right)\), n = 2,…, 6, are respectively one-sided and two-sided errors and are statistically independent. The presence of \(V_{n}\) in (29)–(33) makes the individual equations be regression equations and captures random shocks.

Curvature requires that ODF be convex in outputs and quasi-convex in inputs. See, for example, Färe and Grosskopf (1994) and O’Donnell and Coelli (2005). For ODF to be convex in outputs it is sufficient that all the principal minors of the Hessian matrix are non-negative. In our two-output case this requires that:

$$\frac{{\partial^{2} D_{o} }}{{\partial y_{2}^{2} }} = \frac{{\partial^{2} \ln D_{o} }}{{\partial \left( {\ln y_{2} } \right)^{2} }}\frac{1}{{y_{2} }}\frac{{D_{o} }}{{y_{2} }} + \left( {\frac{{\partial \ln D_{o} }}{{\partial \ln y_{2} }}} \right)^{2} \frac{{D_{o} }}{{y_{2}^{2} }} - \frac{{\partial \ln D_{o} }}{{\partial \ln y_{2} }}\frac{{D_{o} }}{{y_{2}^{2} }} \ge 0,$$

or, equivalently:

$$a_{22} + \frac{{\partial \ln D_{o} }}{{\partial \ln y_{2} }}\left( {\frac{{\partial \ln D_{o} }}{{\partial \ln y_{2} }} - 1} \right) \ge \, 0,$$
(34)

since \(D_{o} /y_{2}^{2} \ge 0\). The necessary condition of (34) is \(a_{22} \ge 0\), because:

$$\frac{{\partial \ln D_{o} }}{{\partial \ln y_{2} }}\left( {\frac{{\partial \ln D_{o} }}{{\partial \ln y_{2} }} - 1} \right) \le 0.$$

Equation (28) can be transformed into an equality:

$$a_{22} + \frac{{\partial \ln D_{o} }}{{\partial \ln y_{2} }}\left( {\frac{{\partial \ln D_{o} }}{{\partial \ln y_{2} }} - 1} \right) = \, U_{7} + V_{7} .$$
(35)

Quasi-convexity in inputs will be achieved if and only if all the principal minors of the following bordered Hessian matrix are non-positive:

$$F = \left( {\begin{array}{*{20}l} 0 \hfill & \quad {f_{1} } \hfill & \quad {f_{2} } \hfill & \quad {f_{3} } \hfill \\ {f_{1} } \hfill & \quad {f_{11} } \hfill & \quad {f_{12} } \hfill & \quad {f_{13} } \hfill \\ {f_{2} } \hfill & \quad {f_{12} } \hfill & \quad {f_{22} } \hfill & \quad {f_{23} } \hfill \\ {f_{3} } \hfill & \quad {f_{13} } \hfill & \quad {f_{23} } \hfill & \quad {f_{33} } \hfill \\ \end{array} } \right) ,$$

where \(f_{n} = \partial D_{o} /\partial x_{n} = \left( {D_{o} /x_{n} } \right)\partial \ln D_{o} /\partial \ln x_{n}\), n = 1, 2, 3; \(f_{mn} = \partial^{2} D_{o} /\partial x_{m} \partial x_{n}\), m, n = 1, 2, 3. There are seven principal minors in total to be considered. Among them, the following three conditions obviously hold for certainty and hence can be ignored, i.e.:

$$F_{11} = \left| {\left( {\begin{array}{*{20}c} 0 & {f_{1} } \\ {f_{1} } & {f_{11} } \\ \end{array} } \right)} \right| = - \left( {f_{1} } \right)^{2} \le 0,$$
$$F_{22} = \left| {\left( {\begin{array}{*{20}c} 0 & {f_{2} } \\ {f_{2} } & {f_{22} } \\ \end{array} } \right)} \right| = - \left( {f_{2} } \right)^{2} \le 0,$$
$$F_{33} = \left| {\left( {\begin{array}{*{20}c} 0 & {f_{3} } \\ {f_{3} } & {f_{33} } \\ \end{array} } \right)} \right| = - \left( {f_{3} } \right)^{2} \le 0.$$

However, the remaining four conditions must be considered:

$$F_{312} = \left| {\left( {\begin{array}{*{20}l} 0 \hfill & {f_{1} } \hfill & {f_{2} } \hfill \\ {f_{1} } \hfill & {f_{11} } \hfill & {f_{12} } \hfill \\ {f_{2} } \hfill & {f_{12} } \hfill & {f_{22} } \hfill \\ \end{array} } \right)} \right| = 2f_{1} f_{12} f_{2} - \left( {f_{2} } \right)^{2} f_{11} - \left( {f_{1} } \right)^{2} f_{22} \le 0,$$
$$F_{313} = \left| {\left( {\begin{array}{*{20}l} 0 \hfill & {f_{1} } \hfill & {f_{3} } \hfill \\ {f_{1} } \hfill & {f_{11} } \hfill & {f_{13} } \hfill \\ {f_{3} } \hfill & {f_{13} } \hfill & {f_{33} } \hfill \\ \end{array} } \right)} \right| = 2f_{1} f_{13} f_{3} - \left( {f_{3} } \right)^{2} f_{11} - \left( {f_{1} } \right)^{2} f_{33} \le 0,$$
$$F_{323} = \left| {\left( {\begin{array}{*{20}l} 0 \hfill & {f_{2} } \hfill & {f_{3} } \hfill \\ {f_{2} } \hfill & {f_{22} } \hfill & {f_{23} } \hfill \\ {f_{3} } \hfill & {f_{23} } \hfill & {f_{33} } \hfill \\ \end{array} } \right)} \right| = 2f_{2} f_{23} f_{3} - \left( {f_{3} } \right)^{2} f_{22} - \left( {f_{2} } \right)^{2} f_{33} \le 0,$$
$$F = \left| {\left( {\begin{array}{*{20}l} 0 \hfill & {f_{1} } \hfill & {f_{2} } \hfill & {f_{3} } \hfill \\ {f_{1} } \hfill & {f_{11} } \hfill & {f_{12} } \hfill & {f_{13} } \hfill \\ {f_{2} } \hfill & {f_{12} } \hfill & {f_{22} } \hfill & {f_{23} } \hfill \\ {f_{3} } \hfill & {f_{13} } \hfill & {f_{23} } \hfill & {f_{33} } \hfill \\ \end{array} } \right)} \right| = - f_{1} \left| {\begin{array}{*{20}c} {f_{1} } & {f_{12} } & {f_{13} } \\ {f_{2} } & {f_{22} } & {f_{23} } \\ {f_{3} } & {f_{23} } & {f_{33} } \\ \end{array} } \right| + f_{2} \left| {\begin{array}{*{20}c} {f_{1} } & {f_{11} } & {f_{13} } \\ {f_{2} } & {f_{12} } & {f_{23} } \\ {f_{3} } & {f_{13} } & {f_{33} } \\ \end{array} } \right| - f_{3} \left| {\begin{array}{*{20}c} {f_{1} } & {f_{11} } & {f_{12} } \\ {f_{2} } & {f_{12} } & {f_{22} } \\ {f_{3} } & {f_{13} } & {f_{23} } \\ \end{array} } \right| \le 0.$$

Let \(S_{n} = \partial \ln D_{o} /\partial \ln x_{n}\), n = 1, 2, 3. We simplify and re-formulate the foregoing four inequalities into equalities:

$$2S_{1} S_{2} h_{12} - S_{2}^{2} h_{11} - S_{1}^{2} h_{22} = V_{8} - U_{8} ,$$
(36)
$$2S_{1} S_{3} h_{13} - S_{3}^{2} h_{11} - S_{1}^{2} h_{33} = V_{9} - U_{9} ,$$
(37)
$$2S_{2} S_{3} h_{23} - S_{3}^{2} h_{22} - S_{2}^{2} h_{33} = V_{10} - U_{10} ,$$
(38)
$$\begin{aligned} & - S_{1}^{2} h_{22} h_{33} - 2h_{12} h_{23} S_{1} S_{3} - 2S_{1} S_{2} h_{13} h_{23} + 2h_{13} h_{22} S_{1} S_{3} + 2h_{12} h_{33} S_{1} S_{2} + h_{23}^{2} S_{1}^{2} \hfill \\ & \quad + 2h_{11} h_{23} S_{2} S_{3} + h_{13}^{2} S_{2}^{2} - 2h_{12} h_{13} S_{2} S_{3} - h_{11} h_{33} S_{2}^{2} - h_{11} h_{22} S_{3}^{2} + h_{12}^{2} S_{3}^{2} = V_{11} - U_{11} . \hfill \\ \end{aligned}$$
(39)

Here, \(h_{11} = S_{1}^{2} + b_{11} - S_{1}\), \(h_{22} = S_{2}^{2} + b_{22} - S_{2}\), \(h_{33} = S_{3}^{2} + b_{33} - S_{3}\), \(h_{12} = S_{1} S_{2} + b_{12}\),\(h_{13} = S_{1} S_{3} + b_{13}\), and \(h_{23} = S_{2} S_{3} + b_{23}\).

As for the error components in (35)–(39) we assume that \(U_{n} \sim \left| {N\left( {0,\sigma_{Un}^{2} } \right)} \right|\) and \(V_{n} \sim N\left( {0,\sigma_{Vn}^{2} } \right)\), n = 7,…, 11, are respectively one-sided and two-sided errors and are statistically independent. The presence of \(V_{n}\) in the five equations makes the individual equations be regression equations and captures statistical noise.

1.3 Estimation Procedure

Equations (26), (29)–(33), and (35)–(39) form a system of regression equations with composed errors. Since (29) and (30) are linearly dependent we arbitrarily delete (30) from the system of equations and suggest estimating the remaining ten equations simultaneously by ML. In this manner, we can similarly impose monotonicity and curvature on the ODF of (26) to the usage of a Bayesian approach. It is noteworthy that here we impose five more curvature restrictions on ODF compared to DODF, whose curvature conditions involve merely unknown parameters.

Following the case of DODF, we assume that all of the eleven composed errors, either v + u or v − u, in those equations are statistically independent. Their individual pdf’s are similar to (14) in the text and the corresponding likelihood function can be readily derived. We therefore omit their derivation. After getting the coefficient estimates, we apply the following formula, proposed by Battese and Coelli (1988), to directly calculate the measure of technical efficiency:

$$TE_{o} = E\left( {e^{{ - U_{1} }} |\omega_{1} } \right) = \frac{{\varPhi \left( {\frac{{\mu_{*1} - \sigma_{*1}^{2} }}{{\sigma_{*} }}} \right)}}{{\varPhi \left( {\frac{{\mu_{*1} }}{{\sigma_{*} }}} \right)}}\exp \left( {0.5\sigma_{*1}^{2} - \mu_{*1} } \right),$$
(40)

where \(\omega_{1} = V_{1} + U_{1}\), \(\sigma_{1}^{2} = \sigma_{U1}^{2} + \sigma_{V1}^{2}\), \(\mu_{*1} = - \omega_{1} \sigma_{U1}^{2} /\sigma_{1}^{2}\), and \(\sigma_{*1}^{2} = \sigma_{U1}^{2} \sigma_{V1}^{2} /\sigma_{1}^{2}\). Term \(TE_{o}\) must lie between 0 and 1.

An analogous caveat to the case of DODF is worth citing, i.e., the resulting set of slope coefficient estimates may not be able to ensure that all monotonicity and curvature conditions are satisfied for all data points. This is attributed to the addition of the disturbances Vj’s, j = 2,…, 11, to Eqs. (29)–(33) and (35)–(39) and appears to be an advantage over the Bayesian approach. The number of violating observations is expected to be relatively small. We shall discuss this in the empirical study section below.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Huang, TH., Chiu, YH. & Mao, CY. Imposing Regularity Conditions to Measure Banks’ Productivity Changes in Taiwan Using a Stochastic Approach. Asia-Pac Financ Markets 28, 273–303 (2021). https://doi.org/10.1007/s10690-020-09319-z

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10690-020-09319-z

Keywords

JEL Classification

Navigation