Skip to main content

Shephard’s Input and Output Distance Functions: Cost and Revenue Efficiency Decompositions

  • Chapter
  • First Online:
Benchmarking Economic Efficiency

Abstract

In this chapter, we present the classic approach to calculate and decompose cost and revenue efficiency based on Shephard’s radial input and output distance functions. These decompositions follow closely the presentation done in Chap. 2 where both economic efficiency measures can be separated into technical and allocative components, i.e., expressions (2.33) and (2.44). However, rather than resorting to Farrell’s input and output technical efficiency measures, the decomposition is formalized in terms of the distance function and duality theory. At the time of publishing his seminal paper in 1957, Farrell did not seem to be aware of the work by Shephard, printed initially in his 1953 book titled Theory of Cost and Production Functions. There, he formalized the duality between the cost function and the input distance function, constituting the theoretical base for the decomposition of economic efficiency. Farrell cites Debreu’s (1951) “coefficient of resource utilization” as a source of inspiration, although, had he been aware of it, he could have relied equally on Shephard’s contribution, which is specific to production theory. Nevertheless, Shephard never introduced the concept of overall economic efficiency nor that of allocative efficiency, being one step short of proposing this decomposition explicitly.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 119.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 159.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 159.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    Interestingly, a precursor of the input distance function was also introduced independently by Malmquist in the same year but in a consumer context (Malmquist, 1953).

  2. 2.

    For a discussion on the relevance of considering either the equality approach or that based on duality recovering Fenchel-Mahler inequalities, we refer the reader to Sect. 2.4.5 of Chap. 2 and Chap. 13 related to the so-called essential property.

  3. 3.

    General notation: For a vector v of dimension D, \( v\in {\mathrm{\mathbb{R}}}_{++}^D \) means that each element of v is positive; \( v\in {\mathrm{\mathbb{R}}}_{+}^D \) means that each element of v is nonnegative; v > 0D means v ≥ 0D but v ≠ 0D; and, finally, 0D denotes a zero vector of dimension D. Given two vectors v and u, vuvdud, d ∈ {1, …, D}, v < uvd < ud, d ∈ {1,…, D}, and v <* uvd < ud or vd = ud = 0 and d ∈ {1, …, D}. The inner—or scalar—product of two D dimensional vectors v ≡ [v1, …,vD] and u ≡ [u1, …, uD] is denoted as \( v\cdot u\equiv {\sum}_{d=1}^D{v}_d{u}_d \).

  4. 4.

    In these definitions, at least one element of the comparison vectors, x′ and y′, could be equal or, respectively, smaller or greater than those in the strongly efficient reference vectors x and y.

  5. 5.

    This is one of the regularity conditions imposed on the production function or, more generally, the transformation function, so they are well behaved, i.e., they are characterized by a decreasing (increasing) marginal of substitution (transformation) between inputs (outputs) resulting from strictly convex (concave) isoquants.

  6. 6.

    In these definitions, all the elements of the comparison vectors, x´ and y´, are, respectively, smaller or greater than those in the weakly efficient reference vectors x and y.

  7. 7.

    An exception can be found in the field of environmental economics, where the technological trade-off between desirable and undesirable outputs, including the null-jointness axiom, is usually modelled through their weak disposability; see Färe et al. (1989) and Zofío and Prieto (2001).

  8. 8.

    Again, this issue is not of concern in the parametric approach based on well-behaved production functions satisfying the desirable regularity conditions. The usual functional forms present equal strong, weak, and isoquant efficiency subsets. An example is the Cobb-Douglas function, whose input set corresponds to \( L(y)=\left\{{\prod}_{m=1}^M{x}_m^{\alpha_m}\ge y,\kern0.5em {\alpha}_m>0\right\} \).

  9. 9.

    It is relevant to recall that the weakly efficient subsets WLS(y) and WPS(x), generated under the DEA assumption of strong disposability of inputs and outputs, are not to be confused with the production possibility sets LW(y) and PW(x), defined under the assumption of weakly disposable inputs and outputs.

  10. 10.

    DEA started with the seminal paper of Charnes et al. (1978) where the CRS versions of the input-oriented and output-oriented radial models where proposed. Having been formulated as linear programs, they had, from the beginning on, two alternative dual formulations: the envelopment form, or primal linear program, and the multiplier form, or dual linear program.

  11. 11.

    The second-stage additive model is a particular case of the third oldest DEA model (Charnes et al., 1985). It corresponds to its primal or envelopment form where the original right-hand side term of the first M restrictions, xmο, m = 1, …, M, have been replaced by θxmο, m = 1, …, M, being θ the optimal value of the first-stage model (3.23). It can be refined searching for the strong efficient projection that minimizes the same objective function, at the expense of solving a more complex nonlinear program (see Aparicio et al., 2007).

  12. 12.

    Program (3.24) can be refined and substituted by a more complex nonlinear program that has the advantage of identifying the closest projection (minimizing the objective function) instead of the “farthest” projection, which means that the new optimal slacks can be much smaller; see Aparicio et al. (2007, 2017c).

  13. 13.

    This procedure can also be refined, by reducing the set of possible peers from the subset of strong efficient firms to the subset of extreme strong efficient firms, those that cannot be deleted without modifying its frontier facet. For VRS technologies, the last subset guarantees that any projection accepts a single representation as convex linear combination of extreme strong efficient firms.

  14. 14.

    The subset of efficient peers are all the efficient firms that belong to the mentioned optimal hyperplane. For large problems, it is interesting to consider as peers only the subset of “extreme efficient firms,” that is, firms that cannot be expressed as a convex linear combination of other efficient firms.

  15. 15.

    This results was named after Shephard by Diewert; see Fox (2018: 514).

  16. 16.

    Shephard himself did not introduce the concept of cost or revenue efficiency, neither proposed their decompositions.

  17. 17.

    The proofs of the propositions included in this chapter can be found in these references.

  18. 18.

    This is a standard assumption in the parametric approach to economic efficiency measurement where the preferred functional forms impose homotheticity, for example, the case of the Cobb-Douglas specification. Even when relying on flexible functional forms such as the translog or the quadratic specifications, homotheticity is routinely imposed by restricting the value of the associated parameters, normally without testing whether this restriction holds empirically or not. For a discussion on how to decompose cost efficiency relying on the parametric approach under non-homotheticity, we refer the reader to Aparicio and Zofío (2017).

  19. 19.

    Färe et al. (2019) rely on this property in their proposal to decompose profit efficiency in a multiplicative way.

  20. 20.

    The allocative efficiency term corresponding to the reverse approach, \( {AE}_{R(I)}^R\left(x,\hat{y},w\right) \), can be expressed in terms of the disparity between shadow prices and market prices at \( {\partial}^WL\left(\hat{y}\right) \), as in (3.46).

  21. 21.

    Now, the reverse allocative efficiency term corresponding \( {AE}_{R(O)}^R\left(\hat{x},y,p\right) \) can be expressed in terms of the disparity between shadow prices and market prices at \( {\partial}^WP\left(\hat{x}\right) \), as in (3.58).

  22. 22.

    Both options can be easily changed to constant returns and weak disposability. The syntax necessary to calculate both options simultaneously is the following:

  23. 23.

    We refer the reader to Sect. 2.6.1 in Chap. 2 for the installation of the “Benchmarking Economic Efficiency” package. All Jupyter Notebooks implementing the different economic models in this book can be downloaded from the reference site: http://www.benchmarkingeconomicefficiency.com.

  24. 24.

    When calculating the decomposition of revenue efficiency as presented in Table 3.5, the function reports the inverse of these efficiency scores.

Bibliography

  • Aigner, D., Lovell, C. K., & Schmidt, P. (1977). Formulation and estimation of stochastic frontier production functions models. International Economic Review, 17, 377–396.

    Article  Google Scholar 

  • Ali, A. I., & Seiford, L. M. (1993). Computational accuracy and infinitesimals in data envelopment analysis. Infor, 31(4), 290–297.

    Google Scholar 

  • Aparicio, J., & Zofío, J. L. (2017). Revisiting the decomposition of cost efficiency for non-homothetic technologies: A directional distance function approach. Journal of Productivity Analysis, 48(2), 133–146.

    Article  Google Scholar 

  • Aparicio, J., Ruiz, J. L., & Sirvent, I. (2007). Closest targets and minimum distance to the Pareto-efficient frontier in DEA. Journal of Productivity Analysis, 28(3), 209–218.

    Article  Google Scholar 

  • Aparicio, J., Borras, F., Pastor, J. T., & Vidal, F. (2015a). Measuring and decomposing firm′ s revenue and cost efficiency: The Russell measures revisited. International Journal of Production Economics, 165, 19–28.

    Article  Google Scholar 

  • Aparicio, J., Pastor, J. T., & Zofio, J. L. (2015b). How to properly decompose economic efficiency using technical and allocative criteria with non-homothetic DEA technologies. European Journal of Operational Research, 240(3), 882–891.

    Article  Google Scholar 

  • Aparicio, J., Cordero, J. M., & Pastor, J. T. (2017c). The determination of the least distance to the strongly efficient frontier in data envelopment analysis oriented models: Modelling and computational aspects. Omega, 71, 1–10.

    Article  Google Scholar 

  • Atkinson, S. E., & Tsionas, M. G. (2016). Directional distance functions: Optimal endogenous directions. Journal of Econometrics, 190(2), 301–314.

    Article  Google Scholar 

  • Balk, B. M. (1998). Industrial price, quantity, and productivity indices: The micro-economic theory and an application. Kluwer Academic Publishers.

    Book  Google Scholar 

  • Banker, R. D., & Thrall, R. M. (1992). Estimation of returns to scale using data envelopment analysis. European Journal of Operational Research, 62(1), 74–84.

    Article  Google Scholar 

  • Banker, R. D., Charnes, A., & Cooper, W. W. (1984). Some models for estimating technical and scale inefficiencies in data envelopment analysis. Management Science, 30(9), 1078–1092.

    Article  Google Scholar 

  • Bogetoft, P., Färe, R., & Obel, B. (2006). Allocative efficiency of technically inefficient production units. European Journal of Operational Research, 168(2), 450–462.

    Article  Google Scholar 

  • Boles, J. N. (1967). Efficiency squared-efficient computation of efficiency indexes. In Western farm economic association, proceedings 1966 (pp. 137–142). Pullman.

    Google Scholar 

  • Boles, J. N. (1971, February). The 1130 Farrell efficiency system-multiple products, multiple factors. Giannini Foundation of Agricultural Economics.

    Google Scholar 

  • Boussemart, J.-P., Briec, W., Peypoch, N., & Tavéra, C. (2009). α-Returns to scale in multi-output technologies. European Journal of Operational Research, 197(1), 332–339.

    Article  Google Scholar 

  • Camanho, A. S., & Dyson, R. G. (2008). A generalisation of the Farrell cost efficiency measure applicable to non-fully competitive settings. Omega, 36, 147–162.

    Article  Google Scholar 

  • Charnes, A., Cooper, W. W., & Rhodes, E. (1978). Measuring the efficiency of decision making units. European Journal of Operational Research, 2, 429–444.

    Article  Google Scholar 

  • Charnes, A., Cooper, W. W., Golany, B., Seiford, L. M., & Stutz, J. (1985). Foundations of data envelopment analysis for Pareto-Koopmans efficient empirical production functions. Journal of Econometrics, 30(1–2), 91–107.

    Article  Google Scholar 

  • Debreu, G. (1951, July). The coefficient of resource utilization. Econometrica, 19(3), 273–292.

    Article  Google Scholar 

  • Diewert, W. E. (1971). An application of the Shephard duality theorem: A generalized Leontief production function. Journal of Political Economy, 79(3), 481–507.

    Article  Google Scholar 

  • Econometric Software Inc. (2014). LIMDEP 11. Plainview. http://www.limdep.com/.

  • Färe, R. (1988). Fundamentals of production theory. Springer.

    Book  Google Scholar 

  • Färe, R., & Primont, D. (1995). Multi-output production and duality: Theory and applications. Kluwer Academic Publishers.

    Book  Google Scholar 

  • Färe, R., Grosskopf, S., & Lovell, C. A. K. (1985). The measurement of efficiency of production. Kluwer-Nijhoff.

    Book  Google Scholar 

  • Färe, R., Grosskopf, S., Lovell, C. A., & Pasurka, C. (1989). Multilateral productivity comparisons when some outputs are undesirable: A nonparametric approach. The Review of Economics and Statistics, 71(1), 90–98.

    Article  Google Scholar 

  • Färe, R., He, X., Li, S., & Zelenyuk, V. (2019). A unifying framework for Farrell profit efficiency measurement. Operations Research, 67(1), 183–197.

    Article  Google Scholar 

  • Farrell, M. J. (1957). The measurement of productive efficiency. Journal of Royal Statistical Society, Series A, 120(III), 253–290.

    Article  Google Scholar 

  • Fenchel, W. (1949). On conjugate convex functions. Canadian Journal of Mathematics, 1(1), 73–77.

    Article  Google Scholar 

  • Fox, K. (2018). The ET interview: Professor W. Erwin Diewert. Econometric Theory, 34, 509–542.

    Article  Google Scholar 

  • Greene, W. (2008). The econometric approach to efficiency analysis. In H. Fried, C. A. Lovell, & S. S. Schmidt (Eds.), The measurement of productive efficiency and productivity growth. Oxford University Press.

    Google Scholar 

  • Juo, J.-C., Fu, T.-T., Yu, M.-M., & Lin, Y.-H. (2015). Profit-oriented productivity change. Omega, 57, 76–87.

    Article  Google Scholar 

  • Kumbhakar, S. C., Hung-Jen, W., & Horncastle, A. P. (2015). A practitioner’s guide to stochastic frontier analysis using Stata. Cambridge University Press.

    Book  Google Scholar 

  • Madden, P. (1986). Concavity and optimization in microeconomics. Blackwell Publishers.

    Google Scholar 

  • Mahler, K. (1939). Ein übertragungsprinzip für konvexe körper. Časopis pro pěstování matematiky a fysiky, 68(3), 93–102.

    Article  Google Scholar 

  • Malmquist, S. (1953). Index numbers and indifference surfaces. Trabajos de Estadística, 4(1), 209–242.

    Article  Google Scholar 

  • McFadden, D. (1978). Cost, revenue, and profit functions. In M. Fuss & D. McFadden (Eds.), Production economics: A dual approach to theory and applications. North-Holland.

    Google Scholar 

  • MirHassani, S. A., & Alirezaee, M. (2005). An efficient approach for computing non-Archimedean ε in DEA based on integrated models. Applied Mathematics and Computation, 166(2), 449–456.

    Article  Google Scholar 

  • Parmeter, C. F., & Kumbhakar, S. C. (2014). Efficiency analysis: A primer on recent advances. Foundations and Trends in Econometrics, 7(3–4), 191–385.

    Article  Google Scholar 

  • Pastor, J. T., Aparicio, J., Alcaraz, J., Vidal, F., & Pastor, D. (2016). The reverse directional distance function. In Advances in efficiency and productivity (pp. 15–57). Springer.

    Chapter  Google Scholar 

  • Portela, M. C. A. S., & Thanassoulis, E. (2014). Economic efficiency when prices are not fixed: Disentangling quantity and price efficiency. Omega, 47, 36–44.

    Article  Google Scholar 

  • Russell, R. R. (1988). On the axiomatic approach to the measurement of technical efficiency. In Measurement in economics (pp. 207–217). Physica.

    Chapter  Google Scholar 

  • Shephard, R. W. (1953). Cost and production functions. Princeton University Press.

    Google Scholar 

  • Shephard, R. W. (1970). Theory of cost and production functions. Princeton University Press.

    Google Scholar 

  • Zofio, J. L., & Prieto, A. M. (2001). Environmental efficiency and regulatory standards: The case of CO2 emissions from OECD industries. Resource and Energy Economics, 23(1), 63–83.

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Rights and permissions

Reprints and permissions

Copyright information

© 2022 Springer Nature Switzerland AG

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Pastor, J.T., Aparicio, J., Zofío, J.L. (2022). Shephard’s Input and Output Distance Functions: Cost and Revenue Efficiency Decompositions. In: Benchmarking Economic Efficiency. International Series in Operations Research & Management Science, vol 315. Springer, Cham. https://doi.org/10.1007/978-3-030-84397-7_3

Download citation

Publish with us

Policies and ethics