Skip to main content

The Enhanced Russell Graph Measure (ERG=SBM): Economic Inefficiency Decompositions

  • Chapter
  • First Online:
Benchmarking Economic Efficiency

Abstract

The enhanced Russell graph measure, ERG, (Pastor et al., 1999) was designed as a new global efficiency measure to overcome the computational difficulties of the Russell graph measure of technical efficiency, RG (Färe et al., 1985). Historically, Farrell (1957) implemented the first measure of technical efficiency, while Färe and Lovell (1978), after suggesting some desirable properties that an ideal technical efficiency measure should satisfy, proposed the so-called Russell input measure of technical efficiency. An output version and a graph version were presented in the first book by Färe et al. (1985). These three initial Russell measures have been already presented in Chap. 5. The difficulty for solving the graph version motivated, in the last 1990s, the search for a new formulation easier to handle and to solve, and the proposed solution was the ERG measure. In the literature, two papers can be found, the first one by Pastor et al. (1999) and the second one by Tone (2001). Both proposed exactly the same linear fractional programming model for solving the Russell graph measure through the same reformulation. Hence, the ERG model was consecutively published twice with a gap of 2 years, which is something that occurs quite seldom in research. Pastor et al. (1999) published it first in the European Journal of Operational Research, and Tone (2001) published exactly the same models in the same journal under a different name, SBM, and related them also to the RG measure. It shows that two groups of researchers, without any interaction, may developed the same idea in a quasi-contemporary way without being aware of it. This is the reason why we refer to both measures, which are the same, as ERG=SBM and call it the “enhanced Russell graph slack-based measure.”

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.

    Actually, according to Russell and Schworm (2009), the “Russell measure” corresponds to the Färe-Lovell index introduced by Färe and Lovell (1978), who named it after the first author in recognition of his teaching activity.

  2. 2.

    With respect to the output-oriented measure, we only want to point out that, in Chap. 5 and for reasons of simplicity, instead of minimizing \( 1/\frac{1}{N}\sum \limits_{n=1}^N{\phi}_n \), we proposed directly to maximize its inverse \( \frac{1}{N}\sum \limits_{n=1}^N{\phi}_n \).

  3. 3.

    As usual, just by modifying the equality in the convexity constraint, \( \sum \limits_{j=1}^J{\lambda}_j=1 \), we get the NIRS (≤) and the NDRS (≥). If we want to get the CRS version, as originally formulated by Pastor et al. (1999), we just need to delete the convexity constraint. Model (7.1) was the only one not considered by Tone (2001).

  4. 4.

    Equalities (7.2) and the CRS version of model (7.3) appeared in Pastor et al. (1999) referenced as (3) and (4). Moreover, the CRS version of model (7.3) appeared in Tone (2001) referenced as (7). It should be clear that from the beginning we must assume that all input and output values are positive.

  5. 5.

    The new variable is equal to the inverse of the denominator of the objective function; see next section.

  6. 6.

    The inequality β ≥ 0 is a must for having a well-defined linear program as (7.5), although we know that, due to its definition, β is always a positive number.

  7. 7.

    Geometrically, the intuition is confirmed: when using any type of slack-based model, the distance to the frontier can be measured along the L1 path defined by the corresponding slacks.

  8. 8.

    In Aparicio et al. (2017a), the normalization procedure was not derived directly, as explained later on. Moreover, until now, the usual way of relating profit inefficiency with technical inefficiency has been obtaining a suitable Fenchel-Mahler inequality based on duality results. Here, we are proposing a more direct approach taking advantage of the final inequality obtained in the last-mentioned paper. However, developing this process has inspired us for proposing an alternative way of getting a profit inefficiency decomposition, as proposed later on in Chap. 13 based on Pastor et al. (2021b).

  9. 9.

    The inequality is an equality if, and only if, the projection \( \left({\hat{x}}_o,{\hat{y}}_o\right) \) achieves maximum profit.

  10. 10.

    The normalized profit inefficiency was originally called by Chambers et al. (1998) the Nerlovian profit inefficiency in honor of the economist Marc Nerlove, who was the first to suggest that profit inefficiency must be normalized in order to obtain a dimensionless number that is comparable to technical inefficiency, which, as already shown, is also a pure number.

  11. 11.

    According to Proposition 1 in Aparicio et al. (2017a), the other decomposition was achieved realizing that the numerator of TIERG = SBM(xo, yo), \( \sum \limits_{n=1}^N\frac{1}{Ny_{on}}{s}_n^{+}+\sum \limits_{m=1}^M\frac{1}{Mx_{om}}{s}_m^{-} \), is a weighted additive function. The inequality obtained was similar to (7.9) but less tightened since the normalizing denominator was simply \( {\delta}_{\left({x}_o,{y}_o,p,w\right)} \). In the last-mentioned paper, this first decomposition was used for deriving the second decomposition (7.9) that we have derived here directly.

  12. 12.

    Not all the firms generate multiple optimal solutions, as our next example shows.

  13. 13.

    It is well known that the profit-maximizing point(s) must be located close to the NW corner of the sample, which intuitively means “less inputs and more outputs.” As a consequence, a non-efficient point can never be a profit maximizing point.

  14. 14.

    An output-weak efficient point is a point whose projection on the strong efficient frontier by means of the additive model only identifies nonzero output slacks.

  15. 15.

    It may happen that the projection identified by either of the last two models gets an optimal projection with an output value \( \sum \limits_{j=1}^J{\lambda}_j^{\ast }{y}_j\ge {y}_o \) but different from yo. But since \( \left({\hat{x}}_o,\sum \limits_{j=1}^J{\lambda}_j^{\ast }{y}_j\right)\ge \left({\hat{x}}_o,{y}_o\right) \) and we are sure that the obtained projection belongs to T, the properties of T guarantee that \( \left({\hat{x}}_o,{y}_o\right) \) also belongs to T, and consequently, it belongs to L(yo).

  16. 16.

    Since the measure we are using is a strong efficient measure, we can use to classify the firms the—simplest—additive model or, alternatively, draw a picture in the two-input plane to see geometrically the position of the firms; see Fig. 7.2.

  17. 17.

    Let us point out that the normalizing factor of the traditional approach does only depend on the input components of the firm being rated and on their market costs, which shows that in case of multiple projections, any of them will give rise to the same normalized cost decomposition.

  18. 18.

    Obviously, when dealing with single-value inefficiency measures, the negative conclusion cannot be formulated.

  19. 19.

    An input-weak efficient point is a point whose projection on the strong efficient frontier by means of the additive model only identifies nonzero input slacks.

  20. 20.

    It may happen that the projection identified by model (7.13) gets an optimal projection with an input value \( \sum \limits_{j=1}^J{\lambda}_j^{\ast }{x}_j\le {x}_o \) but different from xo. But since \( \left(\sum \limits_{j=1}^J{\lambda}_j^{\ast }{x}_j,{\hat{y}}_o\right)\le \left({x}_o,{\hat{y}}_o\right) \) and the obtained projection belong to T, the properties of T guarantee that \( \left({x}_o,{\hat{y}}_o\right) \) also belongs to T, and consequently, it belongs to P(xo).

  21. 21.

    Since we are assuming a VRS technology, each optimal slack is upper bounded. Being more precise, for firm (xo, yo), its vector of optimal output slack s+∗ is upper bounded by Y − yo, where Y is the vector of maximum output values over the sample of firms.

  22. 22.

    Since the measure we are using is a strong efficient measure, we can use to classify the firms the additive model or even draw a picture, like Fig. 7.3, in the two output planes to see geometrically the location of the firms.

  23. 23.

    Since the normalizing factor is independent of the projection, we can choose any of them for the two mentioned firms with double projections. Hence, the uniqueness of the revenue inefficiency decompositions is guaranteed in this case, being independent of the assigned projection. We have add an asterisk to the projection of each of this two units in Table 7.3 to indicate that a second projection is possible.

  24. 24.

    This property is always satisfied by the general direct approach, as we are going to show very soon.

  25. 25.

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

Bibliography

  • 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.

    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.

    Google Scholar 

  • Aparicio, J., Ortiz, L., & Pastor, J. T. (2017a). Measuring and decomposing profit inefficiency through the slacks-based measure. European Journal of Operational Research, 260(2), 650–654.

    Article  Google Scholar 

  • Aparicio, J., Pastor, J. T., & Zofio, J. L. (2017b). Can Farrell’s allocative efficiency be generalized by the directional distance function approach? European Journal of Operational Research, 257(1), 345–351.

    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.

    Google Scholar 

  • Chambers, R. G., Chung, Y., & Färe, R. (1998, August). Profit, directional distance functions, and Nerlovian efficiency. Journal of Optimization Theory and Applications, 98(2), 351–364.

    Article  Google Scholar 

  • Charnes, A., & Cooper, W. W. (1962). Programming with linear fractional functionals. Naval Research Logistics Quarterly, 9(3–4), 181–186.

    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.

    Google Scholar 

  • Choi, Y., Zhang, N., & Zhou, P. (2012). Efficiency and abatement costs of energy-related CO2 emissions in China: A slacks-based efficiency measure. Applied Energy, 98, 198–208.

    Article  Google Scholar 

  • Cooper, W. W., Pastor, J. T., Borras, F., Aparicio, J., & Pastor, D. (2011b). BAM: A bounded adjusted measure of efficiency for use with bounded additive models. Journal of Productivity Analysis, 35(2), 85–94.

    Google Scholar 

  • Färe, R., & Lovell, C. K. (1978). Measuring the technical efficiency of production. Journal of Economic Theory, 19(1), 150–162.

    Google Scholar 

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

    Google Scholar 

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

    Google Scholar 

  • Halická, M., & Trnovská, M. (2018). The Russell measure model: Computational aspects, duality, and profit efficiency. European Journal of Operational Research, 268(1), 386–397.

    Article  Google Scholar 

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

    Google Scholar 

  • Lovell, C. A. K., & Pastor, J. T. (1995). Units invariant and translation invariant DEA models. Operations Research Letters, 18, 147–151.

    Google Scholar 

  • Pastor, J. T., Ruiz, J. L., & Sirvent, I. (1999). An enhanced DEA Russell graph efficiency measure. European Journal of Operational Research, 115(3), 596–607.

    Google Scholar 

  • Pastor, J. T., Zofío, J. L., Aparicio, J., & Pastor, D. (2021b). A new general direct approach for decomposing economic inefficiency. Working paper. Erasmus Research Institute of Management (ERIM), Erasmus University Rotterdam, The Netherlands.

    Google Scholar 

  • Russell, R. R., & Schworm, W. (2009). Axiomatic foundations of efficiency measurement on data-generated technologies. Journal of Productivity Analysis, 31, 77–86.

    Google Scholar 

  • Sharp, J. A., Meng, W., & Liu, W. (2007). A modified slacks-based measure model for data envelopment analysis with ‘natural’ negative outputs and inputs. Journal of the Operational Research Society, 58, 1672–1677.

    Article  Google Scholar 

  • Sueyoshi, T., & Sekitani, K. (2007a). Computational strategy for Russell measure in DEA: Second-order cone programming. European Journal of Operational Research, 180(1), 459–471.

    Article  Google Scholar 

  • Sueyoshi, T., & Sekitani, K. (2007b). Measurement of returns to scale using a non-radial DEA model: A range-adjusted measure approach. European Journal of Operational Research, 176(3), 1918–1946.

    Article  Google Scholar 

  • Tone, K. (2001). A slacks-based measure of efficiency in data envelopment analysis. European Journal of Operational Research, 130(3), 498–509.

    Google Scholar 

  • Tone, K. (2004). Malmquist productivity index. In Handbook on data envelopment analysis (pp. 203–227). Springer.

    Chapter  Google Scholar 

  • Tone, K., & Tsutsui, M. (2009). Network DEA: A slacks-based measure approach. European Journal of Operational Research, 197(1), 243–252.

    Article  Google Scholar 

  • Tone, K., & Tsutsui, M. (2010). Dynamic DEA: A slacks-based measure approach. Omega, 38(3–4), 145–156.

    Article  Google Scholar 

  • Tone, K., & Tsutsui, M. (2014). Dynamic DEA with network structure: A slacks-based measure approach. Omega, 42(1), 124–131.

    Article  Google Scholar 

  • Tone, K., Chang, T. S., & Wu, C. H. (2020). Handling negative data in slacks-based measure data envelopment analysis models. European Journal of Operational Research, 282(3), 926–935.

    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). The Enhanced Russell Graph Measure (ERG=SBM): Economic Inefficiency 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_7

Download citation

Publish with us

Policies and ethics