Skip to main content
Log in

Cone ratio models with shared resources and nontransparent allocation parameters in network DEA

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

Abstract

Many studies have examined the performance of production systems with shared resources through the application of data envelopment analysis (DEA). The present models are based on the multiplier-type frameworks and resource allocation variables (RAVs) in simple yet edifying network settings, such as multi-component and two-stage structures. Two issues associated with RAVs, however, are relevant to both the theory and practice. First, the existing models with RAVs are nonlinear in general. Second, a potential conflict of interest between a central evaluator (CE) and the managers of decision-making units (DMUs), due to the unknown allocation parameters, has not been addressed. The current study contributes to the resolution of these issues by presenting conflict free models (CFMs). Two striking features of the proposed models include their convenient transformation into linear programs and that they are conflict free. Thus, the manager of a DMU has no basis to argue against the evaluation results by simply claiming that faulty information regarding the split of shared resources has been used, as no such information relating to the CE’s preference or pertaining to RAVs is included in the model formulation. Furthermore, we investigate the relations between CFMs and the existing models, strengthening the interpretations of the existing models. Finally, the proposed models incorporate partial ordering preferences expressed as the cone ratio constraints, which are suitable for a wide range of real life applications. A dataset extracted from literature is used to illustrate the main concept that drives this research.

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
Fig. 2
Fig. 3

Similar content being viewed by others

References

  • Allen R, Athanassopoulos AD, Dyson RG, Thanassoulis E (1997) Weights restrictions and value judgments in DEA: evolution, development and future directions. Ann Oper Res 73:13–34

    Article  Google Scholar 

  • Athanassopoulos AD (1998) Decision support for target-based resource allocation of public services in multiunit and multi-level systems. Manag Sci 44(2):173–187

    Article  Google Scholar 

  • Avis D, Rosenberg GD, Savani R, Stengel BV (2010) Enumeration of Nash equilibria for two-player games. Econ Theor 42:9–37

    Article  Google Scholar 

  • Banker RD, Charnes A, Cooper WW (1984) Some models for the estimation of technical and scale inefficiencies in data envelopment analysis. Manag Sci 30(9):1078–1092

    Article  Google Scholar 

  • Beasley JE (1995) Determining teaching and research efficiencies. J Oper Res Soc 46:441–452

    Article  Google Scholar 

  • Ben-Tal A, Eiger G, Gershovitz V (1994) Global minimization by reducing the duality gap. Math Progr 63:193–212

    Article  Google Scholar 

  • Charnes A, Cooper WW, Rhodes E (1978) Measuring the efficiency of decision making units. Eur J Oper Res 2:429–444

    Article  Google Scholar 

  • Charnes A, Cooper WW, Huang ZM, Sun DB (1990) Polyhedral cone-ratio DEA models with an illustrative application to large commercial banks. J Econom 46(1–2):73–91

    Article  Google Scholar 

  • Charnes A, Cooper WW, Huang ZM, Sun DB (1991) Relations between half-space and finitely generated cones in polyhedral cone-ratio DEA models. Int J Syst Sci 22(1):2057–2077

    Article  Google Scholar 

  • Chen Y, Liang L, Yang F, Zhu J (2006) Evaluation of information technology investment: a data envelopment analysis approach. Comput Oper Res 33(5):1368–1379

    Article  Google Scholar 

  • Chen Y, Cook WD, Li N, Zhu J (2009a) Additive efficiency decomposition in two-stage DEA. Eur J Oper Res 196:1170–1176

    Article  Google Scholar 

  • Chen Y, Liang L, Zhu J (2009b) Equivalence in two-stage DEA approaches. Eur J Oper Res 193:600–604

    Article  Google Scholar 

  • Chen Y, Cook WD, Zhu J (2010) Deriving the DEA frontier for two-stage processes. Eur J Oper Res 202:138–142

    Article  Google Scholar 

  • Cherchye L, De Rock B, Dierynck B, Roodhooft F, Sabbe J (2013) Opening the “black box” of efficiency measurement: input allocation in multioutput settings. Oper Res 61(5):1148–1165

    Article  Google Scholar 

  • Cook WD, Hababou M, Tuenter HJH (2000) Multicomponent efficiency measurement and shared inputs in data envelopment analysis: an application to sales and service performance in bank branches. J Prod Anal 14:209–224

    Article  Google Scholar 

  • David G, Lindrooth R, Helmchen LA, Burns LR (2011) Do hospitals cross subsidize? NBER working paper no. 17300

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

    Article  Google Scholar 

  • Färe R, Grosskopf R (2000) Network DEA. Socio Econ Plan Sci 34:35–49

    Article  Google Scholar 

  • Farrell MJ (1957) The measurement of productive efficiency. J R Stat Soc Ser A 120(3):253–290

    Article  Google Scholar 

  • Faulhaber GR (1975) Cross-subsidization: pricing in public enterprises. Am Econ Rev 65(5):966–977

    Google Scholar 

  • Harris M, Kriebel CH, Raviv A (1982) Asymmetric information, incentives and intrafirm resource allocation. Manag Sci 28(6):604–620

    Article  Google Scholar 

  • Kao Q (2009a) Efficiency decomposition in network data envelopment analysis: a relational model. Eur J Oper Res 192:949–962

    Article  Google Scholar 

  • Kao Q (2009b) Efficiency measurement for parallel production systems. Eur J Oper Res 196:1107–1112

    Article  Google Scholar 

  • Kao Q, Hwang SN (2008) Efficiency decomposition in two-stage data envelopment analysis: an application to non-life insurance companies in Taiwan. Eur J Oper Res 185:418–429

    Article  Google Scholar 

  • Koopmans TC (1951) Activity analysis of production and allocation. Cowles Commission Monograph no. 13. Wiley, New York

    Google Scholar 

  • Kung LC, Chen YJ (2012) The resource allocation problem for demand forecasting and performance measurement. Oper Res Lett 40:61–68

    Article  Google Scholar 

  • Lewis HF, Sexton TR (2004) Network DEA: efficiency analysis of organizations with complex internal structure. Comput Oper Res 31:1365–1410

    Article  Google Scholar 

  • Liang L, Cook WD, Zhu J (2008) DEA models for two-stage processes: game approach and efficiency decomposition. Naval Res Logist 55:643–653

    Article  Google Scholar 

  • Rajan MV, Reichelstein S (2004) A perspective on “asymmetric information, incentives and intrafirm resource allocation”. Manag Sci 50(12):1615–1623

    Article  Google Scholar 

  • Rousseau JJ, Semple JH (1995) Two-person ratio efficiency games. Manag Sci 41:435–441

    Article  Google Scholar 

  • Talluri S, Yoon KP (2000) A cone-ratio DEA approach for AMT justification. Int J Prod Econ 66(2):119–129

    Article  Google Scholar 

  • Thompson RG, Singleton FD, Thrall RM, Smith BA (1986) Comparative site evaluations for locating a high-energy physics lab in Texas. Interfaces 16(6):35–49

    Article  Google Scholar 

  • Thompson RG, Langemeier LN, Lee CT, Lee E, Thrall RM (1990) The role of multiplier bounds in efficiency analysis with application to Kansas farming. J Econom 46(1–2):93–108

    Article  Google Scholar 

  • Tone K, Tsutsui M (2009) Network DEA: a slacks-based measure approach. Eur J Oper Res 197:243–252

    Article  Google Scholar 

  • Wong YH, Beasley JE (1990) Restricting weight flexibility in data envelopment analysis. J Oper Res Soc 41(9):829–835

    Article  Google Scholar 

Download references

Acknowledgments

The authors are indebted to Professor Podinovski and two anonymous reviewers for their thoughtful comments and suggestions. This research is supported by National Natural Science Funds of China (No. 71301155), National Natural Science Funds of China for Innovative Research Groups (No.71121061), and the Fundamental Research Funds for the Central Universities (J2014hgbz0172).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jingjing Ding.

Appendix

Appendix

Theorem 1

\(\theta_{CRM}^{*} = \mathop {\hbox{max} }\limits_{{\beta_{i} \in [LB_{i} ,UB_{i} ]}} \quad \theta_{CRD}^{*} ({\varvec{\upbeta}})\), where \(\theta_{CRM}^{*}\) is the optimal value of model ( 3 ), and \(\theta_{CRD}^{*} ({\varvec{\upbeta}})\) is the optimal value of model ( 6 ) with given \({\varvec{\upbeta}}\).

Proof

Note that θ * CRM can be secured by the decomposition approach, i.e.,\(\theta_{CRM}^{*} = \mathop {\hbox{max} }\limits_{{\beta_{i} \in [LB_{i} ,UB_{i} ]}} \, \theta_{CMP}^{*} ({\varvec{\upbeta}})\), where \(\theta_{CMP}^{*} ({\varvec{\upbeta}})\) is the optimal value of the inner problem of model (5). This theorem holds since the primal–dual gap between \(\theta_{CMP}^{*} ({\varvec{\upbeta}})\) and \(\theta_{CMD}^{*} ({\varvec{\upbeta}})\) is zero. Q.E.D.

Lemma 1

\(f_{i} ({\varvec{\upbeta}},{\varvec{\uplambda}}_{1} ,{\varvec{\uplambda}}_{2} )(i = 1, \ldots ,t)\) are quasi-convex with respect to \({\varvec{\upbeta}}\).

Proof

Suppose \({\varvec{\upbeta}}_{1}\),\({\varvec{\upbeta}}_{2}\) in [LU 1UB 1] × [LU 2UB 2] × ··· × [LU l UB l ]. For any λ ∊ [0, 1], we have

$$\begin{array}{l} f_{i} (\lambda {\varvec{\upbeta}}_{1} + (1 - \lambda ){\varvec{\upbeta}}_{2} ,{\varvec{\uplambda}}_{1} ,{\varvec{\uplambda}}_{2} ) = {{{\mathbf{a}}^{iT} \left( \begin{array}{l} \lambda {\varvec{\upbeta}}_{1} + (1 - \lambda ){\varvec{\upbeta}}_{2} X^{s} {\varvec{\uplambda}}_{1} \hfill \\ ({\mathbf{E}} - \lambda {\varvec{\upbeta}}_{1} + (1 - \lambda ){\varvec{\upbeta}}_{2} ){\mathbf{X}}^{s} {\varvec{\uplambda}}_{2} \hfill \\ {\mathbf{X}}^{1} {\varvec{\uplambda}}_{1} \, \hfill \\ \, {\mathbf{X}}^{2} {\varvec{\uplambda}}_{2} \hfill \\ \end{array} \right)} \mathord{\left/ {\vphantom {{{\mathbf{a}}^{iT} \left( \begin{array}{l} \lambda {\varvec{\upbeta}}_{1} + (1 - \lambda ){\varvec{\upbeta}}_{2} X^{s} {\varvec{\uplambda}}_{1} \hfill \\ ({\mathbf{E}} - \lambda {\varvec{\upbeta}}_{1} + (1 - \lambda ){\varvec{\upbeta}}_{2} ){\mathbf{X}}^{s} {\varvec{\uplambda}}_{2} \hfill \\ {\mathbf{X}}^{1} {\varvec{\uplambda}}_{1} \, \hfill \\ \, {\mathbf{X}}^{2} {\varvec{\uplambda}}_{2} \hfill \\ \end{array} \right)} {{\mathbf{a}}^{iT} \left( \begin{array}{l} \lambda {\varvec{\upbeta}}_{1} + (1 - \lambda )\beta_{2} X_{o}^{s} \hfill \\ ({\mathbf{E}} - \lambda {\varvec{\upbeta}}_{1} + (1 - \lambda ){\varvec{\upbeta}}_{2} ){\mathbf{X}}_{o}^{s} \hfill \\ {\mathbf{X}}_{o}^{1} \hfill \\ {\mathbf{X}}_{o}^{2} \hfill \\ \end{array} \right)}}} \right. \kern-0pt} {{\mathbf{a}}^{iT} \left( \begin{array}{l} \lambda {\varvec{\upbeta}}_{1} + (1 - \lambda )\beta_{2} X_{o}^{s} \hfill \\ ({\mathbf{E}} - \lambda {\varvec{\upbeta}}_{1} + (1 - \lambda ){\varvec{\upbeta}}_{2} ){\mathbf{X}}_{o}^{s} \hfill \\ {\mathbf{X}}_{o}^{1} \hfill \\ {\mathbf{X}}_{o}^{2} \hfill \\ \end{array} \right)}} \hfill \\ \quad = \rho {{{\mathbf{a}}^{iT} \left( \begin{array}{l} {\varvec{\upbeta}}_{1} {\mathbf{X}}^{s} {\varvec{\uplambda}}_{1} \hfill \\ ({\mathbf{E}} - {\varvec{\upbeta}}_{1} ){\mathbf{X}}^{s} {\varvec{\uplambda}}_{2} \hfill \\ {\mathbf{X}}^{1} {\varvec{\uplambda}}_{1} \, \hfill \\ \, {\mathbf{X}}^{2} {\varvec{\uplambda}}_{2} \hfill \\ \end{array} \right)} \mathord{\left/ {\vphantom {{{\mathbf{a}}^{iT} \left( \begin{array}{l} {\varvec{\upbeta}}_{1} {\mathbf{X}}^{s} {\varvec{\uplambda}}_{1} \hfill \\ ({\mathbf{E}} - {\varvec{\upbeta}}_{1} ){\mathbf{X}}^{s} {\varvec{\uplambda}}_{2} \hfill \\ {\mathbf{X}}^{1} {\varvec{\uplambda}}_{1} \, \hfill \\ \, {\mathbf{X}}^{2} {\varvec{\uplambda}}_{2} \hfill \\ \end{array} \right)} {{\mathbf{a}}^{iT} \left( \begin{array}{l} {\varvec{\upbeta}}_{1} {\mathbf{X}}_{o}^{s} \hfill \\ ({\mathbf{E}} - {\varvec{\upbeta}}_{1} ){\mathbf{X}}_{o}^{s} \hfill \\ {\mathbf{X}}_{o}^{1} \hfill \\ {\mathbf{X}}_{o}^{2} \hfill \\ \end{array} \right)}}} \right. \kern-0pt} {{\mathbf{a}}^{iT} \left( \begin{array}{l} {\varvec{\upbeta}}_{1} {\mathbf{X}}_{o}^{s} \hfill \\ ({\mathbf{E}} - {\varvec{\upbeta}}_{1} ){\mathbf{X}}_{o}^{s} \hfill \\ {\mathbf{X}}_{o}^{1} \hfill \\ {\mathbf{X}}_{o}^{2} \hfill \\ \end{array} \right)}} + (1 - \rho ){{{\mathbf{a}}^{iT} \left( \begin{array}{l} {\varvec{\upbeta}}_{2} {\mathbf{X}}^{s} {\varvec{\uplambda}}_{1} \hfill \\ ({\mathbf{E}} - {\varvec{\upbeta}}_{2} ){\mathbf{X}}^{s} {\varvec{\uplambda}}_{2} \hfill \\ {\mathbf{X}}^{1} {\varvec{\uplambda}}_{1} \, \hfill \\ \, {\mathbf{X}}^{2} {\varvec{\uplambda}}_{2} \hfill \\ \end{array} \right)} \mathord{\left/ {\vphantom {{{\mathbf{a}}^{iT} \left( \begin{array}{l} {\varvec{\upbeta}}_{2} {\mathbf{X}}^{s} {\varvec{\uplambda}}_{1} \hfill \\ ({\mathbf{E}} - {\varvec{\upbeta}}_{2} ){\mathbf{X}}^{s} {\varvec{\uplambda}}_{2} \hfill \\ {\mathbf{X}}^{1} {\varvec{\uplambda}}_{1} \, \hfill \\ \, {\mathbf{X}}^{2} {\varvec{\uplambda}}_{2} \hfill \\ \end{array} \right)} {{\mathbf{a}}^{iT} \left( \begin{array}{l} {\varvec{\upbeta}}_{2} {\mathbf{X}}_{o}^{s} \hfill \\ ({\mathbf{E}} - {\varvec{\upbeta}}_{2} ){\mathbf{X}}_{o}^{s} \hfill \\ {\mathbf{X}}_{o}^{1} \hfill \\ {\mathbf{X}}_{o}^{2} \hfill \\ \end{array} \right)}}} \right. \kern-0pt} {{\mathbf{a}}^{iT} \left( \begin{array}{l} {\varvec{\upbeta}}_{2} {\mathbf{X}}_{o}^{s} \hfill \\ ({\mathbf{E}} - {\varvec{\upbeta}}_{2} ){\mathbf{X}}_{o}^{s} \hfill \\ {\mathbf{X}}_{o}^{1} \hfill \\ {\mathbf{X}}_{o}^{2} \hfill \\ \end{array} \right)}} \hfill \\ \end{array}$$

where

$$\rho = {{{\mathbf{a}}^{iT} \left( \begin{array}{l} {\varvec{\upbeta}}_{1} {\mathbf{X}}_{o}^{s} \hfill \\ ({\mathbf{E}} - {\varvec{\upbeta}}_{1} ){\mathbf{X}}_{o}^{s} \hfill \\ {\mathbf{X}}_{o}^{1} \hfill \\ {\mathbf{X}}_{o}^{2} \hfill \\ \end{array} \right)} \mathord{\left/ {\vphantom {{{\mathbf{a}}^{iT} \left( \begin{aligned} {\varvec{\upbeta}}_{1} {\mathbf{X}}_{o}^{s} \hfill \\ ({\mathbf{E}} - {\varvec{\upbeta}}_{1} ){\mathbf{X}}_{o}^{s} \hfill \\ {\mathbf{X}}_{o}^{1} \hfill \\ {\mathbf{X}}_{o}^{2} \hfill \\ \end{aligned} \right)} {{\mathbf{a}}^{iT} \left( \begin{array}{l} \lambda {\varvec{\upbeta}}_{1} + (1 - \lambda ){\varvec{\upbeta}}_{2} {\mathbf{X}}_{o}^{s} \hfill \\ ({\mathbf{E}} - \lambda {\varvec{\upbeta}}_{1} + (1 - \lambda ){\varvec{\upbeta}}_{2} ){\mathbf{X}}_{o}^{s} \hfill \\ {\mathbf{X}}_{o}^{1} \hfill \\ {\mathbf{X}}_{o}^{2} \hfill \\ \end{array} \right)}}} \right. \kern-0pt} {{\mathbf{a}}^{iT} \left( \begin{array}{l} \lambda {\varvec{\upbeta}}_{1} + (1 - \lambda ){\varvec{\upbeta}}_{2} {\mathbf{X}}_{o}^{s} \hfill \\ ({\mathbf{E}} - \lambda {\varvec{\upbeta}}_{1} + (1 - \lambda ){\varvec{\upbeta}}_{2} ){\mathbf{X}}_{o}^{s} \hfill \\ {\mathbf{X}}_{o}^{1} \hfill \\ {\mathbf{X}}_{o}^{2} \hfill \\ \end{array} \right)}}.$$

Therefore, it follows that \(f_{i} (\lambda {\varvec{\upbeta}}_{1} + (1 - \lambda ){\varvec{\upbeta}}_{2} ,{\varvec{\uplambda}}_{1} ,{\varvec{\uplambda}}_{2} ) \le \hbox{max} \{ f_{i} ({\varvec{\upbeta}}_{1} ,{\varvec{\uplambda}}_{1} ,{\varvec{\uplambda}}_{2} ),f_{i} ({\varvec{\upbeta}}_{2} ,{\varvec{\uplambda}}_{1} ,{\varvec{\uplambda}}_{2} )\}\). Q.E.D.

Theorem 2

The CFM is equivalent to the Model (14).

Proof

The constraints \(\theta \ge f_{i} ({\varvec{\upbeta}},{\varvec{\uplambda}}_{1} ,{\varvec{\uplambda}}_{2} ) ,\quad i = 1, \ldots ,t,\beta_{i} \in [LB_{i} ,UB_{i} ]\) in model (13) are equivalent:

$$\begin{gathered} \theta \ge \mathop {\hbox{max} }\limits_{{\beta_{i} \in [LB_{i} ,UB_{i} ]}} \{ f_{1} ({\varvec{\upbeta}},{\varvec{\uplambda}}_{1} ,{\varvec{\uplambda}}_{2} ), \ldots ,f_{t} ({\varvec{\upbeta}},{\varvec{\uplambda}}_{1} ,{\varvec{\uplambda}}_{2} )\} \hfill \\ \quad = \hbox{max} \left\{ {\mathop {\hbox{max} }\limits_{{\beta_{i} \in [LB_{i} ,UB_{i} ]}} f_{1} ({\varvec{\upbeta}},{\varvec{\uplambda}}_{1} ,{\varvec{\uplambda}}_{2} ), \ldots ,\mathop {\hbox{max} }\limits_{{\beta_{i} \in [LB_{i} ,UB_{i} ]}} f_{t} ({\varvec{\upbeta}},{\varvec{\uplambda}}_{1} ,{\varvec{\uplambda}}_{2} )} \right\} \hfill \\ \end{gathered}$$

Note, \(f_{i} ({\varvec{\upbeta}},{\varvec{\uplambda}}_{1} ,{\varvec{\uplambda}}_{2} )\) (i = 1, …, t) are quasi-convex with respect to \({\varvec{\upbeta}}\) over the feasible region [LU 1UB 1] × [LU 2UB 2] × ··· × [LU l UB l ] according to Lemma 1. The maximal value of \(\mathop {\hbox{max} }\limits_{{\beta_{i} \in [LB_{i} ,UB_{i} ]}} f_{i} ({\varvec{\upbeta}},{\varvec{\uplambda}}_{1} ,{\varvec{\uplambda}}_{2} )\) is attained at some extreme point in EXT. Therefore, the constraints can be alternatively written as \(\theta \ge f_{i} ({\varvec{\upbeta}},{\varvec{\uplambda}}_{1} ,{\varvec{\uplambda}}_{2} ),\quad {\varvec{\upbeta}} \in \varvec{EXT},i = 1, \ldots ,t.\) Q.E.D.

Proposition 1

The efficiency of a DMU measured by CFM is not less than that measured by CRM.

Proof

Suppose the optimal RAV of CRM is β * i . Then the optimal value of CRM equals \(\theta_{CRD}^{*} (\beta_{i}^{*} )\), where \(\theta_{CRD}^{*} (\beta_{i}^{*} )\) is the optimal value of model (6) with β * i . Note that any feasible solution of model (6), given β * i , is a feasible solution of model (14). This is because the set of constraints of model (6), given β * i , are a subset of the set of constraints of model (14). Therefore, the optimal value of model (14), i.e., CFM is not less than \(\theta_{CRD}^{*} (\beta_{i}^{*} )\). Hence, the proposition follows.  Q.E.D.

Proposition 2

If U and V are nonnegative orthants, CRM is equivalent to (20), and the optimal RAVs can be any nonzero value in [LB 1UB 1] × ··· × [LB l UB l ].

Proof

It suffices to show that the optimal value of model (20) equals \(\mathop {\hbox{max} }\limits_{{\beta_{i} \in [LU_{i} ,UB_{i} ]}} \theta_{CRD}^{*} ({\varvec{\upbeta}})\) (Theorem 1). For any nonzero value in [LB 1UB 1] × ··· × [LB l UB l ], it is sufficient to show that the feasible region of model (20) is the smallest possible, or the feasible region of model (20) is contained in the feasible region of model (6) with arbitrary feasible \({\varvec{\upbeta}}\). Note, the constraints related to \({\varvec{\upbeta}}\) of model (6) are as follows:

$$\left. \begin{array}{l} \beta_{i} \left( {\theta x_{io}^{s} - \sum\limits_{j = 1}^{n} {\lambda_{j}^{1} x_{ij}^{s} } } \right) \ge 0 \hfill \\ (1 - \beta_{i} )\left( {\theta x_{io}^{s} - \sum\limits_{j = 1}^{n} {\lambda_{j}^{2} x_{ij}^{s} } } \right) \ge 0 \hfill \\ \end{array} \right\}\quad i = 1, \ldots ,l$$

Further note that if β i is set to zero, the consequence is to delete the associated constraints for model (6). In other words, the closer β i is to zero, the smaller the number of effective constraints remain in model (6), i.e., the feasible region becomes the smallest when β i is nonzero for i = 1, …, l. Therefore, the optimal value of the model (20) is \(\mathop {\hbox{max} }\limits_{{\beta_{i} \in [LU_{i} ,UB_{i} ]}} \, \theta_{CRD}^{*} ({\varvec{\upbeta}})\). Q.E.D.

Proposition 3

If U and V are nonnegative orthants, the reduced CRM is equal to ( 21 ). Therefore, the optimal values are bound-dependent.

Proof

To see this, we first note that, if U and V are nonnegative orthants, the reduced CRM can be transformed into the linear equivalent model through substitution of variables. For the sake of space, only the dual of the linear equivalent program is presented here:

$$\begin{array}{*{20}l} {\hbox{min} } \hfill & \theta \hfill \\ {s.t.} \hfill & {\sum\limits_{j = 1}^{n} {\lambda_{j}^{1} y_{rj}^{1} } \ge y_{ro}^{1} } \hfill \\ {} \hfill & {\sum\limits_{j = 1}^{n} {\lambda_{j}^{2} y_{rj}^{2} } \ge y_{ro}^{2} } \hfill \\ {} \hfill & { \, \sum\limits_{j = 1}^{n} {\lambda_{j}^{1} x_{ij}^{1} } \le \theta x_{io}^{1} } \hfill \\ {} \hfill & {\sum\limits_{j = 1}^{n} {\lambda_{j}^{2} x_{ij}^{2} } \le \theta x_{io}^{2} } \hfill \\ {} \hfill & {b_{i} - a_{i} + \sum\limits_{j = 1}^{n} {\lambda_{j}^{2} (x_{ij}^{s} )} = \sum\limits_{j = 1}^{n} {\lambda_{j}^{1} (x_{ij}^{s} )} } \hfill \\ {} \hfill & {b_{i} UB_{i} - a_{i} LB_{i} + \sum\limits_{j = 1}^{n} {\lambda_{j}^{2} (x_{ij}^{s} )} \le \theta x_{io}^{s} } \hfill \\ {} \hfill & {\lambda_{j}^{1} ,\lambda_{j}^{2} ,a_{i} ,b_{i} \ge 0.} \hfill \\ \end{array}$$
(23)

For any i (i = 1, …, l), the optimal β * i is obtained as: β * i  = v s* i /v * i , where v s* i and v * i are the shadow prices associated with constraint \(b_{i} - a_{i} + \sum\nolimits_{j = 1}^{n} {\lambda_{j}^{2} \left( {x_{ij}^{s} } \right)} = \sum\nolimits_{j = 1}^{n} {\lambda_{j}^{1} \left( {x_{ij}^{s} } \right)}\) and \(b_{i} UB_{i} - a_{i} LB_{i} + \sum\nolimits_{j = 1}^{n} {\lambda_{j}^{2} \left( {x_{ij}^{s} } \right) \le \theta x_{io}^{s} }\), respectively, if v * i  ≠ 0, or β * i can take any value in[LB i UB i ].

By Proposition 1, the optimal value of model (21), denote θ * CFM , is greater or equal to the optimal value of the reduced CRM, denoteθ * RCRM , orθ * CFM  ≥ θ * RCRM  = θ * A1 , where θ * A1 is the optimal value of mode (23). It remains to be shown that θ * CFM  ≤ θ * RCRM  = θ * A1 . We first present the following fact summarized in Lemma A1.

Lemma (A1)

There exists an optimal solution (λ 1* j λ 2* j a * i b * i θ * A1 ) to model ( 23 ) with a * i b * i  = 0.

Proof

To see this, suppose a * i b * i  ≠ 0. Let a * i  − b * i  = c i , where c i can be positive, negative or zero. Assume, without loss of generality, that c i  < 0 for illustrative purpose. Let \(\bar{b} = b_{i}^{*} - a_{i}^{*} = c_{i} ,\bar{a} = 0\). Then,

$$\left\{ \begin{gathered} \bar{b} - \bar{a} + \sum\nolimits_{j = 1}^{n} {\lambda_{j}^{2*} (x_{ij}^{s} )} = \sum\nolimits_{j = 1}^{n} {\lambda_{j}^{1*} \left( {x_{ij}^{s} } \right)} \hfill \\ (\bar{b} - \bar{a})UB_{i} + \bar{a}(UB_{i} - LB_{i} ) + \sum\nolimits_{j = 1}^{n} {\lambda_{j}^{2} \left( {x_{ij}^{s} } \right)} \le \hfill \\ (b_{i}^{*} - a_{i}^{*} )UB_{i} + a_{i}^{*} (UB_{i} - LB_{i} ) + \sum\nolimits_{j = 1}^{n} {\lambda_{j}^{2} \left( {x_{ij}^{s} } \right)} \le \theta^{*} x_{io}^{s} \hfill \\ \end{gathered} \right.$$

Note, \((\lambda_{j}^{1*} ,\lambda_{j}^{2*} ,\bar{a},\bar{b},\theta_{A1}^{*} )\) is a feasible solution to model (23), and so is an optimal solution. Thus, Lemma A1 holds. Q.E.D.

Based on Lemma A1, we have \(\left. \begin{array}{l} b_{i}^{*} + \sum\nolimits_{j = 1}^{n} {\lambda_{j}^{2*} \left( {x_{ij}^{s} } \right)} = \sum\nolimits_{j = 1}^{n} {\lambda_{j}^{1*} \left( {x_{ij}^{s} } \right)} \\ b_{i}^{*} UB_{i} + \sum\nolimits_{j = 1}^{n} {\lambda_{j}^{2*} \left( {x_{ij}^{s} } \right)} \le \theta^{*} x_{io}^{s} \\ \end{array} \right\}\) if a * i  = 0; otherwise, \(\left. \begin{gathered} - a_{i}^{*} + \sum\nolimits_{j = 1}^{n} {\lambda_{j}^{2*} \left( {x_{ij}^{s} } \right)} = \sum\nolimits_{j = 1}^{n} {\lambda_{j}^{1*} \left( {x_{ij}^{s} } \right)} \hfill \\ - a_{i}^{*} LB_{i} + \sum\nolimits_{j = 1}^{n} {\lambda_{j}^{2*} \left( {x_{ij}^{s} } \right)} \le \theta^{*} x_{io}^{s} \hfill \\ \end{gathered} \right\}\) if b * i  = 0; which means:

$$\begin{gathered} \sum\limits_{j = 1}^{n} {\lambda_{j}^{2*} \left( {x_{ij}^{s} } \right) - \left( {\sum\limits_{j = 1}^{n} {\lambda_{j}^{2*} \left( {x_{ij}^{s} } \right) - \sum\limits_{j = 1}^{n} {\lambda_{j}^{1*} \left( {x_{ij}^{s} } \right)} } } \right)UB_{i} } \le \theta^{*} x_{io}^{s} \quad\quad {\text{if}}\;\sum\limits_{j = 1}^{n} {\lambda_{j}^{2*} \left( {x_{ij}^{s} } \right) - \sum\limits_{j = 1}^{n} {\lambda_{j}^{1*} \left( {x_{ij}^{s} } \right)} } \le 0 \hfill \\ \sum\limits_{j = 1}^{n} {\lambda_{j}^{2*} \left( {x_{ij}^{s} } \right) - \left( {\sum\limits_{j = 1}^{n} {\lambda_{j}^{2*} \left( {x_{ij}^{s} } \right)} - \sum\limits_{j = 1}^{n} {\lambda_{j}^{1*} \left( {x_{ij}^{s} } \right)} } \right)LB_{i} } \le \theta^{*} x_{io}^{s} \quad\quad {\text{if}}\;\sum\limits_{j = 1}^{n} {\lambda_{j}^{2*} \left( {x_{ij}^{s} } \right) - \sum\limits_{j = 1}^{n} {\lambda_{j}^{1*} \left( {x_{ij}^{s} } \right)} } \ge 0 \hfill \\ \end{gathered}$$

Therefore, (λ 1* j λ 2* j θ * A1 ) is a feasible solution to model (21). To see this, note that

(i) If \(\sum\nolimits_{j = 1}^{n} {\lambda_{j}^{2*} \left( {x_{ij}^{s} } \right) - \sum\nolimits_{j = 1}^{n} {\lambda_{j}^{1*} \left( {x_{ij}^{s} } \right)} } \le 0\), then

\(\sum\nolimits_{j = 1}^{n} {\lambda_{j}^{2*} \left( {x_{ij}^{s} } \right) - \left( {\sum\limits_{j = 1}^{n} {\lambda_{j}^{2*} \left( {x_{ij}^{s} } \right) - \sum\limits_{j = 1}^{n} {\lambda_{j}^{1*} \left( {x_{ij}^{s} } \right)} } } \right)LB_{i} } \le \sum\nolimits_{j = 1}^{n} {\lambda_{j}^{2*} \left( {x_{ij}^{s} } \right) - \left( {\sum\nolimits_{j = 1}^{n} {\lambda_{j}^{2*} \left( {x_{ij}^{s} } \right)} - \sum\nolimits_{j = 1}^{n} {\lambda_{j}^{1*} \left( {x_{ij}^{s} } \right)} } \right)UB_{i} } \le \theta^{*} x_{io}^{s}\), (ii) If \(\sum\nolimits_{j = 1}^{n} {\lambda_{j}^{2*} \left( {x_{ij}^{s} } \right) - \sum\nolimits_{j = 1}^{n} {\lambda_{j}^{1*} \left( {x_{ij}^{s} } \right) \ge 0} }\), then \(\sum\nolimits_{j = 1}^{n} {\lambda_{j}^{2*} \left( {x_{ij}^{s} } \right) - \left( {\sum\nolimits_{j = 1}^{n} {\lambda_{j}^{2*} \left( {x_{ij}^{s} } \right)} - \sum\nolimits_{j = 1}^{n} {\lambda_{j}^{1*} \left( {x_{ij}^{s} } \right)} } \right)UB_{i} } \le \sum\nolimits_{j = 1}^{n} {\lambda_{j}^{2*} \left( {x_{ij}^{s} } \right) - \left( {\sum\nolimits_{j = 1}^{n} {\lambda_{j}^{2*} \left( {x_{ij}^{s} } \right)} - \sum\nolimits_{j = 1}^{n} {\lambda_{j}^{1*} \left( {x_{ij}^{s} } \right)} } \right)LB_{i} } \le \theta^{*} x_{io}^{s} .\)

It is then clear that the constraints related to the shared inputs can be fully satisfied.

Thus, θ * CFM  ≤ θ * A1 . Q.E.D.

Proposition 4

If β 1 = β 2 = ··· = β l  ∊ [LBUB], or l = 1, \(\mathop {\hbox{max} }\limits_{{\beta_{i} \in [LB,UB]}} \, \theta_{CRD}^{*} ({\varvec{\upbeta}})\) is attained at some end point LB or UB.

Proof

Suppose (λ 11* j λ 21* j θ * CRD (β 1)) and (λ 12* j λ 22* j θ * CRD (β 2)) are optimal solutions to model (6), respectively, corresponding to β = β 1 and β = β 2. Now let us construct λ 1* j and λ 2* j as follows:

$$\left\{ \begin{array}{l} \lambda_{j}^{1*} = \frac{{\lambda \beta_{1} \lambda_{1}^{11*} + (1 - \lambda )\beta_{2} \lambda_{1}^{12*} }}{{\lambda \beta_{1} + (1 - \lambda )\beta_{2} }} \hfill \\ \lambda_{j}^{2*} = \frac{{\lambda (1 - \beta_{1} )\lambda_{1}^{21*} + (1 - \lambda )(1 - \beta_{2} )\lambda_{1}^{22*} }}{{1 - \lambda \beta_{1} - (1 - \lambda )\beta_{2} }} \hfill \\ \end{array} \right..$$

where λ is an arbitrary value in [0,1]. Obviously, λ 1* j and λ 2* j are nonnegative. Now substitute them into the model (6), we have

$$\begin{gathered} \sum\limits_{j = 1}^{n} {\lambda_{j}^{1*} y_{rj}^{1} } \ge y_{ro}^{1} \quad r = 1, \ldots ,s_{1} \hfill \\ \sum\limits_{j = 1}^{n} {\lambda_{j}^{2*} y_{rj}^{2} } \ge y_{ro}^{2} \quad r = 1, \ldots ,s_{2} \hfill \\ \left[ {\beta^{*} \sum\limits_{j = 1}^{n} {\lambda_{j}^{1*} X_{j}^{s} } { + (1} - \beta^{*} )\sum\limits_{j = 1}^{n} {\lambda_{j}^{2*} X_{j}^{s} } } \right] \le \left( {\lambda \theta_{CRD}^{*} (\beta_{1} ) + (1 - \lambda )\theta_{CRD}^{*} (\beta_{2} )} \right)X_{o}^{s} \quad i = 1, \ldots ,l \, \hfill \\ \sum\limits_{j = 1}^{n} {\lambda_{j}^{1*} x_{ij}^{1} } \le \left( {\frac{{\lambda \beta_{1} }}{{\lambda \beta_{1} + (1 - \lambda )\beta_{2} }}\theta_{CRD}^{*} (\beta_{1} ) + \frac{{(1 - \lambda )\beta_{2} }}{{\lambda \beta_{1} + (1 - \lambda )\beta_{2} }}\theta_{CRD}^{*} (\beta_{2} )} \right)x_{io}^{1} \quad i = 1, \ldots ,m_{1} \hfill \\ \sum\limits_{j = 1}^{n} {\lambda_{j}^{2*} x_{ij}^{2} } \le \left( {\frac{{\lambda (1 - \beta_{1} )}}{{1 - \lambda \beta_{1} - (1 - \lambda )\beta_{2} }}\theta_{CRD}^{*} (\beta_{1} ) + \frac{{(1 - \lambda )(1 - \beta_{2} )}}{{1 - \lambda \beta_{1} - (1 - \lambda )\beta_{2} }}\theta_{CRD}^{*} (\beta_{2} )} \right)x_{io}^{2} \quad i = 1, \ldots ,m_{2} \hfill \\ \end{gathered},$$

where β * = λβ 1 + (1 − λ)β 2.

Let,

$$\begin{gathered} \theta^{*} = \hbox{max} \left\{ {\lambda \theta_{CRD}^{*} (\beta_{1} ) + (1 - \lambda )\theta_{CRD}^{*} (\beta_{2} ),\frac{{\lambda \beta_{1} }}{{\lambda \beta_{1} + (1 - \lambda )\beta_{2} }}\theta_{CRD}^{*} (\beta_{1} ) + \frac{{(1 - \lambda )\beta_{2} }}{{\lambda \beta_{1} + (1 - \lambda )\beta_{2} }}\theta_{CRD}^{*} (\beta_{2} ),} \right. \hfill \\ \quad \left. {\frac{{\lambda (1 - \beta_{1} )}}{{1 - \lambda \beta_{1} - (1 - \lambda )\beta_{2} }}\theta_{CRD}^{*} (\beta_{1} ) + \frac{{(1 - \lambda )(1 - \beta_{2} )}}{{1 - \lambda \beta_{1} - (1 - \lambda )\beta_{2} }}\theta_{CRD}^{*} (\beta_{2} )\} } \right\} \hfill \\ \end{gathered}$$

It then follows that (λ 1* j λ 2* j θ *) is a feasible solution to model (6) with β = β *. Therefore, the optimal solution with respect to β *, θ * CRD (β *) should be less than or equal to θ *.

Given that θ * is a convex combination of θ * CRD (β 1) and θ * CRD (β 2), we have θ * CRD (β *) ≤ max {θ * CRD (β 1), θ * CRD (β 2)}, where β * = λβ 1 + (1 − λ)β 2.

Since λ is an arbitrarily chosen convex multiplier, we conclude that \(\mathop {\hbox{max} }\limits_{{\beta_{{}} \in [LB,UB]}} \, \theta_{CRD}^{*} (\beta )\) is attained at LB or UB. Q.E.D.

Proposition 5

If LB i  = 0, UB i  = 1, CRM, the reduced CRM, and the associated CFMs (20) and (21) are all equivalent.

Proof

The equivalence between CFMs (20) and (21) is straightforward, if we substitute LB i  = 0, UB i  = 1in (21). Following Proposition 2 and 3, the equivalence across CRM, reduced CRM, and CFMs (20) and (21) is established. Q.E.D.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ding, J., Feng, C., Bi, G. et al. Cone ratio models with shared resources and nontransparent allocation parameters in network DEA. J Prod Anal 44, 137–155 (2015). https://doi.org/10.1007/s11123-014-0420-0

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11123-014-0420-0

Keywords

JEL classification

Navigation