Skip to main content
Log in

Sorting the good guys from bad: on the optimal audit structure with ex-ante information acquisition

  • Research Article
  • Published:
Economic Theory Aims and scope Submit manuscript

Abstract

In a costly state verification model under commitment, the principal may acquire a costly public and imperfectly revealing signal before or after contracting. If the project remains profitable after all signal realizations, optimally the signal is collected, if at all, only after contracting, and it may be acquired randomly or deterministically. Moreover, audit is deterministic and targeted on some signal-state combinations. The paper provides a detailed characterization of the optimal contract and performs a comparative static analysis of signal acquisition strategy and payoffs with respect to enforcement costs and informativeness of the signal. We explore the robustness of the results, including commitment issues. The results are interpreted in light of the observed features of financial contracts, showing that the optimality of the standard debt contract with deterministic audit extends to a setting with random audit.

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
Fig. 4
Fig. 5
Fig. 6
Fig. 7

Similar content being viewed by others

Notes

  1. We will henceforth refer to the ex-ante public information collected by the principal as a signal about the state, although it does not have the usual meaning of a privately informed person signalling (but see Sect. 8.2 for a discussion of this case).

  2. The NRP has replaced the Taxpayer Compliance Measurement Program (TCMP), which consisted in more thorough audits of stratified random sample of taxpayers (for a description of both programs, see Andreoni et al. 1998; Erard and Feinstein 2010). Similar methods are used also in other countries.

  3. By contrast, Carlier and Renou (2005) find that exogenous fixed differences in beliefs between lender and borrower may mean that the optimal contract does not resemble debt.

  4. The no-commitment case is dealt with in Sect. 8.1.

  5. We model a financial contracting relationship, but the setup is general and can be adapted to apply to a variety of different contexts. Mookherjee and Png (1989) provide an interesting interpretation of a similar problem in a context of taxation.

  6. This can be seen by adding (2) and (3).

  7. This is then consistent either with the role of the principal as a utilitarian regulator who just wants to achieve an efficient monitoring system (e.g., a tax authority with a fixed revenue requirement who wishes to minimize the enforcement cost), or of a principal that maximizes her own expected payoff but, due to competition between principals, is driven to offer contracts yielding her a zero expected payoff. Border and Sobel (1987) point out that for many objectives of the principal (the utilitarian case referred to above, minimizing the expected audit cost or maximizing the achievable expected return to the principal) an optimal scheme will be audit efficient.

  8. The implications of relaxing this assumption and letting the principal get all the surplus are discussed in Sect. 8.3.

  9. Since repayments are non-decreasing with the state (see above), any report of the high state must be truthful and hence it is a dominant strategy never to audit the high state report.

  10. The cost saving to audit depends on how the informativeness of the signal leads to a differential between monitoring probabilities in good and bad signal states and can be written as \(\rho c_{m}\left( m_{L}^{G}-m_{L}^{B}\right) \). This reduces to \(\rho c_{m}\) as \( m_{L}^{G}=1,m_{L}^{B}=0.\) Consequently, the higher the auditing cost relative to the signal acquisition cost, the higher the gain from signal acquisition.

  11. The propositions also hold if it is the agent who can collect the signal.

  12. Full details of these statements are available in the working paper (CSEF WP n. 201, www.csef.it/WP/wp201.pdf).

  13. Where \(\rho =(\pi _{HG}+\pi _{MG})\pi _{LB}-(\pi _{HB}+\pi _{MB})\pi _{LG},\) defined in (4).

  14. i.e., for \(c_{m}\) no less than the value at which \(EPC_{\sigma }\) and \( NC_{a} \) intersect: \(c_{m}\ge \widehat{c}_{m}.\)

  15. i.e., for \(c_{a}\le \widehat{c}_{a},\) the value at which \(EPC_{\sigma }\) and \(NC_{a}\) intersect.

  16. The literature which does consider signalling highlights this role for ex-ante information acquisition (Crémer and Khalil 1992).

  17. The ambiguity remains even in the extreme case in which the signal is perfectly informative.

  18. The principal could cheat on the realization of the random device, either by using a public device with the wrong odds or playing the device privately.

  19. Another way round the no-commitment problem has the two parties engaging in some bargaining after the state is realized, e.g., collude to find an ex-post efficient repayment combined with no audit (Dewatripont 1988), or noncooperatively trade-off the threat of audit against the threat of cheating on the report (Khalil 1997; Khalil and Parigi 1998; Menichini and Simmons 2006). Knowing that some such process could occur, the contract would have to include a renegotiation-proof constraint, so that only those reporting and audit schemes that will actually be implemented subsequently are feasible. This leads to deterministic audit again (Krasa and Villamil 2000; Hvide and Leite 2010).

  20. We have assumed that ex-ante competition between principals leads to a contract maximizing the agent’s expected payoff. However, ex-post, once the contract is agreed, the agent is locked in and the principal acts in his own interest.

  21. For this to be negative, we need \(EP_{B}=\left( \pi _{HB}+\pi _{MB}\right) \left( f_{M}-f_{L}\right) -\pi _{LB}c_{m}>0\) at the value of \(c_{m}\) at which \(EPC_{\sigma }=0\). This follows from assumption 1 and \( EP=EP_{B}+EPC_{\sigma }\).

References

  • Andreoni, J., Erard, B., Feinstein, J.: Tax compliance. J. Econ. Lit. 36, 818–860 (1998)

    Google Scholar 

  • Border, K.C., Sobel, J.: Samurai accounting: a theory of auditing and plunder. Rev. Econ. Stud. 54, 525–540 (1987)

    Article  Google Scholar 

  • Carlier, G., Renou, L.: A costly state verification model with diversity of opinions. Econ. Theory 25, 497–504 (2005)

    Article  Google Scholar 

  • Chander, P., Wilde, L.L.: A general characterization of optimal income tax enforcement. Rev. Econ. Stud. 65, 165–183 (1998)

    Article  Google Scholar 

  • Chew, M.C.: On pairing observations from a distribution with monotone likelihood ratio. Ann. Stat. 1(3), 433–445 (1973)

    Article  Google Scholar 

  • Crémer, J., Khalil, F.: Gathering information before signing a contract. Am. Econ. Rev. 82, 566–578 (1992)

    Google Scholar 

  • Dewatripont, M.: Commitment through renegotiation-proof contracts with third parties. Rev. Econ. Stud. 55(3), 377–389 (1988)

    Article  Google Scholar 

  • Erard, B., Feinstein, J.: Econometric models for multi-stage audit processes: an application to the IRS national research program. In: Alm, J., Martinez-Vazquez, J., Torgler, B. (eds.) Developing Alternative Frameworks for Explaining Tax Compliance, pp. 113–137. Routledge, London (2010)

    Google Scholar 

  • Gale, D., Hellwig, M.: Incentive-compatible debt contracts: the one period problem. Rev. Econ. Stud. 52, 647–663 (1985)

    Article  Google Scholar 

  • Hart, O.: Firms, Contracts and Financial Structure. Clarendon Press, Oxford (1995)

    Book  Google Scholar 

  • Hvide, H.K., Leite, T.E.: Optimal debt contracts under costly enforcement. Econ. Theory 44, 149–165 (2010)

    Article  Google Scholar 

  • Khalil, F.: Auditing without commitment. RAND J. Econ. 28, 629–640 (1997)

    Article  Google Scholar 

  • Khalil, F., Parigi, B.: Loan size as a commitment device. Int. Econ. Rev. 39, 135–150 (1998)

    Article  Google Scholar 

  • Krasa, S., Villamil, A.P.: Optimal multilateral contracts. Econ. Theory 4, 167–187 (1994)

    Article  Google Scholar 

  • Krasa, S., Villamil, A.P.: Optimal contracts when enforcement is a decision variable. Econometrica 68, 119–134 (2000)

    Article  Google Scholar 

  • Krasa, S., Sharma, T., Villamil, A.P.: Debt contracts with cooperative deviations. J. Math. Econ. 41, 856–874 (2005)

    Article  Google Scholar 

  • Lewis, T.R., Sappington, D.E.M.: Information management in incentive problems. J. Political Econ. 105, 796–821 (1997)

    Article  Google Scholar 

  • Macho-Stadler, I., Pérez-Castrillo, D.: Auditing with signals. Economica 69, 1–20 (2002)

    Article  Google Scholar 

  • Maskin, E., Tirole, J.: The principal-agent relationship with an informed principal: the case of private values. Econometrica 58, 379–409 (1990)

    Article  Google Scholar 

  • Maskin, E., Tirole, J.: The principal-agent relationship with an informed principal, II: common values. Econometrica 60, 1–42 (1992)

    Article  Google Scholar 

  • Menichini, A.M.C., Simmons, P.: Liars and inspectors: optimal financial contracts when monitoring is non-observable. Contrib. Theor. Econ. 6, 1–19 (2006)

    Article  Google Scholar 

  • Mookherjee, D., Png, I.: Optimal auditing, insurance and redistribution. Q. J. Econ. 104, 399–415 (1989)

    Article  Google Scholar 

  • Reinganum, J.F., Wilde, L.L.: Equilibrium verification and reporting policies in a model of tax compliance. Int. Econ. Rev. 27, 739–760 (1986)

    Article  Google Scholar 

  • Scotchmer, S.: Audit classes and tax enforcement policy. Am. Econ. Rev. 77, 229–233 (1987)

    Google Scholar 

  • Thuronyi, V.: Presumptive taxation. In: Thuronyi, V. (ed.) Tax Law Design and Drafting, vol. 1. International Monetary Fund, Washington, DC (1996)

  • Townsend, R.: Optimal contracts and competitive markets with costly state verification. J. Econ. Theory 21, 265–293 (1979)

    Article  Google Scholar 

  • Williamson, S.D.: Costly monitoring, financial intermediation and equilibrium credit rationing. J. Monet. Econ. 18, 159–179 (1986)

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Anna Maria C. Menichini.

Additional information

We are grateful to the Editor Nicholas Yannelis, two anonymous referees, Sandeep Baliga, Alberto Bennardo, Gabriella Chiesa, Marcello D’Amato, Paolo Garella, Tracy Lewis, Dilip Mookherjee, Marco Pagano, Nicola Persico, Ailsa Roell, Maria Grazia Romano, Joel Sobel, Francesca Toscano, Zaifu Yang, the participants of the CSEF-IGIER Symposium on Economics and Institutions in Capri, the ASSET Conference in Padua, the RES Conference in Surrey, the SIE Conference in Rome, the EARIE Conference in Toulouse, as well as the seminar participants at the universities of Bologna and Naples. The usual disclaimer applies.

Appendix

Appendix

1.1 Signal after the contract

Proof of Proposition 1

The proof proceeds in steps. We first show that the incentive constraints (20) and (21) bind. Then, that \( m_{M}^{N}=m_{M}^{\sigma }=0\). Further that also the incentive constraints ( 24) and (25) bind. Last, we prove that (22) and (23) are slack. This allows us to set the contract problem as one in two states, where the top two, \(H\) and \(M\), are pooled into a single state.

  1. 1.

    The incentive constraints (20) and (21) bind. It follows from Border and Sobel (1987) that the bottom non-monitored state has a binding incentive constraint with the top monitored state. So if \( m_{M}^{N},m_{M}^{\sigma }>0,\) (20) and (21) must bind. Conversely, if \(m_{M}^{N}=m_{M}^{\sigma }=0,\) repayments must be pooled in the top two states and \(R_{\hat{H}}^{N}=R_{\hat{ M}}^{N},R_{\hat{H}}^{\sigma }=R_{\hat{M}}^{\sigma }\).

  2. 2.

    \(m_{M}^{N}=m_{M}^{\sigma }=0\) (this is proved below for the no-signal case, but the proof is analogous for the with-signal case).

    The rationale is that when \(m_{M}^{N}>0\) and \(R_{\hat{M}}^{N}<f_{M}\), there are savings to be made on audit cost by reducing \(m_{M}^{N}\) and raising \(R_{ \hat{M}}^{N}\) in a way which has a neutral effect on the participation constraint (keeping \(R_{\hat{M}}^{N}\left( 1-m_{M}^{N}\right) \) constant), but slackens (24), reducing the enforcement cost. This goes on until we hit one of the two corners, i.e., either \(m_{M}^{N}=0,\) or \(R_{ \hat{M}}^{N}=f_{M}\) or both.

    Suppose we first reach \(m_{M}^{N}=0\) and \(R_{\hat{M}}^{N}<f_{M}.\) Then, \(R_{ \hat{M}M}^{N}\) is irrelevant and from (20) binding \(R_{\hat{M} }^{N}=R_{\hat{H}}^{N}.\) If in constraint (24) we replace \(R_{\hat{H} }^{N}\) with \(R_{\hat{M}}^{N}\) and rearrange, this becomes \(R_{\hat{M} }^{N}\le \left( 1-m_{L}^{N}\right) f_{L}+m_{L}^{N}f_{M}.\) If we hit \(R_{ \hat{M}}^{N}=f_{M}\) with \(m_{M}^{N}>0,\) we can reduce both \(R_{\hat{H}}^{N}\) and \(m_{M}^{N},\) preserving the participation constraint. Doing this, we can attain \(m_{M}^{N}=0.\) To see this, replace \(R_{\hat{H}}^{N}\) by \(f_{M},\) \(R_{ \hat{M}}^{N}\) by \(f_{M}\) and \(m_{M}^{N}\) by zero in the participation constraint (19), to get

    $$\begin{aligned} \left( \pi _{H}+\pi _{M}\right) f_{M}+\pi _{L}f_{L}-I-\left( \alpha \Sigma _{\sigma }\pi _{L\sigma }m_{L}^{\sigma }+\left( 1-\alpha \right) \pi _{L}m_{L}^{N}\right) c_{m}-\alpha c_{a}\ge EP. \end{aligned}$$

    This is positive by assumption 1 and shows that there exists a value of \(R_{\hat{H}}^{N}\ge f_{M}\) with \(m_{M}^{N}=0,R_{\hat{M}}^{N}=f_{M}\), which satisfies the participation constraint.

  1. 3.

    The incentive constraints (24) and (25) bind (this is proved below for the no-signal case, but the proof is analogous for the with-signal case).

    Consider constraints (22) and (24). Suppose we have an optimum with (24) slack and, as we know from Border and Sobel (1987), since the lowest state always has positive monitoring, \( m_{L}^{N}>0.\) The optimum must satisfy both the above inequalities and the participation constraint.

    Keep all the \(R_{\hat{H}}^{N},R_{\hat{M}}^{N}\) fixed, so the right-hand sides of the incentive constraints are all fixed. Reducing \(m_{L}^{N}\) will keep the second inequality valid for small changes but could violate the first inequality (if it was binding initially). Suppose it was binding initially. Then, it is possible to reduce \(R_{\hat{H}}^{N}\) too to keep (22) binding. This will also slacken (20). The effect, if possible, is to reduce agency cost. Reducing \(R_{\hat{H}}^{N}\) and \( m_{L}^{N} \) also has implications for the participation constraint. Can you keep doing this until (24) binds at the given \(R_{\hat{M} }^{N}\)? If so to keep (22) binding, there will be finite changes \( \Delta R_{\hat{H}}^{N},\Delta m_{L}^{N}\) such that

    $$\begin{aligned} \left( f_{H}-f_{L}\right) \left( 1-m_{L}^{N}-\Delta m_{L}^{N}\right)&= f_{H}-R_{\hat{H}}^{N}-\Delta R_{\hat{H}}^{N}, \end{aligned}$$
    (51)
    $$\begin{aligned} \left( f_{M}-f_{L}\right) \left( 1-m_{L}^{N}-\Delta m_{L}^{N}\right)&= f_{M}-R_{\hat{M}}^{N}. \end{aligned}$$
    (52)

    Equality (51) will certainly hold if \(\left( f_{H}-f_{L}\right) \Delta m_{L}^{N}\ge \Delta R_{\hat{H}}^{N}\) or \(\Delta R_{\hat{H} }^{N}/\Delta m_{L}^{N}\le f_{H}-f_{L}.\) Within a given signal situation, the participation constraint is \(\pi _{H}R_{\hat{H}}^{N}+\pi _{M}R_{\hat{M} }^{N}+\pi _{L}\left( f_{L}-m_{L}^{N}c_{m}\right) .\) To prevent this falling with simultaneous reductions in \(R_{\hat{H}}^{N}\) and \(m_{L}^{N}\) needs \( \pi _{H}\Delta R_{\hat{H}}^{N}\ge \pi _{L}c_{m}\Delta m_{L}^{N}\) or \( \Delta R_{\hat{H}}^{N}/\Delta m_{L}^{N}\ge \pi _{L}c_{m}/\pi _{H}.\) So we need both \(\Delta R_{\hat{H}}^{N}/\Delta m_{L}^{N}\le f_{H}-f_{L}\) and \( \Delta R_{\hat{H}}^{N}/\Delta m_{L}^{N}\ge \pi _{L}c_{m}/\pi _{H}.\) We can find such simultaneous changes \(\Delta R_{\hat{H}}^{N},\Delta m_{L}^{N}\) if \( \pi _{H}(f_{H}-f_{L})\ge \pi _{L}c_{m}.\) We can reduce \(m_{L}^{N}\) and \(R_{ \hat{H}}^{N}\) to reach a solution where (16) binds if \(\pi _{H}(f_{H}-f_{L})\ge \pi _{L}c_{m}\) so long as there are \(0\le m_{L}^{N^{*}}\le 1,R_{\hat{H}}^{N^{*}}<f_{H}\) such that

    $$\begin{aligned} \left( f_{H}-f_{L}\right) \left( 1-m_{L}^{N^{*}}\right)&\le f_{H}-R_{ \hat{H}}^{N^{*}} \\ \left( f_{M}-f_{L}\right) \left( 1-m_{L}^{N^{*}}\right)&= f_{M}-R_{\hat{ M}}^{N} \end{aligned}$$

    which would imply

    $$\begin{aligned} \left( f_{M}-f_{L}\right) \left( 1-m_{L}^{N^{*}}\right)&= f_{M}-R_{\hat{ M}}^{N} \\&\le f_{M}-f_{L}\le f_{H}-f_{L} \end{aligned}$$

    But this is always true. So it is always possible to reduce \(m_{L}^{N}\) sufficiently to make (24) bind.

  1. 4.

    The incentive constraints (22) and (23) are slack. Setting \(m_{M}^{N},m_{M}^{\sigma }=0\), \(R_{\hat{H}}^{N}=R_{\hat{M}}^{N}\) and (24) binding, the remaining incentive constraints (22) and (24) become, respectively,

    $$\begin{aligned} R_{\hat{M}}^{N}-f_{L}&\le m_{L}^{N}\left( f_{H}-f_{L}\right) ,\quad R_{\hat{M} }^{\sigma }-f_{L}\le m_{L}^{\sigma }\left( f_{H}-f_{L}\right) \\ R_{\hat{M}}^{N}-f_{L}&= m_{L}^{N}\left( f_{M}-f_{L}\right) ,\quad R_{\hat{M} }^{\sigma }-f_{L}=m_{L}^{\sigma }\left( f_{M}-f_{L}\right) \end{aligned}$$

    whence we deduce that if the last set is binding, the first must be slack.

\(\square \)

The rest of the proof proceeds as follows. We have four variables to choose \( m_{L}^{G},m_{L}^{B},m_{L}^{N},\alpha .\) We first show that one set of optimal solution values for \(m_{L}^{G},m_{L}^{B}\) is \( m_{L}^{G}=1,m_{L}^{B}=0\) for any admissible values of \(0\le \alpha ,m_{L}^{N}\le 1.\) In fact if \(c_{a}>0\), these are the only optimal values when \(\alpha >0\) (Proposition 2). When \(\alpha =0\), they are admissible but so are any others (in fact, \(m_{L}^{G},m_{L}^{B}\) are then irrelevant). Using \(m_{L}^{G}=1,m_{L}^{B}=0\) then gives information about the optimal values of \(\alpha ,m_{L}^{N}\) (through the trade-off between \(\alpha \) and \(m_{L}^{N}\) in the objective and constraint). Proposition 3 shows that if \(NC_{a}>0\) and \(c_{a}>0\), then optimally \(\alpha =0\) and \( m_{L}^{N}\) assumes the value of the no-signal contract. Proposition 4 shows that if \(NC_{a}<0\) and \(c_{a}>0,\) then optimally \(\alpha >0\), and depending on the probability distribution of revenues after signal collection, either \(m_{L}^{N}=0\) or \(m_{L}^{N}=1,\) or \(\alpha =1\) and \(m_{L}^{N}\) is irrelevant.

Proof of Proposition 2

Suppose that \(0<\alpha <1\) and \(c_{a}>0.\)

  1. 1.

    An optimum cannot have all \(m_{L}^{\sigma }=m_{L}^{N}=1,\sigma =G,B.\) If it did, the participation constraint (35) would be slack \((\left( \pi _{H}+\pi _{M}\right) f_{M}+\pi _{L}f_{L}-I-\pi _{L}c_{m}-\alpha c_{a}>0\) by assumption 1) and a rent would be needlessly left to the principal. Similarly, it cannot have all \(m_{L}^{\sigma }=m_{L}^{N}=0,\sigma =G,B,\) since then the expected repayments would not meet the reservation level of the principal \((f_{L}-I-\alpha c_{a}<0)\). Thus, at least one audit probability must be positive.

  2. 2.

    Suppose that there are \(m_{L}^{\sigma },m_{L}^{N}>0,\sigma =G,B,\) and \(m_{L}^{G}<1\) such that the participation constraint (35) holds. This is not an optimum because it is possible to lower \(m_{L}^{B}\) (i.e., lower \(R_{\hat{M}}^{B}\)) and increase \(m_{L}^{G}\) (i.e., increase \(R_{ \hat{M}}^{G}\)) reducing the expected cost (34) and slackening the participation constraint (35). Indeed, because \(\pi _{HG}+\pi _{MG}>\pi _{HB}+\pi _{MB}\), the increase in \(m_{L}^{G}\, (R_{\hat{M}}^{G})\) and reduction in \(m_{L}^{B}\, (R_{\hat{M}}^{B})\) allow an increase in the expected revenues after a good signal larger than the fall in revenues after the bad signal. Similarly, this shift reduces the with-signal expected audit cost since \(\pi _{LG}<\pi _{LB}.\) This can be seen in the following expression:

    $$\begin{aligned}&\frac{1}{f_{H}-f_{L}}\left\{ \left[ \left( \pi _{HG}+\pi _{MG}\right) \left( f_{M}-f_{L}\right) -\pi _{LG}c_{m}\right] dR_{\hat{M}}^{G}\right. \\&\left. +\left( \left( \pi _{HB}+\pi _{MB}\right) \left( f_{M}-f_{L}\right) -\pi _{LB}c_{m}\right) dR_{\hat{M}}^{B}.\right\} \end{aligned}$$

    Since \(\left[ \left( \pi _{HG}\!+\!\pi _{MG}\right) \left( f_{H}\!-\!f_{L}\right) \!-\!\pi _{LG}c_{m}\right] \!-\!\left[ \left( \pi _{HB}\!+\!\pi _{MB}\right) \left( f_{H}-f_{L}\right) -\pi _{LB}c_{m}\right] \) \(>0,\) the participation constraint is slackened by an increase in \(m_{L}^{G}\) matched by an equal reduction in \( m_{L}^{B}.\) This allows a further reduction in \(m_{L}^{B}\) and thus a reduction in the objective function. So an optimum cannot have \(m_{L}^{G}<1\) and \(m_{L}^{B}>0\) and must at least entail one of \(m_{L}^{G}=1\) or \( m_{L}^{B}=0.\)

  3. 3.

    Suppose that \(m_{L}^{G}=1\) but \(1>m_{L}^{B}>0.\) Then, if \(m_{L}^{N}<1,\) it is possible to vary \(m_{L}^{N}\) and \(m_{L}^{B}\) keeping the participation constraint constant

    $$\begin{aligned} \mathrm{d}m_{L}^{N}=-\frac{\alpha \left[ \left( \pi _{HB}+\pi _{MB}\right) \left( f_{M}-f_{L}\right) -\pi _{LB}c_{m}\right] }{\left( 1-\alpha \right) \left[ \left( \pi _{H}+\pi _{M}\right) \left( f_{M}-f_{L}\right) -\pi _{L}c_{m} \right] }\mathrm{d}m_{L}^{B}<0 \end{aligned}$$
    (53)

    Thus, \(m_{L}^{B}\) and \(m_{L}^{N}\) must vary in opposite directions to satisfy the participation constraint. The effect of such variations in the objective function (34) is \(dEC=\alpha c_{m}\left[ -\pi _{LG}\mathrm{d}m_{L}^{N}+\pi _{LB}(\mathrm{d}m_{L}^{B}-\mathrm{d}m_{L}^{N})\right] +\pi _{L}c_{m}\mathrm{d}m_{L}^{N},\) which, using (53), becomes \(dEC=\alpha c_{m}\frac{\rho (f_{M}-f_{L})}{\left( \pi _{H}+\pi _{M}\right) \left( f_{M}-f_{L}\right) -\pi _{L}c_{m}}\mathrm{d}m_{L}^{B},\) where \(\rho \) is the correlation index (4). Since the coefficient of \(\mathrm{d}m_{L}^{B}\) is positive, it is best to reduce \(m_{L}^{B}\) and raise \(m_{L}^{N}\) as far as possible.

    1. (a)

      One possibility is to hit \(m_{L}^{B}=0\) with \(m_{L}^{N}>0.\) This proves the result.

    2. (b)

      Alternatively, \(m_{L}^{N}=1\) with \(m_{L}^{B}>0\) still. In this last case, the optimization problem becomes

      $$\begin{aligned}&\min \alpha \left[ c_{a}-\pi _{LB}c_{m}\left( 1-m_{L}^{B}\right) \right] +\pi _{L}c_{m} \\&\text {s.t. }-\alpha \left\{ \left( 1-m_{L}^{B}\right) \left[ \left( \pi _{HB}+\pi _{MB}\right) (f_{M}-f_{L})-\pi _{LB}c_{m}\right] +c_{a}\right\} \\&\quad +\left( \left( \pi _{H}+\pi _{M}\right) f_{M}+\pi _{L}f_{L}-\pi _{L}c_{m}\right) =I \end{aligned}$$

      It is then possible to vary \(\alpha \) and \(m_{L}^{B}\) keeping the participation constraint constant

      $$\begin{aligned} \mathrm{d}\alpha =\frac{\alpha \left\{ \left[ \left( \pi _{HB}+\pi _{MB}\right) (f_{M}-f_{L})-\pi _{LB}c_{m}\right] \right\} }{\left( 1-m_{L}^{B}\right) \left[ \left( \pi _{HB}+\pi _{MB}\right) (f_{M}-f_{L})-\pi _{LB}c_{m}\right] +c_{a}}\mathrm{d}m_{L}^{B}>0 \qquad \end{aligned}$$
      (54)

      The effect on the objective function is \(dEC=\left[ c_{a}-\pi _{LB}c_{m}\left( 1-m_{L}^{B}\right) \right] \mathrm{d}\alpha +\alpha \pi _{LB}c_{m}\mathrm{d}m_{L}^{B},\) whence using (54) gives

      $$\begin{aligned} dEC=\frac{\alpha c_{a}\left( \pi _{HB}+\pi _{MB}\right) \left( f_{M}-f_{L}\right) }{\left( 1-m_{L}^{B}\right) \left[ \left( \pi _{HB}+\pi _{MB}\right) (f_{M}-f_{L})-\pi _{LB}c_{m}\right] +c_{a}}\mathrm{d}m_{L}^{B}>0. \end{aligned}$$

      It is possible to reduce \(m_{L}^{B}\) to zero while keeping \(\alpha >0\) since, with \(m_{L}^{G}=m_{L}^{N}=1,\,m_{L}^{B}=0,\) the participation constraint becomes \(\alpha EPC_{\sigma }+\left( 1-\alpha \right) \left[ \left( \pi _{H}+\pi _{M}\right) f_{M}+\pi _{L}f_{L}-I-\pi _{L}c_{m}\right] \ge 0.\) Since \(\left( \pi _{H}+\pi _{M}\right) f_{M}+\pi _{L}f_{L}-I-\pi _{L}c_{m}>0,\) this can be satisfied with \(\alpha >0.\)

  4. 4.

    Alternatively, suppose the optimum has \(m_{L}^{B}=0\) but \(0<m_{L}^{G}<1\) and \(m_{L}^{N}>0.\) It is then possible to reduce \(m_{L}^{N},\) raise \( m_{L}^{G}\) so as to keep the participation constraint constant: \(\mathrm{d}m_{L}^{N}=- \frac{\alpha \left\{ \left( \pi _{HG}+\pi _{HB}\right) (f_{M}-f_{L})-\pi _{LG}c_{m}\right\} }{(1-\alpha )\left\{ \left( \pi _{H}+\pi _{M}\right) (f_{M}-f_{L})-\pi _{L}c_{m}\right\} }\mathrm{d}m_{L}^{G}.\) The change in objective ( 34) is \(dEC=\alpha c_{m}\frac{-\rho (f_{H}-f_{L})}{\left( \pi _{H}+\pi _{M}\right) (f_{M}-f_{L})-\pi _{L}c_{m}}\mathrm{d}m_{L}^{G}<0.\) So it is best to raise \(m_{L}^{G}\) to maximum and reduce \(m_{L}^{N}.\)

    1. (a)

      One possibility is that we hit \(m_{L}^{G}=1\) with \(m_{L}^{N}>0,\) which proves the result.

    2. (b)

      The other possible case has \(m_{L}^{N}=0\) with \(m_{L}^{G}<1\) still. In this last case, the optimization problem becomes

      $$\begin{aligned}&\displaystyle \min \alpha \left( c_{a}+\pi _{LG}m_{L}^{G}c_{m}\right) \\&\displaystyle \text {s.t. }\alpha \left\{ \left[ \left( \pi _{HG}+\pi _{HB}\right) \left( f_{M}-f_{L}\right) -\pi _{LG}c_{m}\right] m_{L}^{G}-c_{a}\right\} =I-f_{L}. \end{aligned}$$

      Varying \(m_{L}^{G}\) and \(\alpha \) so as to keep the participation constraint constant gives \(\mathrm{d}m_{L}^{G}=-\frac{\left( \left( \pi _{HG}+\pi _{HB}\right) \left( f_{M}-f_{L}\right) -\pi _{LG}c_{m}\right) m_{L}^{G}-c_{a}}{\alpha \left[ \left( \pi _{HG}+\pi _{HB}\right) \left( f_{M}-f_{L}\right) -\pi _{LG}c_{m}\right] }\mathrm{d}\alpha .\) Using the resulting \(\mathrm{d}m_{L}^{G}\) in the objective function \(dEC=\left( c_{a}+c_{m}\pi _{LG}m_{L}^{G}\right) \mathrm{d}\alpha +\alpha c_{m}\pi _{LG}\mathrm{d}m_{L}^{G}\) gives \(dEC=\frac{\left( \pi _{HG}+\pi _{HB}\right) \left( f_{M}-f_{L}\right) c_{a}}{\left( \pi _{HG}+\pi _{HB}\right) \left( f_{M}-f_{L}\right) -\pi _{LG}c_{m}}\mathrm{d}\alpha >0.\) So it is best to reduce \(\alpha \) and raise \(m_{L}^{G}\) as far as possible. In fact, we can attain \(m_{L}^{G}=1\) with \(\alpha >0.\)

Hence, the optimum must involve \(m_{L}^{G}=1,m_{L}^{B}=0\).\(\square \)

Proposition 2 is valid if \(\alpha >0.\) If \(\alpha =0,m_{L}^{\sigma }\) have no impact on participation constraint or objective and so without loss of generality, we can also set them to these values if \( \alpha =0.\)

To prove Propositions (3) and (4), we use the following lemma.

Lemma 1

When \(0\le \alpha <1\) and \(0<m_{L}^{N}<1\) are optimal, if \(EPC_{\sigma }\ne 0,\) a simultaneous variation in \(\alpha \) and \(m_{L}^{N}\) keeping the participation constraint constant changes expected enforcement cost according to

$$\begin{aligned} dEC\!=\!\frac{\left\{ \left( \pi _{H}\!+\!\pi _{M}\right) c_{a}\!-\!\left[ \pi _{LB}\left( \pi _{HG}\!+\!\pi _{MG}\right) \!-\pi _{LG}\left( \pi _{HB}\!+\!\pi _{MB}\right) \right] c_{m}\right\} \left( f_{M}\!-\!f_{L}\right) }{\left( \pi _{H}\!+\!\pi _{M}\right) \left( f_{M}\!-\!f_{L}\right) \!-\!\pi _{L}c_{m}}\mathrm{d}\alpha , \end{aligned}$$

whose sign depends on the sign of \(\left( \pi _{H}+\pi _{M}\right) c_{a}-\rho c_{m}\equiv NC_{a}.\)

Proof

Consider Program \(\mathcal {P}_{EC-\mathrm{after}}^{\prime }.\) The participation constraint (35) is

$$\begin{aligned}&\alpha \left\{ \left( \pi _{HG}+\pi _{MG}\right) f_{M}+\left( 1-\pi _{HG}-\pi _{MG}\right) f_{L}-I-\pi _{LG}c_{m}-c_{a}\right\} \\&\quad +\left( 1-\alpha \right) \left\{ \left( \pi _{H}+\pi _{M}\right) R_{\hat{M} }^{N}+\pi _{L}\left( f_{L}-m_{L}^{N}c_{m}\right) -I\right\} =0 \end{aligned}$$

which can also be written as:

$$\begin{aligned}&\alpha \left\{ \left( \pi _{HG}+\pi _{MG}\right) f_{M}+\left( 1-\pi _{HG}-\pi _{MG}\right) f_{L}-I-\pi _{LG}c_{m}-c_{a}\right\} +\left( 1-\alpha \right) \\&+\left\{ \left( \pi _{H}+\pi _{M}\right) \left( f_{M}-f_{L}\right) m_{L}^{N}+\left( \pi _{H}+\pi _{M}\right) f_{L}+\pi _{L}\left( f_{L}-m_{L}^{N}c_{m}\right) -I\right\} =0 \end{aligned}$$

Suppose \(0\le \alpha <1\) and \(0<m_{L}^{N}<1\) are optimal. When \(EPC_{\sigma }\ne 0\) simultaneous variation in \(\alpha ,m_{L}^{N}\) to keep the participation constraint constant must satisfy

$$\begin{aligned} \left[ EPC_{\sigma }-\left( PC_{N}-I\right) \right] \mathrm{d}\alpha +\left( 1-\alpha \right) \left[ \left( \pi _{H}+\pi _{M}\right) \left( f_{M}-f_{L}\right) -\pi _{L}c_{m}\right] \mathrm{d}m_{L}^{N}=0 \end{aligned}$$
$$\begin{aligned} \frac{\mathrm{d}m_{L}^{N}}{\mathrm{d}\alpha }=-\frac{EPC_{\sigma }-\left( PC_{N}-I\right) }{ \left( 1-\alpha \right) \left[ \left( \pi _{H}+\pi _{M}\right) \left( f_{M}-f_{L}\right) -\pi _{L}c_{m}\right] } \end{aligned}$$
(55)

While the sign of the denominator is positive by Assumption 5, the sign of the numerator is ambiguous. However, it can be inferred from the sign of \(EPC_{\sigma }.\) For \(\alpha \in \left( 0,1\right) ,\) to satisfy the participation constraint with equality, \(PC_{N}-I\lesseqgtr 0\) if \( EPC_{\sigma }\gtreqless 0.\) Thus, if

  • \(EPC_{\sigma }>0,\) \(\frac{\mathrm{d}m_{L}^{N}}{\mathrm{d}\alpha }|_{PC}<0;\)

  • \(EPC_{\sigma }<0,\) \(\frac{\mathrm{d}m_{L}^{N}}{\mathrm{d}\alpha }|_{PC}>0.\)

The simultaneous variation in \(m_{L}^{N},\alpha \) changes the expected enforcement cost by \(\left( c_{a}+\pi _{LG}c_{m}-\pi _{L}m_{L}^{N}c_{m}\right) \mathrm{d}\alpha +\left( 1-\alpha \right) \pi _{L}c_{m}\mathrm{d}m_{L}^{N},\) i.e., using 55,

$$\begin{aligned} dEC=\frac{\left\{ \left( \pi _{H}\!+\!\pi _{M}\right) c_{a}\!-\!\left[ \pi _{LB}\left( \pi _{HG}\!+\!\pi _{MG}\right) \!-\!\pi _{LG}\left( \pi _{HB}\!+\!\pi _{MB}\right) \right] c_{m}\right\} \left( f_{M}\!-\!f_{L}\right) }{\left( \pi _{H}\!+\!\pi _{M}\right) \left( f_{M}\!-\!f_{L}\right) \!-\!\pi _{L}c_{m}}\mathrm{d}\alpha .\nonumber \\ \end{aligned}$$
(56)
  • If \(EPC_{\sigma }=0,\,PC=(1-\alpha )\left\{ m_{L}^{N}\left[ \left( \pi _{H}+\pi _{M}\right) \left( f_{M}-f_{L}\right) -\pi _{L}c_{m}\right] -I+f_{L}\right\} \) and given \(\alpha <1\) we must have \(m_{L}^{N}\left[ \left( \pi _{H}+\pi _{M}\right) \left( f_{M}-f_{L}\right) -\pi _{L}c_{m} \right] -I+f_{L}=0.\) In this case, no marginal compensation between \( m_{L}^{N},\alpha \) is possible.

\(\square \)

Proof of Proposition 3

If \(NC_{a}>0,\) from (56), enforcement costs are minimized by reducing \(\alpha .\) However, the change in \(m_{L}^{N}\) following a change in \(\alpha \) necessary to preserve the participation constraint depends on the sign of \(EPC_{\sigma }.\) Three cases can arise:

  1. 1.

    If \(EPC_{\sigma }>0,\) \(\frac{\mathrm{d}m_{L}^{N}}{\mathrm{d}\alpha }|_{PC}<0.\) Thus, \( m_{L}^{N}\) and \(\alpha \) vary in opposite directions to satisfy the participation constraint. So, to minimize the enforcement cost, it is best to raise \(m_{L}^{N}\) and decrease \(\alpha \). However, an optimum cannot have \(m_{L}^{N}=1\) and \(\alpha >0\), because the participation constraint would be slack (\(\alpha EPC_{\sigma }+(1-\alpha )\left( PC_{N}-I\right) =0.\) If \( m_{L}^{N}=1,\,PC_{N}-I=\left( \pi _{H}+\pi _{M}\right) f_{M}+\pi _{L}f_{L}-I-\pi _{L}c_{m},\) which is positive by Assumption 1. Given \(EPC_{\sigma }>0,\,PC>0\)). The firm can then reduce \(m_{L}^{N}\) until the participation constraint binds, giving \(PC_{N}-I<0\). But because \(\frac{ dEC}{\mathrm{d}m_{L}^{N}}=-\frac{\left( 1-\alpha \right) \left( f_{M}-f_{L}\right) \left( \left( \pi _{H}+\pi _{M}\right) c_{a}-\rho c_{m}\right) }{EPC_{\sigma }-\left( PC_{N}-I\right) }<0,\) it is best to raise \(m_{L}^{N}\) and lower \( \alpha .\) However, \(m_{L}^{N}\) only has to rise to balance \(PC\) and we know that the no-signal solution with \(\alpha =0\) is achievable with \( m_{L}^{N}<1.\) If \(\alpha \) were positive but small, there would be further gains to be realized. Thus, optimally \(\alpha =0,\) with \(R_{\hat{M}}^{N}=R_{ \hat{M}}^{N}|_{\alpha =0}\) (40).

  2. 2.

    If \(EPC_{\sigma }<0,\,\frac{\mathrm{d}m_{L}^{N}}{\mathrm{d}\alpha }|_{PC}>0.\) Now, \( \alpha ,m_{L}^{N}\) must both fall to balance \(PC\). Again, we know that we can achieve \(\alpha =0\) with \(PC=0\) at the no-signal solution. And if we left \(\alpha >0\), there would still be gains to be realized from reducing both \(\alpha ,m_{L}^{N}\). Thus, the solution must have \(m_{L}^{N}>0,\,R_{ \hat{M}}^{N}=R_{\hat{M}}^{N}|_{\alpha =0}\) (40), and \(\alpha =0.\)

  3. 3.

    If \(EPC_{\sigma }=0,\,PC=(1-\alpha )(PC_{N}-I).\) One possibility is \(\alpha =1,\) in which case \(m_{L}^{N}\) is immaterial, the other is \( PC_{N}=I,\) in which case, since \(c_{a}>0,\) it is best to set \(\alpha =0\) and \(R_{\hat{M}}^{N}=R_{\hat{M}}^{N}|_{\alpha =0}\) (40). Of these two possibilities, if \(NC_{a}>0\), it is best to set \(\alpha =0.\)

In either case, enforcement costs are \(EC_{\alpha =0}=\pi _{L}m_{L}^{N}c_{m}\) (41).\(\square \)

Proof of Proposition 4

If \(NC_{a}<0,\) from (56) \(\frac{dEC}{\mathrm{d}\alpha }<0\), enforcement costs are minimized by raising \(\alpha .\) However, the change in \(m_{L}^{N}\) following a change in \(\alpha \) necessary to preserve the participation constraint depends on the sign of \(EPC_{\sigma }.\) Three cases can arise:

  1. 1.

    If \(EPC_{\sigma }>0,\) from (55) \(\frac{\mathrm{d}m_{L}^{N}}{\mathrm{d}\alpha } |_{PC}<0.\) Thus, \(m_{L}^{N}\) and \(\alpha \) vary in opposite directions to satisfy the participation constraint. To minimize the enforcement cost, it is best to reduce \(m_{L}^{N}\) as far as possible resulting in \(m_{L}^{N}=0\) ( \(R_{\hat{M}}^{N}=f_{L}\)) and raise \(\alpha \) just far enough to satisfy the participation constraint (\(\alpha =\underline{\alpha }\)). Setting \(\alpha =1\) (and \(m_{L}^{N}\) interior) is not optimal as it would give a slack participation constraint. The enforcement costs are \(\underline{\alpha } \left( c_{a}+\pi _{LG}c_{m}\right) \).

  2. 2.

    If \(EPC_{\sigma }=0,\) since \(NC_{a}<0\), now it is best to set \(\alpha =1\) and then \(m_{L}^{N}\) is immaterial. Enforcement costs are \(c_{a}+\pi _{LG}c_{m}\).

  3. 3.

    If \(EPC_{\sigma }<0,\,\frac{\mathrm{d}m_{L}^{N}}{\mathrm{d}\alpha }|_{PC}>0.\) Thus, \( m_{L}^{N}\) and \(\alpha \) vary in the same direction to satisfy the participation constraint and \(\alpha \) has to be increased to minimize the enforcement cost. Setting \(\alpha =1\) is not optimal as it would violate participation constraint. Thus, \(m_{L}^{N}\) must be raised as far as possible, giving \(m_{L}^{N}=1\) and \(\alpha \) must be raised sufficiently to satisfy the participation constraint \((\alpha =\bar{\alpha })\). The enforcement costs are \(\bar{\alpha }\left( c_{a}+\pi _{LG}c_{m}\right) +\left( 1-\bar{\alpha }\right) \pi _{L}c_{m}\).

To see why policing involves concentrated audit with \(m_{L}^{G}=1\) and \(\alpha \) interior, consider the enforcement costs when \(m_{L}^{G}\) is interior: \(EC=\alpha \left( c_{a}+\pi _{LG}m_{L}^{G}c_{m}\right) .\) Proportional and opposite changes in \(m_{L}^{G}\) or \(\alpha \) have identical expected audit cost effects, but in addition the expected signal acquisition cost is \(\alpha c_{a}\). Thus, the extra cost of a proportional increase in the audit probability \(m_{L}^{G}\) is more than offset by the saving induced by a reduction in signal acquisition frequency \(\alpha \). It is then optimal to increase the audit probability all the way until \( m_{L}^{G}=1\) and reduce the chance of signal acquisition until the participation constraint binds. Thus, it is the direct cost of collecting the signal that ultimately makes the signal acquisition strategy more expensive than auditing and leads to deterministic audits.

If the project is not ex-ante profitable after a signal \((EPC_{\sigma }<0)\), then high state repayments in no-signal states must be increased to satisfy the participation constraint, thus necessitating audit of no-signal states too. Signal contingent deterministic monitoring is still optimal \(( m_{L}^{G}=1,m_{L}^{B}=0),\) but now signal acquisition has effects on the relative expected audit costs of good signal and no-signal low reports. The optimal combination \(\alpha ,m_{L}^{N}\) minimizes \(EC=\alpha \left( c_{a}+\pi _{LG}c_{m}\right) +\left( 1-\alpha \right) \pi _{L}m_{L}^{N}c_{m}.\) In trading off \(\alpha \) and \(m_{L}^{N},\) each has a specific cost \( (c_{a}+\pi _{LG}c_{m})\) for \(\alpha \) and \(\pi _{L}m_{L}^{N}c_{m}\) for \( m_{L}^{N}\), as well as a common saving of \(\alpha \pi _{L}m_{L}^{N}c_{m},\) since, when there is a signal, \(m_{L}^{N}\) is not applied. From the fact that \(NC_{a}<0\) and the signal is relatively cheap, it turns out that the cost saving from reducing \(\alpha \) exceeds that from reducing \(m_{L}^{N}\) and it is optimal to increase the audit probability all the way until \( m_{L}^{N}=1\) and use \(\alpha \) just to satisfy the participation constraint.\(\square \)

Proof of Proposition 5

Let \(c_{m}^{*},c_{a}^{*}\) (44) solve \(EP=EPC_{\sigma }\). Similarly, let \(\widehat{c}_{m},\widehat{c}_{a}\) (45) solve \( EPC_{\sigma }=NC_{a}.\) Feasibility requires \(c_{a},c_{m}\) are below the \(EP\) line The difference \(c_{a}^{*}-\widehat{c}_{a}\) (which has the opposite sign of \(c_{m}^{*}-\widehat{c}_{m}\)) is equal to \(\rho \left( \pi _{HG}+\pi _{MG}\right) \left( f_{M}-f_{L}\right) -\left( \pi _{H}+\pi _{M}\right) \pi _{LB}(I-f_{L})\). It is increasing in \(\rho \) and zero when \( \rho =\rho ^{*}\equiv \frac{\pi _{LB}\left( \pi _{H}+\pi _{M}\right) \left( I-f_{L}\right) }{\left( \pi _{HG}+\pi _{MG}\right) \left( f_{M}-f_{L}\right) }.\) Hence, if \(\rho \le \rho ^{*}\), we have \( c_{a}^{*}\le \widehat{c}_{a}\) and all three parts of Proposition 4 apply. Conversely, if \(\rho >\rho ^{*},c_{a}^{*}> \widehat{c}_{a}\) and \(c_{m}^{*}<\widehat{c}_{m}.\) In this case, the only feasible points \(c_{a},c_{m}\) below the \(EP\) line and with \(NC_{a}\le 0\) must entail \(EPC_{\sigma }>0\) and hence, whenever \(NC_{a}\le 0,\) the signal strategy can only be that described by \(MS_{a},\) in Proposition 4. \(\square \)

Comparative statics: Signal strategy as a function of acquisition and audit cost. We first plot \(\alpha \) against \(c_{a}\) and \(c_{m}.\) When \(NC_{a}<0\) and \( EPC_{\sigma }\ge 0,\, \tfrac{\partial \underline{\alpha }}{\partial c_{m}}= \tfrac{\pi _{LG}\left( I-f_{L}\right) }{\left( \left( 1-\pi _{LG}\right) \left( f_{M}-f_{L}\right) -\pi _{LG}c_{m}-c_{a}\right) ^{2}}>0,\,\tfrac{ \partial ^{2}\underline{\alpha }}{\partial c_{m}^{2}}=\tfrac{2\pi _{LG}^{2}\alpha }{\left( \left( 1-\pi _{LG}\right) \left( f_{M}-f_{L}\right) -\pi _{LG}c_{m}-c_{a}\right) ^{2}}>0.\)

When \(EPC_{\sigma }<0,\,\tfrac{\partial \bar{\alpha }}{\partial c_{m}}=- \tfrac{\pi _{L}-\alpha \pi _{LB}}{\left( \left( 1-\pi _{LB}\right) \left( f_{M}-f_{L}\right) -\pi _{LB}c_{m}+c_{a}\right) ^{2}}<0,\,\tfrac{\partial ^{2}\bar{\alpha }}{\partial c_{m}^{2}}=-\tfrac{2\pi _{LB}\left( \pi _{L}-\alpha \pi _{LB}\right) }{\left( \left( 1-\pi _{LB}\right) \left( f_{M}-f_{L}\right) -\pi _{LB}c_{m}+c_{a}\right) ^{2}}<0.\) At its maximum, the slope of \(\alpha \left( c_{m}\right) \) changes from \(\frac{\partial \alpha }{\partial c_{m}}=\frac{\pi _{LG}}{I-f_{L}}\) to \(\frac{\partial \alpha }{\partial c_{m}}=\frac{\pi _{LG}^{2}}{\rho \left( f_{M}-f_{L}\right) -\pi _{LB}\left( I-f_{L}\right) -\pi _{L}c_{a}}<0\).Footnote 21

When \(EPC_{\sigma }\ge 0,\,\tfrac{\partial \underline{\alpha }}{\partial c_{a}}=\tfrac{I-f_{L}}{\left( \left( \pi _{HG}+\pi _{MG}\right) \left( f_{M}-f_{L}\right) -\pi _{LG}c_{m}-c_{a}\right) ^{2}}>0,\) \(\tfrac{\partial ^{2}\underline{\alpha }}{\partial c_{a}^{2}}=\tfrac{2\left( I-f_{L}\right) }{ \left( \left( \pi _{HG}+\pi _{MG}\right) \left( f_{M}-f_{L}\right) -\pi _{LG}c_{m}-c_{a}\right) ^{3}}>0.\)

When \(EPC_{\sigma }<0,\, \tfrac{\partial \bar{\alpha }}{\partial c_{a}}=- \tfrac{Ef-I-\pi _{L}c_{m}}{\left( \left( \pi _{HB}+\pi _{MB}\right) \left( f_{M}-f_{L}\right) -\pi _{LB}c_{m}+c_{a}\right) ^{2}}<0,\,\tfrac{\partial ^{2}\bar{\alpha }}{\partial c_{a}^{2}}=\tfrac{2\left( Ef-I-\pi _{L}c_{m}\right) }{\left( \left( \pi _{HB}+\pi _{MB}\right) \left( f_{M}-f_{L}\right) -\pi _{LB}c_{m}+c_{a}\right) ^{3}}>0.\)

Since utility is given by \(U=Ef-I-EC=\pi _{H}f_{H}+\pi _{M}f_{M}+\pi _{L}f_{L}-I-EC,\) where \(EC\) is the cost function (34), we can easily plot \(U\) against \(c_{a}\) and \(c_{m}.\) For \(c_{a}>0\) and increasing \(c_{m},\) optimally \(\alpha =0\) and utility is given by \( U_{\alpha =0}=\frac{\left( \pi _{H}+\pi _{M}\right) \left( f_{M}-f_{L}\right) \left( Ef-I-\pi _{L}c_{m}\right) }{\left( \pi _{H}+\pi _{M}\right) \left( f_{M}-f_{L}\right) -\pi _{L}c_{m}}\). Then, \(\tfrac{ \partial U_{\alpha =0}}{\partial c_{m}}=-\tfrac{\left( \pi _{H}+\pi _{M}\right) \pi _{L}\left( f_{M}-f_{L}\right) \left( I-f_{L}\right) }{\left( \left( \pi _{H}+\pi _{M}\right) \left( f_{M}-f_{L}\right) -\pi _{L}c_{m}\right) ^{2}}<0;\) \(\tfrac{\partial U_{\alpha =0}^{2}}{\partial c_{m}^{2}}=-\tfrac{2\left( \pi _{H}+\pi _{M}\right) \pi _{L}^{2}\left( f_{M}-f_{L}\right) \left( I-f_{L}\right) }{\left( \left( \pi _{H}+\pi _{M}\right) \left( f_{M}-f_{L}\right) -\pi _{L}c_{m}\right) ^{3}}<0.\)

For higher \(c_{m},\) optimally \(\alpha >0\) and utility is given by \( U_{MS_{a}}\,=\frac{\left( f_{M}-f_{L}\right) \left[ \left( \pi _{HB}+\pi _{MB}\right) \left( I-f_{L}\right) +\left( \pi _{H}+\pi _{M}\right) EPC_{\sigma }\right] }{\left( \pi _{HG}+\pi _{MG}\right) \left( f_{M}-f_{L}\right) -\pi _{LG}c_{m}-c_{a}}\) and \(\tfrac{\partial U_{MS_{a}}}{ \partial c_{m}}=-\tfrac{\pi _{LG}\left( \pi _{HG}+\pi _{MG}\right) \left( f_{M}-f_{L}\right) \left( I-f_{L}\right) }{\left( \left( \pi _{HG}+\pi _{MG}\right) \left( f_{M}-f_{L}\right) -\pi _{LG}c_{m}-c_{a}\right) ^{2}}<0;\) \(\tfrac{\partial U_{MS_{a}}^{2}}{\partial c_{m}^{2}}=-\tfrac{2\pi _{LG}^{2}\left( \pi _{HG}+\pi _{MG}\right) \left( f_{M}-f_{L}\right) \left( I-f_{L}\right) }{\left( \left( \pi _{HG}+\pi _{MG}\right) \left( f_{M}-f_{L}\right) -\pi _{LG}c_{m}-c_{a}\right) ^{2}}<0.\)

Last, for sufficiently high \(c_{m},\) optimally still \(\alpha >0,\) but utility is given by \(U_{MS_{b}}=\frac{\left( Ef-I-\pi _{L}c_{m}\right) \left( \pi _{HB}+\pi _{MB}\right) \left( f_{M}-f_{L}\right) }{\left( \pi _{HB}+\pi _{MB}\right) \left( f_{M}-f_{L}\right) -\pi _{LB}c_{m}+c_{a}}\,\) and \(\tfrac{\partial U_{MS_{b}}}{\partial c_{m}}=\tfrac{\left( \pi _{HB}+\pi _{MB}\right) \left( f_{M}-f_{L}\right) \left\{ \pi _{LB}EPC_{\sigma }-\pi _{LG}\left[ \left( \pi _{HB}+\pi _{MB}\right) \left( f_{M}-f_{L}\right) -\pi _{LB}c_{m}+c_{a}\right] \right\} }{\left( \left( \pi _{HB}+\pi _{MB}\right) \left( f_{M}-f_{L}\right) -\pi _{LB}c_{m}+c_{a}\right) ^{2}}<0;\,\tfrac{ \partial U_{MS_{b}}^{^{2}}}{\partial c_{m}^{2}}=-\tfrac{2\pi _{LB}\left( \pi _{HB}+\pi _{MB}\right) \left( f_{M}-f_{L}\right) \left\{ \pi _{LB}EPC_{\sigma }-\pi _{LG}\left[ \left( \pi _{HB}+\pi _{MB}\right) \left( f_{M}-f_{L}\right) -\pi _{LB}c_{m}+c_{a}\right] \right\} }{\left( \left( \pi _{HB}+\pi _{MB}\right) \left( f_{M}-f_{L}\right) -\pi _{LB}c_{m}+c_{a}\right) ^{3}}<0.\)

The pattern of utility for varying \(c_{a}\) and \(c_{m}\) is depicted in Figs. 8 and 9.

Fig. 8
figure 8

\(U\left( c_{m}\right) \) for \(c_{a}>0\)

Fig. 9
figure 9

\(U\left( c_{a}\right) \) for \(c_{m}>0\)

1.2 Signal before the contract

Proof of Proposition 6

When \(U^{B}>0,\) using program \(\mathcal {P}_{b4}^{^{\prime }}\), solving the participation constraint for \(R_{\hat{M}}^{\sigma }\) (48) and substituting out both in \(m_{L}^{\sigma }\) and in the utility function \( U^{\sigma }\) gives \(m_{L|b4}^{\sigma }\) and \(EU_{b4}\) reported in Proposition 6 (47 and 49), respectively). Notice that \(m_{L|b4}^{G}-m_{L|b4}^{B}=\tfrac{-\left( I-f_{L}\right) \rho \left( f_{H}-f_{L}+c_{m}\right) }{\left( \pi _{HG}\left( f_{H}-f_{L}\right) -\pi _{LG}c_{m}\right) \left( \pi _{HB}\left( f_{H}-f_{L}\right) -\pi _{LB}c_{m}\right) }<0.\) When \(U^{B}\le 0,\) the principal knows he will make zero profit at best from proceeding after a bad signal and generally will abandon the project. The outcomes after a good signal are not affected, so the expected return to the agent arises only from the good signal outcome ( 50).   \(\square \)

1.3 Globally optimal signal strategy

Proof of Proposition 7

Let \(R_{b4}\) be the optimal repayments in the left-hand (LH) branch (signal before the contract), \(\alpha ,R_\mathrm{after}\) the optimal signal strategy and repayments in the right-hand (RH) branch (signal after the contract).

\(R_{b4}\) satisfies the truthtelling constraint (TT), the participation constraint (PC) in each signal state \(E_{s|\sigma }PC^{\sigma }=0,\sigma =G,B,\) and the relevant limited liability (LL) constraints (18).

The maximum payoff in the LH branch is

$$\begin{aligned} EU_{b4}(R_{b4})=\sum _{\sigma }\pi _{\sigma }E_{s|\sigma }U^{\sigma }(R_{b4}^{\sigma }). \end{aligned}$$

The RH contract has the same TT constraints and LL constraints, and the principal participation constraint is \(\alpha \sum _{\sigma }E_{s|\sigma }PC^{\sigma }+(1-\alpha )E_{s|N}PC^{N}.\) The maximum payoff on RH branch is

$$\begin{aligned} EU_\mathrm{after}(\alpha ,R_\mathrm{after})=\alpha \sum _{\sigma }\pi _{\sigma }E_{s|\sigma }U^{\sigma }\left( R_\mathrm{after}\right) +(1-\alpha )EU^{N}(R_\mathrm{after}). \end{aligned}$$

In more detail,

$$\begin{aligned} EU_{b4}(R_{b4})&= \pi _{G}E_{s|G}U^{G}+\pi _{B}E_{s|B}U^{B} \\&= \pi _{G}\max _{R^{G}}\left[ E_{s|G}U^{G}|E_{s|G}PC^{G},TT^{G},LL^{G}\right] \\&\quad +\pi _{B}\max _{R^{B}}\left[ E_{s|B}U^{B}|E_{s|B}PC^{B},TT^{B},LL^{B}\right] \\&= \pi _{G}E_{s|G}U^{G}(R_{b4}^{G})+\pi _{B}E_{s|B}U^{B}(R_{b4}^{B}) \\&\le \max _{R^{G},R^{B}}\Bigg [\left( \pi _{G}E_{s|G}U^{G}+\pi _{B}E_{s|B}U^{B}\right) |\left( \pi _{G}E_{s|G}PC^{G}\right. \nonumber \\&\quad \left. +\pi _{B}E_{s|B}PC^{B}\right) ,TT^{G},LL^{G},TT^{B},LL^{B}\Bigg ] \\&\equiv EU_\mathrm{after}(1,R^{G}\left( 1\right) ,R^{B}\left( 1\right) ) \\&\le \underset{\alpha }{\max }\left[ EU_\mathrm{after}(\alpha ,R^{G}\left( \alpha \right) ,R^{B}\left( \alpha \right) )|0\le \alpha \le 1\right] \end{aligned}$$

where \(R^{\sigma }\left( \alpha \right) \) is the optimal RH branch repayment at the fixed value of \(\alpha \). In fact, the first inequality is strict since for \(\alpha =1\), there is cross-subsidization in the participation constraint with the principal making a loss in the bad signal state and a gain in the good signal state. \(\alpha =1\) is optimal if \(EPC_{\sigma }=0.\) In other cases, \((EPC_{\sigma }\lessgtr 0)\) optimally \(\alpha <1\) in which case the second inequality is then also strict. Incidentally, in these cases, there is also cross-subsidization in the different parts of the ex-ante participation constraint but now between with-signal and no-signal states. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Menichini, A.M.C., Simmons, P.J. Sorting the good guys from bad: on the optimal audit structure with ex-ante information acquisition. Econ Theory 57, 339–376 (2014). https://doi.org/10.1007/s00199-014-0818-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00199-014-0818-y

Keywords

JEL Classification

Navigation