INFORMATION, RISK SHARING, AND INCENTIVES IN AGENCY PROBLEMS

Published date01 February 2017
AuthorJia Xie
DOIhttp://doi.org/10.1111/iere.12212
Date01 February 2017
INTERNATIONAL ECONOMIC REVIEW
Vol. 58, No. 1, February 2017
INFORMATION, RISK SHARING, AND INCENTIVES IN AGENCY PROBLEMS
BYJIA XIE1
Bank of Canada, Canada
This article studies the use of information for incentives and risk sharing in agency problems. When the principal is
risk neutral or the outcome is contractible, risk sharing is unnecessary or dealt with by a contract on the outcome, so
information systems are used for incentives only. When the outcome is noncontractible, a risk-averse principal relies on
imperfect information for both incentives and risk sharing. Under the first-order approach, this article relaxes Gjesdal’s
criterion for ranking information systems and finds conditions justifying the first-order approach when the principal is
risk averse and the outcome is noncontractible.
1. INTRODUCTION
The principal–agent problem consists of two stages: the first stage is the principal’s choice of
an information system, which generates contractible (i.e., commonly observable and verifiable)
signals or performance measures. The second stage is the optimal design of a contract based on
the information system chosen in the first stage. Much of the literature focuses on the second
stage, assuming that the information system is given. This article focuses on the first stage and
studies the choice of information systems by the principal.
Comparing information systems in agency problems was first raised by Holmstr ¨
om (1979)
and further studied by Kim (1995); Jewitt (1997, 2007); Dewatripont et al. (1999); Demougin
and Fluet (2001); Fagart and Sinclair-Desgagne (2007); and Xie (2011). These studies assume
either a risk-neutral principal or a contractible outcome. In these cases, information systems
are used to provide incentives to the agent’s action, and two main predictions are derived: (i)
An information system is valuable if it is informative about the agent’s action,2and (ii) the
statistical condition for being “informative” can be relaxed under the first-order approach, by
which the principal can predict the agent’s action using the agent’s first-order conditions alone.
In many cases, however, the principal is strictly risk averse and the outcome is noncontractible.
Gjesdal (1982) gives three main reasons for the outcome being noncontractible.3First, the
outcome may be unobservable at the time when the agent is paid. For instance, the manager of
a firm may be paid irreversibly before the outcome of his action is observed. Second, contracting
on outcome may be too costly. For instance, perfect auditing of income tax returns is expensive.
Finally, outcomes are often imperfect estimates of the “real” outcomes. For instance, the quality
measure of a project is an imperfect estimate of the “real” quality.
When the principal is strictly risk averse and the outcome is noncontractible, information
systems are used not only to provide incentives but also to allocate risk in the outcome of the
Manuscript received February 2013; revised April 2015.
1The analysis and conclusions set forth are those of the author and do not indicate concurrence by the Bank of
Canada. I am especially indebted to Frank Page and Michael Rauh for advice and encouragement, and I am grateful
to the editor and two referees for insightful comments that greatly improve the article. For comments and discussions,
I thank Eric Rasmusen, Michael Baye, Lars Stole, James Walker, Jason Allen, Hanna Halaburda, and Brian Peterson
as well as seminar and conference participants at Indiana University, Washington University in St. Louis, University
of Texas at Dallas, and the Midwest Economics Association. Needless to say, any mistakes are mine. Please address
correspondence to: Jia Xie, The Department of Real Estate Management, Ted Rogers School of Management, Ryerson
University, 55 Dundas St. W, Toronto, ON M5G 2C3 Canada. E-mail: jia.xie@ryerson.ca.
2See also Shavell (1979), Laffont and Martimort (2002), Tirole (2006), and Bolton and Dewatripont (2005).
3See also Mirrlees (1976), Baker (2002), and Maskin (2002).
157
C
(2017) by the Economics Department of the University of Pennsylvania and the Osaka University Institute of Social
and Economic Research Association
158 XIE
agent’s action. Following this line of reasoning, Gjesdal (1982) develops a criterion—composed
of an incentive condition and a risk-sharing condition—that compares information systems by
how informative they are about the agent’s action and the outcome, separately. Gjesdal (1982),
however, does not assume the first-order approach. Therefore, it is natural to wonder if his
criterion can be weakened under the first-order approach and if the first-order approach can be
justified for the case with a strictly risk-averse principal and a noncontractible outcome.
The main purpose of this article is to provide the first answers in the literature to these
questions. First, we find that the incentive condition in Gjesdal (1982) can be relaxed under
the first-order approach, because only local incentive compatibility constraints matter, and
therefore one only needs to focus on a neighborhood of the action in question. The risk-sharing
condition, on the other hand, is the same because one has to consider all possible values of the
outcome, whether the first-order approach applies or not.
Second, the first-order approach can be justified by conditions ensuring that the agent’s utility
is concave in his action.4Finding these conditions for the case with a strictly risk-averse prin-
cipal and a noncontractible outcome, however, presents technical difficulties: The interaction
between the two roles of information systems may cause the agent’s utility to be nonconcave
in his action. Nevertheless, we show that, by imposing restrictions on the information structure
and both parties’ risk aversions, we can justify the first-order approach.
The rest of the article is organized as follows: I set up the model in Section 2. In Section 3,
I study the two cases where risk sharing is not a challenge, so the only concern is providing
incentives. In Section 4, I study the case where information systems are ranked for both incen-
tive and risk-sharing purposes. I justify the first-order approach in Section 5, and conclude in
Section 6.
2. THE MODEL
A principal faces a set of contractible—that is, commonly observable and verifiable—
information systems, each of which can be represented by a random vector ˜x.5After choosing
an ˜x, the principal makes a take-it-or-leave-it contract, s˜x(·)[s,¯s], with an agent, who has an
outside reservation utility of 0. I abuse notation by dropping the subscript in s˜xif the context
makes it clear which information system sis based on. The parameters sand ¯sare the lower
and upper bounds of the agent’s payments, respectively.6If the contract is accepted, the agent
chooses an unobservable real-valued action aR+and incurs private cost c(a)withc>0
and c 0. The action astochastically generates a real-valued outcome ˜
bon a fixed support
b[b,¯
b], and ˜
bis imperfectly correlated with the information system ˜x.
The principal’s utility v(bs) is defined on her residual. To avoid negative residuals, I
assume b¯s. The agent derives utility from the received payment minus the private cost of
action, u(s)c(a). I assume that v>0, v 0, u>0, and u <0. In particular, the agent is
strictly risk averse, whereas the principal could be risk neutral or strictly risk averse.
All distribution and density functions are denoted by Fand f, respectively, with the subscript
indicating which random variables are intended. For instance, F˜
b(b|a) is the marginal distri-
bution function of ˜
b, given a, and f(˜
b,˜x)(b,x|a) is the joint density of ˜
band ˜x, given a. I abuse
notation by dropping the subscript if the context makes things clear; in this case, the arguments
indicate which random vectors are intended. All density functions are positive and continuous
4See Rogerson (1985), Jewitt (1988), Sinclair-Desgagne (1994), and, more recently, Conlon (2009) for justification
of the first-order approach in cases where the principal is risk neutral and/or the outcome is contractible.
5I therefore exclude the cases considered by Maskin and Tirole (1999) and Maskin (2002), where information systems
are either observable by one party but not the other or are commonly observable but nonverifiable. The principal’s
problems in these cases are mechanism design problems mixed with moral hazard, which are not the focus of this article.
6As explained by Jewitt et al. (2008), contracts are always bounded due to legal and other constraints on payments.
Page (1987) also introduces the upper and lower bounds to avoid nonexistence problems. However, more recently,
Kadan et al. (2011) prove the existence of an optimal contract without relying on a compact contract space.

To continue reading

Request your trial

VLEX uses login cookies to provide you with a better browsing experience. If you click on 'Accept' or continue browsing this site we consider that you accept our cookie policy. ACCEPT