venture capital investment under uncertainty and asymmetric

50
Venture Capital Investment under Uncertainty and Asymmetric Beliefs: A Continuous-Time, Stochastic Principal-Agent Model * Yahel Giat , Steven T. Hackman and Ajay Subramanian § October 21, 2008 Abstract We develop a continuous-time, stochastic principal-agent model to investigate the effects of asymmetric beliefs and agency conflicts on the characteristics and valuation of venture capital projects. In our model, a venture capitalist (VC) and an entrepreneur (EN) have imperfect information and differing beliefs about the intrinsic quality of a project in addition to having asymmetric attitudes towards its risk. We characterize the equilibrium of the stochastic dynamic game in which the VC’s dynamic investments, the EN’s effort choices, the dynamic compensation contract between the VC and EN, and the project’s termination time are derived endogenously. Consistent with observed contractual structures, the equilibrium dynamic contracts feature both equity-like and debt-like components, the staging of investment by the VC, the progressive vesting of the EN’s stake, and the presence of inter-temporal performance targets or milestones that must be realized for the project to continue. We numerically implement the model and calibrate it to aggregate data on VC projects. Our analysis of the calibrated model shows that EN optimism significantly enhances the value that venture capitalists derive. Entrepreneurial optimism explains the discrepancy between the discount rates used by VCs (40%), which adjust for optimistic payoff projections by ENs, and the average expected return of VC projects (15%). Our results show how the “real option” value of venture capital investment is affected by the presence of agency conflicts and asymmetric beliefs. Key Words: Dynamic Principal-Agent Models, Stochastic Dynamic Games, Incentive Con- tracts, Imperfect Information, Heterogeneous Beliefs. * We gratefully acknowledge financial support from the Kauffman Foundation under the “Roadmap for an En- trepreneurial Economy” initiative. We thank two anonymous referees and seminar audiences at the the 2007 Stan- ford Institute for Theoretical Economics (SITE) workshop on “Dynamic Financing and Investment”, the 2007 North American Summer Meeting of the Econometric Society (Duke University, Durham, NC), the 2007 Real Options Con- ference (Berkeley, CA), the 2008 Chicago-Minnesota Theory Conference (University of Chicago), the Fields Institute for Mathematical Sciences (Toronto, Canada), the University of Paris-Dauphine (Paris, France), and ESSEC (Paris, France) for valuable comments. The usual disclaimers apply. Department of Industrial Engineering and Management, Jerusalem College of Technology, Jerusalem, Israel School of Industrial and Systems Engineering, Georgia Institute of Technology, Atlanta, GA 30332 § Department of Risk Management and Insurance, J. Mack College of Business, Georgia State University, Atlanta, GA 30303

Upload: freddy56

Post on 18-Dec-2014

402 views

Category:

Documents


3 download

DESCRIPTION

 

TRANSCRIPT

Venture Capital Investment under Uncertainty and

Asymmetric Beliefs: A Continuous-Time, Stochastic

Principal-Agent Model∗

Yahel Giat†, Steven T. Hackman‡ and Ajay Subramanian§

October 21, 2008

Abstract

We develop a continuous-time, stochastic principal-agent model to investigate the effects ofasymmetric beliefs and agency conflicts on the characteristics and valuation of venture capitalprojects. In our model, a venture capitalist (VC) and an entrepreneur (EN) have imperfectinformation and differing beliefs about the intrinsic quality of a project in addition to havingasymmetric attitudes towards its risk. We characterize the equilibrium of the stochastic dynamicgame in which the VC’s dynamic investments, the EN’s effort choices, the dynamic compensationcontract between the VC and EN, and the project’s termination time are derived endogenously.Consistent with observed contractual structures, the equilibrium dynamic contracts feature bothequity-like and debt-like components, the staging of investment by the VC, the progressivevesting of the EN’s stake, and the presence of inter-temporal performance targets or milestonesthat must be realized for the project to continue. We numerically implement the model andcalibrate it to aggregate data on VC projects. Our analysis of the calibrated model shows thatEN optimism significantly enhances the value that venture capitalists derive. Entrepreneurialoptimism explains the discrepancy between the discount rates used by VCs (∼ 40%), whichadjust for optimistic payoff projections by ENs, and the average expected return of VC projects(∼ 15%). Our results show how the “real option” value of venture capital investment is affectedby the presence of agency conflicts and asymmetric beliefs.

Key Words: Dynamic Principal-Agent Models, Stochastic Dynamic Games, Incentive Con-tracts, Imperfect Information, Heterogeneous Beliefs.

∗We gratefully acknowledge financial support from the Kauffman Foundation under the “Roadmap for an En-trepreneurial Economy” initiative. We thank two anonymous referees and seminar audiences at the the 2007 Stan-ford Institute for Theoretical Economics (SITE) workshop on “Dynamic Financing and Investment”, the 2007 NorthAmerican Summer Meeting of the Econometric Society (Duke University, Durham, NC), the 2007 Real Options Con-ference (Berkeley, CA), the 2008 Chicago-Minnesota Theory Conference (University of Chicago), the Fields Institutefor Mathematical Sciences (Toronto, Canada), the University of Paris-Dauphine (Paris, France), and ESSEC (Paris,France) for valuable comments. The usual disclaimers apply.

†Department of Industrial Engineering and Management, Jerusalem College of Technology, Jerusalem, Israel‡School of Industrial and Systems Engineering, Georgia Institute of Technology, Atlanta, GA 30332§Department of Risk Management and Insurance, J. Mack College of Business, Georgia State University, Atlanta,

GA 30303

1 Introduction

Real-world productive activities are typically characterized by decentralized decision-making. The

agents who control various aspects of production often have different objectives, access to informa-

tion, and beliefs about project qualities (see Chen and Zenios, 2005, Gibbons, 2005). For example,

entrepreneurs (ENs) are often more optimistic about the success of their start-up firms than the

experienced venture capitalists (VCs) who provide capital (see Baker et al, 2005 for a recent sur-

vey). Further, VCs are usually well-diversified and less exposed to firm-specific risk than the less

diversified ENs who have significant human capital invested in their firms. As a result, VCs and

ENs have differing attitudes towards the risks of projects, which leads to agency conflicts that affect

the financing and operation of start-up firms. The interests of ENs (the “agents”) are aligned with

those of VCs (the “principals”) through incentive contracts that are affected by the VC’s and EN’s

heterogeneous beliefs about the outcomes of projects as well as their differing risk attitudes.

We develop a dynamic, stochastic principal-agent model of venture capital investment to ex-

amine the impact of asymmetric beliefs on the characteristics of venture capital projects—their

values, the structures of dynamic contracts between VCs and ENs, the durations of VC projects,

and the manner in which VC investment is staged over time. In our model, VCs and ENs have

asymmetric beliefs about the intrinsic qualities of projects as well as asymmetric attitudes towards

their risk. We characterize the equilibrium of the stochastic dynamic game in which the VC’s

dynamic investments, the EN’s effort choices, the dynamic compensation contract for the EN, and

the project’s termination time are derived endogenously. We calibrate the model parameters by

matching the return distributions of VC projects predicted by the model to their observed values in

the data. We show that the degree of EN optimism is significant enough to explain the discrepancy

between the discount rates used by VCs to value projects (∼ 40%), which adjust for optimistic

payoff projections by ENs, and the average expected return of VC projects (∼ 15%). EN optimism

is a key determinant of the durations and economic values of VC relationships, and could explain

features of observed contracts between VCs and ENs.

Model Overview: In our continuous-time stochastic model, a cash-constrained EN with a

project approaches a VC for funding. The project generates potential value through physical capital

1

investments by the VC and human capital (effort) investments by the EN. We model the evolution

of the project’s termination payoff at each date, which is the total payoff (present value of future

earnings) if the project is terminated at that date. The termination payoff evolves as a Gaussian

process and is contractible. The variance of the termination payoff process is the project’s intrinsic

risk, which remains invariant through time. The drift of the termination payoff process has two

components: a fixed, non-discretionary component that represents the project’s intrinsic quality,

and a discretionary component that is determined by the VC’s investment and the EN’s effort. The

discretionary component is observable, but non-verifiable and, therefore, non-contractible.

The VC and the EN have imperfect information about the project’s intrinsic quality and could

have differing, normally distributed priors. Their respective beliefs are, however, common knowl-

edge, that is, they “agree to disagree” about their respective mean assessments of project quality,

the difference of which represents the degree of asymmetry in beliefs. We consider the general sce-

nario in which the VC’s and EN’s mean assessments of project quality could differ from its true

mean. Further, the EN could be either optimistic or pessimistic relative to the VC. The common

variance of the VC’s and EN’s respective assessments of the project’s quality is the project’s tran-

sient risk. The transient risk is resolved over time as the VC and EN update their assessments of

the project’s quality based on observations of the project’s termination payoff.

The VC has linear preferences whereas the EN is risk-averse with CARA preferences. The VC

offers the EN a long-term contract that specifies her dynamic investment policy, the termination

time (a stopping time) of the project, and the EN’s payoff. The EN dynamically chooses his

effort to maximize his expected utility. The contractually specified payoffs of the VC and EN, the

investment policy, the EN’s effort policy, and the termination time are derived endogenously in

equilibrium of the dynamic game between the VC and EN.

The Equilibrium Contracts: We derive the incentive efficient dynamic contracts between the

VC and EN. Under an optimal contract, the change in the EN’s stake in the project or her promised

payoff (his “certainty equivalent” expected future utility) evolves as an Ito process. The change in

the EN’s stake has a performance-sensitive component that depends on the change in the project’s

termination payoff and a performance-invariant component that does not. The key contractual

parameters—the VC’s investments, the EN’s effort, and his compensation—are determined by the

2

EN’s pay-performance sensitivities, that is, the sensitivities of the change in the EN’s stake to

the change in the project’s termination payoff. Conditional on the project’s continuation, the VC’s

optimal investments and the EN’s pay-performance sensitivities are deterministic functions of time.

The performance-invariant component of the change in the EN’s stake over the period is, however,

stochastic and depends on the project’s termination payoff history through its effect on the VC’s

and EN’s updated assessments of the project’s intrinsic quality.

Consistent with observed contractual structures, (i) the VC’s payoff structure has “debt” and

“equity” components; (ii) the VC optimally stages her investment; (iii) the EN’s stake in the

project progressively vests over time; and (iv) the project is continued if and only if inter-temporal

milestones or performance targets are realized (see Gompers, 1995, Kaplan and Stromberg, 2003).

The Dynamics of Equilibrium Contracts: The time-paths of the VC’s investments and the

EN’s pay-performance sensitivities depend on the relative magnitudes of the degree of asymmetry

in beliefs and the costs of risk-sharing between the VC and EN. If the EN is pessimistic, then

the pay-performance sensitivities and investments increase over time. If the EN is “reasonably

optimistic,” i.e., the EN is optimistic, but the degree of his optimism is below a threshold relative

to the costs of risk-sharing, then the pay-performance sensitivities and investments decrease over

time. If, however, the EN is “exuberant,” i.e., the degree of EN optimism is above this threshold,

then the pay-performance sensitivities decrease over time and investments increase in early periods

and decrease in later periods. Hence, depending on the relative magnitudes of risk-sharing costs

and asymmetry in beliefs, the VC’s investment policy could become more aggressive over time,

less aggressive over time, or vary non-monotonically. The EN’s compensation could become either

more or less sensitive to performance over time.

The intuition for the above results hinges on the interplay among (i) the EN’s effort that is

positively (negatively) affected by his optimism (pessimism); (ii) the costs of risk-sharing due to

the EN’s risk aversion; and (iii) the complementary effects of the VC’s investment and the EN’s

effort on output. The passage of time lowers the degrees of optimism (or pessimism) as successive

project realizations cause the VC and the EN to revise their initial assessments of project quality.

Hence, the beneficial (detrimental) effects of optimism (pessimism) in mitigating the agency costs

of risk-sharing between the VC and the EN decline over time.

3

If the EN is optimistic, the VC can increase the performance-sensitive component of the EN’s

compensation because the EN overvalues this component. Hence, the EN’s pay-performance sen-

sitivity and effort are initially high. As the EN’s optimism declines over time, the positive effect

of optimism on the power of incentives that can be provided to the EN declines so that the EN’s

pay-performance sensitivity and effort decrease. (The opposite implications hold true when the EN

is pessimistic.) If the EN is optimistic, but his degree of optimism is below a threshold, the VC’s

investment also declines over time. Because investment and effort are complementary, the decline

in the power of incentives to the EN and, therefore, his effort over time causes the VC to also lower

her investment over time. If the EN’s optimism is above a threshold, the VC exploits the EN’s

exuberance by initially increasing her investment to compensate for the decrease in effort of the

EN. After a certain point in time when the VC’s investment attains its maximum, the decreasing

effort of the EN makes it optimal for the VC to also lower her investments.

The effects of risk on the investment path also depend on the degree of asymmetry in beliefs. If

the EN is either pessimistic or reasonably optimistic, the time path of optimal investments decreases

pointwise with the EN’s risk aversion, the project’s intrinsic risk, and its initial transient risk. If

the EN is exuberant, however, the investment path is, in general, non-monotonic and increases with

the EN’s risk aversion as well as the project’s intrinsic and transient risks in early periods, and

decreases in later periods. In contrast with traditional real options models in which all decisions are

made by monolithic agents (Dixit and Pindyck, 1994), the interaction between asymmetric beliefs

and agency conflicts could lead to a positive or negative relation between risk and investment.

Calibration and Numerical Analysis: We numerically implement the model and calibrate

the baseline values of its parameters, which include the average intrinsic quality of VC projects, the

degree of asymmetry in beliefs between VCs and ENs, the EN’s risk aversion, and his disutility of

effort. We estimate these parameters by matching the predicted distributions of “round by round”

returns of VC projects to their observed values reported in Cochrane (2005).

Consistent with anecdotal evidence, our “indirect inference” approach shows that ENs are,

indeed, significantly optimistic relative to VCs. (Recall that we do not assume that the EN is

optimistic a priori.) Because the VC exploits EN optimism through the provision of more powerful

incentives, EN optimism significantly enhances the value to the VC. Interestingly, however, EN

4

optimism lowers overall project value because the VC prolongs the project and over-invests to take

advantage of the EN’s optimism and enhance her own value at the expense of project value.

We examine how changes to the degree of EN optimism, the project’s intrinsic risk, and its

transient risk affect the project value and the VC’s stake. The project’s intrinsic and transient

risks have opposing effects on the “speed of learning” about project quality and, therefore, the rate

at which the degree of EN optimism declines over time. As a result, they have differing effects

on the project value and the VC’s stake: the project value and the VC’s stake decrease with the

project’s intrinsic risk, but vary non-monotonically with its transient risk.

Prior empirical and anecdotal literature documents that VCs use discount rates around 40%

to value projects even though the average expected return of VC projects is approximately 15%

(Cochrane, 2005). It has been suggested that higher discount rates could be a mechanism that VCs

use to “adjust” optimistic projections by ENs. Previous research, however, has not ascertained

whether EN optimism is, in fact, significant enough to generate such a large discrepancy between

VC discount rates and the average expected returns of VC projects. We define the implied discount

rate (IDR) as the rate at which the VC would discount the EN’s projections of the project’s payoffs

to conform to her own valuation of the project’s payoffs. The IDRs for a wide range of parameter

values predicted by the model lie between 30% and 50%, which is consistent with the range of

VC discount rates reported in prior empirical research (e.g. Sahlman, 1990, Cochrane, 2005). Our

study, therefore, confirms that entrepreneurial “optimism premia” are indeed high enough to justify

the discount rates used by VCs in reality.

Related Literature: Our study belongs to the growing body of literature that analyzes

dynamic principal-agent models. In a seminal study, Holmstrom and Milgrom (1987) present a

continuous-time principal-agent framework in which the principal and agent have CARA prefer-

ences and payoffs are normally distributed. They show that the optimal contract for the agent is

affine in the project’s performance. Schattler and Sung (1993) and Sung (1995) provide a rigorous

development of the first-order approach to the analysis of continuous-time principal-agent prob-

lems with exponential utility using martingale methods. Following Spear and Srivastava (1987),

a significant stream of the literature applies dynamic principal-agent models to study executive

compensation (Spear and Wang, 2005, Cvitanic et al, 2005, Cadenillas et al, 2006, Sannikov, 2007)

5

and financial contracting (DeMarzo and Fishman, 2006, Biais et al, 2007).

We contribute to this literature by developing and analyzing a dynamic principal-agent model

with imperfect public information and heterogeneous beliefs. The optimal dynamic contract reflects

the effects of Bayesian learning and the resultant dynamic variation of the degree of asymmetry

in beliefs in addition to the usual tradeoff between risk-sharing and incentives. Further, both the

principal and the agent take productive actions in our model.

In the specific context of venture capital, a strand of the literature investigates the importance

of staging in the mitigation of VC-EN agency conflicts. Using a deterministic model, Neher (1999)

shows that staging is essential to overcome the hold-up problem. As in Neher (1999), the manner

in which VC investment is staged over time as well as the number of stages are determined en-

dogenously in our framework. Our framework is, however, stochastic and incorporates asymmetric

beliefs between the VC and EN.1

Another strand of the literature on venture capital analyzes the features of the optimal contracts

that emerge in “double-sided” two-period moral hazard models in which the VC and EN exert effort

(Casamatta, 2003, Cornelli and Yosha, 2003, Schmidt, 2003, Repullo and Suarez, 2004). We too

develop a model in which the VC and EN take value-enhancing actions. Similar to these studies,

the optimal contracts predicted by our analysis have “debt” and “equity” features consistent with

observed contractual structures. Our study focuses on the effects of asymmetric beliefs on the

characteristics of VC-EN relationships in a dynamic principal-agent model.

The principal-agent paradigm is also applied to various operations management contexts. Fol-

lowing the early work of Atkinson (1979), recent studies examine the inefficiencies arising from

either hidden information (for example, Cachon and Lariviere, 2001, Ha, 2001) or hidden action

(for example, Lal and Srinivasan, 1993, Plambeck and Zenios, 2000, 2003, Chen, 2005) in supply

chain contracting. We contribute to this line of research by developing and analyzing a dynamic

principal-agent model with heterogeneous beliefs, and in which both the principal and agent make

value-enhancing decisions over time. Our framework could potentially be applied in supply chain

contexts as well as scenarios such as venture capital investment and R&D in which heterogeneous1Kockesen and Ozerturk (2004) argue that some sort of EN “lock-in” is essential for staged financing to occur.

Egli et al (2006) argue that staging can be used to build an EN’s credit rating. Berk et al (2004) develop an R&Dmodel with a single, monolithic agent in which staging is exogenous.

6

beliefs and agency conflicts play important roles.

In summary, we contribute to the literature by developing and analyzing a dynamic, stochastic

principal-agent model of venture capital investment. The model is parsimonious, yet realistic

enough to be taken to the data to yield quantitative assessments of the effects of the salient aspects

of VC projects, namely, risky payoffs, agency conflicts, uncertainty about project quality and

asymmetric beliefs. The tractability of the model, coupled with the fact that it is able to match

disparate empirical data on the payoff distributions of projects as well as the discount rates used

to value them, suggests that it could be useful as a tool to value risky ventures.

2 The Model

We develop a continuous time framework with time horizon [0, T ]. At date zero, a cash-constrained

entrepreneur (hereafter, EN) with a project approaches a venture capitalist (hereafter, VC) for

funding. The project generates value through physical capital investments by the VC and human

capital (effort) investments by the EN. Both the VC and the EN have imperfect information about

the project and differ, in general, in their initial assessments of the project’s quality.

If the VC agrees to invest in the project, she offers the EN a long-term contract that describes

her subsequent investments in the project, the EN’s compensation, and the termination time of the

relationship. The VC’s investments are made continuously over time. The termination time could

be a random stopping time.

The key state variable in the model is the project’s termination payoff Vt, which is the total

payoff if the VC-EN relationship is terminated at date t. The termination payoff is the only

economic variable that is contractible. For simplicity, we assume the project does not generate any

intermediate cash flows so that all payoffs occur upon termination.

2.1 The Termination Payoff Process

All stochastic processes are defined on an underlying probability space (Ω,F , P ) on which is defined

a standard Brownian motion B. The initial termination payoff of the project is V0. The incremental

termination payoff, that is, the change in the termination payoff over the infinitesimal period

7

[t, t + dt], dVt, is the sum of a base output, a Gaussian process that is unaffected by the actions

of the VC and EN, and a discretionary output, a deterministic component that depends on the

physical capital investments by the VC and the human capital (effort) by the EN. It is given by

dVt =

base output︷ ︸︸ ︷(Θ− lt)dt + sdBt +

discretionary output︷ ︸︸ ︷Acα

t ηβt dt. (1)

The first component, Θ, of the base output represents the project’s core output growth rate,

which we hereafter refer to as the project’s intrinsic quality. The VC and EN have imperfect

information about Θ and could also differ in their beliefs about its value. The second component of

the base change, lt, represents “operating costs,” which could include wages to salaried employees,

depreciation expenses, decline in revenues due to increased competition, fixed costs arising from

increases in the scale of the project, etcetera. These costs are deterministic and increasing over

time, which ensure that termination occurs in finite time almost surely. The third component of the

base change, sdBt, where s > 0 is a constant, represents the “intrinsic” component of the project’s

risk in period [t, t + dt]. It is the component of the project’s risk that remains invariant over time,

and is independent of Θ.

The discretionary output in period [t, t+dt] is a direct result of the VC’s capital investment rate

ct and the EN’s effort ηt, and is described by a Cobb-Douglas production function. The discretionary

output is observable to the VC and the EN. However, as in the literature on incomplete contracting

(see Chapter 6 of Laffont and Martimort, 2002), the discretionary output is non-verifiable and,

therefore, non-contractible. Because the discretionary output is non-contractible, the EN must be

indirectly provided with appropriate incentives to exert effort through her explicit contract with

the VC that can only be contingent on the termination payoff process.

The uncertainty in the value of Θ is the project’s transient risk. The VC’s and EN’s initial priors

on Θ are normally distributed with Θ ∼ N(µV C0 , σ2

0) and Θ ∼ N(µEN0 , σ2

0), respectively. Their

respective beliefs are, however, common knowledge, that is, they agree to disagree (see Morris,

1995, Allen and Gale, 1999). Because the equilibrium does not depend on how the EN’s and VC’s

mean assessments of project quality relate to its true mean, we make no assumptions about the

true mean of the project quality distribution. We consider the most general scenario in which the

8

EN could be optimistic or pessimistic relative to the VC, that is, µEN0 could be greater or less than

µV C0 . While the VC and EN disagree on the mean of the project’s intrinsic quality, they agree on

its variance, σ20.

2

The transient risk is resolved over time as the VC and the EN update their priors on Θ in a

Bayesian manner based on observations of the project’s performance. Define

dξt := dVt − (Φ(ct, ηt)− lt)dt = Θdt + sdBt. (2)

It follows from well-known formulae (Oksendal 2003) that the posterior distribution on Θ for each

date t ≥ 0 is N(µ`t, σ

2t ), ` = V C, EN , where

σ2t =

s2σ20

s2 + tσ20

, µ`t =

s2µ`0 + σ2

0ξt

s2 + tσ20

, ` = V C, EN. (3)

Note that the σt tend to zero. Let

∆t := µENt − µV C

t =s2∆0

s2 + tσ20

=∆0

σ20

σ2t (4)

denote the degree of asymmetry in beliefs at date t. It is resolved deterministically and monoton-

ically over time, and its absolute value, | ∆t |, also declines over time. Consequently, if the EN is

more optimistic (pessimistic) than the VC, the degree of optimism (pessimism) declines.

2.2 VC-EN Interaction

The contract between the VC and the EN describes the VC’s capital investments over time, the

EN’s effort, the termination date, and the EN’s payoff upon termination. The termination time

is, in general, a random stopping time that is contingent on the project’s performance history. We

follow the traditional principal-agent literature by having the contract also specify the EN’s effort

and requiring that the contract be incentive compatible with respect to the specified effort of the2The literature on behavioral economics (see Baker et al, 2005) distinguishes between optimism and overconfidence.

The EN is “optimistic” if his assessment of the mean (the first moment) of the project quality distribution is higherthan that of the VC, while he is “overconfident” if his assessment of the variance (the second moment) of the projectquality distribution is lower than that of the VC. In the terminology of the behavioral economics literature, therefore,the EN could be optimistic, but not overconfident in our framework.

9

EN (see Holmstrom and Milgrom, 1987).

Let Ft denote the information filtration generated by the history of termination payoffs, the

VC’s investments and the project’s discretionary outputs. A contract is described by the quadruple

(Pτ , c, η, τ), where c and η are Ft-adapted stochastic processes, τ is an Ft-stopping time, and

Pτ is a nonnegative Fτ-measurable random variable. Pτ is the EN’s contractually promised payoff

and Vτ −Pτ is the VC’s payoff at the termination time τ of the contractual relationship. In period

[t, t + dt], the VC’s investment rate is ct and the EN’s effort is ηt.

The VC offers the EN a long-term contract at date zero. The VC is risk-neutral whereas the

EN is risk-averse with inter-temporal CARA preferences described by a negative exponential utility

function. Their discount rates are equal and set to zero to simplify the notation. We extend the

model to incorporate nonzero discount rates when we calibrate it to the data in Section 6.

The EN’s expected utility at date zero from a contract (Pτ , c, η, τ) is

−EEN0

[exp

− λ

(Pτ −

∫ τ

tkηγ

t dt)]

. (5)

In (5), EEN0 denotes the expectation with respect to the EN’s beliefs at date zero and the parameter

λ ≥ 0 characterizes the EN’s risk aversion. The EN’s disutility from effort in period [t, t + dt] is

given by kηγt dt with k > 0, γ > 0. For future reference in the derivation of the equilibrium, we

follow Holmstrom and Milgrom (1987) by defining the EN’s certainty equivalent expected future

utility, Pt, from the contract at any date t as

exp(−λPt) := EENt exp

(− λ

(Pτ −

∫ τ

tkηγ

udu))

, (6)

where the notation EENt denotes the EN’s expectation conditioned on the information available at

date t, that is, the σ-field Ft. Note that the EN’s certainty equivalent future expected utility at the

contractual termination date τ is his contractually promised terminal payoff Pτ . For expositional

convenience, we hereafter refer to the EN’s certainty equivalent expected future utility process

Pt, t ≥ 0 as his promised payoff process.

The allocation of bargaining power between the VC and the EN is determined by the certainty

equivalent reservation utility or promised payoff that the EN must be guaranteed at date zero. We

10

allow for all possible allocations of bargaining power that are indexed by different values of P0.

A contract (Pτ , c, η, τ) is feasible if and only if it is incentive compatible for the EN with respect

to his effort choices, that is, given the terminal payoff, Pτ , the VC’s investment policy, c, and

the termination time τ , it is optimal for the EN to exert effort described by the process η. The

risk-neutral VC’s optimal contract choice is a feasible contract that maximizes her expected payoff

net of her investments, i.e., a feasible contract (Pτ , c, η, τ) is optimal if and only if it solves

(Pτ , c, η, τ) = arg max(Pτ ′ ,c′,η′,τ ′)

EV C0

[Vτ ′ − Pτ ′ −

∫ τ ′

tc′tdt

], (7)

where EV C0 denotes the expectation with respect to the VC’s beliefs at date zero and the maxi-

mization is over feasible contracts.

3 The Equilibrium

We assume the following condition on the parameters for the remainder of the paper:

Assumption 1 (1− α)γ/β > 2.

This condition implies that the EN faces decreasing returns to scale from the provision of effort.

Further, the EN’s disutility from his effort is sufficiently pronounced relative to his positive contri-

bution to output that an equilibrium contract between the VC and the EN exists.

3.1 Structure of Optimal Contract

The following two theorems characterize the optimal contract. Proofs are provided in Appendix A.

Theorem 1 (The EN’s Promised Payoff Process)

The EN’s promised payoff evolves as dPt = atdt + btdVt, where the contractual parameters at ∈IR, bt ∈ IR++ are Ft-progressively measurable.

The parameter bt is the EN’s pay-performance sensitivity. It represents the sensitivity of the change

in the promised payoff to performance during the infinitesimal period [t, t + dt]. The parameter at

is the EN’s performance-invariant compensation. It determines the component of the change in the

promised payoff that does not depend on performance during the infinitesimal period [t, t + dt].

11

In light of Theorem 1, a contract is completely specified by the performance-invariant compen-

sation and pay-performance sensitivity parameters, at, bt, the VC’s investment rate, ct, the EN’s

effort, ηt, at each time t, and the termination time τ .

3.2 Existence and Characterization of Equilibrium

We briefly outline the arguments involved in the derivation of the optimal contract, which is formally

characterized in Theorem 2 below and proved in Appendix A.

Fix date t ≥ 0. The derivation proceeds in four steps:

Step 1. The EN’s incentive compatible effort. For a given EN’s pay-performance sensitivity bt and

the VC’s investment rate ct at date t, we show that the EN’s incentive compatible effort is

η(bt, ct) :=(Aβcα

t bt

γk

) 1γ−β

. (8)

Step 2. The EN’s performance-invariant compensation. The VC optimally chooses her investment

rate ct and the EN’s pay-performance sensitivity bt incorporating the EN’s incentive compat-

ible effort given by (8). The performance-invariant compensation parameter at is chosen to

satisfy the “promise-keeping” constraint, that is, the EN’s promised payoff is actually deliv-

ered by the contract. The promise-keeping constraint pins down the contactual parameter at

as a function of the other two contract parameters bt and ct. In particular,

at = at(bt, ct) := 0.5λs2b2t + kη(bt, ct)γ − bt

(Acα

t η(bt, ct)β − lt + µEN

t

). (9)

Step 3. The optimal investment and pay performance sensitivity. Incorporating the EN’s incentive

compatible effort (8) and the functional form for at in (9), the change in the VC’s continuation

value (her expected future payoffs) at date t is

CVt = Λt(bt, ct)dt, where (10)

Λt(bt, ct) := ∆tb− 0.5λs2b2 + φ(b)cα γγ−β − c + µV C − lt. (11)

12

In (11), ∆t is the degree of asymmetry of beliefs at date t, defined in (4), and

φ(b) :=

γ−β

(1k

) βγ−β

((βbγ

) βγ−β

(1− βb

γ

)), if 0 ≤ b ≤ γ/β,

0, otherwise.

(12)

The function Λt(bt, ct) is the rate-of-change of the VC’s continuation value; hereafter, we

shall refer to it simply as the continuation rate. The VC chooses the investment rate ct and

the EN’s pay-performance sensitivity bt to maximize the continuation rate. Assumption 1

guarantees a unique solution b∗t , c∗t to this maximization problem. In particular, it implies

that Λt(bt, ·) is strictly concave in ct, since the exponent on ct is less than one. Consequently,

given the pay-performance sensitivity, there is a unique investment rate. We show that

c∗t = c(b∗t ) :=( αγ

γ − β

) γ−β(1−α)γ−β φ(bt)

γ−β(1−α)γ−β , (13)

b∗t = arg max0<bt

Λt(bt, c(bt)), where (14)

Λt(bt, c(bt)) :=[∆tbt − 0.5λs2b2

t +γ − β − αγ

αγc(bt)

]+ (µV C

t − lt). (15)

We refer to the function c(·) as the optimal investment function. We discuss the properties

of the optimal investment function, which plays a central role in our analysis, in Section 3.4.

Step 4. Determination of the optimal termination time. The optimal termination time of the contract

is the solution to the optimal stopping problem

τ∗ = arg maxτ≤T

EV C0

∫ τ

0Λt(b∗t , c

∗t )dt, (16)

where the maximization is over all Ft-stopping times τ ≤ T .

Theorem 2 (Characterization of Equilibrium)

(a) Conditional on the project not being terminated prior to date t ∈ [0, T ]:

– The EN’s pay-performance sensitivity parameter, b∗t , solves (14).

– The VC’s equilibrium investment rate is c∗t = c(b∗t ), where c(·) is defined in (13).

13

– The EN’s performance-invariant compensation parameter is a∗t := at(b∗t , c∗t ), where at(·, ·)is defined in (9).

– The EN’s effort level is η∗t := η(b∗t , c∗t ), where η(·, ·) is defined in (8).

(b) The termination time of the relationship solves the optimal stopping problem (16).

3.3 The VC’s Controllable Rate Function

Let

Ft(b) := ∆tb− 0.5λs2b2 +γ − β − αγ

αγc(b) (17)

denote the “controllable” portion of the continuation rate; we hereafter refer to it as the VC’s

controllable rate function. As summarized in Theorem 2, the equilibrium contract at date t is

determined by b∗t , the solution to (14). An examination of (15) shows that b∗t is also the solution to

b∗t = arg max0<b

Ft(b). (18)

By Theorem 2 and (18), the EN’s pay-performance sensitivity b∗t , the VC’s investment rate c∗t ,

and the EN’s effort η∗t are deterministic functions of time (conditional on the project’s continuation).

The performance-invariant compensation parameter a∗t is, however, stochastic and depends, in

particular, on the EN’s current mean assessment µENt of the project’s intrinsic quality. The proof

of the theorem shows that this parameter adjusts stochastically to ensure that the EN’s promise

keeping constraints are satisfied at each date and state.

The equilibrium contract critically depends on the VC’s controllable rate function, Ft(b). This

function consists of three components:

• Economic rent (cost) from the EN’s optimism (pessimism). When ∆t > 0, the term, ∆tb,

reflects the rents that the VC extracts from the EN by exploiting his optimism about the

project’s intrinsic quality. When ∆t < 0, ∆tb is the cost that the VC must bear to compensate

the EN for his pessimism about the project’s intrinsic quality.

• Cost of risk. The term, 12λs2b2, reflects the VC’s costs of risk-sharing with the risk-averse

EN. We refer to λs2 as the price of risk ; hereafter, we denote it by p.

14

• Return on investment. The “return on investment” term, γ−β−αγαγ c(b), reflects the VC’s

expected return as a result of her investment and the EN’s effort.

The interplay among these three “forces” determines the equilibrium dynamics.

3.4 Uniqueness and Stability of Equilibrium

The characteristics of the contract depend on the optimal investment function c(·) given in (13).

The following properties of c(·) play a central role in our subsequent analysis (see Figure 1).

Proposition 1

(a) The function c(·) is strictly positive and strongly unimodal3 on [0, γβ ], satisfies c(0) = c( γ

β ) = 0,

and achieves its maximum at b = 1.

(b) The function c(·) is strictly concave on [0, bM ] and strictly convex on [bM , γβ ], where bM ∈ (1, γ

β )

is the unique minimum of the function c′(·).

Proof. The proof of this, and all subsequent results in the paper, are provided in Appendix B.

The intuition for the non-monotonicity of the function is that an increase in the agent’s pay-

performance sensitivity affects the principal’s investment in two distinct but opposite ways. On

the positive side, the agent increases his effort. Because investment and effort are complementary,

the increase in the agent’s effort provides an incentive for the principal to increase her investment.

On the negative side, since the agent’s disutility of effort increases, the principal’s cost to maintain

the agent’s participation also increases. For lower values of the pay-performance sensitivity, the

complementarity of investment and effort causes the benefits of increased output to dominate.

Hence, the principal finds it beneficial to increase her investment. However, beyond a threshold

level of pay-performance sensitivity, the costs of inducing high effort from the agent are so high

that the principal lowers her investment. In other words, it is optimal for the principal to allow

output to be dominated by the agent’s effort.

The ratio of the absolute value of the initial degree of asymmetry of beliefs to the price of risk,

namely, | ∆0 | /p, provides an a priori bound on the equilibrium pay-performance sensitivity b∗t .3A function f(·) is strongly unimodal on the interval [a, b], a < b, if there exists an x∗ ∈ (a, b) such that f(·) is

increasing on [a, x∗] and f(·) is decreasing on [x∗, b]. Obviously, the value x∗ maximizes f(·) on [a, b].

15

Figure 1: Optimal investment function

6

- b

c(b)

1

qc′(1) = 0

γβ

c(0) = c( γβ

) = 0 bM

.

...........................

.........................

.......................

......................

....................

...................

.................

................

.............................................................................. ........ ....... ....... ....... ....... ...... ....... ....... ....... ....... ........ ........ ........ ........ ......... ........ ......... ........ .......... ........ ........... ........ ....................... ......... ......... ....................... ........ ........... ......... .......... ......... ......... ........ ......... ........

................

............... ............... .............. ............. ...... ....... .........

c′′(bM ) = 0q¾ convex -

¾ c o n c a v e-

Proposition 2

An optimal solution to (18) is always less than or equal to max∆0p , 1 if ∆0 ≥ 0 and is less than

1 if ∆0 < 0.

In our subsequent analysis we assume that the initial degree of asymmetry in beliefs, ∆0, is below

a threshold relative to the price of risk, p. The assumption ensures that the equilibrium is stable

and the contractual parameters are continuous functions of the primitives of the model.

Assumption 2 ∆0/p ≤ bM .

(The parameter bM above is defined in Proposition 1.) It follows immediately from Proposition

2 and Assumption 2 that a solution to (18) must lie in the interval [0, bM ). By Proposition 1,

the optimal investment function c(·) is strictly concave on the interval [0, bM ]. It follows from

(17) that the VC’s controllable rate function, Ft(b), is also strictly concave on [0, bM ] and hence

strongly unimodal. Consequently, there exists a unique solution b∗t to (18). Moreover, it must also

be positive, since the proof of Proposition 1 shows that the marginal optimal investment c′(0) is

infinite. We summarize these observations with the following proposition.

Proposition 3

Under Assumptions 1 and 2, the function Ft(·) is strictly concave on [0, bM ]. Further, the solution

to (18) is strictly positive and less than bM .

4 Equilibrium Dynamics

We investigate the dynamics of the EN’s compensation, his effort, and the VC’s investment condi-

tional on continuation of the project. Since the degree of asymmetry in beliefs, ∆t, and variance,

σ2t , are deterministic functions of time (see (4)), it follows from Theorem 2 and Proposition 1 that

16

the equilibrium values for the pay-performance sensitivity, investment and effort at each point in

time (conditional upon continuation) are also deterministic. The only component of the contract

that is stochastic and is adjusted based on realizations of the termination payoff Vt of the project

is the performance-invariant compensation parameter a∗t .

Let

Ft(b) = F (b) := −0.5λs2b2 +γ − β − αγ

αγc(b) (19)

be the principal’s controllable rate function in the benchmark scenario in which beliefs are sym-

metric. Since F (.) is time-independent, the agent’s equilibrium pay-performance sensitivities, the

principal’s investments and the agent’s effort are all constant. Let b∗p, c∗p and η∗p denote their values.

It follows from (4) and (17) that the VC’s controllable rate function can be expressed as

Ft(b) =∆0

σ20

σ2t b + F (b), (20)

where F (.) is defined in (19). Since σt → 0, it follows from Berge’s Theorem of the Maximum

that b∗t → b∗p, and thus (c∗t , η∗t ) → (c∗p, η∗p) by continuity where (b∗p, c∗p, η∗p) are the equilibrium pay-

performance sensitivity, investment, and effort in the benchmark scenario with symmetric beliefs.

We now describe the manner in which these economic variables converge to their asymptotic values.

Theorem 3 (The Dynamics of the Equilibrium—Optimistic EN)

Suppose that the EN is more optimistic than the VC so that ∆0 ≥ 0.

(a) The EN’s pay-performance sensitivity b∗t decreases monotonically with t and approaches b∗p

as t →∞.

(b) The value

t∗ := (∆0

p− 1)

s2

σ20

=∆0

λσ20

− s2

σ20

(21)

is the point in time at which the EN’s pay-performance sensitivity and effort, and the VC’s

investment rate equal their values in the “no agency” benchmark scenario. (This interpreta-

tion of t∗ only applies if t∗ ≥ 0, which holds if and only if the initial degree of asymmetry of

beliefs is at least as large as the price of risk.)

17

(c) The EN’s pay-performance sensitivity b∗t exceeds 1 if t < t∗, equals 1 at t = t∗, and less than

1 if t > t∗.

(d) The VC’s investment rate c∗t increases until time t∗ and then decreases monotonically towards

c∗p as t →∞.

(e) For t ≥ t∗, η∗t decreases monotonically towards η∗p as t →∞ .

Note that if ∆0 ≤ p so that t∗ ≤ 0, then the EN’s pay-performance sensitivity, his effort, and the

VC’s investment rate all decrease monotonically over time. Figures 2 and 3 illustrate the results of

Theorem 3. They describe the trajectories of the EN’s pay-performance sensitivity and the VC’s

investment for two different values of the initial degree of asymmetry in beliefs, ∆0.

The intuition for the results of Theorem 3 hinges on the interplay among the positive effect

of optimism on the EN’s effort, the costs of risk-sharing due to the EN’s risk aversion that are

negatively affected by the project’s intrinsic risk, and the complementary effects of investment

and effort on output. The passage of time lowers the degree of optimism as successive project

realizations cause the VC and the EN to revise their initial assessments of project quality.

If the EN is optimistic, the VC can increase the performance-sensitive component of the EN’s

compensation because the EN overvalues this component. Hence, the EN’s pay-performance sen-

sitivity and effort are initially high. The negative effect of the evolution of time on the EN’s

optimism, however, causes the EN’s pay-performance sensitivity and effort to decline over time.

Due to the previously discussed non-monotonic relation between the VC’s investment and the EN’s

pay-performance sensitivity, the VC’s investment initially increases to “compensate” for the de-

crease in effort of the EN. After a certain point in time when investment attains its maximum, the

decreasing effort of the EN makes it optimal for the VC to also lower her capital investments.

Theorem 4 (The Dynamics of the Equilibrium—Pessimistic EN)

Suppose that the EN is more pessimistic than the VC so that ∆0 < 0.

(a) The EN’s pay-performance sensitivity b∗t increases monotonically with t and approaches b∗p as

t →∞.

(b) The VC’s investment rate c∗t increases monotonically towards c∗p as t →∞.

18

6

- t

b∗t

0 2 4 6 8 10 12 14

....................

............................................ ........... .............. ................. ............ ................ .................... ............. ................ ................... ............. ................ .................. ............. ............... .................. .............. ............... ................. ..................... ........................ ............. ............... .................. ............................................. ............. ............... ................. ..................... ........................ ............................................. ............. ............... ................. .............................................

∆0 < 0

b∗p∆0 = 0

.............

.............

......................

....................

......... ............ ............... .................. ............. ................. .................... ............. ................ ................... ............. ................ .................. ............. ................ .................. .............. ................ ................. .............. ............... ................. ...................... ........................ ..................... ........................ ............. ............... ................. ............................................. ...................... ........................ ............. ............... .................. .............................................

∆0 > 0

Figure 2: Possible equilibrium pay-performance sensitivity paths

(c) The EN’s effort η∗t increases monotonically towards η∗p as t →∞ .

If the EN is pessimistic, he under-values the performance-sensitive portion of his compensation

relative to the VC. Hence, the power of incentives that can be provided to the EN is initially

low so that his pay-performance sensitivity and effort as well as the VC’s investment are initially

low. With the evolution of time, the degree of pessimism declines, which has a positive effect on

the power of incentives to the EN so that his pay-performance sensitivity, effort, and the VC’s

investment all increase.

Theorems 3 and 4 describe the paths of the EN’s pay-performance sensitivity and effort, and

the VC’s investment rate conditional on the project’s continuation. Depending on the relationship

between the degree of asymmetry in beliefs and the price of risk, it follows from the theorems

that the VC’s investments until termination (these are the investments that are actually observed

because there is no investment after termination) could either increase, decrease, or vary non-

monotonically (initially increase and then decrease).

4.1 Sensitivity of Equilibrium Dynamics

In light of Theorems 3 and 4, the manner in which the equilibrium dynamics are affected by changes

in the underlying parameters critically depends on the initial value of the degree of asymmetry in

beliefs ∆0. In what follows the EN is said to be pessimistic if ∆0 < 0, reasonably optimistic if

∆0 ∈ [0, p) and exuberant if ∆0 ∈ (p, pbM ], where p = λs2 is the price of risk. (Assumption 2

guarantees that ∆0 ≤ pbM .)

19

6

- t

c∗t

2 4 6 8 10 12 14 16

c(1) “no agency” investment level

.....................

............................................. ........... .............. ................. ............ ................ .................... ............. ................ ................... ............. ................ .................. ............. ............... .................. .............. ............... ................. ............. ............... ................. ...................... ....................... ............ ............... ................... ........................ ..................... ...................... ........................ ............ ............... .................. ............................................. .............................................

∆0 < 0

c∗p∆0 = 0

...............

.....

...................

.................. ......... ............ .............. ................ ............. ................ .................... ............. ................ ................... ............. ................ .................. .............. ................ .................. .............. ................ ................. .............. ............... ................. ..................... ........................ .............. ............... ................. ............................................. ..................... ......................... ...................... ........................ ..................... ........................ .............................................

∆0 < p

.

.................................

..............................

............................

.........................

......................

...................

...................................................... .................. ..... ....... .......... ............ ............... ......... .......... .............. .................. ................ ............. .......... ....... ...... ......................... ........................ ........................ ......................... ....................... ......................... ....................... ......................... ....................... ......................... ....................... ........................ ...................... ......................... .............................................. ..................... ......................... ..............................................

∆0 > p

Figure 3: Possible equilibrium investment paths

The following theorem characterizes the effects of the EN’s risk aversion, λ, the initial transient

risk, σ20, and the intrinsic risk, s2, on the equilibrium dynamics.

Theorem 5 (Effects of Intrinsic Risk, Transient Risk and EN’s Risk Aversion)

(a) If the EN is pessimistic, the paths of the EN’s pay-performance sensitivity and the VC’s

investment are pointwise decreasing in the EN’s risk aversion, pointwise increasing in the

initial transient risk, and pointwise decreasing in the intrinsic risk.

(b) If the EN is optimistic, the path of the EN’s pay-performance sensitivity is pointwise decreas-

ing in the EN’s risk aversion and the initial transient risk.

(c) If the EN is reasonably optimistic, then the path of the VC’s equilibrium investment is

pointwise decreasing in the EN’s risk aversion and the initial transient risk.

(d) If the EN is exuberant, the VC’s investment path changes as depicted in Figure 4 as a result

of a change in the EN’s risk aversion and the initial transient risk. More precisely, let λ1 < λ2

and σ1 < σ2 be two possible values of the EN’s risk aversion and the initial transient risk,

respectively. There exist t∗(λ1, λ2) and t∗∗(σ1, σ2) such that the VC’s investments when the

EN’s risk aversion is λ1 (the initial transient risk is σ1) are higher than her investments when

the EN’s risk aversion is λ2 (the initial transient risk is σ2) for t < t∗(λ1, λ2) (t < t∗∗(σ1, σ2))

and lower for t > t∗(λ1, λ2) (t > t∗∗(σ1, σ2)).

(e) If ∆0 ≤ 4p, parts (b)-(d) hold for the intrinsic risk. If ∆0 > 4p, the effects of intrinsic risk on

20

the pay-performance sensitivity and investment paths is ambiguous.4

Figure 4 demonstrates that the path of equilibrium investment converges to different limiting values

depending on the EN’s risk aversion.

The EN’s pay-performance sensitivity, b∗t , declines with his risk aversion because an increase in

the EN’s risk aversion increases the costs of risk-sharing. An increase in the transient risk lowers

the degree of optimism or pessimism at each date because the “signal to noise ratio” is increased

so that the VC and EN “learn faster.” An increase in the intrinsic risk increases the costs of risk-

sharing and also increases the degree of optimism or pessimism at each date because the “signal to

noise ratio” decreases so that the VC and EN “learn more slowly.”

When the EN is pessimistic, the pointwise decline in the degree of pessimism with the transient

risk increases the power of incentives to the EN so that the EN’s pay-performance sensitivity and

the VC’s investment increase at each date. On the other hand, the pointwise increase in the degree

of pessimism and the costs of risk-sharing with the intrinsic risk decreases the power of incentives

to the EN so that his pay-performance sensitivity and the VC’s investment decrease at each date.

When the EN is optimistic, the decline of the degree of optimism with the initial transient risk

causes the economic rents to the VC in each period from the EN’s optimism to be lowered relative

to the costs of risk-sharing. Hence, the EN’ pay-performance sensitivity declines. Intrinsic risk,

however, has conflicting effects on the power of incentives to the EN. An increase in the intrinsic

risk increases the degree of optimism at each date, which has a positive effect on the power of

incentives. However, an increase in the intrinsic risk also increases the costs of risk-sharing, which

has a negative effect on the power of incentives. When the EN is optimistic and ∆0 ≤ 4p, the

costs of risk-sharing outweigh the benefits of the EN’s optimism so that the EN’s pay-performance

sensitivity also decreases with intrinsic risk. When ∆0 > 4p, the conflicting effects of optimism and

risk-sharing costs cause the effects of intrinsic risk to be ambiguous.

The change in the VC’s investment path when the EN is optimistic critically depends on whether

the EN is reasonably optimistic or exuberant. If the EN is reasonably optimistic, then the costs of

risk-sharing outweigh the benefits of the EN’s optimism so that the VC’s investment path declines4The condition is trivially satisfied when ∆0 < 0. Under Assumption 2, the condition ∆0 ≤ 4p is automatically

satisfied when bM ≤ 4. Since b∗t < bM , this condition implies that the EN’s optimal pay-performance sensitivityshould be less than four, which is easily satisfied in reasonable parametrizations of the model.

21

6

- t

c∗t

t∗2 t∗1

c(1) “no agency” investment level..................................... ........ .......... ...... ........ .......... ...... ......... ........... ...... ......... ........... ........ ........ .......... ............. ......... .......... .............. ........... ............. ............ ............ ............. ......................... ......................... ........................ ............ ............. ........................ ........... ............. ........................ ........................ ........... ............. ........................ ............. ............ ........... ............. ............. ............ ........................ ........... ............. ............. ........... ........... ............. ....................... ........... ............. ............. ........... .......................

c2(t)

.

...............................

............................

........................

.........................

..........................

................

............................................ ........... .............. ........... ............. .......... ............. .......... .............. ........... ............. ........... ............. ........ .......... .............. ................. ................. .............. .......... .............. ........ .......... .......... ........ ....... ............ ............. ........................ ............. ............ ........... ............. ............. ........... ........... ............. ............. ........... ........... ............. ............. ........... ...... ........ .......... .......... ........ ...... ...... ........ .......... .......... ........ ...... ........... ............. .......................

c1(t)

Figure 4: Sensitivity of the equilibrium investment path to a change in the EN’s risk aversion, theinitial transient risk, or the intrinsic risk. Path c2(·) corresponds to an increase in λ, σ2

0 or s2.

with the EN’s risk aversion as well as the project’s intrinsic and transient risk. If the EN is

exuberant, then an increase in intrinsic or transient risk increases the costs of risk-sharing, thereby

partially offsetting the VC’s rents from the EN’s optimism. Early in the project, it is beneficial for

the VC to compensate for the resulting decline in the EN’s effort by increasing investment. As time

passes, however, the EN’s degree of optimism declines thereby reducing the rents to the VC. The

costs of risk-sharing, therefore, dominate in later in the project so that an increase in risk results

in a decline in the VC’s investment.

In stark contrast with traditional real options models with monolithic agents (e.g. Dixit and

Pindyck, 1992), the results of Theorem 5 show that the interactive effects of optimism and agency

conflicts could lead to a positive or negative relation between risk and investment.

Theorem 6 (Effects of Degree of Asymmetry in Beliefs)

(a) The path of the EN’s equilibrium pay performance sensitivity is pointwise increasing in the

initial degree of asymmetry in beliefs.

(b) If the EN is pessimistic or reasonably optimistic, then the path of the VC’s equilibrium

investment is pointwise increasing in the initial degree of asymmetry in beliefs.

(c) If the EN is exuberant, then the path of equilibrium investment by the VC changes as in

Figure 5 as a result of a change in the initial degree of asymmetry in beliefs—the time-path

of investment shifts “to the right” if the initial degree of asymmetry increases.

22

6

- t

c∗t

t∗1 t∗2

c(1) “no agency” investment level.............................. ....... ........ .......... ...... ........ .......... ...... ........ ........... ...... ......... ........... ....... ........ ........... ........... ........ ....... .............. ........... ............. ............ ........................ ........................ ............ ............. ........................ ............ ............. ........................ ........... ............. ........................ ............ ............ ........... ............. ............. ........... .......... .............. ............. ........... .......... .............. ............. .......... ........... ............ ....................... .......... .............. .......... ........ ...... .......... .............. ............. ........... .......................

c1(t)

.

...............................

............................

........................

.........................

..........................

................

............................................ ........... .............. ........... ............. .......... ............. .......... .............. ........... ............. ........................ ...................................... ........ ........... .............. .................. ..................... ........................ ............................ ...... ........ ........... .......... ........ ....... ............ ............. ........................ ............. ............ ........... ............. ............. ........... ........... ............. ............. ........... ........... ............. ............. ........... ...... ........ .......... .......... ........ ...... ...... ........ .......... .......... ........ ...... ........... ............. ....................... c2(t)

Figure 5: Sensitivity of equilibrium investment path to the initial degree of asymmetry in beliefs.Path c2(·) corresponds to an increase in ∆0.

An increase in the initial degree of asymmetry in beliefs increases the power of incentives that

can be provided to the EN so that his pay-performance sensitivity increases at each date. When

the EN is pessimistic or reasonably optimistic, the VC increases her investment at each date. When

the EN is exuberant, however, the investment path is non-monotonic. The intuition for the effects

of risk on investment discussed earlier is reversed so that the degree of asymmetry in beliefs affects

the investment path as described in part (c) of the theorem.

5 Project Duration

The following proposition shows that there exists a trigger level of the project’s mean quality at

each date such that it is optimal for the VC to continue the project if and only if her current

assessment of the project’s quality exceeds the trigger.

Proposition 4 (The Optimal Termination Policy)

The optimal stopping policy for the VC is a trigger policy: there exist µ∗t such that the VC

terminates the project only if µV Ct < µ∗t .

23

Let Y ∗t dt := (c∗t

αη∗tβ − lt)dt denote the equilibrium net discretionary output in period [t, t+ dt].

Since dVt = Y ∗t dt + ξtdt, it follows that

Vt − V0 =∫ t

0dVu =

(∫ t

0Y ∗

u du)

+(∫ t

0ξudu

).

Given the formula for µt given in (3), we may conclude that µt ≥ µ∗t if and only if Vt ≥ V ∗t , where

V ∗t := V0 +

(∫ t

0Y ∗

u du)

+(s2 + tσ2

0)µ∗t − s2µ0

σ20

.

The sequence of the V ∗t may be thought of as the performance targets the project must reach at

each date to prevent termination.

An increase in the EN’s initial degree of optimism about project quality increases the rents to

the VC from the EN’s optimism thereby increasing her expected continuation value at each point

in time. Hence, an increase in the EN’s optimism prolongs the project’s duration. An increase

in the EN’s risk aversion or cost of effort, however, increases the costs of risk-sharing for the VC,

thereby lowering her continuation value at each point in time.

Proposition 5 (Comparative Statics of Project Duration)

The project duration τ increases with the initial degree of asymmetry in beliefs, decreases with the

EN’s risk aversion, and decreases with the EN’s cost of effort.

6 Numerical Analysis

We numerically explore further implications of the model using a discrete-time approximation of

the continuous-time model. We describe the details of the numerical implementation in the on-line

Appendix C. We directly model the evolution of the VC’s current assessment of project quality µV Ct

because, as explained in Section 5, it determines her continuation decision at any date t. In the

first stage of the numerical implementation, we approximate the evolution of µV Ct using a discrete

lattice and derive the termination triggers µ∗t . In the second stage, given the triggers obtained from

the first stage, we use Monte Carlo simulation to model the evolution of µV Ct and to obtain the

key output variables of interest. Gompers (1995) reports that the average length of a round of VC

24

financing is approximately one year. Accordingly, we set the time period between successive dates

in the discrete lattice to one year and assume that it corresponds to a single round of financing.

6.1 Calibration

To obtain a reasonable set of “baseline” parameter values for our numerical analysis, we calibrate

the model to actual aggregate data on the distribution of round by round returns of venture

capital projects reported in Cochrane (2005). We classify the parameters of the model into two

groups: “direct” parameters whose baseline values can be set using guidance from previous empirical

research, and “indirect” parameters whose values are estimated by matching statistics predicted

by the model to their observed values in the data.

In our numerical implementation, we incorporate a nonzero discount rate for the VC (and EN).

Cochrane (2005) finds that the average expected return on venture capital investment in his sample

is 15%. Accordingly, we set the discount rate, Rb, to 15%. Further, we assume that the VC has

experience so that her prior assessment of the project quality distribution is correct. We assume a

production technology with constant returns to scale so that β = 1 − α. We assume a quadratic

form l(t) = l1tl2 for the loss function. We normalize the initial seed capital V0 to one and set P0

to this value. We estimate the remaining parameters of the model (see Table 1) by matching its

predictions to data.

In our estimation, we use statistics on the round by round returns and standard deviations of

VC projects in each of the first four rounds of financing reported in Table 4 of Cochrane (2005).

As the length of a single round of financing is set to one year, the round-by-round returns of a VC

project in the model are Vt−Vt−1−ct−1

ct−1, 1 ≤ t ≤ 4. Cochrane (2005) also reports the overall mean

and standard deviation of the VC project returns and the average number of rounds of financing.

The statistics used for our estimation are displayed in the first rows of the two panels of Table

2. We estimate the values of the indirect parameters of the model by matching the predicted

values of the statistics in Table 2 to their observed values. The standard errors of the estimates are

determined by parametric bootstrapping (see the on-line Appendix C).

As shown in Table 2, the model is able to closely match the observed statistics. The estimated

value of the degree of asymmetry in beliefs ∆0 = 0.504, while the VC’s assessment of the mean

25

project quality µV C0 = 0.113. The data, therefore, suggest that the level of entrepreneurial optimism

is very significant. The baseline values of the average intrinsic risk s and transient risk σ0 are high,

which confirms anecdotal and empirical evidence that venture capital is risky and is characterized

by significant uncertainty about project quality.

6.2 Numerical Results

We first analyze the model when the parameters take their baseline values in Table 1. We then

explore various comparative static relationships by varying parameters about their baseline value.

In our numerical analyses, we compare the actual scenario in which there are asymmetric beliefs

and agency conflicts with two benchmark scenarios: the no agency scenario in which beliefs are

symmetric, and both the VC and EN have linear preferences; and the symmetric beliefs scenario

in which the VC and EN have symmetric beliefs, but the EN has CARA preferences.

We compute two output variables in each of the three scenarios. The

Project V alue := EV C0

[e−RbτVτ −

τ−1∑

t=0

e−Rbtct

]. (22)

is the expected total payoffs to the firm less the capital investments discounted at the rate Rb. The

V C V alue := Project Value− EV C0

[e−RbτPτ

]. (23)

is the Project Value less the termination payoff to the EN discounted at the rate Rb. The expecta-

tions in (22) and (23) are with respect to the VC’s beliefs about project quality, which are assumed

to be correct. The termination payoff process evolves as in (1) with the contractual parameters,

(a∗, b∗, c∗), the EN’s effort, η∗, and the performance targets, V ∗, set to their equilibrium values

for the specific economic scenario (no agency, symmetric beliefs or actual) being analyzed.

6.2.1 Baseline Analysis

Table 3 reports the Project Value and VC Value in the actual scenario and the two benchmark

scenarios. The difference between the project values (VC values) in the no agency and symmetric

benchmark scenarios represent the deadweight agency costs of risk sharing between the VC and

26

EN from the perspective of the firm (the VC fund). The difference between the project values (VC

values) in the actual and symmetric scenarios reflect the effects of EN optimism on the project

value (VC value). We see that the VC significantly benefits from EN optimism. The project value,

however, is lower in the actual scenario than in the symmetric beliefs scenario. This is the due to

the fact that the VC exploits EN optimism by over-investing and prolonging the project’s duration

to increase her value at the expense of the overall value of the project.

Table 4 reports the EN’s pay-performance sensitivities and the investments for the first four

rounds. Consistent with Theorem 3, the EN’s pay-performance sensitivity and the VC’s investments

decline over time. The EN’s pay-performance sensitivity decreases sharply across the four periods.

Successive capital infusions by the VC, therefore, rapidly reduce the EN’s stake in the firm.

6.2.2 Comparative Statics

Figure 6 (a) shows that the project value and VC value both vary non-monotonically with the

initial transient risk—they initially decrease and then increase. To understand the intuition for the

effects of transient risk, note that, by (4), the degree of EN optimism at any date t declines with

the initial transient risk. This has a negative effect on the power of incentives to the EN, his effort,

and the economic rents the VC can extract from EN optimism. From (2) and (3), the standard

deviation σµt of the evolution of the mean assessment of project quality is

σµt =

sσ20

s2 + tσ20

. (24)

From (24), an increase in the initial transient risk increases the standard deviation of the evolution

of the mean assessment of project quality and, therefore, the likelihood of both high and low

realizations. Since the VC can limit her downside by terminating the relationship if intermediate

signals of project quality are sufficiently poor, the “real option value” of continuing the project

increases with the initial transient risk. The interaction between the negative effects of transient

risk on the degree of EN optimism and its positive effects on the real option value of continuation

causes the project value and VC Value to vary non-monotonically with the initial transient risk.

27

Figure 6 (b) shows that the project value and VC value decline with the intrinsic risk, s.5 From

(24), the standard deviation of the evolution of the mean assessment of project quality decreases

with intrinsic risk above a threshold. Hence, the option value of continuing the relationship in any

period also declines. An increase in the intrinsic risk also increases the costs of risk-sharing, which

has a negative effect on the EN’s effort and the VC’s investment. The project value and VC value,

therefore, decline with intrinsic risk.

Figure 6 (c) shows that the VC value significantly increases with the degree of asymmetry in

beliefs, which illustrates the benefits to the VC from exploiting the EN’s optimism by providing

more powerful incentives. The project value, however, initially increases and then decreases with

the degree of asymmetry in beliefs. When the degree of EN optimism is low, the positive effects

of increased optimism on the EN’s effort, the VC’s investment and the project’s output cause the

project value to increase with optimism. When the degree of EN optimism is above a threshold,

however, the VC exploits the EN’s optimism by sub-optimally (from the standpoint of project

value) prolonging the project’s duration and over-investing in the project, thereby increasing her

value at the expense of project value.

6.2.3 Implied Discount Rates

There is considerable empirical and anecdotal evidence that VCs typically use high discount rates

in the range between 35% and 50% to value projects (see Sahlman, 1990, Gladstone and Gladstone,

2002). Sahlman (1990) suggests that high discount rates could be a mechanism that VCs use to

adjust optimistic projections by ENs. To the best of our knowledge, however, it has not been

ascertained whether optimism could indeed explain the high discount rates used by VCs. Do the

levels of EN optimism predicted by our model lead to the discount rates observed in reality?

We calculate the implied discount rate (IDR) of a project as the rate the VC would use to

discount the EN’s projections of the project’s payoffs to equal her own valuation of the project

defined in (23). In other words, the IDR is the discount rate the VC would use to obtain her

valuation of the project if the project’s intrinsic quality were (hypothetically) distributed according5We allow for the discount rate to vary with the intrinsic risk as follows: Rb(s) := r + ((0.15 − r)/s)s, where

r = 0.068 is the risk-free rate reported in Cochrane (2005) and s is the baseline value of s.

28

to the EN’s beliefs. The IDR βV C solves:

E0

[e−βV Cτ (Vτ − Pτ )−

τ−1∑

t=0

e−βV CtctVt

∣∣∣ Θ ∼ N(µEN0 , σ2

0)]

= VC value. (25)

Figure 7 reports the IDR’s for varying values of α, σ0, s, ∆0. The range of IDR’s is consistent with

the discount rates that VC’s use to assess the value of a new venture (see Sahlman, 1990). Our

results, therefore, suggest that entrepreneurial optimism, indeed, explains the discount rates used

by VCs in reality.

7 Conclusions

We develop a continuous-time, stochastic principal-agent model to investigate the effects of asym-

metric beliefs and agency conflicts on the characteristics and valuation of venture capital projects.

We characterize the equilibrium of the stochastic dynamic game in which the VC’s dynamic in-

vestments, the EN’s effort choices, the dynamic compensation contract between the VC and EN,

and the project’s termination time are derived endogenously. Consistent with observed contractual

structures, the equilibrium dynamic contracts feature both equity-like and debt-like components,

the staging of investment by the VC, the progressive vesting of the EN’s stake, and the presence of

inter-temporal milestones or performance targets that must be realized for the project to continue.

We numerically implement the model and calibrate it to aggregate data on VC projects. Our

numerical analysis shows that EN optimism significantly enhances the value that venture capitalists

derive. Entrepreneurial optimism explains the huge discrepancy between the discount rates used

by VCs (∼ 40%), which adjust for optimistic payoff projections by ENs, and the average expected

return of VC projects (∼ 15%). Our results show how the “real option” value of venture capital

investment is affected by the presence of agency conflicts and asymmetric beliefs. Permanent and

transitory components of projects’ risks have differing effects on their values and durations.

29

References

1. Allen, F. and Gale, D. (1999), “Diversity of Opinion and the Financing of New Technologies,”Journal of Financial Intermediation, 8, 68-89.

2. Atkinson, A. 1979. Incentives, Uncertainty and Risk in the Newsboy Problem. DecisionSciences 10 341-353.

3. Baker, M., Ruback, R. and Wurgler, J. (2005), “Behavioral Corporate Finance: A Survey,”Handbook of Corporate Finance: Empirical Corporate Finance (ed: E. Eckbo), Part III,Chapter 5, Elsevier/North-Holland.

4. Berk, J., Green, R. and Naik, V. (2004), “Valuation and Return Dynamics of New Ventures,”The Review of Financial Studies, 17, 1-35.

5. Biais, B., Mariotti, T., Plantin, G., Rochet, J. (2007), “Dynamic Security Design: Conver-gence to Continuous Time and Asset Pricing Implications,” Review of Economic Studies, 74,345-390.

6. Cachon, G.P., M.A. Lariviere. 2001. “Contracting to Assure Supply: How to Share DemandForecasts in a Supply Chain,” Management Sci. 47(5) 629-646.

7. Cadenillas, Abel, Jaksa Cvitanic, and Fernando Zapatero, 2006. “Optimal risk-sharing witheffort and project choice,” Journal of Economic Theory, forthcoming.

8. Casamatta, C. (2003), “Financing and Advising: Optimal Financial Contracts with VentureCapitalists,” Journal of Finance, 58, 2059-2086.

9. Chen, F. 2005. Salesforce Incentives, Market Information, and Production/Inventory Plan-ning. Management Sci. 51(1) 60-75.

10. Chen F., S.A. Zenios. 2005. Introduction to the Special Issue on Incentives and Coordinationin Operations Management. Management Sci. 51(1) 1.

11. Cochrane, J. (2005), “The Risk and Return of Venture Capital,” Journal of Financial Eco-nomics, 75, 3-52.

12. Cornelli, F. and Yosha, O. (2003), “Stage Financing and the Role of Convertible Securities,”Review of Economic Studies, 70, 1-32.

13. Cvitanic, J., Wan. X and Zhang, J. 2005, “Continuous-Time Principal-Agent Problems withHidden Action and Lump-Sum Payment”, Working paper.

14. Davison, A.C. and Hinkley, D.V. (1997). Bootstrap Methods and their Application, CambridgeSeries in Statistical and Probabilistic Mathematics.

15. DeMarzo, P. and Fishman, M. (2007), “Optimal Long-Term Financial Contracting with Pri-vately Observed Cash Flows,” Review of Financial Studies, forthcoming.

16. Dixit, A., R. Pindyck. (1994), Investment Under Uncertainty. Princeton University Press,Princeton, NJ.

17. Egli, D. and Ongena, S. and Smith, D.C. (2006), “On the Sequencing of Projects, ReputationBuilding, and Relationship Finance,” Finance Research Letters, 3, 23-39.

30

18. Gelderen, M. and Thurik, R. and Bosma, N. (2005), “Success and Risk Factors in the Pre-Startup Phase,” Small Business Economics, 24, 365-380.

19. Gibbons, R. 2005. Incentives between Firms (and Within). Management Sci. 51(1) 2-17.

20. Gibbons, R., Murphy, K.J. (1992), “Optimal Incentive Contracts in the Presence of CareerConcerns: Theory and Evidence,” Journal of Political Economy 100(3) 468-505.

21. Gladstone, D. and Gladstone, L. (2002), Venture Capital Handbook: An Entrepreneur’s Guideto Raising Venture Capital, Prentice Hall, Upper Saddle River, New Jersey.

22. Gompers, P. (1995), “Optimal Investment, Monitoring, and the Staging of Venture Capital,”The Journal of Finance, 50, 1461-1489.

23. Gompers, P. and Lerner, J. (1996), “The Use of Covenants: An Empirical Analysis of VenturePartnership Agreements,” Journal of Law and Economics, 39, 463-498.

24. Ha, A.Y. 2001. Supplier-Buyer Contracting: Asymmetric Cost Information and Cutoff LevelPolicy for Buyer Participation. Naval Res. Logist. 48(1) 41-64.

25. Holmstrom, B. (1999), “Managerial Incentive Problems: A Dynamic Perspective,” Review ofEconomic Studies 66, 169-182.

26. Holmstrom, B., Milgrom, P. (1987), “Aggregation and Linearity in the Provision of Intertem-poral Incentive,” Econometrica 55(2) 303-328.

27. Kaplan, S. and Stromberg, P. (2003), “Financial Contracting Theory Meets the Real World:An Empirical Analysis of Venture Capital Contracts,” Review of Economic Studies, 70, 281-315.

28. Kockesen, L. and Ozerturk, S. (2004), “Exclusivity and Over-investment: A Model of Rela-tionship Financing,” Working Paper, Columbia University.

29. Kulatilaka, N., E.C. Perotti. 1998. Strategic Growth Options. Management Sci. 44(8)1021-1031.

30. Lal, R., V. Srinivasan. 1993. Compensation Plans for Single and Multi-Product Salesforces:An Application of the Holmstrom-Milgrom Model. Management Sci. 39(7) 777-793.

31. Morris, S. (1995), “The Common Prior Assumption in Economic Theory,” Economics andPhilosophy, 11, 227-253.

32. Neher, D. (1999), “Staged Financing: An Agency Perspective,” Review of Economic Studies,66, 255-274.

33. Oksendal, B. (2003), Stochastic Differential Equations: An Introduction with Applications,Sixth Edition, Springer-Verlag, Heidelberg, Germany.

34. Plambeck, E.L., S.A. Zenios. 2000. Performance-Based Incentives in a Dynamic Principal-Agent Model. Manufacturing Service Oper. Management 2(3) 240-263.

35. Plambeck, E.L., S.A. Zenios. 2003. Incentive Efficient Control of a Make-to-Stock ProductionSystem. Oper. Res. 51(3) 371-386.

36. Repullo, R. and Suarez, J. (2004), “Venture Capital Finance: A Security Design Approach,”Review of Finance, 8, 75-108.

31

37. Sahlman, W. (1990), “The Structure and Governance of Venture-Capital Organizations,”Journal of Financial Economics, 27, 473-521.

38. Sannikov, Y. (2007), “A Continuous-Time Version of the VC-EN Problem,” Review of Eco-nomic Studies, forthcoming.

39. Schattler, H. and Sung, J. (1993), “The First Order Approach to the Continuous-Time VC-ENProblem with Exponential Utility,” Journal of Economic Theory 61, 331-371.

40. Schmidt, K. (2003), “Convertible Securities and Venture Capital Finance,” Journal of Fi-nance, 58, 1139-1166.

41. Spear, S. and Srivastava, S. (1987), “On Repeated Moral Hazard with Discounting,” Reviewof Economic Studies 54(4), 599-617.

42. Spear, S. and Wang, C. (2005), “When to Fire a CEO: Optimal Termination in DynamicContracts,” Journal of Economic Theory 120(2), 239-256.

43. Sung, J. (1995), “Linearity with Project Selection and Controllable Diffusion Rate in Continuous-Time Principal-Agent Problems,” RAND Journal of Economics 26, 720-743.

44. Trigeorgis, L., 1999, Real Options: Managerial Flexibility and Strategy in Resource Allocation,The MIT Press, Cambridge, MA.

32

Appendix A

Proof of Theorem 1

A rigorous proof of the theorem requires a precise interpretation of equation (1), which describesthe evolution of the termination payoff process. As in the traditional principal-agent literature(see Holmstrom and Milgrom, 1987), we consider the termination payoff process V (·) to be agiven random process on a probability space with investment and effort altering the probabilitydistribution of this process.

The Termination Payoff Process

We consider an underlying probability space (Ω,F) with probability measures Q`, ` ∈ V C, EN,representing the VC’s and EN’s beliefs. Θ is a normal random variable with variance σ2

0 andmean µ`

0 under measure Q` and B is a standard Brownian motion. The complete and augmentedfiltration of the probability space generated by the Brownian motion B(·) is denoted by Ft.Consider the process V (·) = sB(·) where s2 is the intrinsic risk of the project. We will use theGirsanov transformation (see Oksendal, 2003) to obtain new probability measures on (Ω,F) suchthat the process V (·) evolves as in (1).

Suppose that η(·) and c(·) are strictly positive, square-integrable Ft-measurable stochasticprocesses (under the measures QV C and QEN ) defined on the time horizon [0, T ] describing the EN’schoices of effort and the VC’s choices of investments over time. Recall that l(·) is a deterministicprocess describing the operating costs of the firm. Define the processes

ζc,η(t) := exp[∫ t

0(Θ + Ac(u)αη(u)β − l(u))s−1dB(u)− 1

2

∫ t

0(Θ + Ac(u)αη(u)β − l(u))2s−2du] (26)

Bc,η(t) := B(t)−∫ t

0(Θ + Ac(u)αη(u)β − l(u))s−1du . (27)

The process ζc,η(·) is a positive, square-integrable martingale.6 Define the new measure Π`c,η by

dΠ`c,η

dQ`= ζc,η(T ). (28)

By Girsanov’s theorem (see Oksendal, 2003), the process Bc,η(·) is a Brownian motion under the

6The processes are assumed to satisfy the Novikov condition (see Oksendal, 2003):

E` exp[1

2

∫ T

0

(Θ + Ac(u)αη(u)β − l(u))2s−2du] < ∞, ` ∈ V C, EN .

Because the equilibrium investment and effort processes described in Theorem 1 are deterministic and Θ is a normalrandom variable, the Novikov condition is satisfied by these processes.

In fact, we do not need to assume that feasible (not necessarily optimal) investment and effort processes satisfy theNovikov condition for our analysis to be valid; we only require that they be square-integrable. In this case, the processζc,η(·) is only guaranteed to be a local martingale and the measure Π`

c,η is a finite measure, but not necessarily aprobability measure. Our analysis, however, only requires that Π`

c,η be a finite measure. Since, as mentioned earlier,the Novikov condition is satisfied by the equilibrium investment and effort processes, the measure corresponding tothe equilibrium processes is a probability measure.

33

measure Π`c,η. Further, under this measure, the process V (·) evolves as

dV (t) = [Θ + Ac(t)αη(t)β − l(t)]dt + sdBc,η(t) . (29)

Equation (29) describes the evolution of the termination payoff process and is identical to equation(1), but with the Brownian motion and the probability measures representing the VC’s and EN’sbeliefs depending on the investment and effort processes. It is important to keep in mind that V (·)is a fixed process whose sample paths are not affected by investment and effort. Investment andeffort, however, alter the probability distribution of the sample paths of V (·).

For future reference, we make an important observation. The process

dWc,η(t) := s−1[dV (t)− (Ac(t)αη(t)β − l(t))dt− µ`tdt] (30)

is an Ft-Brownian motion with respect to the probability measure Π`c,η. Moreover, the complete

and augmented filtration generated by this Brownian motion is Ft. The EN’s and VC’s meanassessments of project quality Θ at date t, µEN

t , µV Ct are given by (2) and (??).

Utility Related Processes

Let τ ≤ T be an Ft-stopping time denoting the termination time of the VC-EN relationship.Let c(·), η(·) and η(·) be strictly positive Ft-adapted square-integrable processes on [0, τ ].7 Acontract is represented by the quadruple (Pτ , c(·), η(·), τ). We now define some processes that areused frequently in the sequel.

• Cumulative value process of the EN. This is the conditional expected future utility to the EN atany date including the sunk disutilities of prior effort, from a given contract (Pτ , c(·), η(·), τ).Formally,

UP,c,η,τ (t) := EENc,η [− exp(−λ[P (τ)−

∫ τ

0kη(u)γdu]) | Ft] . (31)

Here, E`c,η[· | Ft]; ` ∈ Pr,Ag denotes conditional expectation at date t under the probability

measure Π`c,η defined in (28). For future reference, we note that the cumulative value process

of the EN is a square-integrable Ft-martingale under the measure ΠENc,η .

• Promised payoff process for the EN. The EN’s promised payoff process corresponding to agiven contract (Pτ , c(·), η(·), τ) is defined by (6).

• Adjusted cumulative value process of the EN. This process represents the cumulative valueprocess of the EN from a contract where his effort is η(s); s ≤ t and effort η(s); s ≥ t. Formally,

YP,c,τ (η(·); t; η(·)) := EENc,η [− exp(−λ[P (τ)−

∫ t

0kη(u)γdu−

∫ τ

tkη(u)γdu]) | Ft] . (32)

• EN’s maximum conditional expected utility process. This process represents the EN’s maxi-mum conditional expected utility at date t given that he has exerted effort η(s); s ≤ t and his

7These processes are assumed to satisfy the Novikov condition—see footnote 6.

34

terminal payoff, the VC’s investment process, and the termination time are (Pτ , c(·), and τ),respectively. Formally, we define

ZPτ ,c,η,τ (t) := supη(·)YP,c,τ (η(·); t; η(·)) . (33)

To simplify the subsequent notation, we drop the subscripts denoting the dependence of the pro-cesses defined in (31)-(33) on the contract wherever there is no danger of confusion.

Structure of Incentive Compatible Contracts

A contract (Pτ , c(·), η(·), τ) is incentive compatible with respect to the EN’s effort if and only if,

given the terminal payoff, Pτ , the VC’s investment process, c(·), and the termination time, τ ,

the EN’s optimal effort choices are η(·). The following lemma, from which the theorem follows,

characterizes incentive compatible contracts.

Lemma 1 (Incentive Compatible Contracts)

A contract (Pτ , c(·), η(·), τ) is incentive compatible only if the EN’s promised payoff process P (·)satisfies the following stochastic differential equation:

dP (t) = a(t)dt + b(t)dV (t) (34)

where

b(t) =γk

Aβc(t)αη(t)γ−β (35)

and

a(t) :=λ

2b(t)2s2 + kη∗(t)γ − b(t)

(Ac(t)αη(t)β − l(t) + µEN

t

). (36)

Proof. Define the process η(.) as follows:

η(s) = η(s) for s 6= t, η(t) = η′(t), (37)

where η′(t) is any candidate (possibly sub-optimal) effort choice of the EN at date t. By theprinciple of optimality of dynamic programming (Oksendal, 2003), the effort η(t) is optimal for theEN at date t when his prior effort choices are η(·) only if

η(t) = argmaxη′(t)EENc,η [Z(η(·); t + dt)− Z(η(·); t) | Ft] = argmaxη′(t)Ec,η[dZ(η(·); t) | Ft] . (38)

In what follows, we derive the infinitesimal change dZ(η(·); t) and then use (38) to establish thestatements of the Lemma. It follows from (32), (33), and (37) that

dZ(η(·); t) = dZ(η(·); t) + Z(η(·); t)kλ(η′(t)γ − η(t)γ)dt . (39)

35

Since the process η(·) represents the EN’s optimal effort choices by hypothesis, it follows that

Z(η(·), t) = U(η(·), t), (40)

and hence the process Z(η(·), ·) is a square-integrable Ft-martingale under the measure ΠENc,η .

By the martingale representation theorem (see Oksendal, 2003), there exists a square-integrable,Ft-adapted process ω(·) such that8

dZ(η(·); t) = ω(t)dWc,η(t) = ω(t)s−1[dV (t)− (Ac(t)αη(t)β − l(t))dt− µENt dt] . (41)

Since the expectation in the dynamic programming equation (38) is taken under the measure ΠENc,η ,

it follows from (30) and (41) that Z(η(·); t) evolves under this measure as

dZ(η(·); t) = ω(t)s−1Ac(t)α(η′(t)β − η(t)β)dt + ω(t)dWc,η(t) . (42)

Substituting (42) in (39) yields

dZ(η(·); t) =[ω(t)s−1Ac(t)α(η′(t)β − η(t)β) + kλZ(η(·); t)(η′(t)γ − η(t)γ)

]dt

+ω(t)dWc,η(t) . (43)

Having derived the requisite expression for dZ(η(·); t), we substitute it in (38) to obtain

η(t) = argmaxη′(t)[ω(t)s−1Ac(t)αη′(t)β + kλZ(η(·); t)η′(t)γ ] . (44)

It follows that the effort η(t) is optimal over the interval [t, t + dt] if and only if

ω(t)Z(η(·); t) = − kλs

Ac(t)α

γ

βη(t)γ−β . (45)

From the definition of the promised payoff process in (6), and using (40), we have

P (t) = − logZ(η(·); t)λ

+∫ t

0kη(u)γdu . (46)

Using Ito’s Lemma and (41), we obtain

dP (t) = −ω(t)s−1[dV (t)− (Ac(t)αη(t)β − l(t))dt− µENt dt]

λZ(η(·); t) +ω(t)2

2λZ(η(·); t)2 dt + kη(t)γdt. (47)

Now substituting the expression (45) for the quantity ω(t)Z(η(·);t) in (47) yields

dP (t) = a(t)dt + b(t)dV (t), (48)8Identity (41) is an almost sure relation that holds under all equivalent probability measures on the probability

space. It is only under the measure ΠENc,η defined in (28) that the process [dV (t) − (c(t)αη(t)β − l(t))dt − µEN

t dt] isthe increment of a Brownian motion.

36

where b(t) and a(t) are given by (35) and (36), respectively, as claimed. This completes the proofof the lemma.

Because an incentive compatible contract must have the affine form by Lemma 1, it immediatelyfollows that the optimal contract must also have the same affine form. This completes the proof ofTheorem 1.

Proof of Theorem 2

By (35) and (36), a candidate optimal contract is completely described by the investment processc(·), the processes a(·), b(·) describing the performance-invariant and performance-dependent com-ponents of the EN’s promised payoff process, and the termination time τ . By (35), the contract isincentive compatible if and only if the EN’s effort at date t is

ηt = η(bt, ct) :=(Aβcα

t bt

γk

) 1γ−β

. (49)

DefineMa,b,c,τ (0) = EV C

c,η

[(V (τ)− P (τ)−

∫ τ

0c(s)ds)

](50)

as the VC’s expected future payoff at date 0 if she chooses a contract (Pτ , c(·), η(·), τ) ≡ (a(·), b(·), c(·), τ),where η(.) is given by (49). The VC’s contract choice problem is then the following:

(a∗(·), b∗(·), c∗(·), τ∗) = argmax(a,b,c,τ)Ma,b,c,τ (0). (51)

Let the “state” of the system at any date t be described by the ordered pair (t, µV Ct ). We

first restrict consideration to Markov controls where a(t), b(t), c(t) and the decision to terminatethe relationship only depend on the current state (t, µV C

t ). We derive the optimal Markov controlpolicy. We then appeal to the verification theorem of dynamic programming (see Theorem 11.2.3 ofOksendal, 2003) to conclude that the optimal Markov control policy is, in fact, the optimal controlpolicy over the entire space of admissible Ft-adapted controls.

We note from (35) and (36) that the control a(·) is determined by the controls b(·), c(·) andthe state of the system. Hence, a Markov control policy is completely described by (b(·), c(·), τ).For simplicity, we abuse notation by denoting the VC’s continuation value in state (t, µV C

t ) fromadopting the Markov control policy (b(·), c(·), τ) by

Mb,c,τ (t, µV Ct ) = EV C

t;c,η

[(V (τ)− V (t))− (P (τ)− P (t))−

∫ τ

tc(s)ds

], (52)

where η(·) is determined by (49). Let M∗(t, µV Ct ) be the optimal continuation value within the

space of Markov controls and (b∗(·), c∗(·), τ∗) be the optimal Markov control policy (we derive thispolicy in the following).

Suppose that the VC deviates from the optimal policy over the infinitesimal time interval[t, t+ dt] by choosing the controls (b(t), c(t)), Let M(t, µV C

t ) denote the VC’s continuation value at

37

date t under this deviated policy. It follows from Lemma 1 and (52) that

M(t, µV Ct ) = EV C

t;c,η

[− a(t)dt + (1− b(t))dV (t)− c(t)dt + M∗(t + dt, µV C

t+dt)]

. (53)

By (29), we have

M(t, µV Ct ) = EV C

t;c,η

[−a(t)dt+(1−b(t))[µV C

t +Ac(t)αη(t)β−l(t)]dt−c(t)dt+M∗(t+dt, µV Ct+dt)

], (54)

where η(t) is given by (49) with b(t) replacing b(t). Since the VC’s investment and EN’s effort areobservable, the VC’s assessment µV C

t+dt of project quality at date t+dt is independent of the choicesof controls (b(t), c(t)). Hence, the function M∗(t + dt, µV C

t+dt) is also independent of these choices.By the principle of optimality of dynamic programming (see Oksendal, 2003), the optimal controls(b∗(t), c∗(t)) at date t must maximize the “flow” term in (54), that is,

(b∗(t), c∗(t)) = argmaxb(t),c(t)

− a(t)dt + (1− b(t))[µV C

t + Ac(t)αη(t)β − l(t)]dt− c(t)dt

(55)

By (36) and (49), we can show (after some algebra) that

(b∗(t), c∗(t)) = arg maxb(t),c(t)

Λt(b(t), c(t))dt, (56)

whereΛt(b, c) := (∆tb− 0.5λs2b2 + φ(b)cα γ

γ−β − c + µV Ct − lt) (57)

and

φ(b) := Aγ

γ−β

(1k

) βγ−β

((βb

γ

) βγ−β

(1− βb

γ

)). (58)

We first determine the VC’s optimal investment rate c(b) as a function of the EN’s pay-performance sensitivity b and then simultaneously derive the optimal investment rate and pay-performance sensitivity. By (57) and (58), the optimal investment rate is zero if b ≥ γ/β. Forb ∈ (0, γ/β), Assumption 1 guarantees that the function Λt(b, ·) is strictly concave in the invest-ment rate c (the exponent on c is guaranteed to be less than 1). As a consequence, setting thepartial derivative of Λt(b, ·) with respect to c equal to zero implies that the optimal investment asa function of the pay-performance sensitivity b is given by (13). Substitution in (56), we see thatthe VC chooses the pay-performance sensitivity at date t to solve (14).

By the above arguments, the VC receives a “flow” payoff Λt(b∗t , c∗t )dt in each infinitesimaltime period [t, t + dt]. It immediately follows that the risk-neutral VC chooses to terminate therelationship at the stopping time that solves (16).

The Markov control policy derived above trivially satisfies the conditions of the dynamic pro-gramming verification theorem (see Section 11 of Oksendal, 2003). Hence, it is, in fact, the optimalcontrol policy among the space of all square-integrable Ft-adapted controls. This completes theproof of Theorem 2.

38

Appendix B: Proofs of Remaining Results

For each model parameter “Π” (e.g. σ20, s2, λ, ∆0, k) we let bt(π) denote the solution to (18)

at time t, define ct(π) := c(bt(π)), and let b(π) and c(π) denote the corresponding time pathswhen the parameter Π’s value equals π. We write F ′

t(b, π) when we wish to explicitly indicate thefunctional dependence of the derivative of Ft on the parameter value π. For subsequent reference,the derivative of the VC’s controllable rate function (17) is given by

F ′t(b) = ∆t − pb +

γ − β − αγ

αγc′(b) =

s2

s2 + tσ20

∆0 − pb +γ − β − αγ

αγc′(b). (59)

Proof of Proposition 1.

The marginal optimal investment is given by

c′(b) ∝(1

k

) 1(1−α)

γβ−1 br1(γ − b)r2(1− b) (60)

where

r1 :=2− (1− α) γ

β

(1− α) γβ − 1

and r2 :=α γ

β

(1− α) γβ − 1

,

and where the symbol ∝ means “equal up to a positive multiplicative constant”. Under Assumption2, the parameter r2 is positive and the parameter r1 is negative. Since γ

β > 1 (Assumption 1), thestrong unimodality of c(·) easily follows from (60). Since c(0) = c(γ) = 0 and c′(0) = +∞, it alsofollows from (60) that c(·) achieves its maximum at b = 1. Part (a) has been established.

To establish part (b), we note that the second derivative of the optimal investment function is

c′′(b) ∝ br1−1(γ

β− b)r2−1[r1(

γ

β− b)(1− b)− r2b(1− b)− b(

γ

β− b)].

The expression inside the brackets is a strictly convex quadratic function whose value at 1 is nega-tive, whose value at γ

β > 1 is positive, and whose value at 0 is negative since r1 < 0. Consequently,there is exactly one root bM of the quadratic in the interval (1, γ

β ) such that c′′(bM ) = 0. At bM themarginal investment is at its minimum. Moreover, since c′′(·) is negative on [0, bM ) and is positiveon (bM , γ

β ), the function is strictly concave on [0, bM ] and strictly convex on [bM , γβ ].

Proof of Proposition 2.

Suppose ∆0 ≥ 0. It directly follows from (59) that

F ′t(b) ≤ ∆0 − pb +

γ − β − αγ

αγc′(b),

since ∆t ≤ ∆0 for all t. Figure 1 (p. 16) shows that c′(1) = 0 and c′(b) < 0 for all b ∈ (1, γβ ). It is

straightforward then to check that F ′t(b) < 0 for all b > max∆0

p , 1, which proves the claim when

39

∆0 ≥ 0. Suppose ∆0 < 0 so that ∆t < 0 for all t. Then

F ′t(b) < −pb +

γ − β − αγ

αγc′(b),

By the previous arguments, the right hand side above is less than zero for b ≥ 1 so that b < 1.

Proof of Theorem 3.

a) By (59), for any fixed b > 0, F ′t(b) decreases with t because ∆t = s2

s2+tσ20

∆0 decreases with t for∆0 > 0. By definition, F ′

t(b∗t ) = 0. Hence, F ′

s(b∗t ) < 0 for s > t. Since F ′

s(b∗s) = 0 by definition,

we must have b∗s < b∗t by the strong unimodality of Fs(·). The fact that b∗t → b∗p as t →∞ followsfrom the fact that ∆t → 0 as t →∞ and the Theorem of the Maximum. This establishes part (a).b) The derivative F ′

t(1) = s2

s2+tσ20

∆0 − p is zero when t = t∗ and t∗ ≥ 0. Thus, bt∗ = 1 and ct∗ is atits maximum, which coincides with the “no agency” benchmark case, as required for part (b).c) Since F ′

t(1) > 0 if t < t∗ and F ′t(1) < 0 if t > t∗, part (c) follows from the strong unimodality of

each Ft(·).d) Part (a) establishes that the b∗t decrease with time, and so part (d) follows from part (c) andthe fact that the optimal investment function c(·) increases on [0, 1] and decreases on [1, bM ].e) Since b∗t and c∗t both decrease with time on [t∗,∞), part (e) follows immediately from thefunctional form (8).

Proof of Theorem 4.

a) By (59), for any fixed b > 0, F ′t(b) increases with t because ∆t = s2

s2+tσ20

∆0 increases with t for∆0 < 0. By definition, F ′

t(b∗t ) = 0. Hence, F ′

s(b∗t ) > 0 for s > t. Since F ′

s(b∗s) = 0 by definition,

we must have b∗s > b∗t by the strong unimodality of Fs(·). The fact that b∗t → b∗p as t →∞ followsfrom the fact that ∆t → 0 as t →∞ and the Theorem of the Maximum. This establishes part (a).b) By Proposition 2, 0 < b∗t < 1 when ∆0 < 0. By Proposition 1, the optimal investment functionc(·) is strictly increasing on (0, 1). It follows that, because b∗t increases with t, c∗t = c(b∗t ) alsoincreases. This establishes part (b).c) Since η∗t = η(b∗t , c∗t ), part (c) follows directly from (8) and the fact that both b∗t and c∗t increasewith t.

The following Lemma will be used repeatedly in the proofs to follow.

Lemma 2

If F ′t(b, π) is an increasing (decreasing) function of π, then bt(π) is an increasing (decreasing)

function of π.

Proof. Let π1 < π2. Suppose first that F ′t(b, π) is an increasing function of π. By definition,

0 = F ′t(bt(π2), π2) = F ′

t(bt(π1), π1) < F ′t(bt(π1), π2),

40

which immediately implies bt(π1) < bt(π2) by the strong unimodality of Ft. The proof in thedecreasing case is analogous.

Proof of Theorem 5.

a) If ∆0 < 0, F ′t(b) decreases with λ, decreases with s, and increases with σ0 by (59). By Lemma 2,

b∗t decreases with λ, decreases with s and increases with σ0. Since b∗t ∈ (0, 1) by Proposition 2, andc(·) is increasing in (0, 1) by Proposition 1 that c∗t = c(b∗t ) also decreases with λ, decreases with s

and increases with σ0. This establishes part (a).b) Part (b) follows using similar arguments by observing that F ′

t(b) decreases with λ and decreaseswith σ0.c) Since F ′

t(1, π) = ∆t − p < ∆0 − p < 0, it follows from the strong unimodality of Ft(·, π) thatbt(π) ∈ (0, 1) for all t. Since the function c(·) increases on [0, 1], part (c) follows from part (b).d) We turn to part (d). (Please refer to Figure 4.) Suppose π1 < π2 where π = λ or σ2

0. Firstwe note that by part (a) the path b(π1) lies strictly above the path b(π2). Let t∗j , j = 1, 2, denotethe value of t∗ in (21) corresponding to πj . Clearly, t∗1 > t∗2. By Theorem 3(c), in the interval[0, t∗2) both paths b(π1) and b(π2) lie above one. Since b(π1) > b(π2) and since the function c(·)decreases on [1, bM ], it follows that c(π1) < c(π2) in this interval. Analogously, by Theorem 3(c),in the interval (t∗1,∞) both paths b(π1) and b(π2) lie below one. Since b(π1) > b(π2) and sincethe function c(·) increases on [1, bM ], it follows that c(π1) > c(π2) in this interval. By Theorem3(d), we know that in the interval [t∗2, t

∗1] the path c(π1) increases whereas the path c(π2) decreases.

Moreover, by definition of t∗j we have that ct∗j (πj) = c(1), j = 1, 2, and so

ct∗1(π1) = c(1) > ct∗1(π

2) and ct∗2(π1) < ct∗2(π

2) = c(1).

Thus, the trajectories c(π1) and c(π2) cross exactly once in this interval, as claimed. Part (d) hasbeen established.e) We turn our attention to s2. Since the proofs of parts (c) and (d) when π = λ or σ2

0 are basedon part (b) (and Proposition 1 and Theorem 3), it is sufficient to establish part (b) when π = s2.To this end, we shall prove that for each t and π, ∂F ′t (bt(π),π)

∂π < 0, from which the result will followby direct application of Lemma 1. First, we examine the case when bt(π) ≤ 1. Since

F ′t(bt(π), π) =

∆0π

π + tσ20

− λbt(π)π +γ − β − αγ

αγc′(bt(π)) = 0, (61)

it follows that

π∂F ′

t(bt(π), π)∂π

=∆0tσ

20π

(π + tσ20)2

− λbt(π)π

=∆0tσ

20π

(π + tσ20)2

− [ ∆0π

π + tσ20

+γ − β − αγ

αγc′(bt(π))

]by (61)

= − [ ∆0π2

(π + tσ20)2

+γ − β − αγ

αγc′(bt(π))

]

41

which is less than zero since bt(π) ≤ 1 implies that c′(bt(π)) ≥ 0. (The condition ∆0 ≤ 4p is notrequired in this case.) Now consider the case when bt(π) ≥ 1. Here,

∂F ′t(bt(π), π)

∂π=

∆0tσ20

(π + tσ20)2

− λbt(π) <4λtσ2

(π + tσ20)2

− λ since bt(π) ≥ 1 and ∆0 ≤ 4p = 4λπ

= − λ(π − tσ2

0)2

(π + tσ20)2

≤ 0.

The proof is complete.

Proof of Theorem 6.

The proofs of parts (a) and (b) are identical to the proofs of Theorem 5(a, b). The proof of part (c)follows the same arguments given in the proof of Theorem 5(d), except that here part (a) impliesthat the path b(π1) lies strictly below the path b(π2).

Proof of Proposition 4.

Let

φ(t, µt, τ) := EV Ct

[ ∫ τ

t(F ∗

v − lv + Θ)dv]

(62)

denote the VC’s continuation value function at date t given her current project assessment µt anda given (possibly sub-optimal) stopping time τ . In the above, F ∗

v satisfies (18). The VC’s optimalcontinuation value function is

φ∗(t, µt) := supτ≥t

φ(t, µt, τ), (63)

By standard dynamic programming arguments, the optimal termination time τ∗ (if it exists) mustsolve (63) for any t ∈ [0, T ] and µt ∈ (−∞,∞). Further, the VC continues the project at any date t

and project assessment µt if and only if φ∗(t, µt) > 0. The proof proceeds by showing that φ∗(t, ·) ismonotonic (non-decreasing) and lower semi-continuous. It then follows that at each date t ∈ [0, T ]there exists a trigger µ∗t such that the VC continues the project if and only if µt > µ∗t .

We prove the monotonicity and lower semi-continuity of φ∗(t, ·) by considering the sequence ofdiscrete stopping time problems in which for each fixed positive integer N the VC is constrained toterminate the project only at the discrete set of times 0, T

2N , . . . , (2N−1)T2N , T. We show that the

VC’s optimal value functions φ∗N (t, ·) in the discrete problems are continuous and monotonic. Wethen use a convergence argument to show that φ∗(t, ·) is lower semi-continuous and monotonic.

Pick a positive integer N . We use backward induction to show continuity and monotonicity ofφ∗N (t, ·). To establish continuity we further show there exist positive constants κ1

t , κ2t such that

φ∗N (t, µt) ≤ κ1t + κ2

t maxµt, 0. (64)

42

The assertions of continuity, monotonicity and (64) are trivial at date T since φ∗N (T, ·) ≡ 0. Supposethat the assertion is true for t ∈ [t′ + 1

2N , . . . , T ]. We will establish that the assertion is true fort ∈ [t′, t′ + 1

2N ). Consider first any t ∈ (t′, t′ + 12N ]. By the dynamic programming principle,

φ∗N (t, µt) = Et

[ ∫ t′+ 1

2N

t(F ∗

v − lv + Θ)dv + φ∗N (t′ +1

2N, µt′+ 1

2N)]

. (65)

Since F ∗v is bounded and deterministic by (18), (4), (58) and (13) and Θ is normally distributed,

EV C0

[ ∫ T

0((F ∗

v )2 + (lv)2 + (Θ)2)dv]

< ∞. (66)

We can therefore apply Fubini’s theorem to conclude that

φ∗N (t, µt) =[ ∫ t′+ 1

2N

t(F ∗

v − lv + µt)dv + Etφ∗N (t′ +

12N

, µt′+ 1

2N)]

. (67)

We first establish monotonicity of φ∗N (t, ·). The integral on the right-hand side of (67) obvi-ously increases with µt; it remains to show the expectation on the right-hand side of (67) is alsomonotonic in µt. A bit of algebra applied to (2) and (3) shows that µt′+ 1

2N∼ N(µt, σ

2) is normallydistributed. Further, µt′+ 1

2Nmay be expressed in the form ft(µt, Z) where Z ∼ N(0, 1) and ft(·, ·)

is an increasing, linear function of its arguments. The monotonicity of Etφ∗N (t′+ 1

2N , ·) now followsfrom

Etφ∗N (t′ +

12N

, µt′+ 1

2N) = Eφ∗N (t′ +

12N

, ft(µt, Z)), (68)

since the expectation on the right-hand side of (68) is taken with respect to the standard normaldensity, which is independent of the problem parameters, and since both ft(·, z) and φ∗N (t′ + 1

2N , ·)are monotonic in µt (the latter by the inductive assumption).

Next we show continuity of φ∗N (t, ·). Once again, this property obviously holds for the integralon the right-hand side of (67); it remains to show the expectation on the right-hand side of (67)is also continuous in µt. This result will follow from identity (68) if the limit and expectationoperators may be interchanged, since both ft(·, z) and φ∗N (t′ + 1

2N , ·) are continuous in µt (thelatter by the inductive assumption). By the inductive assumption (64) the function φ∗N (t′ + 1

2N , ·)is bounded above by a positive function whose expectation

E[κ1

t′+ 1

2N+ κ2

t′+ 1

2Nmaxµt′+ 1

2N, 0

]= κ1

t′+ 1

2N+ κ2

t′+ 1

2N σt√

2πe−1/2(

µtσt

)2 + µtP (Z > −µt

σt) (69)

is finite, and thus the interchange is justified by the dominated convergence theorem.To complete the inductive argument we must show that (64) holds for t. The integral on the

right-hand side of (67) is bounded above by (t′ + 12N − t)(F ∗

0 + max(µt, 0))—recall the F ∗t decrease

with t. Since (64) holds for t = t′+ 12N , the inequality (69) shows that the expectation on the right-

hand side of (67) is bounded above by (κ1t′+ 1

2N

+ κ2t′+ 1

2N

σt√2π

) + κ2t′+ 1

2N

max(µt, 0). It is therefore

possible to define positive constants κ1t , κ2

t such that (64) holds for t, as required.

43

Finally, we must establish the inductive step for t = t′. We have

φ∗N (t, µt) = max[0,

[ ∫ t′+ 1

2N

t′(F ∗

v − lv + µt′)dv + Et′φ∗N (t′ +

12N

, µt′+ 1

2N)]]

, (70)

where (70) differs from (67) because the VC can terminate at date t′. It should be clear that theprevious arguments still apply, and hence the inductive step is established.

Because 0, 12N , . . . , (2N−1)T

2N , T ⊂ 0, 12N′ , . . . ,

(2N′−1)T

2N′ , T for all N < N ′, it follows thatφ∗N (t, µt) ≤ φ∗N ′(t, µt). For each (t, µt) ∈ [0, T ]× (−∞,∞) we may therefore define

φ(t, µt) := limN→∞

φ∗N (t, µt) . (71)

We claim that φ = φ∗. Fix (t, µt) ∈ [0, T ] × (−∞,∞). Since φ∗(t, µt) ≥ φ∗N (t, µt) for all N ,φ∗(t, µt) ≥ φ(t, µt). Suppose that φ∗(t, µt) > φ(t, µt). Choose any ε < (φ∗(t, µt) − φ(t, µt))/2.There exists a stopping time τ ε such that φ(t, µt) < φ∗(t, µt)− ε < φ(t, µt, τ

ε) where φ(t, µt, τε) is

defined in (62). Define the stopping time τ εN = i

2N 1 i

2N <τε< i+1

2N . There exists N sufficiently large

such that φ(t, µt, τεN ) > φ(t, µt, τ

ε) − ε. It follows that φ(t, µt) < φ(t, µt, τεN ). By definition of the

function φ∗N (t, µt), however, φ(t, µt, τεN ) ≤ φ∗N (t, µt) ≤ φ(t, µt), which is a contradiction. Hence,

φ(t, µt) = φ∗(t, µt). The monotonicity of φ∗(t, ·) easily follows from the monotonicity of φ∗N (t, ·)and the fact that φ∗(t, µt) = limN→∞ φ∗N (t, µt). The lower semi-continuity of φ∗(t, ·) follows fromthe fact that the supremum of continuous functions is lower semi-continuous.

Proof of Proposition 5.

The controllable rate function Ft(·) (17) is an increasing function of ∆0, which implies that F ∗t

is also a increasing function of ∆0. One may proceed exactly as in the proof of Proposition 4 toestablish that each CVt(·) is a pointwise increasing function of ∆0. Consequently, the trigger valueswill decrease. Since a change in ∆0 has no effect on the sample paths, part (a) follows. The proofof (b) is the same, except that each Ft(·) is now a decreasing function of either λ or k, and thusthe trigger values will increase.

44

On-Line Appendix C

Numerical Implementation of the Model

By Proposition 4, the relationship is terminated at date i if and only if µV Ci is less than a trigger µ∗i .

In our numerical implementation, therefore, we focus on directly modeling the evolution of µV Ci ,

which is hereafter denoted by µi to simplify the notation. In the first stage of our implementation,we approximate the evolution of µi using a discrete lattice and derive the triggers µ∗i that determinewhether the relationship is continued or terminated in each period. In the second stage, given thetriggers obtained from the first step, we use Monte Carlo simulation to model the evolution of µi

and derive the key output variables of interest, e.g., the duration of the relationship, the “rationalexpectations” value of the firm, and the continuation value of the VC.

Lattice design

We use a lattice to approximate the possible evolution of µi, i = 0, 1, 2, . . . , T − 1, which is givenby (3). Gompers (1995) finds that the average time between investments for different investmentstages is 1.09 years. We, therefore, set the length of each period δt to one year. Sahlman (1990)provides empirical evidence there are at most 8 investment stages. We, therefore, choose a finitetime horizon T = 10 in our numerical implementation.

At date 0 the project quality is given by µ0. Let n(i) denote the number of states at date i > 0and let µi,j denote the firm’s quality at the jth state at date i, j = 1, . . . , n(i). The lattice is designedso that the minimal and maximal states at date i, µi,1 and µi,n(i), are κ standard deviations belowand above the minimal and maximal states at date i− 1, respectively. More precisely,

µi,n(i) = µi−1,n(i−1) + κσµi−1 (72)

andµi,1 = µi−1,1 − κσµ

i−1 (73)

In (72) and (73), σµi−1 is the standard deviation of the evolution of the assessment of project quality

over the period (i− 1, i). This can be derived from equation (3) as

σµi−1 =

σ2i−1√

s2 + σ2i−1

. (74)

The values for the remaining n(i)−2 states are equally spaced between the minimum and maximumstates. The number of states in the lattice increases linearly from period to period. That is,n(i) = Mi for i > 0. The value of M is set to 25 and the value of κ is set to 2.5.9

9We found that an increase in M or κ or both did not change the optimal trigger values (to within a 3% tolerance),and that the probability of the firm surviving to the 10th period was zero.

45

The VC’s Continuation Value and the Termination Triggers

Let CVi,j denote the VC’s continuation value at state µi,j . At date T − 1 the continuation value isindependent of the future and is given by CVT−1,j = µT−1,j + F ∗

T−1 − lT−1 for each state µT−1,j ,j = 1, 2, . . . , n(T − 1). At earlier dates the continuation values are given by:

CVi,j = µi,j + F ∗i − li +

n(i+1)∑

k=1

pi+1,ki,j max(CVi+1,k, 0) . (75)

In the above, pi+1,ki,j denotes the probability that the assessment of project quality will transition

from state µi,j at date i to state µi+1,k at date i + 1.10

Starting from the last investment period T and working backwards through time we use dy-namic programming to compute the continuation values for all states and dates. Since the truecontinuation value function is continuous and increasing, we complete the approximation to CVi(·)by linear interpolation. We then determine the optimal trigger µ∗i , which solves CVi(µ∗i ) = 0.

In the second stage, given the termination triggers, µ∗i , we directly model the evolution of(3) using Monte Carlo simulation to compute the various economic statistics of interest. We run50, 000 simulations; the key economic statistics that we derive do not change by more than 1% ifthe number of simulation runs is increased beyond 50, 000.

Estimation of Indirect Parameters

We use the simulated method of moments to estimate the indirect (or the “deep” structural)parameters of the model. Let O denote the vector of the 11 aggregate statistics (see Table 2) weare trying to match.

Phase I: Estimation of Indirect Parameter Values

We simulate a large number N of firms and fix this simulated sample. The number N is chosenlarge enough to minimize simulation errors. The simulation error is negligible when N = 50, 000.For a given candidate indirect parameter vector π, we compute the vector V of simulated values ofthe 11 statistics. Let di(Vi, Oi) = Vi−Oi denote the difference between the simulated and observedvalues of ith statistic and d denote the vector of differences. Define

f(π) := dT Σd. (76)10If µi+1,k is within ±κσµ

i from µi,j , we set

pi+1,ki,j := Φ

[(1

2(µi+1,k + µi+1,k+1)− µi)

1

σµi

]− Φ

[(1

2(µi+1,k + µi+1,k−1)− µi)

1

σµi

],

where Φ(·) denotes the cdf of the standard normal distribution. Otherwise, the transition probability is zero.

46

In (76) the matrix Σ is diagonal. The vector π∗ of parameter estimates solves

π∗ = arg minπ

f(π). (77)

Phase II: Bootstrapping to Determine Confidence Intervals for π∗

We use parametric bootstrapping to determine the confidence intervals for the estimated parameters(see Davison and Hinkley, 1997). This method consists of two steps:

Step 1: Generation of bootstrapped statistics.We fix the parameter vector π = π∗. We generate X samples of NC firms denoted as (S1, . . . ,SX).

We set NC = 7765 to incorporate the fact that the 11 Cochrane statistics are calculated from asample of 7765 firms. We use samples Si to compute the ith vector of the 11 statistics, respec-tively. In this manner, we obtain a set of X vectors of “bootstrapped” Cochrane statistics Vj ,j = 1, 2, . . . , X.

Step 2: Repeat Phase I for each bootstrapped vector of statisticsWe replace the vector O of actual values of the statistics of Phase I with the vector Vj of

bootstrapped values. We solve (76) and (77) to obtain a set of X “bootstrapped” estimates ofthe indirect parameter vector (π∗1, . . . , π

∗X). We use these vectors to obtain standard errors for the

estimated parameters π∗.

Phase III: Computation of Output Statistics and Confidence Intervals

We use π∗ as the baseline vector in our numerical analysis. To compute the confidence intervals forall output statistics of interest when parameters are set to their baseline values, we compute thevalues of the statistics for each of the X bootstrapped samples described above with the parametervector set to the corresponding vector π∗i , i = 1, 2, . . . , X, determined above. In our “comparativestatics” analyses, where one or more parameters are varied while keeping the other parametersat their baseline values, we apply the same percentage changes to the parameter values in thebootstrapped collection π∗i , i = 1, 2, . . . , X, to compute confidence intervals. For example, if weanalyze the effect of varying the degree of asymmetry in beliefs ∆0 by 10% from its baseline valueon some output statistic X, we determine the confidence interval for X by computing its values forthe bootstrapped samples where ∆0 is varied by 10% from its corresponding baseline value for thebootstrapped sample.

47

Table 1: Baseline Parameter ValuesThe table displays the baseline parameter values obtained by the model calibration. The point estimates and their

standard errors are shown in the second and third rows, respectively. The standard errors are obtained by generating

bootstrapped samples using the model as described in Appendix C.

Technology Parameters Belief Parameters Preference Parameters

A α l1 l2 µV C0 ∆0 s σ0 λ k γ

0.6834 0.3212 0.0434 2.1726 0.1128 0.5039 0.7308 0.4401 1.5518 0.0480 4.9614(0.001) (0.005) (0.001) (0.021) (0.002) (0.015) (0.113) (0.083) (0.034) (0.001) (0.077)

Table 2: Predicted and Observed StatisticsThe table displays the observed values of the statistics used to calibrate the model, their predicted values from the

model, and the standard deviations of the statistics obtained by generating bootstrapped samples using the model

as described in Appendix C. The first row shows the round by round returns (RTRR) and standard deviations of

VC projects, the mean overall return (Mean All), standard deviation (Stdev All), and mean duration (Duration All)

of VC projects reported in Cochrane (2005, Table 4). The second row shows the model’s predictions, and the third

shows the corresponding standard errors.

Standard Deviation of RTRR Standard Deviation of RTRR Mean StdDev DurationRd 1 Rd 2 Rd 3 Rd 4 Rd 1 Rd 2 Rd 3 Rd 4 All All

0.26 0.20 0.15 0.09 0.90 0.83 0.77 0.84 0.20 0.71 2.100.26 0.19 0.15 0.09 0.86 0.84 0.82 0.81 0.21 0.71 2.08

(0.008) (0.008) (0.010) (0.020) (0.012) (0.012) (0.013) (0.022) (0.014) (0.006) (0.012)

Table 3: Project Value, VC Value, Duration, and Survival ProbabilitiesThe table shows the Project Value, VC Value, Expected Total Investment (Inv), Expected Duration, and the proba-

bilities of termination in successive round (p∗1, p∗2, p∗3, p∗4) in the Actual Scenario with asymmetric beliefs and agency

conflicts, the Symmetric benchmark scenario in which beliefs are symmetric, and the No Agency benchmark scenario

in which beliefs are risk attitudes are symmetric.

Agency Project VC Duration p∗1 p∗2 p∗3 p∗4Scenario Value Value

Actual 0.954 0.385 2.080 0.250 0.455 0.259 0.035Symmetric 0.960 0.177 1.834 0.401 0.388 0.188 0.023No Agency 1.014 0.288 1.961 0.330 0.411 0.226 0.032

Table 4: Contract Parameter Values (first four periods)The table shows the contractual parameters—the EN’s pay-performance sensitivities and the VC’s investments—inthe first four rounds in the Actual Scenario with asymmetric beliefs and agency conflicts, the Symmetric benchmarkscenario in which beliefs are symmetric, and the No Agency benchmark scenario in which beliefs are risk attitudesare symmetric.

Agency b∗1 b∗2 b∗3 b∗4 c∗1 c∗2 c∗3 c∗4Scenario

Actual 0.477 0.397 0.348 0.315 0.097 0.095 0.093 0.092Symmetric 0.141 0.147 0.150 0.152 0.083 0.083 0.083 0.084No Agency 1.000 1.000 1.000 1.000 0.101 0.101 0.101 0.101

48

Figure 6: Variations of Project Value and VC Value with Transient Risk, Intrinsic Risk, and Degree of Asymmetry in Beliefs

a) Effects of Transient Risk σ0

0

0.2

0.4

0.6

0.8

1

1.2

-100 -80 -60 -40 -20 0 20 40 60 80 100Percent Deviation

V-Project V-VC

b) Effects of Intrinsic Risk

s

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

-100 -80 -60 -40 -20 0 20 40 60 80 100Percent Deviation

V-Project

V-VC

c) Effects of Degree of Asymmetry in Beliefs

Δ0

0

0.2

0.4

0.6

0.8

1

-100 -80 -60 -40 -20 0 20 40 60 80 100Percent Deviation

V-Project V-VC

Figure 7: Variations of Implied Discount Rates (IDRs) with Transient Risk, Intrinsic Risk, and Degree of Asymmetry in Beliefs

Implied Discount Rate

0

0.2

0.4

0.6

0.8

-100 -80 -60 -40 -20 0 20 40 60 80 100Percent Deviation from Baseline Values

sigma s_risk

Delta_0