168
Helsinki University of Technology Systems Analysis Laboratory Research Reports A92, August 2005 PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION Janne Gustafsson Dissertation for the degree of Doctor of Technology to be presented with due permission for public examination and debate in Auditorium E at Helsinki University of Technology, Espoo, Finland, on the 26th of August, at 12 o'clock noon. Helsinki University of Technology Department of Engineering Physics and Mathematics Systems Analysis Laboratory

PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

Helsinki University of Technology Systems Analysis Laboratory Research Reports A92, August 2005 PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION Janne Gustafsson Dissertation for the degree of Doctor of Technology to be presented with due permission for public examination and debate in Auditorium E at Helsinki University of Technology, Espoo, Finland, on the 26th of August, at 12 o'clock noon. Helsinki University of Technology Department of Engineering Physics and Mathematics Systems Analysis Laboratory

Page 2: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

Distribution: Systems Analysis Laboratory Helsinki University of Technology P.O. Box 1100 FIN-02015 HUT, FINLAND Tel. +358-9-451 3056 Fax +358-9-451 3096 [email protected] This report is downloadable at www.sal.hut.fi/Publications/r-index.html ISBN 951-22-7790-5 ISSN 0782-2030 Otamedia Oy Espoo 2005

Page 3: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

Title: Portfolio Optimization Models for Project Valuation Author: Janne Gustafsson Cheyne Capital Management Stornoway House 13 Cleveland Row London SW1A 1DH [email protected] Publication: Systems Analysis Laboratory Research Reports A92, August 2005 Abstract: This dissertation presents (i) a framework for selecting and managing a

portfolio of risky multi-period projects, called Contingent Portfolio Programming (CPP), and (ii) an inverse optimization procedure that uses this framework to compute the value of a single project. The dissertation specifically examines a setting where the investor can invest both in private projects and securities in financial markets, but where the replication of project cash flows with securities is not necessarily possible. This setting is called a mixed asset portfolio selection (MAPS) setting. The valuation procedure is based on the concepts of breakeven selling and buying prices, which are obtained by first solving an optimization problem and then an inverse optimization problem.

In the theoretical part of the dissertation, it is shown that breakeven prices are consistent valuation measures, exhibiting sequential consistency, consistency with contingent claims analysis (CCA), and sequential additivity. Due to consistency with CCA, the present approach can be regarded as a generalization of CCA to incomplete markets. It is also shown that, in some special cases, it is possible to derive simple calculation formulas for breakeven prices which do not require the use of inverse optimization. Further, it is proven that breakeven prices for a mean-variance investor converge towards the prices given by the Capital Asset Pricing Model (CAPM) as the investor’s risk tolerance goes to infinity. The numerical experiments show that CPP is computationally feasible for relatively large portfolios both in terms of projects and states, and illustrate the basic phenomena that can be observed in a MAPS setting.

Keywords: Project valuation, Project portfolio selection, Mixed asset portfolio selection,

Multi-period projects, Ambiguity

Page 4: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION
Page 5: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

Academic dissertation Systems Analysis Laboratory Helsinki University of Technology Portfolio Optimization Models for Project Valuation Author: Janne Gustafsson Supervising professor: Professor Ahti Salo, Helsinki University of Technology Preliminary examiners: Professor Don N. Kleinmuntz, University of Illinois, USA

Professor Harry M. Markowitz, Harry Markowitz Company, USA

Official opponent: Professor James S. Dyer, University of Texas at Austin, USA Publications The dissertation consists of the present summary article and the following papers: [I] Gustafsson, J., A. Salo (2005): Contingent Portfolio Programming for the

Management of Risky Projects. Operations Research (to appear). [II] Gustafsson, J., B. De Reyck, Z. Degraeve, A. Salo (2005): Project Valuation in

Mixed Asset Portfolio Selection. Systems Analysis Laboratory Research Report E16. Helsinki University of Technology. Helsinki, Finland.

[III] Gustafsson, J., A. Salo (2005): Project Valuation under Ambiguity. Systems Analysis Laboratory Research Report E17. Helsinki University of Technology. Helsinki, Finland.

[IV] Gustafsson J., A. Salo (2005): Valuing Risky Projects with Contingent Portfolio Programming. Systems Analysis Laboratory Research Report E18. Helsinki University of Technology. Helsinki, Finland.

[V] Gustafsson, J., A. Salo, T. Gustafsson (2001): PRIME Decisions - An Interactive Tool for Value Tree Analysis, Proceedings of the Fifteenth International Conference on Multiple Criteria Decision Making (MCDM) Ankara, Turkey, July 10-14, 2000, Lecture Notes in Economics and Mathematical Systems, Murat Köksalan and Stanley Zionts (eds.), 507 165–176.

Contributions of the author Janne Gustafsson has been the primary author of all of the papers. He initiated Papers [I], [II], and [IV], whereas the topics for Papers [III] and [V] were suggested by Prof. Salo.

Page 6: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION
Page 7: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

PREFACE This dissertation is a compilation of two of my central works on project portfolio

selection and project valuation, Papers [I] and [II], together with three more or less

related papers. My aim has been to improve my own understanding, as well as that

of practitioners and the general academic, about how to value and select projects

and other non-tradable investments, especially when it is possible to invest in

securities as well. This topic is fundamental to many academic fields, including

corporate finance and management science.

First and foremost, my thanks go to my supervisor Professor Ahti Salo for providing

invaluable support and for suggesting me to work on this topic as a part of my

Master’s Thesis. Since that time, the ideas have evolved significantly further and

become what you see in this dissertation. I would especially like to thank Professors

Bert De Reyck and Zeger Degraeve of London Business School for their support with

one of the papers. London Business School was indeed a truly outstanding research

environment.

I would also like to thank Professor Jim Dyer of University of Texas at Austin whose

review comments on Paper [I] led to a substantial improvement in the quality of the

paper. It is unlikely that I will ever receive such quality feedback on any of my present

or possible future papers. My sincere thanks also go to Professor Harry M. Markowitz

for the time he had for reading and commenting on my manuscripts. The central

ideas of his pioneering model and his exemplary writing style have had much

influence on my research and the way in which I try to write articles.

Last, I am grateful to Dr. Hannu Kahra and Mr. Lauri Pietarinen for creating an

exciting work environment at Monte Paschi Asset Management, where I was

employed during the writing of the first version of the summary article. Likewise, I

would like to thank my present colleagues Petteri Barman and Mikko Syrjänen for

providing an opportunity to work in an outstanding team in London.

London, U.K., June 27, 2005,

Janne Gustafsson

Page 8: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION
Page 9: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

1

1 INTRODUCTION 1.1 Context and Background The evaluation and selection of risky projects, such as research and development

(R&D) projects, has attracted substantial interest among both academicians and

practitioners (see, e.g., Martino 1995, and Henriksen and Treynor 1999). Especially

in high-technology firms, the selection of the R&D portfolio can be a major

determinant of the future performance of the company.

The task of selecting risky projects has been widely studied within several scientific

disciplines, most prominently in corporate finance, operations research, and

management science (see, e.g., Brealey and Myers 2000, Ghasemzadeh et al. 1999,

Heidenberger 1996, Smith and Nau 1995, Gear and Lockett 1973). Many of the

developed approaches aim at placing a value on the project; if this value is positive

the project is started; otherwise it is not. For example, in the literature on corporate

finance, the value of a risky project, like that of any other risky investment, is

calculated as the net present value (NPV) of its cash flows, discounted at a discount

rate that reflects the riskiness of the project (Brealey and Myers 2000). It is typically

suggested that this rate should be the expected rate of return of a security that is

“equal in risk” to the project. Under the Capital Asset Pricing Model (CAPM; Sharpe

1964, Lintner 1965), two assets are regarded equally risky if they have the same

beta, and therefore the finance literature suggests that one should use the beta of the

project, or equivalently, the covariance between the project and the financial market

portfolio, to determine the discount rate for the project (Brealey and Myers 2000).

However, the use of the CAPM in project valuation relies on the assumption that the

firm is a public company maximizing its share price; yet, many companies make

decisions about accepting and rejecting projects before their initial public offering.

Several methods have been proposed to value projects of a private firm. These

include decision analysis (French 1986, Clemen 1996) and options pricing analysis,

which is also referred to as contingent claims analysis (CCA; Merton 1973b, Hull

1999, Dixit and Pindyck 1994, Trigeorgis 1996). Still, conventional methods based on

decision analysis, such as decision trees, may lead to biased estimates of project

value, because they do not take into account the opportunity costs imposed by

alternative investment opportunities. Options pricing analysis accounts for the effect

that financial instruments have on project value, but its applicability is limited,

1

Page 10: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

2

because in practice it may be difficult to replicate project cash flows using financial

instruments.

When used to select a project portfolio, most single-project valuation methods are

problematic in that they neglect the effect of project interactions, such as synergies

and diversification, on the overall performance of the portfolio. Also, these methods

do not take into account the firm’s resource constraints, which may strongly limit the

projects that can simultaneously be included into the portfolio. Therefore, such

methods may lead to a suboptimal project portfolio both in terms of expected cash

flows and the risk of the portfolio. For this reason, it is advisable to employ a project

portfolio selection method instead. Such a method can potentially determine the most

valuable portfolio where project synergies and the effect of diversification on the risk

of the portfolio are taken into account, although it may not directly put a value on any

single project. Still, as shown in this dissertation, these methods can also be applied

to value single assets in the firm’s portfolio through a specific inverse optimization

procedure.

Earlier project portfolio methods have, however, suffered from various shortcomings

that have hindered the use of the methods in practice. For example, many of the

currently available methods, such as the method by Gear and Lockett (1973) and

Heidenberger (1996), make restrictive assumptions about the nature of the investor’s

risk aversion, failing to imply diversification, for instance. Further, some methods do

not consider uncertainty, while others fail to properly account for the multi-period

nature of projects.

1.2 Aims and Practical Relevance This dissertation has two primary aims. First, it aims at developing a framework for

selecting a portfolio of risky multi-period projects which is (i) well-founded in the

theories of finance, management science, and operations research, and also (ii)

practically applicable in the sense that it (a) captures most of the phenomena that are

relevant to R&D portfolio selection and (b) is computationally feasible for portfolio

selection problems of realistic size. Second, the dissertation aims at developing a

procedure for project valuation in a setting where the firm can invest both in private

projects and publicly-traded securities, but where replication of project cash flows

with securities is not necessarily possible. This setting is called the mixed asset

portfolio selection (MAPS) setting.

2

Page 11: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

3

While a framework for selecting a portfolio of risky projects enjoys general interest of

practitioners of corporate finance, also MAPS-based project valuation is important in

practice. On the one hand, MAPS-based project valuation is, in principle, called for

when a corporation or an individual makes investments both in securities and

projects, or other lumpy investment opportunities. For example, many investment

banks invest in a portfolio of publicly traded securities and undertake uncertain one-

time endeavors, such as venture capital investments. On the other hand, a

fundamental problem in the literature on corporate finance is the valuation of a single

project while taking into account the opportunity costs imposed by securities (see,

e.g., Brealey and Myers 2000). This is a MAPS setting that includes one project and

several securities.

1.3 Structure of Dissertation The dissertation includes five papers. Papers [I] and [II] form the core of the

dissertation. Paper [I] presents a modeling framework, called Contingent Portfolio

Programming (CPP), for selecting and managing a portfolio of risky multi-period

projects. The framework is taken into use and extended in Paper [II], which examines

the valuation of risky projects in a MAPS setting. The valuation procedure is based

on the concepts of breakeven selling and buying prices (Luenberger 1998, Smith and

Nau 1995, Raiffa 1968), which rely on the comparison of MAPS problems with and

without the project being valued. Paper [II] shows that breakeven buying and selling

prices exhibit several important properties and that they are therefore consistent

valuation measures. Paper [I] constitutes the primary modeling contribution of the

dissertation, whereas Paper [II] contains the dissertation’s main methodological and

theoretical contribution.

The three other papers provide additional contributions. Papers [III] examines a

single-period MAPS setting where the investor is either unable to give probability

estimates or where the estimates are ambiguous. In particular, we concentrate in this

paper on the Choquet-Expected Utility (CEU) model, which is able to capture

ambiguity, and develop two models to solve MAPS models when the investor is a

CEU maximizer. Paper [IV] discusses multi-period project valuation, compares the

present approach to other multi-period approaches in the literature, and produces

further computational results. Paper [V] describes a decision support system based

on an interval value tree method called Preference Ratios In Multi-attribute

Evaluation (PRIME; Salo and Hämäläinen 2001), and presents a case study where it

is used to develop scenarios for the market share of Sonera SmartTrust, a Finnish

3

Page 12: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

4

high-technology company. A similar approach can be used for scenario generation in

project portfolio selection.

2 EARLIER APPROACHES

2.1 R&D Project Selection Models Several methods for the selection of R&D projects have been developed over the

past few decades (see, e.g., Martino 1995, and Henriksen and Traynor 1999). Many

of these methods are based on mathematical optimization. Such methods are

typically focused on capturing some specific characteristics of R&D portfolio selection

such as synergies or follow-up projects, as described in Table 1. An ideal project

selection framework would implement all of the characteristics in Table 1 in a

theoretically rigorous way. However, few methods aim at capturing all of these

features, and many of those that do resort to theoretically questionable approaches

in modeling some of the features.

Optimization-based R&D project selection methods, such as the ones in Table 1, can

be viewed as extensions of standard capital budgeting models (see, e.g., Luenberger

1998). These models capture rather complex problems with project

interdependencies and resources constraints, but they do not usually address

uncertainties associated with the projects’ outcomes, which makes it impossible to

Table 1. Overview of Approaches to R&D Project Selection.

Features Model RN FP VA CO PV RC RD SY Ghasemzadeh et al. 1999 X X X Heidenberger 1996 X X X X Santhanam & Kyparisis 1996 X X X Czajkowski & Jones 1986 X C Fox et al. 1984 X X Mehrez & Sinuary-Stern 1983 X Aaker & Tyebjee 1978 X X X Gear & Lockett 1973 X X X X Gear et al. 1971 Bell et al. 1967 X X Watters 1967 X C Brandenburg & Stedry 1966 X X

Key: CO = correlation or other probabilistic interaction between project outcomes, FP = follow-up projects, PV = project versions, RC = resource constraints, RD = resource dynamics, RN = reaction to new information, SY = synergies (cross terms for project outcomes), VA = variability aversion, X= feature present in basic model, C = chance-constrained model

4

Page 13: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

5

attach risk measures to project portfolios. Also, these models do not usually offer

possibilities for reacting to new information. In Table 1, the scarcity of X’s in columns

“VA”, “CO”, and “RN” highlights these shortcomings.

Even though some methods do deal with project uncertainties and the investor’s risk

aversion, they often do so by resorting to unrealistically restrictive assumptions or

theoretically unfounded approaches. For example, the method by Mehrez and

Sinuany-Stern (1983) relies on restrictive assumptions about the investor’s utility

function while Czajkowski and Jones (1986) employ chance-constraints that may

lead to preference models that are inconsistent with expected utility theory and other

well-founded preference frameworks. Also, the models of Gear and Lockett (1973)

and Heidenberger (1996) do not account for the variability of portfolio returns, even

though they allow the investor to react to new information.

Another limitation in some optimization models (e.g., Gear and Lockett 1973,

Czajkowski and Jones 1986, and Ghasemzadeh et al. 1999) is that project inputs are

separated from outputs, wherefore projects cannot produce inputs for other projects,

for instance. These models typically also assume that there exists a predefined,

static supply of resources in each time period (see, e.g., Ghasemzadeh et al. 1999

and Gear and Lockett 1973) which makes it impossible to invest profits for later or

immediate use. There is consequently a need for dynamic modeling of resources;

early attempts into this direction have been presented by Brandenburg and Stedry

(see Gear et al. 1971).

In summary, there appears to be need for an optimization method that rigorously

captures (i) project uncertainties, (ii) the investor’s risk preferences, and (iii) dynamic

production and consumption of resources.

2.2 Stochastic Programming

Stochastic programming models analogous to R&D portfolio selection models have

appeared in investment planning as well as in asset-liability management (e.g.,

Bradley and Crane 1972, Kusy and Ziemba 1986, Mulvey and Vladimirou 1989, Birge

and Louveaux 1997, Mulvey et al. 2000). These two problem contexts share

similarities with the selection of R&D projects in that (i) the investor seeks to

maximize the value of a portfolio of risky assets in a multi-periodic setting and (ii)

5

Page 14: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

6

there are several asset categories which parallel the multiple resource types

consumed and produced by R&D projects.

A key difference between R&D portfolio selection models and the financial stochastic

programming models is that in financial optimization, the (dis)investment decisions

are unconstrained quantities that do not restrict the investor’s future decision

opportunities (e.g., security trading). In contrast, R&D project selection involves

“go/no go”-style decisions (Cooper 1993) where the “go”-decision leads to later

project management decisions while the “no go”-decision terminates the project

without offering further decision opportunities. Table 2 contrasts the key

characteristics of selected financial models of stochastic programming to CPP, which

is developed in Paper [I]. Among these, CPP has close parallels with the dynamic

model of Bradley and Crane (1972), as well as the models of Birge and Louveaux

(1997) and Mulvey et al. (2000) which employ state (scenario) trees.

2.3 Project Valuation Methods The literature on corporate finance contains a large number of apparently rivaling

methods for the valuation of risky projects. The most popular approaches include (i)

decision trees (Hespos and Strassman 1965, Raiffa 1968), (ii) expected utility theory

(von Neumann and Morgenstern 1947, Raiffa 1968), (iii) the risk-adjusted NPV

Table 2. Comparison of Some Stochastic Programming Approaches to Portfolio Selection

Bradley and Crane (1972)

Kusy and Ziemba (1986)

Mulvey and Vladimirou (1989)

Birge and Louveaux (1997, §1.2) CPP (Paper [I])

Model type Linear Linear Quadratic or non-linear, network

Linear Linear, mixed integer

Multiple time periods Yes Yes but only two stages

Yes but only two stages Yes Yes

Model of uncertainty State (event) tree

States for second stage; for external cash flows only

States for second stage State tree State tree

Objective Expected value of terminal wealth

Expected discounted net revenues

Mean-variance model or Expected utility of terminal wealth

Expected utility of terminal wealth

Utility of terminal wealth; mean-risk model

Model of risk aversion / risk measure

Loss constraints

Penalty from constraint violations

Variance or power utility function

Piecewise linear utility function

LSAD or EDR

Type of decisions Quantitative (asset trading)

Quantitative (asset trading)

Quantitative (asset trading)

Quantitative (asset trading)

Choices between actions

Decisions influence future decision possibilities

No No No No Yes

6

Page 15: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

7

method (see, e.g., Brealey and Myers 2000), (iv) real options (Dixit and Pindyck

1994, Trigeorgis 1996), (v) Robichek and Myers’ (1966) certainty equivalent method

(see also Brealey and Myers 2000, Chapter 9), (vi) Hillier’s (1963) method, and (vii)

Smith and Nau’s (1995) method. These methods are summarized in Table 3. The

three columns under the heading “Purpose” indicate the purpose to which the

method is intended. As Table 3 describes, decision trees and expected utility theory

are complementary techniques to the other methods, which aim at calculating the

present value of an investment. In a complete project valuation framework, there is a

specific method addressing each of the three purposes.

Table 3. Methods for the valuation of risky multi-period investments.

Purpose Method CE PV ST Formula / explanation

Risk-adjusted NPV X [ ]

1 (1 )

Tt

tt adj

E cNPV I

r=

= − ++∑

Decision tree X A chart with decision and chance nodes

Expected utility theory X [ ] [ ]( )1 ( )CE X u E u X−=

Contingent claims analysis X 0

n

i ii

NPV I S x ∗

=

= − + ∑

Robichek and Myers (1966) X [ ]

1 (1 )

Tt

tt f

CE cNPV I

r=

= − ++∑

Hillier (1963) X 1 (1 )

Tt

tt f

cNPV I CEr=

⎡ ⎤= − + ⎢ ⎥

+⎢ ⎥⎣ ⎦∑

Smith and Nau (1995) X

NPV = breakeven selling or buying price Preference model for cash flow streams: ( ) ( )1 2 1 2[ , ,..., ] , ,...,T TU c c c E u c c c∗= ⎡ ⎤⎣ ⎦

MAPS (Paper [II]) X NPV = breakeven selling or buying price Preference model for terminal wealth levels

Key: CE = Certainty equivalent for a risky alternative, PV = Present value of a risky cash flow stream, ST = Structuring of decision opportunities and uncertainties, I = investment cost, ct = risky cash flow at time t, radj = risk-adjusted discount rate, u = utility function, Si = price of security i, ix ∗ = amount of security i in the replicating portfolio, rf = risk-free interest rate, u* = intertemporal (multi-attribute) utility function.

The MAPS valuation approach developed in Paper [II] is most closely related to the

method by Smith and Nau (1995), the main difference being in the employed

preference model. It is also consistent with contingent claims analysis, and in some

special cases, with Hillier’s (1963) method, as discussed in Papers [II] and [IV]. On

the other hand, the CPP framework, which is developed in Paper [I] and

7

Page 16: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

8

consequently used in Papers [II] and [IV], implements a decision-uncertainty

structure similar to decision trees, and also allows the use of a wide range of

preference models, including the expected utility model. Paper [III] employs a

Choquet-expected utility model, which is an alternative to expected utility theory.

In the following, the methods in Table 3 and their limitations and possible uses are

discussed in more detail.

2.3.1 Decision Trees and Related Approaches A decision tree describes the points at which decisions can be made and the way in

which these points are related to unfolding uncertainties. Conventionally, decision

trees have been utilized together with expected utility theory (EUT; von Neumann

and Morgenstern 1947) so that each end node of the decision tree is associated with

the utility implied by the earlier actions and the uncertainties that have resolved

earlier. This decision tree formulation does not explicitly include the time axis or

provide guidelines for accounting for the time value of money.

In corporate finance, decision trees are used to describe how project management

decisions influence the cash flows of the project (see, e.g., Brealey and Myers 2000,

Chapter 10). Here, decision trees are typically applied together with the risk-adjusted

NPV method, whereby an explicitly defined time axis is also constructed. However,

the selection of an appropriate discount rate for NPV is often problematic, mainly

because the rate is influenced by three confounding factors, (i) the risk of the project,

which depends on the project’s correlation with other investments, (ii) the opportunity

costs imposed by alternative investment opportunities, and (iii) the investor’s risk

preferences.

Several methods for determining the discount rate have been proposed in the

literature. However, most of them have problematic limitations. For example, the

weighted average cost of capital (WACC) is appropriate only for average-risk

investments in a firm, whereas discount rates based on expected utility theory do not

account for the opportunity costs imposed by securities in financial markets. The real

options literature suggests the use of contingent claims analysis (CCA) to derive the

appropriate discount rate by constructing replicating portfolios using market-traded

securities. Still, it may be difficult to construct replicating portfolios for private projects

in practice. Last, the use of a CAPM discount rate is appropriate only for public

companies whose all shares are traded in markets that satisfy the CAPM

8

Page 17: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

9

assumptions. Even when these assumptions are satisfied, a CAPM discount rate is

applicable for single projects only when there are no synergies between projects.

2.3.2 Robichek and Myers’ and Hillier’s Methods Robichek and Myers’ (1966) and Hillier’s (1963) methods are two alternative ways of

determining a risk-adjustment to a discount rate in a multi-period setting. These

methods have been widely discussed in the literature on corporate finance (see, e.g.,

Keeley and Westerfield 1972, Chen and Moore 1982, Ariel 1998, and Brealey and

Myers 2000). They both employ expected utility theory or a similar preference model

to derive a certainty equivalent (CE) for a risky prospect.

In Robichek and Myers’ method, the investor first determines a CE for the cash flow

of each period, and then discounts it back to its present value at the risk-free interest

rate. Yet, because CEs are taken separately for each cash flow, the method does not

account for the effect of cash flows’ temporal correlation on the cash flow stream’s

aggregate risk; hence, it may lead to an unnecessarily large risk-adjustment. In

contrast, in Hillier’s (1963) method, we first determine the cash flow streams that can

be acquired with the project in different scenarios and then calculate the NPVs of

these streams using the risk-free interest rate. The result is a probability distribution

for risk-free-discounted NPV, for which a CE is then determined. However, the use of

the risk-free interest rate essentially means that any money received before the end

of the planning horizon is invested in the risk-free asset. Yet, it might be more

advantageous to invest the funds in risky securities instead. Therefore, Hillier’s

method is, strictly speaking, applicable only in settings, where the investor cannot

invest in risky securities.

2.3.3 Smith and Nau’s Method The idea behind Smith and Nau’s (1995) method, which Smith and Nau call “full

decision tree analysis,” is to explicitly account for security trading in each decision

node of a decision tree. The main advantage of the approach is that it appropriately

accounts for the effect that the possibility to invest in securities has on the discount

rate of a risky project. Incorporating decision trees, a preference model, and security

trading, Smith and Nau’s method is one of the most complete project valuation

methods to date. However, it does not consider alternative projects, which impose an

opportunity cost on the project being valued, wherefore it is applicable only in a

setting where the investor can invest in a single project and several securities. In a

multi-project setting, the method is subject to the usual shortcomings of single-project

9

Page 18: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

10

valuation methods; in particular, it fails to account for the effect of diversification and

project synergies.

Also, the practically appealing form of the method, the integrated rollback procedure,

relies on several restrictive assumptions: (i) additive independence (Keeney and

Raiffa 1976), (ii) constant absolute risk aversion (CARA), and (iii) partial

completeness of markets. Yet, as pointed out by Keeney and Raiffa (1976), additive

independence entails possibly unrealistic preferential restrictions. The CARA

assumption may also be questionable, because it leads to an exponential utility

function with utility bounded from above. This is known to result in an unrealistic

degree of risk aversion at high levels of outcomes (see, e.g., Rabin 2000). In

practice, it may also be difficult to create a replicating portfolio for market-related

cash flows of a project, as it is assumed in partially complete markets.

In view of the limitations of Smith and Nau’s (1995) method, it appears that there is a

need for project selection framework that (i) considers all projects in the portfolio,

implements (ii) decision trees and (iii) security trading, and allows for (iv) a realistic

array of risk preferences without resorting to overly restrictive assumptions.

3 PROJECT PORTFOLIO SELECTION MODEL The Contingent Portfolio Programming (CPP) framework presented in Paper [I] is the

underlying modeling framework that is used throughout this dissertation, except in

Paper [V]. CPP allows risks to be managed both through diversification (Markowitz

1952, 1959) and staged decision making (Cooper 1993), and accounts for the firm’s

resource constraints. The framework has also the advantage that, when the firm’s

risk measure satisfies a linearity property, it leads to linear programming models,

which can readily solved for portfolio selection models of realistic size.

3.1 Framework In CPP, projects are regarded as risky investment opportunities that consume and

produce several resources over multiple time periods. Analogously to Gear and

Lockett (1973) and the Stage-Gate process of Cooper (1993), the staged nature of

R&D projects is captured through project-specific decision trees, which support

managerial flexibility by allowing the investor to take stepwise decisions on each

project in view of most recent information (Dixit and Pindyck 1994, Trigeorgis 1996,

Brandao and Dyer 2004, Brandao, Dyer, and Hahn 2004). Uncertainties are modeled

using a state tree, representing the structure of future states of nature, as shown in

10

Page 19: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

11

the leftmost chart in Figure 1. In general, the state tree is a multinomial tree that can

have different probability distributions in its branches.

0 1 2

Stat

es

50%

30%

70%

40%

60%

50%

15%

35%

20%

30%

ω0

ω1

ω2

ω12

ω21

ω22

ω11

Start? Continue?Yes

No

Yes

No

0 1 2Time

0

Start?Yes

No

2

Continue?Yes

No

Yes

No

1Time

ω11

ω12ω11

ω12ω21

ω22ω21

ω22

ω1

ω2

Time

Figure 1. A state tree, a decision sequence, and a decision tree for a project.

Projects are modeled using decision trees that span over the state tree. The two

rightmost charts in Figure 1 describe how project decisions, when combined with the

state tree, lead to project-specific decision trees. The specific feature of these

decision trees is that the chance nodes are shared by all projects, since they are

generated using the common state tree. This allows for taking into account the

correlations between projects.

Also securities can be straightforwardly incorporated into the CPP framework in order

to develop a MAPS setting. While Paper [I] briefly mentions this possibility, a proper

development of the resulting CPP MAPS model is given in Papers [II] and [IV]. In

such a model, security trading is implemented through state-specific trading

variables, which are similar to the ones used in financial models of stochastic

programming (e.g. Mulvey et al. 2000) and in Smith and Nau’s (1995) method. In

addition to introducing security trading, Paper [II] presents a risk-constrained mean-

risk version of the CPP model, which is not discussed in Paper [I]. Paper [III] employs

a one-period version of this model, formerly called “general deviation-based mean-

risk model” in a previous version of Paper [II], which appears in my Licentiate Thesis

(Gustafsson 2004).

Four types of constraints are imposed on a CPP model: (i) resource (budget)

constraints, (ii) decision consistency constraints, (iii) risk constraints, which apply to

risk-constrained models only, and (iv) deviation constraints. Resource constraints

11

Page 20: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

12

ensure that there is a nonnegative amount of cash in each state. Decision

consistency constraints implement the projects’ decision trees. These constraints

require that (i) at each decision point reached, only one action is selected, and that

(ii) at each decision point that is not reached, no action is taken. Deviation constraints

are needed in the formulation of many deviation-based risk measures.

3.2 Objective Function The preference models in CPP can be further classified into two classes: (1)

preference functional models, such as the expected utility model, and (2) bi-criteria

optimization models. In general, we can refer by a “mean-risk model” either to a

preference functional model (like in Paper [I]) or to a bi-criteria optimization model

that uses optimization constraints (like in Papers [II] – [IV]). I adopt here the latter

terminology.

In a preference functional model, the investor seeks to maximize the utility of the

terminal resource position

max [ ]U X ,

where U is the investor’s preference functional and X is the random value of the

resource position in period T. Under expected utility theory, the preference functional

is given by [ ] [ ]( )=U X E u X , where u is the investor’s von Neumann-Morgenstern

utility function. A risk-constrained mean-risk model can be formulated as follows:

max [ ]E X ,

subject to

[ ]X Rρ ≤ ,

where ρ denotes the risk constraint and R the investor’s risk tolerance.

One of the main differences between Papers [I]–[IV] is the objective function used

within the CPP framework. Paper [I] concentrates on linear preference functionals,

mean-lower semi-absolute deviation (mean-LSAD) model and mean-expected

downside risk (mean-EDR) model, which lead to linear CPP models. Such models

can typically be solved when there is a reasonably large number of states and

projects (e.g., 100 projects and 200 states), as shown by the numerical experiments

in Paper [I]. Lower semi-absolute deviation (LSAD; Ogryczak and Ruszczynski 1999,

Konno and Yamazaki 1991) and expected downside risk (EDR; Eppen et al. 1989)

are defined as

12

Page 21: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

13

LSAD: ( ) ( ) ( )X X

X X X X Xx dF x x dF xµ µ

δ µ µ−∞ −∞

= − = −∫ ∫ and

EDR: ( ) ( ) ( )X X XEDR x dF x x dF xτ τ

τ τ−∞ −∞

= − = −∫ ∫ ,

where Xµ is the mean of random variable X, τ is some constant target value, and FX

is the cumulative density function of X. The mean-LSAD model exhibits linear pricing

and consistency with stochastic dominance (Levy 1992, Ogryczak and Ruszczynski

1999), whereas the mean-EDR model is consistent with expected utility theory (von

Neumann and Morgenstern 1947, Fishburn 1977) and hence also dynamically

consistent (Machina 1989). Also, as long as the preference model and the underlying

random variables satisfy the assumptions of Dyer and Jia’s (1997) relative risk-value

models, in particular the relative risk independence condition and non-negativity of

outcomes, the preference model exhibits also the properties observed with relative

risk-value models, such as the decomposition to the relative risk-value form.

In contrast, Paper [II] focuses on the risk-constrained mean-variance model, because

hereby it is possible to contrast the results with the CAPM, which is based on the

Markowitz (1952) mean-variance model. Similarly, Paper [IV] employs the risk-

constrained mean-variance model in its numerical experiments, although it is not

otherwise limited to this model. The numerical experiments in Paper [IV] involve also

expected utility maximizers exhibiting constant absolute risk aversion (CARA;

Keeney and Raiffa 1976).

Paper [III] uses the Choquet-Expected Utility (CEU) model (Choquet 1953, Gilboa

1987, Schmeidler 1989, Wakker 1990, Camerer and Weber 1992), which under

stochastic dominance reduces to the Rank Dependent Expected Utility (RDEU;

Quiggin 1982, 1993) model. This model is given by

[ ] [ ] [ ] ϕ∞

−∞

′= = = −∫( ) ( ) (1 ( )) ( )c X XU X CEU X E u X u x F x dF x ,

where ϕ is the transformation function describing the investor’s ambiguity aversion.

Paper [III] focuses on two special cases of the transformation function, quadratic and

exponential functions, and presents two alternative formulations of a CEU MAPS

model, the binary variable model and the rank-constrained model. Also, Paper [III]

presents a MAPS model using Wald’s maximin criterion (Wald 1950), which is a

limiting special case of the CEU model.

13

Page 22: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

14

4 PROJECT VALUATION While Paper [I] provides a framework for determining the value of a project portfolio,

the methodology for calculating the value of a single project within a portfolio is

developed in Paper [II]. With a single exception in Section 5.2 of Paper [II], examined

project valuation settings in Papers [II] – [IV] are MAPS settings with at least the risk-

free asset available. That is, the investor is able to invest in both (i) securities, which

can be bought and sold in any quantities, and (ii) projects, which are lumpy all-or-

nothing type investments.

4.1 Breakeven Buying and Selling Prices Since a project is a non-tradable investment opportunity, the value of a project is

defined as the amount of money at present that is equally desirable to the project,

which corresponds to the conceptual definition of NPV in the literature on corporate

finance. Still, the procedure obtained in Paper [II] is quite different from the

conventional NPV formula found in course books on corporate finance (e.g., Brealey

and Myers 2000). Nevertheless, as shown in Paper [IV], the approach coincides with

many of the conventional project valuation approaches when the investor can invest

only in a single project and the risk-free asset.

In a portfolio context, the above definition for project value can be interpreted so that

the investor is indifferent between the following two portfolios: (A1) a portfolio with the

project and (B1) a portfolio without the project and cash equal to the value of the

project. We may alternatively define the value of a project as the indifference

between the following two portfolios: (A2) a portfolio without the project and (B2) a

portfolio with the project and a debt equal to the value of the project. The project

values obtained in these two ways will not, in general, be the same. The first type of

value is called the “breakeven selling price” (BSP), as the portfolio comparison can

be understood as a selling process, and the second type of value the “breakeven

buying price” (BBP).

As discussed in Paper [II], finding a BSP and BBP is an inverse optimization problem

(see, e.g., Ahuja and Orlin 2001): one has to find a change in the budget so that the

optimal value of the second portfolio optimization problem matches the optimal value

of the first problem. Such problems can be classified into two groups: (i) finding an

optimal value for the objective function, and (ii) finding a solution vector. The problem

of finding a BSP or BBP falls within the first class; inverse optimization problems of

this class can be solved by finding a root to a strictly increasing function. To solve

14

Page 23: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

15

such root-finding problems, we can use usual root-finding algorithms, such as the

bisection method and the secant method.

4.2 Theoretical Results 4.2.1 Consistency of Breakeven Prices As shown in Paper [II], breakeven prices exhibit sequential consistency, consistency

with CCA, and sequential additivity. Due to sequential consistency, the investor will

behave rationally in sequential decision problems. Technically, this means that when

an investor first buys a project and then sells it, or vice versa, his/her (sequential)

buying and selling prices will be equal to each other.

Consistency with CCA refers to the property of the breakeven prices that, whenever

CCA is applicable, i.e. whenever there exists a replicating portfolio for the project, the

breakeven buying and selling prices are equal to each other and yield the same

result as CCA (see also Smith and Nau 1995). Due to this property, the breakeven

prices can be regarded as a generalization of CCA.

Finally, sequential additivity states that the (sequential) BSPs / BBPs of two or more

projects will always add up to the BSP / BBP of the portfolio composed of the same

projects. This is a result of the fact that BSPs and BBPs are added values; if valued

non-sequentially, projects’ breakeven prices are non-additive in general.

4.2.2 Equality of Prices and Valuation Formulas Paper [II] also shows that breakeven prices exhibit two important properties for a

broad class of risk-constrained mean-risk investors: (i) the breakeven prices are

equal to each other, and (ii) they can be solved through a pair of optimization

problems without resorting to possibly laborious inverse optimization.

Papers [III] and [IV] produce analogous valuation formulas for other settings. Each of

these formulas implies that, under the specified circumstances, the breakeven prices

will be equal to each other and that they can be solved through a pair of optimization

problems without using inverse optimization. Paper [III] develops the formula for

expected utility maximizers exhibiting CARA and investors using Wald’s maximin

criterion. Both of these preference models have been widely used in the literature.

On the other hand, Paper [IV] develops a formula for BSP when the investor can

invest only in a single project and in the risk-free asset. A similar formula holds for

both BSP and BBP when the investor exhibits CARA; further, it is now possible to

15

Page 24: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

16

use the certainty equivalent operator directly on the project’s future value distribution.

The paper also shows that when the investor exhibits linear pricing the breakeven

prices coincide with the Hillier’s (1963) method. In contrast, the prices will almost

never give the same result as Robichek and Myers’ (1966) method, because this

method typically overestimates the risk of the project.

4.2.3 Relationship to Capital Asset Pricing Model Paper [II] also derives analytical formulas for BSP and BBP in the case where (a) the

investor is a mean-variance optimizer, (b) the optimal mixed asset portfolio at present

is known, and (c) projects are uncorrelated with securities. Using first these formulas

and then generalizing the result, the paper proves that the breakeven prices of a

mean-variance investor will converge, as risk tolerance goes to infinity, towards the

price that the CAPM would place on the project. This result is valid regardless of the

correlation of the project with market securities or other projects, as long as the

optimal project portfolio in the limit is the same with and without the project. If the

portfolios differ, the value of the project will converge towards the CAPM price of the

difference of portfolios with and without the project.

4.3 Valuation of Real Options An interesting extension to the breakeven price methodology is the valuation of

opportunities, especially that of real options. The term “real option” originates from

the fact that management’s flexibility to adapt later decisions to unexpected future

developments shares similarities with financial options (Dixit and Pindyck 1994,

Trigeorgis 1996, Copeland and Antikarov 2001, Black and Scholes 1973, Merton

1973, Hull 1999). For example, possibilities to expand production when the markets

are up, to abandon a project under bad market conditions, and to switch operations

to alternative production facilities can be seen as options embedded in a project.

Typically, the real options literature employs CCA to value real options, which

requires that project cash flows can be replicated using financial securities. When

replication is possible for all assets, markets are said to be complete. However, it can

be difficult to construct replicating portfolios for private projects in practice, especially

when the projects are developing innovative new products that do not resemble

existing market-traded assets. Therefore, it is relevant to examine how real options

could be priced in incomplete markets. Since real options of a project have

conventionally been valued in the presence of securities, which is a MAPS setting, it

is natural to consider the application of the breakeven price methodology to real

option valuation.

16

Page 25: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

17

Because a real option gives the investor an opportunity but not the obligation to take

an action, we need for the valuation of real options concepts that rely on comparing

the situations where the investor can and cannot take an action instead of does and

does not, as breakeven prices do. Such prices are called opportunity selling and

buying prices. Opportunity prices are always non-negative, because an opportunity

cannot lower the value of the investor’s portfolio. It is also straightforward to show

that the opportunity prices can be obtained by taking a maximum of zero and the

respective breakeven price. Since breakeven prices are consistent with CCA, also

opportunity prices have this property, and hence they can be regarded as a

generalization of the standard CCA real option valuation procedure to incomplete

markets.

5 RESULTS FROM NUMERICAL EXPERIMENTS Numerical experiments are conducted in all of the papers. Paper [I] carries out an

extensive numerical study on the computational performance of CPP models. These

experiments indicate that CPP models of realistic size can be solved in a reasonable

time. When solved as linear programming models, where integer variables are left

continuous, CPP models of about a hundred five-staged projects and several

hundreds of states can be solved in a reasonable time. Mixed integer programming

models with a couple of tens of three-staged projects and less than a hundred states

have usually an acceptable solution time. A similar computational experiment is

conducted in Paper [V], indicating that PRIME models can be solved in a relatively

short time using the PRIME Decisions software.

Also Papers [II] – [IV] conduct numerical experiments. These aim at demonstrating

the properties of breakeven prices in different contexts. Because it is not immediately

obvious how generalizable the results are, the conclusions are necessarily limited to

a rather general level. Overall, two main points can be highlighted.

5.1 Effect of Alternative Investment Opportunities The results from the numerical experiments indicate that alternative investment

opportunities, both projects and securities, do have a major impact on the value of a

project. Beginning from the setting where it is possible to invest only in the project

being valued, the experiments in Paper [II] show that introduction of new projects to

the set of available investment opportunities typically lowers project values. This is

understandable in view of the definition of breakeven prices: the value of the optimal

17

Page 26: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

18

portfolio without the project will increase, which lowers the value of the project. Only

in the case where the value of the optimal portfolio with the project increases more

than the portfolio without the project will the value of the project grow. This might be

the case when the project is negatively correlated with other projects.

Similarly, securities can also lower project values by increasing opportunity costs (i.e.

by increasing the value of the portfolio without the project), and they can also raise

them by providing better diversification of risk (i.e. by raising the value of the portfolio

with the project). The total effect depends on the relative magnitude of these two

phenomena. We also see in the numerical experiments that, as predicted by the

related proposition in Paper [II], when a project has a replicating portfolio, its value is

consistent with the project’s CCA value at all risk levels.

5.2 Pricing Behavior as a Function of Risk Tolerance Another set of insights from numerical experiments is related to the behavior of

breakeven prices when the investor’s risk tolerance increases. First, the experiments

show that, when securities are not available, project values can rise non-

monotonically as the risk tolerance is increased. This is because, without securities,

opportunity costs are imposed in a lumpy manner, and the price of a project is

determined by the projects that fit into the portfolio (limited by the risk constraint) with

and without the project. This may result in a lumpy up and down movement in

breakeven prices.

In contrast, when securities are available, project values change monotonically by

risk tolerance, because securities can be bought in a continuous manner. However,

the project values can either rise or decrease by risk tolerance, depending on the

correlation of the project with the rest of the mixed asset portfolio. Indeed, it is the

correlation with the rest of the mixed asset portfolio that determines the limit

behavior, not that with the security portfolio only. Therefore, one cannot use the

project’s beta to make estimates about the sign of change in the project value when

the investor’s risk tolerance increases.

The limit behavior of the breakeven prices is also examined in the numerical

experiments. As predicted by the related propositions in Paper [II], it is observed that

the breakeven prices for a mean-variance investor converge towards the CAPM price

of the project as the investor’s risk tolerance goes to infinity. Similarly, we see in

Paper [III] that when an expected utility maximizer becomes less risk averse, project

18

Page 27: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

19

values approach values that are close to the projects’ CAPM prices. Indeed, in

neither case do the prices converge towards the values given by a risk-neutral

investor. Finally, it is also observed, as expected, that when the investor becomes

increasingly averse to risk or ambiguity, project values approach the values given by

a maximin investor, the investor using Wald’s (1950) maximin criterion.

6 IMPLICATIONS AND FUTURE RESEARCH DIRECTIONS The methodology developed in this dissertation provides a way to calculate

theoretically justifiable values for risky projects and portfolios of risky projects so that

the investor’s risk preferences, the opportunity costs of alternative investment

opportunities, and project interactions are properly accounted for. In particular, it is

now possible to determine the theoretically appropriate discount rate for a risky

project, even when the project is owned by a private company or when there are

synergies between projects. In addition, the methodology makes it possible to solve

project portfolio selection problems where risks are managed using both

diversification and staged decision making. Uncertainties can also be modeled with

relative accuracy, because the models can be formulated using linear programming,

which permits the use of a large number of states.

This dissertation opens up avenues for further work both in practical and theoretical

areas. On the practical side, applications of the methodology are called for. For

example, the analysis of oil field investments appears to be a promising area for the

present methodology, because many oil companies possess portfolios of oil fields

and the associated uncertainties are mostly related to external sources. Another

interesting application area is the valuation of collateralized debt obligations (CDOs),

which are portfolios of assets with a possibly complex tranche structure. In addition,

in order to ease the use of the methodology in practice, the development of a

dedicated software application to solve CPP models is called for.

On the theoretical side, several extensions of the methodology can be made. It is, for

example, possible to include transactions costs and capital gains tax into a CPP

MAPS model by using several resource types, each indicating the number of specific

assets held, and state-specific decision trees that implement the trading decisions for

each asset bought in the associated state. Further work is also needed in the

modeling of synergies and follow-up project structure of R&D projects. For example,

some synergies can be modeled by describing each project as a sequence of tasks,

19

Page 28: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

20

each modeled as an individual project in CPP’s sense, that are interconnected so

that a task cannot be started before the task preceding it is finished. The synergy

then arises from that a single task is used in the task sequence of several projects.

In general terms, the dissertation provides several new research topics for the related

academic disciples. In corporate finance, where the limitations of CAPM discount

rates have hopefully become more apparent than what they were previously, we

need more research to develop procedures to value projects of public companies

when all shareholders are not CAPM investors. This is relevant, for instance, with

partly state-owned companies, because in order to maintain the necessary voting

power the state has to possess a large amount of shares in the company, and

therefore it cannot efficiently diversify its investment portfolio, as the CAPM requires.

A similar situation may arise also in other cases where the investor is interested both

in the financial return and the voting right provided by the share. For researchers of

operations research and management science, the present work shows that decision

trees, real options, and project portfolio selection models can be profitably combined

together into a unified framework, where project values can be determined using

inverse optimization. I am confident that these three project selection methodologies

and inverse optimization have much to offer to each other and in terms of future

research possibilities.

REFERENCES Aaker, D. A., T. T. Tyebjee. 1978. A Model for the Selection of Interdependent R&D

Projects. IEEE Transactions on Engineering Management EM25(2) 30–36.

Ahuja, R. K., J. B. Orlin. 2001. Inverse optimization. Operations Research 49 771–

783.

Ariel, R. 1998. Risk adjusted discount rates and the present value of risky costs. The

Financial Review 33 17–30.

Birge, J. R., F. Louveaux. 1997. Introduction to Stochastic Programming. Springer,

New York, NY.

Black, F., M. S. Scholes. 1973. The pricing of options and corporate liabilities.

Journal of Political Economy 81 637-654.

Bradley, S. P., D. B. Crane. 1972. A Dynamic Model for Bond Portfolio Management.

Management Science 19(2) 139–151.

20

Page 29: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

21

Brandao, L. and Dyer, J. 2004. Decision Analysis and Real Options: A Discrete Time

Approach to Real Option Valuation. Forthcoming in Annals of Operations

Research.

Brandao, L., Dyer, J., and Hahn, W. 2004. Using Decision Analysis to Solve Real

Option Valuation Problems – Part 1: Building a Generalized Approach. Working

Paper. The University of Texas at Austin.

Brealey, R., S. Myers. 2000. Principles of Corporate Finance. McGraw-Hill, New

York, NY.

Camerer, C., M. Weber. 1992. Recent Developments in Modeling Preferences:

Uncertainty and Ambiguity. Journal of Risk and Uncertainty 5 325–370.

Chen, S–N, W. T. Moore. 1982. Investment Decisions under Uncertainty: Application

of Estimation Risk in the Hillier Approach. Journal of Financial and Quantitative

Analysis 17(3) 425–440.

Choquet, G. 1953. Theory of Capacities. Annales de l’Institut Fourier 5 131–295.

Clemen, R. T. 1996. Making Hard Decisions – An Introduction to Decision Analysis.

Duxbury Press, Pacific Grove.

Cooper, R. G. 1993. Winning at New Products: Accelerating the Process from Idea to

Launch. Addison-Wesley, Reading, Mass.

Copeland, T., V. Antikarov. 2001. Real Options: A Practitioner’s Guide. Texere, New

York.

Czajkowski, A. F., S. Jones. 1986. Selecting Interrelated R&D Projects in Space

Technology Planning. IEEE Transactions on Engineering Management EM33(1)

17–24.

Dixit, A. K., R. S. Pindyck. 1994. Investment Under Uncertainty. Princeton University

Press, Princeton.

Dyer, J. S., J. Jia. 1997. Relative risk-value models. European Journal of Operational

Research 103 170–185.

Eppen G. D., Martin R. K., Schrage L. 1989. A Scenario Based Approach to Capacity

Planning. Operations Research 37 517-527.

Fishburn, P. C. 1977. Mean-Risk Analysis with Risk Associated with Below-Target

Returns. Amererican Economic Review 67(2) 116–126.

21

Page 30: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

22

Fox, G. E., N. R. Baker, J. L. Bryant. 1984. Economic Models for R and D Project

Selection in the Presence of Project Interactions. Management Science 30(7)

890–902.

French, S. 1986. Decision Theory – An Introduction to the Mathematics of

Rationality. Ellis Horwood Limited, Chichester.

Gear, A. E., A. G. Lockett, A. W. Pearson. 1971. Analysis of Some Portfolio Selection

Models for R&D. IEEE Transactions on Engineering Management EM18(2) 66–

76.

Gear, A. E., A. G. Lockett. 1973. A Dynamic Model of Some Multistage Aspects of

Research and Development Portfolios. IEEE Transactions on Engineering

Management EM20(1) 22–29.

Ghasemzadeh, F., N. Archer, P. Iyogun. 1999. A zero-one model for project portfolio

selection and scheduling. Journal of Operational Research Society 50(7) 745–

755.

Gilboa, I. 1987. Expected Utility without Purely Subjective Non-Additive Probabilities.

Journal of Mathematical Economics 16 65–88.

Gustafsson, J. 2004. Valuation of Projects and Real Options in Incomplete Markets.

Licentiate Thesis. Helsinki University of Technology.

Heidenberger, K. 1996. Dynamic project selection and funding under risk: A decision

tree based MILP approach. European Journal Operational Research 95(2) 284–

298.

Henriksen, A., A. Traynor. 1999. A Practical R&D Project-Selection Scoring Tool.

IEEE Transactions on Engineering Management 46(2) 158–170.

Hespos, R. F., P. A. Strassmann. 1965. Stochastic Decision Trees for the Analysis of

Investment Decision. Management Science 11(10) 244–259.

Hillier, F. S. 1963. The Deviation of Probabilistic Information for the Evaluation of

Risky Investments. Management Science 9(3) 443–457.

Hull, J. C. 1999. Options, Futures, and Other Derivatives. Prentice Hall, New York,

NY.

Keeley, R. H., R. Westerfield. 1972. A Problem in Probability Distribution Techniques

for Capital Budgeting. Journal of Finance 27(3) 703–709.

Keeney, R. L., H. Raiffa. 1976. Decisions with Multiple Objectives: Preferences and

Value Tradeoffs. John Wiley & Sons, New York, NY.

22

Page 31: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

23

Konno, H., H. Yamazaki. 1991. Mean-Absolute Deviation Portfolio Optimization and

Its Applications to the Tokyo Stock Market. Management Science 37(5) 519–531.

Kusy, M. I., W. T. Ziemba. 1986. A Bank Asset and Liability Management Model.

Operations Research 34(3) 356–376.

Levy, H. 1992. Stochastic Dominance and Expected Utility: Survey and Analysis.

Management Science 38(4) 555–593.

Lintner, J. 1965. The Valuation of Risk Assets and the Selection of Risky Investments

in Stock Portfolios and Capital Budgets. Review of Economics and Statistics

47(1) 13–37.

Luenberger, D. G. 1998. Investment Science. Oxford University Press, New York,

NY.

Machina, M. 1989. Dynamic Consistency and Non-Expected Utility Models of Choice

Under Risk. Journal of Economic Literature 27(4) 1622–1668.

Markowitz, H. M. 1952. Portfolio Selection. Journal of Finance 7(1) 77–91.

–––––. 1959. Portfolio Selection: Efficient Diversification of Investments. Cowles

Foundation, Yale.

Martino, J. P. 1995. Research and Development Project Selection. John Wiley &

Sons, New York, NY.

Mehrez, A., Z. Sinuany-Stern. 1983. Resource Allocation to Interrelated Risky

Projects Using a Multiattribute Utility Function. Management Science 29(4) 430–

439.

Merton, R. C. 1973. Theory of rational option pricing. Bell Journal of Economics and

Management Science 4(1) 141-183.

Mulvey, J. M., H. Vladimirou. 1989. Stochastic Network Optimization Models for

Investment Planning. Annals of Operations Research 20 187–217.

Mulvey, J. M., G. Gould, C. Morgan. 2000. An Asset and Liability Management Model

for Towers Perrin-Tillinghast. Interfaces 30(1) 96–114.

Ogryczak, W., A. Ruszczynski. 1999. From stochastic dominance to mean-risk

models: Semideviations as risk measures. European Journal of Operational

Research 116(1) 33–50.

Quiggin, J. 1982. A Theory of Anticipated Utility. Journal of Economic Behavior and

Organization 3 323–343.

23

Page 32: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

24

––––––. 1993. Generalized Expected Utility Theory: The Rank-Dependent Model.

Kluwer Academic Publishers, Dordrecht, Holland.

Rabin, M. 2000. Risk Aversion and Expected-Utility Theorem: A Calibration Theorem.

Econometrica 68(5) 1281–1292.

Raiffa, H. 1968. Decision Analysis – Introductory Lectures on Choices under

Uncertainty. Addison-Wesley Publishing Company, Reading, MA.

Robichek, A. A., S. C. Myers. 1966. Conceptual Problems in the Use of Risk-

Adjusted Discount Rates. Journal of Finance 21(4) 727–730.

Salo, A, R.P. Hämäläinen. 2001. Preference Ratios in Multiattribute Evaluation

(PRIME) - Elicitation and Decision Procedures under Incomplete Information.

IEEE Transactions on Systems, Man, and Cybernetics - Part A: Systems and

Humans 31(6) 533-–545.

Santhanam, R., G. J. Kyparisis. 1996. A decision model for interdependent

information system project selection. European J. Oper. Res. 89(2) 380–399.

Schmeidler, D. 1989. Subjective Probability and Expected Utility without Additivity.

Econometrica 57 571–587.

Sharpe, W. F. 1964. Capital Asset Prices: A Theory of Market Equilibrium under

Conditions of Risk. Journal of Finance 19(3) 425–442.

–––––. 1970. Portfolio Theory and Capital Markets. McGraw-Hill, New York, NY.

Smith, J. E., R. F. Nau. 1995. Valuing Risky Projects: Option Pricing Theory and

Decision Analysis. Management Science 41(5) 795–816.

Souder, W. E., T. Mandakovic. 1986. R&D project selection models. Res.

Management 29(4) 36–42.

Triantis, A., A. Borison. 2001. Real Options: State of the Practice. Journal of Applied

Corporate Finance 14(2) 8–24.

Trigeorgis, L. 1996. Real Options: Managerial Flexibility and Strategy in Resource

Allocation. MIT Press, Massachussets.

von Neumann, J., O. Morgenstern. 1947. Theory of Games and Economic Behavior,

Second edition. Princeton University Press, Princeton.

Wakker P. 1990. Under Stochastic Dominance Choquet-Expected Utility and

Anticipated Utility Are Identical. Theory and Decision 29 119–132.

Wald, A. 1950. Statistical Decision Functions. John Wiley & Sons, New York.

24

Page 33: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

25

Yu, P.-L. 1985. Multiple-Criteria Decision Making: Concepts, Techniques, and

Extensions. Plenum Press, New York, NY.

25

Page 34: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

26

Page 35: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

Forthcoming in Operations Research

i

!"$#%& %' () * +-,.0/ 213 456,

798;:<:>=?A@CBEDF8HGIBJBJK<:• LNM D9OQP8RSK

Systems Analysis Laboratory, Helsinki University of Technology,

Otakaari 1M, P.O. Box 1100, 02015 HUT, Finland

[email protected], [email protected]

Abstract: Methods for selecting a research and development (R&D) project

portfolio have attracted considerable interest among practitioners and aca-

demics. This notwithstanding, the industrial uptake of these methods has

remained limited, partly due to the difficulties of capturing relevant concerns

in R&D portfolio management. Motivated by these difficulties, we develop

Contingent Portfolio Programming (CPP) which extends earlier approaches in

that it (i) uses states of nature to capture exogenous uncertainties, (ii) mod-

els resources through dynamic state variables, and (iii) provides guidance for

the selection of an optimal project portfolio that is compatible with the deci-

sion maker’s risk attitude. Although CPP is presented here in the context of

R&D project portfolios, it is applicable to a variety of investment problems

where the dynamics and interactions of investment opportunities must be

accounted for.

Keywords: Research and development, project selection; Decision analysis,

theory; Programming, linear, applications

27

Page 36: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

1

T UQV3WQXZY\[$]_^WQ`aYbV

The selection of research and development (R&D) projects has attracted considerable in-

terest in the literatures on technology management and operations research (OR). These

projects involve many characteristics – such as uncertainties and interdependent resource

constraints – that are potentially amenable to analysis by OR techniques. Indeed, there

exists a variety of related methods, ranging from scoring methods such as value trees

(Keeney and Raiffa 1976, French 1986) to optimization models (see, e.g., Gear and Lockett

1973, Heidenberger 1996, Ghasemzadeh et al. 1999) and dynamic programming methods

such as decision trees and real options (French 1986, Dixit and Pindyck 1994, Smith and

Nau 1995, Trigeorgis 1996). Yet, despite the plethora of methodological approaches, these

methods have not enjoyed widespread industrial use, possibly due to the difficulties of

capturing the full range of phenomena that are relevant to the problem of selecting and

managing R&D projects.

Building on the literatures on decision analysis, R&D management, and portfolio optimi-

zation, we develop Contingent Portfolio Programming (CPP) as a modeling framework which

accommodates most of the characteristics that are relevant to the selection of risky pro-

jects. In CPP, projects are regarded as risky investment opportunities that consume and

produce several resources over multiple time periods. The staged nature of R&D projects

is captured through project-specific decision trees (cf. Gear and Lockett 1973) which sup-

port managerial flexibility by allowing the decision maker (DM) to take stepwise decisions

on each project in view of most recent information (Trigeorgis 1996). Uncertainties, on the

other hand, are modeled through a state tree in the spirit of stochastic programming (see,

e.g., Birge and Louveaux 1997).

While CPP permits a wide range of risk attitudes, we focus on a class of objective func-

tions that are a combination of a mean-risk model (Markowitz 1959, 1987) and a multi-

attribute value function (Keeney and Raiffa 1976). In particular, we consider two objective

functions which lead to linear programming models, permitting the solution of relatively

large-scale project portfolios. The first one is a mean-lower semi-absolute deviation model,

28

Page 37: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

2

(mean-LSAD model; Ogryczak and Ruszsynski 1999) which is a special case of generalized

disappointment models with a standard measure of risk (Jia and Dyer’s 1996, Jia et al.

2001). The second is a mean-expected downside risk model (mean-EDR model; Eppen et

al. 1989) which is consistent with expected utility theory (Fishburn 1977).

The rest of the paper is structured as follows. §2 provides a brief overview of earlier ap-

proaches and §3 presents an introductory example. A formal development of CPP is given

in §4, followed by an analysis of computational complexity in §5. §6 concludes the paper

with suggestions for future research directions.

c defXhgi`aj$Xlk)m\m_XZY\e^Anojp

Several methods for the selection of R&D projects have been developed over the past few

decades (for a review, see Martino 1995 and Henriksen and Traynor 1999). These methods

can be categorized into three aggregate groups: (1) scoring models, (2) optimization mod-

els, and (3) dynamic programming models. Among these, the two latter groups are more

relevant to CPP, although some ideas of scoring models (e.g., consideration of multiple at-

tributes) are also included in CPP.

Optimization models for project selection can be viewed as extensions of standard capital

budgeting models (see, e.g., Luenberger 1998). These models capture project interdepend-

encies and resources constraints, but they do not usually address uncertainties associ-

ated with the projects’ outcomes, which makes it impossible to attach risk measures to

project portfolios. There are some approaches based on utility functions, fuzzy set theory,

and chance-constraints, but the resulting models are problematic as they make restrictive

assumptions about the nature of uncertainty or the DM’s risk preferences.

Stochastic optimization models analogous to R&D portfolio selection models have ap-

peared in investment planning as well as asset and liability management (e.g., Birge and

Louveaux 1997, pp. 20–28, and Mulvey et al. 2000). These two problem contexts share

similarities with the selection of R&D projects in that (i) the DM seeks to maximize the

value of a portfolio of risky assets in a multi-periodic setting and (ii) there are several asset

categories which parallel the multiple resource types consumed and produced by R&D

29

Page 38: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

3

projects. A key difference, however, is that in financial optimization the (dis)investment

decisions are unconstrained quantities that do not restrict the DM’s future decision op-

portunities (e.g., security trading). In contrast, R&D project selection involves “go /no go”-

style decisions where the “go”-decision leads to later project management decisions while

the “no go”-decision terminates the project without offering further decision opportunities.

The staged nature of R&D projects has motivated the development of dynamic program-

ming approaches, most notably (1) decision trees based on decision analysis (see French

1986) and (2) real options (see, e.g., Dixit and Pindyck 1994, Trigeorgis 1996); for a com-

parative analysis of these two approaches, we refer to Smith and Nau (1995). Dynamic

programming methods capture the structure of consecutive decisions and uncertainties of

an R&D project, but they do not explicitly address projects interactions or resource con-

straints. In consequence, researchers have sought to link decision trees with portfolio

models (e.g., Heidenberger 1996 and Gear and Lockett 1973); this notwithstanding, the

resulting models have failed to capture many relevant phenomena in R&D portfolio selec-

tion, such as risk aversion and resource dynamics.

The ability to yield theoretically defensible discount rates is often stated as a major advan-

tage of the real options approach over decision trees (Trigeorgis 1996). However, this ap-

proach assumes that the cash flows from the project can be replicated with financial in-

struments for all states of nature, which may be unrealistic if the project results in inno-

vative products that are not similar to market-traded assets. Furthermore, much of the

real options literature employs continuous stochastic processes in the modeling of uncer-

tainties, whereas the uncertainties of R&D projects often relate to discrete events. These

features may make it difficult to use the real options approach in practical project selec-

tion problems.

30

Page 39: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

4

q roXJj$gi`tsu`iV\e$XJvwdAx\e$s"m\gaj

In CPP, the DM makes decisions about which projects are started, when they are started,

what resources are allocated to them, and in what situations they are terminated or ex-

panded, among others. The decisions are subjected to relevant constraints (e.g., availabil-

ity of resources), and they influence the resource flows that are acquired from the project

portfolio.

For each project, the resource flows depend on the future states of nature. The resource

flows associated with a given portfolio management strategy are consequently uncertain,

which means that the final resource position at the end of the planning horizon is risky. It

is assumed that the DM seeks to maximize the utility (or equivalently, certainty equiva-

lent) of her final resource position.

For illustrative purposes, let us assume that the DM can invest in projects A and B in two

phases (see Figure 1). In period zero, she can start either one or both of the projects. If a

project is started, she can make an additional investment in period one, in which case the

project generates a positive cash flow in period two; otherwise, the project is terminated in

which case it yields no further cash flows. Any surplus that is not invested can be depos-

ited at an 8% risk-free interest rate. The initial budget is $9 million.

Uncertainties are captured through a state tree with seven states, of which two are asso-

ciated with period one and four with period two (see Figure 2). Taken together, Figures 1

and 2 correspond to the decision trees in Figures 3 and 4 where the projects’ cash flows

are shown as a function of the indexed action variables X (X = 1, if action is selected, X=0

if it is not). In Figures 5 and 6, these cash flows are shown as a function of action vari-

ables, while Figure 7 shows the cash flows from the entire portfolio. The two projects are

negatively correlated so that that if project A performs poorly, project B yields a high re-

turn, and vice versa.

31

Page 40: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

5

Proj

ects

Start? Continue?Yes

No

Yes

No

Start? Continue?Yes

No

Yes

No

B

A

0 1 2 Time

s1

s2

s11

s12

s21

s22

0 1 2

Stat

es

Time

50%

50%

30%

70%

60%

40%

s0

Figure 1 Decisions for projects A and B Figure 2 States of nature

Start?

YesXASY

NoXASN

s1

Continue?YesXACY1

NoXACN1

YesXACY2

NoXACN2

s2

s11

s12

s21

s22

$20m

$0

$0

$0

$0

$0

$5m

$10m

$0XASY = “Start project A Yes”

XASN = “Start project A No”XACY = “Continue project A Yes”XACN = “Continue project A No”

s11

s12

s21

s22

0 1 2 Time

-$1m

-$3m

-$3m

Start?

YesXBSY

NoXBSN

s1

Continue?YesXBCY1

NoXBCN1

YesXBCY2

NoXBCN2

s2

s11

s12

s21

s22

$2.5m

$0

$0

$0

$0

$0

$25m

$1m

$10XBSY = “ Start project B Yes”

XBSN = “ Start project B No”

XBCY = “ Continue project B Yes”XBCN = “ Continue project B No”

s11

s12

s21

s22

0 1 2 Time

-$2m

-$2m

-$2m

Figure 3 Decision tree of project A Figure 4 Decision tree of project B

Based on Figure 7, resource constraints can now be written as

0921 0 =−+⋅−⋅− sBSYASY RSXX

008.123 1011 =−⋅+⋅−⋅− ssBCYACY RSRSXX

008.123 2022 =−⋅+⋅−⋅− ssBCYACY RSRSXX

008.15.220 11111 =−⋅+⋅+⋅ ssBCYACY RSRSXX

008.1110 12111 =−⋅+⋅+⋅ ssBCYACY RSRSXX

008.1255 21222 =−⋅+⋅+⋅ ssBCYACY RSRSXX

008.110 2222 =−⋅+⋅ ssBCY RSRSX ,

where RS’s are nonnegative real-valued variables that denote the resource surplus in each

state (i.e., resource position).

32

Page 41: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

6

-1·XASY

-3·XACY1

-3·XACY2

5·XACY2

0

20·XACY1

10·XACY1s1

s2

s11

s12

s21

s22

0 1 2

Stat

es

Time

s0 -2·XBSY

-2·XBCY1

-2·XBCY2

25·XBCY2

2.5·XBCY1

1·XBCY1s1

s2

s11

s12

s21

s22

0 1 2

Stat

es

Time

s0

10·XBCY2

Figure 5 Cash flows of project A Figure 6 Cash flows of project B

-1·XASY +-2·XBSY

-3·XACY1 +-2·XBCY1

-3·XACY2 +-2·XBCY2

5·XACY2 +25·XBCY2

20·XACY1 +2.5·XBCY1

10·XACY1 +1·XBCY1

s1

s2

s11

s12

s21

s22

0 1 2

Stat

es

Time

s0

0 +10·XBCY2

Figure 7 Cash flows of the project portfolio

For the measurement of risk, a second set of constraints is developed by introducing de-

viation variables sV −∆ ( sV +∆ ) which indicate by how much the value of the DM’s resource

position in a specific terminal state falls short of (or exceeds, respectively) the expected

value of the resource position at the end of the planning horizon. These constraints are

0=∆+∆−− −+sss VVEVV , where sV denotes the value of the resource position in state s,

EV is the expected value of the resource position over all terminal states, and the sum

s sV V+ −−∆ + ∆ measures by how much sV differs from EV . Since there are no other re-

source types except money, EV is given by

33

Page 42: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

7

11 12 21 2250% 30% 50% 70% 50% 40% 50% 60% ,s s s sEV RS RS RS RS= ⋅ ⋅ + ⋅ ⋅ + ⋅ ⋅ + ⋅ ⋅

while the deviation constraint for state s11 is 0111111 =∆+∆−− −+sss VVEVRS . The con-

straints for other terminal states can be expressed similarly. Because positive realizations

of −∆ sV will be penalized in the objective function by a negative coefficient, only one of the

terms +∆ sV and −∆ sV can be positive.

Because continued investments in period 1 are possible only if the project was initially

started, the following consistency constraints apply (cf. Figures 3 and 4):

1

1

=+=+

BSNBSY

ASNASY

XX

XX

BSYBCNBCY

ASYACNACY

XXX

XXX

=+=+

11

11 BSYBCNBCY

ASYACNACY

XXX

XXX

=+=+

22

22

The action variables X’s are nonnegative integers; in fact, they are binary variables due to

the two leftmost consistency constraints.

The DM seeks to maximize the certainty equivalent of her terminal resource position. It is

assumed that this can be approximated in the mean-risk form by deducting a risk term

based on lower semi-absolute deviation (LSAD) from the expected resource position in pe-

riod 2 (see, e.g., Ogryczak and Ruszsynski 1999). For example, if the risk aversion coeffi-

cient for LSAD has been estimated at y = 0.5, the objective function is

Maximize CE = EV – 0.5 z LSAD =

11 12 21 2250% 30% 50% 70% 50% 40% 50% 60% 0.5s s s sRS RS RS RS⋅ ⋅ + ⋅ ⋅ + ⋅ ⋅ + ⋅ ⋅ − ⋅

11 12 21 2250% 30% 50% 70% 50% 40% 50% 60%s s s sV V V V− − − − |

⋅ ⋅ ∆ + ⋅ ⋅∆ + ⋅ ⋅ ∆ + ⋅ ⋅∆ ~.

With this objective function, the optimal strategy is to start both projects, but to terminate

project A in period 1 if state s2 occurs and project B if state s1 occurs (i.e., decision vari-

ables XASY, XACY1, XACN2, XBSY, XBCN1, and XBCY2 are one while all other X’s are zero). In pe-

riod 2, the corresponding expected resource position is EV = $18.80 million and the LSAD

term is $2.95 million, which leads to the optimal CE value $18.80 – 0.5 × $ 2.95 = $17.33

million. In period 2, the resource position attains its lowest level in state s12 at $13.76 mil-

lion, well above the 1.082 × $9 ≈ $10.50 million obtained by depositing the initial budget at

the risk-free interest rate. Thus, even though both projects entail the risk of losing most of

the initial investment when evaluated in isolation, the optimal strategy yields a return that

surely exceeds the risk-free interest rate.

34

Page 43: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

8

Assuming that the same risk-free interest rate is applied to all states, the present value of

the optimal project portfolio can be readily calculated. That is, by discounting the CE of

the final resource position at the 8 % risk-free interest rate and by deducting the initial

budget of $9 million from this value, the net present value of the portfolio is found to be

$5.85 million. The risk-adjusted discount rate ρ that accounts for both time and risk

preferences can be computed from 22 )1/()1/( frCEEV +=+ ρ , which gives

=−+= 1/)1( CEEVr fρ 1.08 18.80 /17.33 1 12.5%⋅ − ≈ .

;Af;a$Abb$A

The constraints and the objective function of a CPP model are defined by resource types,

the state tree, and project-specific decision trees. We first define these three concepts and

then discuss the constraints and the objective function in CPP.

$ ACN ¡¢

Resources are inputs and outputs that are either consumed or produced by projects. They

can be production factors (e.g., money, equipment), intangibles (e.g., intellectual property

rights), or other relevant assets that the DM may be interested in. A resource type is de-

noted by r and the set of all resource types by R .

$i£ ¤A¥¦¥-)§©¨ª¦¥NF

The time-state model of CPP is a state tree which represents the structure of future states

of nature. Each state prevails during one period within the planning horizon 0,...,T . The

set of states in period t is denoted by tS , and the set of all states is « Tt

tSS0=

= . The time

period of state Ss ∈ is denoted by t(s).

The state tree starts with a single base state 0s in period 0. Each state 1' ,0ts S t T−∈ < ≤ is

followed by at least one state ts S∈ . This relationship is modeled by the function

SSB →: which returns the unique (immediate) predecessor 1' −∈ tSs of state , 0,ts S t∈ >

(by convention, 0 0( )B s s= ). The n-th predecessor of ( )ts S t n∈ ≥ is defined recursively by

1( ) ( ( ))n nB s B B s−= , where 0 ( )B s s= . This function can be used to obtain the states on a

35

Page 44: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

9

path from the base state 0s to state s . These states, together with state s , are contained

in ( ) ' | 0 such that ( ) 'B kS s s S k B s s= ∈ ∃ ≥ = .

States result through uncertain events (e.g., “markets went up”). The probability that state

( 0)ts S t∈ > obtains, subject to the assumption that its predecessor ( )B s has occurred, is

given by the conditional probability ( )( )B sp s . The base state 0s occurs with probability

one, i.e., 0( ) 1p s = . Unconditional probabilities for the other states ( 0)ts S t∈ > are com-

puted recursively from the equation ))(()()( )( sBpspsp sB ⋅= .

¬$­i® ¯°²±>³Z´µ>¶E·

¸¹aº!¹t» ¼½>¾H¿ÁÀ¿ÄÃ!ÅÆÇÃÈ¿ÉÅHÊÀ

The DM takes decisions with regard to projects z ∈ Z . Following Gear and Lockett (1973),

the decision opportunities for each project are structured as a decision tree which consists

of decision points: that is, for each project z ∈ Z , there is a set of decision points zD such

that at decision point zDd ∈ , the DM chooses one of the actions in da A∈ . The decision

point at which action a can be taken is ( )d a . The first decision point of project z is the

base decision point 0zd .

At each decision point d , the DM has information about (i) what state Sds ∈)( prevails at

this point, and (ii) what actions she has taken earlier on with regard to project z , if any;

this information is conveyed by the action that immediately preceeds d. For all decisions

points zd D∈ other than the base decision point, this action, called the parent action of

d , is given by the function ( )ap d . It is assumed that the decision points form a consis-

tent tree so that each decision point has a unique parent action.

For each action a there is an action variable Xa that is equal to the number of times that

this action is selected at decision point ( )d a (e.g., 1 if the action is selected once; 0 if the

action is not selected). Apart from binary choices, action variables can be used to model

decisions that correspond to nonnegative integers or continuous real numbers.

A project management strategy zX is defined by the action variables that are associated

36

Page 45: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

10

with the decision points zd D∈ of project z . A portfolio management strategy X is the

DM’s complete plan of action for all projects z ∈ Z and all states s S∈ .

ËÌaÍ!ÌaÎ Ï<ÐCÑJÒÈÓ<ÔtÕJÐÖH×SÒCØÑ

The project management strategy zX induces a resource flow ( , )rzRF szX of resource type

r in state s . Letting )(sc ra denote the flow of resource type r in state s due to action a ,

this flow is given by

:( ) ( )

( , ) ( )z dB

r rz a a

d D a As d S s

RF s c s X∈ ∈∈

= ⋅Ù Ù

zX ,

where the restriction in the summation of decision points ensures that actions influence

resource flows only in the current state and relevant future states. The aggregate resource

flow ( , )rRF sX in state s is obtained by adding the flows for all projects, i.e.,

:( ) ( )

( , ) ( , ) ( )z dB

r r rz a a

z Z z Z d D a As d S s

RF s RF s c s X∈ ∈ ∈ ∈

= = ⋅Ú Ú Ú Ú

zX X .

For the time being, we assume that resource flows are linear in the action variables, which

means that interactions among actions (e.g., synergies) are not accounted for. In principle,

such interactions can be captured through cross-terms for pairs of action variables. In

this case, the aggregate resource flow becomes

,: :

( ) ( ) ( ) ( )

( , ) ( )z d z dB B

r ra a a a

z Z d D a A z Z d D a As d S s s d S s

RF s c s X X′

′ ′′ ′ ′∈ ∈ ∈ ∈ ∈ ∈

′∈ ∈

= ⋅ ⋅Û Û Û Û Û Û

X .

Ü$ÝaÜ Þ\ßNà!á¢âZã²äNåaàâ-á

The four main constraint types in CPP are (i) decision consistency constraints, (ii) resource

constraints, (iii) optional constraints, and (iv) deviation constraints.

æçèæçté êë>ìHíÁîÂíÄï!ð2ñï<ðCîÇíòîEóFë¢ð>ì-ôñï<ðCîóõtö<íèðó²î

The structure of decision points influences how many actions can be selected at a given

decision point d . For instance, if the parent action of d was selected once, the DM ar-

rives at d and chooses one of the actions in dA ; but if the parent action was not selected,

the DM does not arrive at d so that none of the actions at d can be selected. Thus, at

each decision point other than the base decision point, the number of selected actions is

the same as the number of times that its parent action was selected. At the base decision

point 0zd , the DM usually chooses one of the alternative actions; however, multiple project

37

Page 46: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

11

instantiations (i.e., exact copies) can be modeled by allowing the number of selected ac-

tions to be greater than one, say Lz.

The above requirements imply the following decision consistency constraints

0dz

a za A

X L z∈

= ∀ ∈÷

Z (1)

0( ) \

d

a ap d z za A

X X d D d z∈

= ∀ ∈ ∀ ∈ø

Z . (2)

Unconstrained quantitative decisions, which do not form a decision tree, can be modeled

by omitting the constraints (1) and (2). For example, decisions in securities trading can be

modeled by associating with each state a continuous unconstrained action variable which

is equal to the amount of transactions in that state.

ùúèùúaû ü<ýCþJÿJýÿCþ þ

Resource constraints can be employed to ensure that there is a nonnegative stock of re-

sources in each state. They are modeled through resource surpluses that would remain in

state Ss ∈ , if the DM were to choose the portfolio strategy X . The surplus of resource

type r in state Ss ∈ is

( ) ( )

( ) ( , )

( ) ( , )

r rrs r r r r

B s s B s

b s RF sRS

b s RF s RSα →

+

=

+ + ⋅ X

X

0

0

if

if

ss

ss

≠=

,

where )(sb r is the initial endowment of resource r in state Ss ∈ and rssB →)(α is the rate at

which the surplus in state ( )B s is transferred to s . Initial endowments cannot be influ-

enced by the DM and they do not entail any costs. The transfer rate may depend on the

resource type and the state: it may be equal to (1 + risk-free interest rate) for money, while

the rate for perishable goods may be zero.

The resource surplus variables rsRS are continuous, and for a given portfolio strategy X ,

they can be solved using the following resource constraints

00 0( , ) ( )r r rsRF s RS b s r R− = − ∀ ∈X

( ) ( ) 0( , ) ( ) \ r r r r rB s s B s sRF s RS RS b s s S s r Rα →+ ⋅ − = − ∀ ∈ ∀ ∈X .

Resource surplus variables are usually constrained to non-negative values. However,

negative values may also be permitted in order to allow for the possibility to borrow funds,

for instance.

38

Page 47: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

12

!#"%$'&%%!)(*"+!,(

Optional constraints include any other constraints that may apply. The two most com-

monly discussed optional constraints in the literature are prerequisite constraints, which

define relations between follow-up and prerequisite projects, and project version con-

straints which model alternative versions of a project. Additional examples on optional

constraints are given by Ghasemzadeh et al. (1999). Note that in CPP project versions can

be modeled with a single project in which the choice among the project versions is made

at the base decision point so that no dedicated constraints for this purpose are needed.

-/.10 2436587:9#;=<?>7A@CBEDC9F;=<GED

The DM seeks to maximize the utility of the terminal resource position, viz.

max [ ]U X ,

where U is the DM’s preference functional and X is the value of the resource position in

period T. Under expected utility theory, the preference functional is given by

[ ] [ ]( )=U X E u X , where u is the DM’s von Neumann-Morgenstern utility function.

HIJCIK LMN#O#PQRSQOSTOQO)N#UVOXWZY[)O\^]

Because nonlinear utility functions can entail computational challenges in the context of

large-scale portfolios, we focus on two special cases of the objective function that (i) im-

plement a reasonable model of risk aversion and (ii) lead to a linear programming model.

In both cases, we assume that the DM’s preference functional can be approximated as a

mean-risk model; such models have been widely used in the field of portfolio selection (see

Markowitz 1952, 1959).

We also assume that the value of the final resource position is (i) additive with regard to

resource types and (ii) linear with respect to the amount of surplus of each resource (see,

e.g., Keeney and Raiffa 1976). These two latter assumptions are not overly restrictive as

linear pricing is widely employed in financial modeling. For a given state, the total value of

resource surpluses can be obtained by associating state-dependent weights rsw with each

resource type. These weights can be interpreted as unit prices so that the monetary value

of resource surpluses in state s is given by

39

Page 48: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

13

( ) ,r rs s s

r R

V w RS∈

= ⋅_

sRS

where rsw is the unit price of resource type r in state s ∈ ST. The expected (monetary) value

of the terminal resource position is thus given by

( ) ( ) ( ) ( )T T

r rT s s s

s S s S r R

EV p s V p s w RS∈ ∈ ∈

= ⋅ = ⋅ ⋅` ` `

T sRS RS , (3)

where TRS is a vector of all rsRS ’s for which r ∈ R and s ∈ ST.

abcCbd eCf^gh4ikjmlngpoqj)g

Several dispersion statistics have been proposed in the literature on portfolio selection.

Markowitz (1952, 1959) suggests the use of variance and semivariance for the selection of

securities. Expected downside risk (EDR) has been employed in capacity planning (Eppen

et al. 1989), while absolute deviation has been applied in real-time stock market analysis

(Konno and Yamazaki 1991). These last two measures are linear, which make them at-

tractive for large-scale portfolio selection problems. Also, these measures lead to mean-

risk models that are consistent with the first and second degrees of stochastic dominance

(FSD and SSD; Levy 1992, Ogryczak and Ruszczynski 1999, Fishburn 1977), which sug-

gests that the resulting models are theoretically reasonable.

In particular, we employ expected downside risk (EDR; Eppen et al. 1989, Fishburn 1977)

and lower semi-absolute deviation (LSAD) as measures of risk. Specifically, EDR is given

by

all : all :

[ ] ( ) | | ( )( )x x

x t x t

EDR X p x x t p x t x

< <

= − = −r r

,

where ( )p x is the probability mass function of X and t is the target value from which de-

viations are computed. When the target value is equal to the expression t = µX = E[X], we

have

all : all :

[ ] ( ) | | ( )( )

X X

X Xx x

x x

LSAD X p x x p x x

µ µ

µ µ< <

= − = −s s

.

Both measures can be calculated from deviation constraints as follows. Let +∆ sV and −∆ sV

be nonnegative deviation variables which measure how much the total value of the re-

source surpluses in state s ∈ ST (i.e., sV ) differs from the risk measure’s target value t.

For EDR, these variables satisfy the equations

( ) 0 ,+ −− − ∆ + ∆ = ∀ ∈s s s TV t V V s SsRS (4)

40

Page 49: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

14

where only one of the variables +∆ sV and −∆ sV can be positive, because −∆ sV has a nega-

tive coefficient in the objective function. The EDR of the value of the final resource posi-

tion is given by the sum

( )T

ss S

p s V −

∈⋅ ∆

t. (5)

The LSAD measure can be computed by using t = ( )T TEV RS in (4) instead of a fixed target

value t. This leads to

( ) ( ) 0+ −− − ∆ + ∆ = ∀ ∈s T T s s TV EV V V s SsRS RS , (6)

whereafter the LSAD can be obtained from (5).

uvwCvx yz#+|~C^8yZ)mz

The objective function can now be stated in the mean-risk form

max ( ) ( )T TEV RP −−T TRS V , (7)

where EVT is defined by Equation (3), and ( )T TRP −∆V is given by

( ) ( )T

T ss S

RP p s Vλ− −

= ⋅ ⋅ ∆

T

V , (8)

where deviation variables −∆ sV terms are obtained either from (4) (EDR) or (6) (LSAD).

When the mean-risk model gives a certainty equivalent for a random variable, like the

mean-LSAD model does, RPT can be interpreted as the DM’s risk premium. By substituting

Equations (3) and (8) into (7), the objective function becomes

max ( )T

r rs s s

s S r R

p s w RS Vλ −

∈ ∈

⋅ − ⋅ ∆

.

Importantly, the mean-EDR model falls within the scope of expected utility theory1 (Fish-

burn 1977), and can therefore be regarded as an acceptable model of risk aversion; in par-

ticular, it does not suffer from dynamic inconsistencies (Machina 1989), which may occur

with the mean-LSAD model and other non-expected utility models. On the other hand, the

mean-LSAD model can be motivated by its link to disappointment models and standard

measures of risk (Jia and Dyer 1996, Jia et al. 2001) and the properties of constant abso-

1 Since [ ] [ ] ( )all : all :

( ) ( ) ( )x x

x t x t

E X EDR X p x x t x p x xλ λ< ≥

− ⋅ = − − +

, the mean-EDR model is

equivalent to the utility function (1 ) ,

( ),

x t x tu x

x x tλ λ+ − <

=

≥ .

41

Page 50: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

15

lute risk aversion and constant relative risk aversion (see French 1986): if the value of the

final resource position is subjected to a positive affine transformation, the certainty

equivalent undergoes a similar transformation so that CE[a⋅X + b] = a⋅CE[X] + b for con-

stants a > 0 and b. In expected utility theory, no utility function implies both risk aversion

and such a linear pricing property.

¡ £¢¤¦¥¨§ª©«¬¤ª­E ­

®¯,° ±³²:´µE¶:·¸¹)µ

Although CPP is based on linear programming, the required computational effort may be-

come prohibitive if the model is very large. It is therefore instructive to examine the num-

ber of decision variables and constraints in a CPP model (see Tables 1 and 2). Here, ºZz

zDD∈

= is the set of all decision points and » »Zz Dd

dz

AA∈ ∈

= is the set of all actions.

Because deviation variables and resource surplus variables are continuous, there are at

most A integer variables in a CPP model. The number of integer variables can be reduced

by not constraining one of the action variables at each decision point to integer values.

Still, this variable can assume integer values only, because it is related to the other inte-

ger-valued actions at the same decision point through Equations (1) and (2). The upper

bound for the number of integer variables can thus be reduced to A D− .

Table 1 Number of Decision Variables Decision variable Number Type Action variables (X’s) |A| Typically integer Resource surplus variables (RS’s) |S|⋅|R| Continuous Deviation variables (∆V’s) 2⋅|ST| Continuous TOTAL |A| + |S|⋅|R| + 2⋅|ST|

Table 2 Number of Constraints Constraint Number Decision consistency constraints |D| Resource constraints |S|⋅|R| Deviation constraints |ST| Optional constraints O TOTAL |D| + |S|⋅|R| + |ST| + O

42

Page 51: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

16

For example, a CPP model with three resource types, five time periods, a binary state tree

(where each state is split into two further states in the next period) and thirty projects

with four consecutive “go/no go” decisions leads to 15·30 = 450 integer action variables

and equally many continuous action variables, as well as 31·3 = 93 continuous resource

surplus variables and 2·16 = 32 continuous deviation variables (cf. Table 1). This leads to

a total of 450 integer variables and 575 continuous variables. The number of constraints

is 450 + 31·3 + 16 + 0 = 559. This model can be readily solved using standard techniques

of mixed integer programming (MIP).

In comparison, a conventional decision tree for the same portfolio selection problem would

contain an enormous number of decision and chance nodes. The first decision node of the

tree would entail 230 alternative decisions, one for each possible project portfolio, and each

of these decisions would lead to a binary chance node that resolves at the end of the first

period. At the start of the second period, there would be 230·2 decision nodes, each pre-

ceded by the earlier portfolio decision and the chance outcome. Assuming that these and

ensuing decision nodes entail at most 230 decision alternatives each, we obtain an upper

limit of 1 + 230·2 + (230·2)2 + (230·2)3 ¼ 1028 for the number of decision nodes. Apart from

being intractable, the resulting tree would still require that the sufficiency of resources is

verified at each decision node. On the other hand, building a separate conventional deci-

sion tree for each project would not support the consideration of project interactions or

the variability of portfolio returns. Nevertheless, in this case, each decision tree would

contain 15 decision nodes and 15 chance nodes, resulting in a total of 450 decision nodes

and equally many chance nodes.

A challenge with all state tree based approaches, including CPP, is that the size of the

state tree can become excessively large if the outcome of the project portfolio depends on a

large number of risk factors. This is often the case when the outcome of each project de-

pends on a project-specific risk factor, because the number of states then increases expo-

nentially with the number of projects. For example, in a setting where each project either

fails or succeeds and the success of each project is independent of that of other projects,

the number of corresponding terminal states becomes 2n, where n is the number of pro-

43

Page 52: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

17

jects. With a sufficiently large n, this necessarily leads to a CPP model that cannot be

solved in a reasonable time.

The above situation does not arise if CPP is used to capture external uncertainties, such

as market risks and regulatory changes that are not influenced by the projects. This focus

on external uncertainties is natural in CPP, because the state tree is shared by all projects

and project decisions do not influence state probabilities. Hence, CPP may be particularly

suitable for scenario analysis where portfolio management strategies are optimized with

regard to a relatively small number of states.

½¾À¿ ÁÂZÃÅÄÆÈÇÉÈÇ=ÊÂEË:ÉÍÌÏÎÈÐÑÄCÒEÓÊÔÃÕÒEËÈÇÖ

The computational performance of CPP models was tested with a dedicated C++-

application that runs under Windows XP operation system using LP Solve 4.0.1.9 software

package for LP and MIP models (available at ftp://ftp.es.ele.tue.nl/pub/lp_solve/). The

optimizations were run on a laptop computer with 512 MB of memory and 1.06 GHz Pen-

tium III processor.

×CØÙCØÚ ÛÜÈÝSÞßáàâãÞ)äSåçæè'é%Þ#å,êÝ

In our numerical experiments, the number of projects, stages per project, time periods,

and resources varied as described in Tables 3 and 4. For each model type, 30 models with

randomized resource flows and state probabilities were generated. The timeout for the so-

lution algorithm was set to 20 minutes to ensure that the total computation time re-

mained reasonable even in cases where the median solution time was several minutes and

the worst case solution time could have been several hours.

___________________________________________________________________________________

INSERT TABLES 3 AND 4 AROUND HERE

___________________________________________________________________________________

In most cases, we used two resource types, (i) money with the risk-free interest rate of 5%

and (ii) a perishable capacity resource with a zero transfer rate. Any additional resource

types, if present, were perishable capacity resources. The weight (i.e., unit price) of capac-

ity resources in the objective function was 0 and that of money was 1. In the first period,

the initial budget for money was $2 million multiplied by the number of projects, while

44

Page 53: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

18

initial monetary endowments in other periods were zero. In each period, initial endow-

ments for other resources were equal to one unit of resources multiplied by the number of

projects.

The state tree had a binary structure such that each non-terminal state was split into two

states in the next period. The probabilities for the terminal states were computed by gen-

erating real numbers from the uniform distribution over the unit interval and by normaliz-

ing the resulting numbers. The probabilities of the other states were aggregated from

those of the terminal states.

Project decision trees consisted of binary “go/no go”-decisions in two or more stages. The

first decision was taken in period 0, followed by the next decision in every consecutive pe-

riod up to the total number of stages. Each “go”-decision entailed an immediate cost. Each

cost was obtained by (i) deciding on the most likely value of the cost and (ii) by multiplying

this value with a random number from the (0,1)-lognormal distribution, implying a log-

normal distribution for all costs. The most likely values for costs were assumed to rise

linearly with time so that later stages were more expensive to carry out than earlier ones.

We used the value of $1 million for the first stage, implying a cost of $2 million of the sec-

ond stage, $3 million for the third stage, and so on. Costs for other resources were mod-

eled similarly, using a most likely value of 1 resource unit for the first stage and a linear

growth model.

The revenues began one period after the last “go”-decision. Similarly to costs, revenues

were randomized by first selecting a most likely value for the revenue and then multiply-

ing this number by a random number drawn from the (0,1)-lognormal distribution. For

each project, revenues were distributed evenly over time, in the sense that the most likely

revenue was the same for each period. The cumulative sum of the most likely values for

revenues over the time horizon was assumed to be 1.15 times the sum of most likely val-

ues for costs over the time horizon. For example, for a two-staged project the sum of the

most likely values for costs was $3 million and hence the sum of the most likely values for

revenues was $3.45 million. In a model with four periods, this would imply that the most

likely revenue in the last period and in the second last period was $1.725 million. Projects

45

Page 54: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

19

did not yield inflows for other resources.

In most cases, a mean-LSAD model with the risk aversion coefficient 0.5λ = was em-

ployed. We also conducted experiments with the mean-EDR model with 0.5λ = and the

target value of 1.05Tb ⋅ , where b is the initial budget and T denotes the last time period.

These are indicated by the symbol ¤ in Tables 3 and 4.

×CØÙCØÙ ëÞ)ìêèíå,ì

Some CPP models were solved as MIP models (Table 4) and some as LP models where inte-

ger variables were left continuous (Table 3). The results for LP models suggest that CPP

models for realistic portfolio selection problems can be solved in a reasonable time, and

that relatively few projects involve action variables with non-integer values. For example,

CPP models with 250 3-staged projects, 2 resources, and 5 time periods (16 terminal

states) were solved in a median time of 46.7 s. Models with 50 5-staged projects, 2 re-

sources, and 9 time periods (256 terminal states) had a median solution time of 6 min 59

s. Typically, LP models with less than 6,000 variables and 3,000 constraints could be

solved within the 20-minute time-out limit. In most cases, the number of non-integer ac-

tion variables was small, about 3%–10% of all action variables.

In LP models, the possibility to borrow additional resources and the DM’s risk neutrality

led to (i) a lower number of fractional integer variables and (ii) a higher probability of at-

taining an integer solution (Table 3). In the presence of both assumptions, action variables

always assumed integer values, suggesting that fractional project management decisions

were either due to limited resources or the DM’s risk aversion.

Typically, MIP models were much more time-consuming to solve than LP models (Table 4):

for example, an MIP model with 30 3-staged projects, 2 resources, and 5 time periods

could be solved in a median time of 49.4 s, while the LP model took only 0.32 s to solve.

The mean and standard deviation of the solution time seemed to grow exponentially with

the number of integer variables, wherefore models with more than 350 integer variables

could not usually be solved within the 20-minute time-out limit. However, the possibility

to borrow resulted in significantly shorter solution times, as models with 100 3-staged

46

Page 55: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

20

projects, 1 resource, and 5 time periods were solved in a median time of 3.335 s which is

not much more than the median solution time of 2.835 s for LP models. When both the

possibility to borrow and the assumption of risk neutrality were introduced, there was no

significant difference in the solution times of MIP and LP models.

î ïñðóòôòöõ/÷pøùõúkûýüÿþ ú ð þú

The CPP modeling framework presented in this paper is applicable to the portfolio man-

agement of correlated R&D projects and, more generally, to the analysis of investment

problems where the dynamics and interdependencies of risky investment opportunities

must be accounted for. This framework has several appealing characteristics, such as the

explicit consideration of resource dynamics and managerial flexibility. It also accommo-

dates a wide range of risk attitudes, including two risk averse preference models that lead

to linear CPP models.

Our simulations indicate that LP and MIP formulations for CPP models of realistic size can

be solved in a reasonable time. LP models of about a hundred five-staged projects and

several hundreds of states can be solved in a reasonable time by using a standard per-

sonal computer and a public domain C++ LP package. MIP models, on the other hand,

usually take much more time solve; MIP models with a couple of tens of three-staged pro-

jects and less than a hundred states have usually an acceptable solution time.

There are several avenues for further research. On the theoretical side, CPP needs to be

extended to settings where more complex resource dynamics must be accounted for (e.g.,

cost of storage, proactive management of multi-purpose resources) or where the DM’s ac-

tions influence the structure of the state tree. In terms of future applications, CPP seems

particularly useful in settings where separate decision trees for each project can be devel-

oped, but where the optimal decisions are interlinked by resource constraints and the

need for a portfolio management strategy which accounts for the DM’s risk attitude.

÷ /ú

Birge, J. R., F. Louveaux. 1997. Introduction to Stochastic Programming. Springer, New York, NY.

47

Page 56: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

21

Dixit, A. K., R. S. Pindyck. 1994. Investment Under Uncertainty. Princeton University Press, Prince-

ton, NJ.

Eppen, G. D., R. K. Martin, L. Schrage. 1989. A scenario approach to capacity planning. Opera-

tions Research 37(4) 517–527.

Fishburn, P. C. 1977. Mean-Risk Analysis with Risk Associated with Below-Target Returns. Amer.

Econ. Rev. 67(2) 116–126.

French, S. 1986. Decision Theory – An Introduction to the Mathematics of Rationality. Ellis Horwood

Limited, Chichester.

Gear, A. E., A. G. Lockett. 1973. A Dynamic Model of Some Multistage Aspects of Research and

Development Portfolios, IEEE Trans. Eng. Management EM20(1) 22–29.

Ghasemzadeh, F., N. Archer, P. Iyogun. 1999. A zero-one model for project portfolio selection and

scheduling. J. Oper. Res. Soc. 50(7) 745–755.

Heidenberger, K. 1996. Dynamic project selection and funding under risk: A decision tree based

MILP approach. European J. Oper. Res. 95(2) 284–298.

Henriksen, A., A. Traynor. 1999. A Practical R&D Project-Selection Scoring Tool. IEEE Trans. Eng.

Management 46(2) 158–170.

Jia, J. J. S. Dyer. 1996. A standard measure of risk and risk-value models. Management Science

42(12) 1691–1705.

Jia, J., J. S. Dyer, J. C. Butler. 2001. Generalized Disappointment Models. Journal of Risk and Un-

certainty 22(1) 59–78.

Keeney, R. L., H. Raiffa. 1976. Decisions with Multiple Objectives: Preferences and Value Tradeoffs.

John Wiley & Sons, New York, NY.

Konno, H., H. Yamazaki. 1991. Mean-Absolute Deviation Portfolio Optimization and Its Applica-

tions to the Tokyo Stock Market. Management Sci. 37(5) 519–531.

Levy, H. 1992. Stochastic Dominance and Expected Utility: Survey and Analysis. Management Sci-

ence 38(4) 555–593.

Luenberger, D. G. 1998. Investment Science. Oxford University Press, New York, NY.

Machina, M. 1989. Dynamic Consistency and Non-Expected Utility Models of Choice Under Risk.

Journal of Economic Literature 27(4) 1622–1668.

Markowitz, H. M. 1952. Portfolio Selection. J. Finance 7(1) 77–91.

–––––. 1959. Portfolio Selection: Efficient Diversification of Investments. Cowles Foundation, Yale.

Martino, J. P. 1995. Research and Development Project Selection. John Wiley & Sons, New York,

NY.

Mulvey, J. M., G. Gould, C. Morgan. 2000. An Asset and Liability Management Model for Towers

Perrin-Tillinghast. Interfaces 30(1) 96–114.

48

Page 57: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

22

Ogryczak, W., A. Ruszczynski. 1999. From stochastic dominance to mean-risk models: Semidevia-

tions as risk measures. European J. Oper. Res. 116(1) 33–50.

Smith, J. E., R. F. Nau. 1995. Valuing Risky Projects: Option Pricing Theory and Decision Analy-

sis. Management Sci. 41(5) 795–816.

Trigeorgis, L. 1996. Real Options: Managerial Flexibility and Strategy in Resource Allocation. MIT

Press, Massachussets.

úkþ û ò ú

This research has been supported by the Academy of Finland. We would like to thank area

editor James S. Dyer, an anonymous associate editor, and three anonymous referees for

their helpful comments on earlier versions of this paper. This work has also benefited from

discussions with Victor De Miguel, Jaakko Dietrich, Tommi Gustafsson, Toni Jarimo,

Juha Koljonen, Pertti Laininen, Harry M. Markowitz, and Juha Ojala. We would also like

to thank Kjell Eikland for his invaluable help with LP Solve software package. The finan-

cial support of the Finnish Cultural Foundation, and the Foundation of Jenny and Antti

Wihuri is gratefully acknowledged.

49

Page 58: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

23

Table 3 Solution times for LP CPP models. 30 iterations were performed per setting. Solution time (s) No. of fractional integer variables

Projects

Stages

Period

s

Resou

rces

No. of

variab

les

No. of

constrain

ts

No. of in

teger va

riables

Exp

ectation

Stan

dard

devia

tion

25%

qu

artile

Med

ian

75%

qu

artile

Exp

ectation

Stan

dard

devia

tion

25%

qu

artile

Med

ian

75%

qu

artile

% in

teger solu

tions

20 3 5 2 374 218 140 0.184 0.028 0.161 0.18 0.201 13.67 7.24 8 13 18 3 % 30 3 5 2 514 288 210 0.332 0.064 0.3 0.32 0.35 14.07 6.91 10 14 16 0 % 60 3 5 2 934 498 420 1.208 0.232 0.992 1.152 1.312 15.03 6.57 10 13 19 0 % 100 3 5 2 1494 778 700 4.11 0.671 3.645 4.016 4.516 19.03 7.53 12 19 23 0 % 250 3 5 2 3594 1828 1750 47.541 7.042 43.032 46.738 50.923 23.53 5,75 19 23 26 0 % 100 4 5 2 3094 1578 1500 19.353 5.178 15.723 18.015 19.348 47.23 11.92 36 46 58 0 % 100 3 6 2 1590 858 700 7.272 1.352 6.299 6.9 8.322 24.6 7.89 19 24 29 0 % 100 4 6 2 3190 1658 1500 35.672 5.917 30.884 35.451 39.036 58.97 15.51 46 57 65 0 % 100 5 6 2 6390 3258 3100 285.30 77.21 214.92 278.89 335.79 126.9 34.99 100 120 140 0 % 100 4 9 2 4534 2778 1500 481.58 111.30 374.06 470.59 550.49 65.1 16.8 53 64 69 0 % 25 5 9 2 3084 2053 775 145.95 38.66 117.64 141.68 166.22 123.27 31.25 94 117 145 0 % 50 5 9 2 4634 2828 1550 413.54 100.53 318.94 418.52 452.75 152.53 38.19 115 146 181 0 % 30 4 6 5 1279 797 450 4.821 0.791 4.226 4.667 5.198 60.33 17.34 45 57 72 0 % 100 4 6 5 3379 1847 1500 73.554 25.105 53.196 64.153 77.932 65.8 18.72 49 61 81 0% 20 3 5 1 343 187 140 0.146 0.031 0.13 0.14 0.15 12.93 5.35 10 11 15 0 % 100 3 5 1 1463 747 700 3.386 0.598 2.884 3.195 3.745 19.6 6.16 14 19 21 0 % 100* 3 5 1 1463 747 700 3.167 1.017 2.674 2.835 3.004 6.47 4.64 2 7 10 20 % 100# 3 5 1 1463 747 700 2.869 0.402 2.574 2.804 3.145 11.6 3.46 9 11 13 0 % 100*# 3 5 1 1463 747 700 2.45 0.493 2.263 2.323 2.423 0 0 0 0 0 100 % 1000*# 3 5 1 14063 7047 7000 618.18 268.27 323.91 605.19 827.79 0 0 0 0 0 100 % 100¤ 3 5 1 1463 747 700 3.047 0.401 2.704 3.004 3.194 11.87 3.9 9 12 13 0 % 100¤ 3 5 2 1494 778 700 3.581 0.437 3.255 3.545 3.936 12.7 3.45 10 12 14 0 % *: borrowing is allowed, #: risk neutrality, ¤: mean-EDR model.

50

Page 59: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

24

Table 4 Solution times for MIP CPP models. 30 iterations were performed per setting. Solution time (s)

Projects

Stages

Period

s

Resou

rces

No. of

variab

les

No. of

constrain

ts

No. of in

teger va

riables

Exp

ectation

Stan

dard

de-

viation

25%

qu

artile

Med

ian

75%

qu

artile

Tim

eouts

(1200 s)

10 3 5 2 234 148 70 1.626 1.861 0.471 0.952 1.552 0 15 3 5 2 304 183 105 7.807 22.855 0.971 1.622 4.076 0 20 3 5 2 374 218 140 10.371 12.612 2.484 7.05 12.147 0 25 3 5 2 444 253 175 41.841 58.301 6.35 22.442 43.943 0 30 3 5 2 514 288 210 128.988 196.7 7.28 49.391 130.498 0 35 3 5 2 584 324 245 242.688 319.226 30.263 99.162 277.799 1 40 3 5 2 654 358 280 406.089 389.605 90.05 285.33 484.627 4 20 3 5 1 343 187 140 7.394 11.98 0.912 2.203 9.383 0 20* 3 5 1 343 187 140 0.247 0.073 0.21 0.23 0.24 0 60* 3 5 1 903 467 420 1.573 0.648 1.141 1.272 1.773 0 100* 3 5 1 1463 747 700 4.31 2.402 2.904 3.355 4.727 0 200* 3 5 1 2863 1447 1400 26.062 16.252 14.661 18.917 28.772 0 100# 3 5 1 1463 747 700 > 1200 - > 1200 > 1200 > 1200 30 10¤ 3 5 2 234 148 70 1.626 3.15 0.32 0.691 1.412 0 *: borrowing is allowed, #: risk neutrality, ¤: mean-EDR model.

51

Page 60: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

52

Page 61: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

Helsinki University of Technology Systems Analysis Laboratory Research Reports E16, June 2005

PROJECT VALUATION IN MIXED ASSET PORTFOLIO

SELECTION

Janne Gustafsson Bert De Reyck Zeger Degraeve Ahti Salo

AB TEKNILLINEN KORKEAKOULUTEKNISKA HÖGSKOLANHELSINKI UNIVERSITY OF TECHNOLOGYTECHNISCHE UNIVERSITÄT HELSINKIUNIVERSITE DE TECHNOLOGIE D’HELSINKI

53

Page 62: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

Title: Project Valuation in Mixed Asset Portfolio Selection Authors: Janne Gustafsson Cheyne Capital Management Limited Stornoway House, 13 Cleveland Row, London SW1A 1DH, U.K. [email protected] Bert De Reyck London Business School Regent's Park, London NW1 4SA, U.K. [email protected] Zeger Degraeve London Business School Regent's Park, London NW1 4SA, U.K. [email protected] Ahti Salo Systems Analysis Laboratory Helsinki University of Technology P.O. Box 1100, 02015 HUT, FINLAND [email protected] www.sal.hut.fi/Personnel/Homepages/AhtiS.html Date: June, 2005 Status: Systems Analysis Laboratory Research Reports E16 June 2005 Abstract: We examine the valuation of projects in a setting where an investor, such

as a private firm, can invest in a portfolio of projects as well as in securities in financial markets. In project valuation, it is important to consider all alternative investments opportunities, such as other projects and financial instruments, because each imposes an opportunity cost on the project under consideration. Therefore, conventional methods based on decision analysis may lead to biased estimates of project values, because they typically consider projects in isolation from other investment opportunities. On the other hand, options pricing analysis, which does consider financial investment opportunities, assumes that a project’s cash flows can be replicated using financial assets, which may be difficult in practice. The main contribution of this paper is the development of a procedure for project valuation in a setting where the firm can invest in a portfolio of projects and securities, but where exact replication of project cash flows is not necessarily possible. We consider both single-period and multi-period models, and show that the valuation procedure exhibits several important analytical properties. We also investigate the pricing behavior of mean-variance investors through a set of numerical experiments.

Keywords: finance, capital rationing, portfolio selection, decision analysis, decision

analysis, mixed integer programming.

54

Page 63: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

1

1 Introduction In the literature on corporate finance (e.g. Brealey and Myers 2000, Chapter 9), it is proposed that

a project’s cash flows should be discounted at the rate of return of a security that is “equivalent in

risk” to the project. Under the Capital Asset Pricing Model (CAPM; Sharpe 1964, Lintner 1965),

two assets are equivalent in risk if they have the same beta. Therefore, an appropriate discount

rate for a project is [ ]( )f M fr r E r rβ= + − ,

with rf the risk-free interest rate, [ ]cov[ , ] / varp M Mr r rβ = the project’s beta, Mr the random

rate of return of the market portfolio, and pr the random rate of return of the project. The use of

the CAPM relies on the assumption that the firm is a public company maximizing its share price.

When the markets abide by the CAPM assumptions, the share price is maximized when all

projects with a rate of return higher than what is given by the CAPM are started (Rubinstein

1973). Yet, many companies make decisions about accepting and rejecting projects before their

initial public offering (IPO). At this stage, the CAPM assumptions are not typically satisfied,

because shareholders are often entrepreneurs and venture capitalists. Also, since there is no share

price to maximize, such firms are typically assumed to maximize the value of their project

portfolio instead. In such a setting, the firm is regarded as an individual investor with well-

defined preferences over risky assets.

Several methods have been proposed to value projects in a setting where the firm is an individual

investor, including decision analysis (French 1986, Clemen 1996) and options pricing analysis,

also referred to as contingent claims analysis and real options analysis (Dixit and Pindyck 1994,

Trigeorgis 1996). However, conventional methods based on decision analysis may lead to biased

estimates of project values, because they typically consider projects in isolation from other

investment opportunities, neglecting the opportunity costs that these investment opportunities

impose. Options pricing analysis does account for the effect that financial instruments have on

project values, but its applicability is limited, because replicating the project’s cash flows using

financial instruments may be difficult in practice. Smith and Nau (1995) have extended decision

analytic methods to include the possibility of trading securities. They do not, however, consider

other projects in the portfolio.

55

Page 64: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

2

The main contribution of this paper is the development of a procedure for project valuation in a

setting, where the firm can invest in a portfolio of projects and securities, but where replication of

project cash flows with securities is not necessarily possible. We call this setting a mixed asset

portfolio selection (MAPS) setting. The valuation procedure is based on the concepts of

breakeven selling and buying prices (Luenberger 1998, Smith and Nau 1995, Raiffa 1968), which

are the investor’s own selling and buying prices for a project. We formulate the procedure for

both expected utility maximizers and mean-risk optimizers. Using the Contingent Portfolio

Programming (CPP) framework (Gustafsson and Salo 2005), we develop a general multi-period

MAPS model, which allows us to study a broad variety of risk-averse preference models in a

multi-period setting. This model accommodates Markowitz’s (1952) mean-variance (MV) model,

Konno and Yamazaki’s (1991) mean-absolute deviation (MAD) model, the mean-lower semi-

absolute deviation (MLSAD) model (Ogryczak and Ruszczynski 1999), and Fishburn’s (1977)

mean-risk models where risk is associated with deviations below a fixed target value. We also

prove several analytical properties for breakeven selling and buying prices.

MAPS-based project valuation is important in practice as well as in theory. On the one hand,

MAPS-based project valuation is required when a corporation or an individual makes investments

both in securities and projects, or other lumpy investment opportunities. For example, in June

2004, Microsoft had invested nearly $44.6 billion in short-term investments in financial markets

in addition to carrying out a large number of projects at the same time (Microsoft 2004). Also,

many investment banks invest in a portfolio of publicly traded securities and undertake uncertain

one-time endeavors, such as venture capital investments and organization of IPOs. On the other

hand, one of the fundamental problems in the literature on corporate finance is the valuation of a

single project while taking into account the opportunity costs imposed by securities (see e.g.

Brealey and Myers 2000). This is a MAPS setting that includes one project and several securities.

This paper is structured as follows. Section 2 introduces single-period MAPS models and

discusses different formulations of the portfolio selection problem. Section 3 presents the multi-

period MAPS model. In Section 4, we introduce the valuation concepts, and produce theoretical

results on the valuation properties of different types of investors. Section 5 demonstrates the

pricing properties of mean-variance investors through a series of numerical experiments. In

56

Page 65: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

3

Section 6, we summarize our findings and discuss the managerial implications of our results.

2 Mixed Asset Portfolio Selection In a MAPS problem, available investment opportunities are divided into two categories: (1)

securities, which can be bought and sold in any quantities, and (2) projects, lumpy all-or-nothing

type investments. From a technical point of view, the main difference between these two types of

investments is that the projects’ decision variables are binary, while those of the securities are

continuous. Another difference is that the cost, or price, of securities is determined by a market

equilibrium model, such as the CAPM, while a project’s investment cost is an endogenous

property of the project.

Portfolio selection models can be formulated either in terms of rates of return and portfolio

weights, like in Markowitz-type formulations, or by using a budget constraint, expressing the

initial wealth level, and maximizing the investor’s terminal wealth level. When properly applied,

both approaches yield identical results. We use the second approach with MAPS, because it is

more suitable to project portfolio selection. We first formulate single-period MAPS models,

where the investments are made at time 0 and the objective at time 1 is optimized. These models

will allow us to generate several insights and show how MAPS is related to Markowitz (1952)

and the CAPM. We then develop the multi-period MAPS model based on Contingent Portfolio

Programming (CPP, Gustafsson and Salo 2005).

Early portfolio selection formulations (Markowitz 1952) were bi-criteria decision problems

minimizing risk while setting a target for expectation. Later, the mean-variance model was

formulated in terms of expected utility theory (EUT) using a quadratic utility function. However,

there are no similar utility functions for most other risk measures, including the widely used

absolute deviation (Konno and Yamazaki 1991). Therefore, we distinguish between two classes

of portfolio selection models: (1) preference functional models, such as the expected utility

model, and (2) bi-criteria optimization models or mean-risk models.

A single-period MAPS model using a preference functional can be formulated as follows. Let

there be n risky securities, a risk-free asset (labeled as the 0th security), and m projects. Let the

57

Page 66: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

4

price of asset i at time 0 be 0iS and the corresponding (random) price at time 1 is 1

iS . The price

of the risk-free asset at time 0 is 1 and 1 fr+ at period 1, where rf is the risk-free interest rate. The

amounts of securities in the portfolio are denoted by , 0,...,ix i n= . The investment cost of

project k in time 0 is 0kC and the (random) cash flow at time 1 is 1

kC . The binary variable zk

indicates whether project k is started or not. The investor’s budget is b. We can then formulate a

MAPS model using a preference functional U as follows:

(i) maximize utility at time 1: 1 1

, 0 1max

n m

i i k ki k

U S x C z= =

⎡ ⎤+⎢ ⎥⎣ ⎦∑ ∑x z

subject to

(ii) budget constraint at time 0: 0 0

0 1

n m

i i k ki k

S x C z b= =

+ =∑ ∑

(iii) binary variables for projects: 0,1 1,...,kz k m∈ =

(iv) continuous variables for securities: free 0,...,ix i n= .

The budget constraint is formulated as equality, because in the presence of a risk-free asset all of

the budget will be expended at the optimum. In this model and throughout the paper, it is

assumed that there are no transaction costs or capital gains tax, and that the investor is able to

borrow and lend at the risk-free interest rate without limit. These assumptions can be relaxed

without introducing prohibitive complexities. For expected utility theory, the preference functional is [ ] [ ]( )U X E u X= , where u is the investor’s von Neumann-Morgenstern utility

function. When the investor is able to determine a certainty equivalent for any random variable X,

U can be expressed as a strictly increasing transformation of the investor’s certainty equivalent

operator CE. Hence, the objective can also be written as max [ ]CE X , which gives the total value

of the investor’s portfolio.

In addition to preference functional models, mean-risk models have been widely used in the

literature. We concentrate on these models, because much of the modern portfolio theory,

including the CAPM, is based on a mean-risk model, namely the Markowitz (1952) mean-

variance model. Table 1 describes three possible formulations for mean-risk models: risk

minimization, where risk is minimized for a given level of expectation (Luenberger 1998),

expected value maximization, where expectation is maximized for a given level of risk (Eppen et

al. 1989), and the additive formulation, where the weighted sum of mean and risk is maximized

(Yu 1985). In Table 1, ρ is the investor’s risk measure, µ is the minimum level for expectation,

58

Page 67: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

5

and R is the maximum level for risk. The parameters λ are tradeoff coefficients. The Markowitz

(1952) model can be understood as a special case of a mean-variance MAPS model where the

number of projects is zero.

Table 1. Formulations of the mean-risk optimization problem.

Objective Constraints Risk minimization 1 1

, 0 1min

n m

i i k ki k

S x C zρ= =

⎡ ⎤+⎢ ⎥

⎣ ⎦∑ ∑x z

1 1

0 1

n m

i i k ki k

E S x C z µ= =

⎡ ⎤+ ≥⎢ ⎥

⎣ ⎦∑ ∑

0 0

0 1

n m

i i k ki k

S x C z b= =

+ =∑ ∑

Expected value maximization

1 1

, 0 1max

n m

i i k ki k

E S x C z= =

⎡ ⎤+⎢ ⎥⎣ ⎦∑ ∑x z

1 1

0 1

n m

i i k ki k

S x C z Rρ= =

⎡ ⎤+ ≤⎢ ⎥⎣ ⎦∑ ∑

0 0

0 1

n m

i i k ki k

S x C z b= =

+ =∑ ∑

General additive 1 1 1 1

1 2, 0 1 0 1max

n m n m

i i k k i i k ki k i k

E S x C z S x C zλ λ ρ= = = =

⎡ ⎤ ⎡ ⎤⋅ + − ⋅ +⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦∑ ∑ ∑ ∑x z

0 0

0 1

n m

i i k ki k

S x C z b= =

+ =∑ ∑

Sharpe (1970) 1 1 1 1

, 0 1 0 1max

n m n m

i i k k i i k ki k i k

E S x C z S x C zλ ρ= = = =

⎡ ⎤ ⎡ ⎤⋅ + − +⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦∑ ∑ ∑ ∑x z

0 0

0 1

n m

i i k ki k

S x C z b= =

+ =∑ ∑

Ogryczak and Ruszczynski (1999)

1 1 1 1

, 0 1 0 1max

n m n m

i i k k i i k ki k i k

E S x C z S x C zλ ρ= = = =

⎡ ⎤ ⎡ ⎤+ − ⋅ +⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦∑ ∑ ∑ ∑x z

0 0

0 1

n m

i i k ki k

S x C z b= =

+ =∑ ∑

The general additive form can be turned into the model employed by Sharpe (1970) by dividing the additive form by 2λ , provided it is nonzero, and into the form used by Ogryczak and

Ruszczynski (1999) by dividing it by 1λ . Apart from expectation and risk constraints, the Karush-

Kuhn-Tucker (KKT) conditions of all of the formulations are identical and therefore will yield the

same efficient frontiers, as long as optimal solutions in the additive formulation remain bounded.

However, if limitless borrowing and shorting are allowed, the additive formulation can give

unbounded solutions unless a risk constraint is introduced. Yet, in this case the formulation

essentially coincides with expected value maximization in terms of KKT conditions.

Both risk minimization and expected value maximization models have advantages and

59

Page 68: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

6

disadvantages. Risk minimization requires the investor to set a minimum level for expectation, a

readily understandable quantity, while the interpretation of a maximum risk level in expected

value maximization may not always be clear. However, expected value maximization allows us to

include multiple risk constraints, e.g. one for variance, one for expected downside risk, and one

for the probability of getting an outcome below a particular level, so that the risk profile of the

portfolio can be matched to the risk preferences of the investor as closely as possible. Due to this

flexibility we focus on this formulation in the following.

3 A Multi-Period MAPS Model

3.1 Framework We develop a multi-period MAPS model using the Contingent Portfolio Programming (CPP)

framework developed by Gustafsson and Salo (2005). In this framework, uncertainties are

modeled using a state tree, representing the structure of future states of nature, as depicted in the

leftmost chart in Figure 1. The state tree need not be binomial or symmetric; it may also take the

form of a multinomial tree with different probability distributions in its branches. In each non-

terminal state, securities can be bought and sold in any, possibly fractional quantities.

0 1 2

Stat

es

50%

30%

70%

40%

60%

50%

15%

35%

20%

30%

ω0

ω1

ω2

ω12

ω21

ω22

ω11

Start? Continue?Yes

No

Yes

No

0 1 2Time

0

Start?Yes

No

2

Continue?Yes

No

Yes

No

1Time

ω11

ω12ω11

ω12ω21

ω22ω21

ω22

ω1

ω2

Time Figure 1. A state tree, decision sequence, and a decision tree for a project.

Projects are modeled using decision trees that span over the state tree. The two right-most charts

in Figure 1 describe how project decisions, when combined with the state tree, lead to project-

specific decision trees. The specific feature of these decision trees is that the chance nodes are

shared by all projects, since they are generated using the common state tree. Security trading is

implemented through state-specific trading variables, which are similar to the ones used in

60

Page 69: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

7

financial models of stochastic programming (e.g. Mulvey et al. 2000) and in Smith and Nau’s

(1995) method. Similar to a single-period MAPS model, the investor seeks either to maximize the

utility of the terminal wealth level, or the expectation of the terminal wealth level subject to a risk

constraint.

3.2 Model Components The two main components of the model are (i) states and (ii) the investor’s investment decisions,

which imply the cash flow structure of the model.

3.2.1 States

Let the planning horizon be 0,...,T . The set of states in period t is denoted by tΩ , and the set

of all states is 0

T

tt=

Ω = Ω∪ . The state tree starts with base state 0ω in period 0. Each non-terminal

state is followed by at least one state. This relationship is modeled by the function :B Ω→Ω

which returns the immediate predecessor of each state, except for the base state, for which the function gives 0 0( )B ω ω= . The probability of state ω , when ( )B ω has occurred, is given by

( ) ( )Bp ω ω . Unconditional probabilities for each state, except for the base state, can be computed

recursively from the equation ( )( ) ( ) ( ( ))Bp p p Bωω ω ω= ⋅ . The probability of the base state is

0( ) 1p ω = .

3.2.2 Investment Decisions

Let there be n securities available in financial markets. The amount of security i bought in state ω is indicated by trading variable ,ix ω , 1,...,i n= , ω∈Ω , and the price of security i in state ω

is denoted by ( )iS ω . Under the assumption that all securities are sold in the next period, the cash

flow implied by security i in state 0ω ω≠ is ( ), ( ) ,( )i i B iS x xω ωω ⋅ − . In base state 0ω , the cash

flow is 00 ,( )i iS x ωω− ⋅ .

The investor can invest privately in m projects. The decision opportunities for each project, 1,...,k m= , are structured as a decision tree, where we have a set of decision points kD and

function ( )ap d that gives the action leading to decision point 0\k kd D d∈ , where 0kd is the

first decision point of project k. Let dA be the set of actions that can be taken in decision point

kd D∈ . For each action a in dA , a binary action variable az indicates whether the action is

selected or not. Action variables at each decision point d are bound by the restriction that only

61

Page 70: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

8

one az , da A∈ , can be equal to one. The action in decision point d is chosen in state ( )dω .

For a project k, the vector of all action variables az relating to the project, denoted by kz , is

called the management strategy of project k. The vector of all action variables of all projects,

denoted by z, is the project portfolio management strategy. We call the pair ( , )x z , composed of

all trading and action variables, the mixed asset portfolio management strategy.

3.2.3 Cash Flows and Cash Surpluses

Let ( , )pkCF ωkz be the cash flow of project k in state ω with project management strategy kz .

When ( )aC ω is the cash flow in state ω implied by action a , this cash flow is given by

:( ) ( )

( , ) ( )k dB

pk a a

d D a Ad

CF C z

ω ω

ω ω∈ ∈∈Ω

= ⋅∑ ∑kz

where the restriction in the summation of the decision points guarantees that actions yield cash

flows only in the prevailing state and in the future states that can be reached from the prevailing state. The set ( )B ωΩ is defined as ( ) | 0 such that ( )B kk Bω ω ω ω′ ′Ω = ∈Ω ∃ ≥ = , where

1( ) ( ( ))n nB B Bω ω−= is the nth predecessor of ω ( 0 ( )B ω ω= ).

The cash flows from security i in state ω∈Ω are given by

( ),

, ( ) ,

( )( , )

( )i is

ii i B i

S xCF

S x xω

ω ω

ωω

ω

⎧− ⋅⎪= ⎨⋅ −⎪⎩

ix 0

0

if if ω ωω ω=

Thus, the aggregate cash flow ( , , )CF ωx z in state ω∈Ω , obtained by summing up the cash

flows for all projects and securities, is

( )

1 1

, 01 :

( ) ( )

, ( ) , 01 :

( ) ( )

( , , ) ( , ) ( , )

( ) ( ) , if

( ) ( ) , if

k dB

k dB

n ms p

i ki k

n

i i a ai d D a A

d

n

i i B i a ai d D a A

d

CF CF CF

S x C z

S x x C z

ω

ω ω

ω ω

ω ω

ω ω ω

ω ω ω ω

ω ω ω ω

= =

= ∈ ∈∈Ω

= ∈ ∈∈Ω

= +

⎧ − ⋅ + ⋅ =⎪⎪⎪= ⎨⎪ ⋅ − + ⋅ ≠⎪⎪⎩

∑ ∑

∑ ∑ ∑

∑ ∑ ∑

i kx z x z

Together with the initial budget of each state, cash flows define cash surpluses that would result

in state ω∈Ω if the investor chose portfolio management strategy ( , )x z . Assuming that excess

cash is invested in the risk-free asset, the cash surplus in state ω∈Ω is given by

62

Page 71: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

9

( ) ( )

( ) ( , , )( ) ( , , ) (1 )B B

b CFCS

b CF r CSωω ω ω

ω ω

ω ω →

+⎧⎪= ⎨ + + + ⋅⎪⎩

x zx z

0

0

if if ω ωω ω=≠

,

where ( )b ω is the initial budget in state ω∈Ω and ( )Br ω ω→ is the short rate at which cash

accrues interest from state ( )B ω to ω . The cash surplus in a terminal state is the investor’s

terminal wealth level in that state.

3.3 Optimization Model When using a preference functional U, the objective function for the MAPS model can be written

as a function of cash surplus variables in the last time period, i.e. ( )

, ,maxU Tx z CS

CS

where CST denotes the vector of cash surplus variables related to period T. Under the risk-

constrained mean-risk model, the objective is to maximize the expectation of the investor’s

terminal wealth level:

, ,max ( )

T

p CSωω

ω∈Ω

⋅∑x z CS

Table 2. Multi-period MAPS models.

Preference functional model Mean-risk model

Objective function

( ), ,

maxU Tx z CSCS

, ,max ( )

T

p CSωω

ω∈Ω

⋅∑x z CS

Budget constraints

00 0( , , ) ( )CF CS bωω ω− = −x z

( ) ( ) 0( , , ) (1 ) ( ) \ B BCF r CS CS bω ω ω ωω ω ω ω→+ + ⋅ − = − ∀ ∈Ωx z

Decision consistency constraints

0

1 1,...,dk

aa A

z k m∈

= =∑

0( ) \ 1,...,

d

a ap d k ka A

z z d D d k m∈

= ∀ ∈ =∑

Risk constraints

( ), Rρ − +∆ ∆ ≤

( ) 0 TCSω ω ωτ ω+ −− − ∆ + ∆ = ∀ ∈ΩTCS

Variables

0,1 1,...,a d kz a A d D k m∈ ∀ ∈ ∀ ∈ =

, free 1,...,ix i nω ω∀ ∈Ω =

freeCSω ω∀ ∈Ω

0,1 1,...,a d kz a A d D k m∈ ∀ ∈ ∀ ∈ =

, free 1,...,ix i nω ω∀ ∈Ω =

freeCSω ω∀ ∈Ω

0 Tω ω−∆ ≥ ∀ ∈Ω

0 Tω ω+∆ ≥ ∀ ∈Ω

63

Page 72: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

10

Three types of constraints are imposed on the model: (i) budget constraints, (ii) decision

consistency constraints, and (iii) risk constraints, which apply to risk-constrained models only.

The multi-period MAPS model under both a preference functional and a mean-risk model is

given in Table 2. The constraints are explained in more detail in the following sections.

3.3.1 Budget Constraints

Budget constraints ensure that there is a nonnegative amount of cash in each state. They can be implemented using continuous cash surplus variables CSω , which measure the amount of cash in

state ω . These variables lead to the budget constraints

00 0( , , ) ( )CF CS bωω ω− = −x z

( ) ( ) 0( , , ) (1 ) ( ) \ B BCF r CS CS bω ω ω ωω ω ω ω→+ + ⋅ − = − ∀ ∈Ωx z

Note that if CSω is negative, the investor borrows money at the risk-free interest rate to cover a

funding shortage. Thus, CSω can also be regarded as a trading variable for the risk-free asset.

3.3.2 Decision Consistency Constraints

Decision consistency constraints implement the projects’ decision trees. They require that (i) at

each decision point reached, only one action is selected, and that (ii) at each decision point that is

not reached, no action is taken. Decision consistency constraints can be written as

0

1 1,...,dk

aa A

z k m∈

= =∑

0( ) \ 1,...,

d

a ap d k ka A

z z d D d k m∈

= ∀ ∈ =∑ ,

where the first constraint ensures that one action is selected in the first decision point, and the

second implements the above requirements for other decision points.

3.3.3 Risk Constraints

A risk-constrained model includes one or more risk constraints. We focus on the single constraint case. When ρ denotes the risk measure and R the risk tolerance, a risk constraint can be expressed

as ( ) Rρ ≤TCS .

In addition to variance (V), several other risk measures have been proposed in the literature on

portfolio selection. These include semivariance (Markowitz 1959), absolute deviation (Konno

and Yamazaki 1991), lower semi-absolute deviation (Ogryczak and Ruszczynski 1999), and their

64

Page 73: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

11

fixed target value counterparts (Fishburn 1977). Semivariance (SV), absolute deviation (AD) and

lower semi-absolute deviation (LSAD) are defined as

SV: ( )2 ( )X

X X Xx dF xµ

σ µ−∞

= −∫ ,

AD: ( )X X Xx dF xδ µ∞

−∞

= −∫ , and

LSAD: ( ) ( ) ( )X X

X X X X Xx dF x x dF xµ µ

δ µ µ−∞ −∞

= − = −∫ ∫ ,

where Xµ is the mean of random variable X and FX is the cumulative density function of X. The

fixed target value statistics are obtained by replacing Xµ by some constant target value τ. All of

these measures can be formulated in a MAPS program through deviation constraints. In general,

deviation constraints are expressed as ( ) 0 TCSω ω ωτ ω+ −− − ∆ + ∆ = ∀ ∈ΩTCS ,

where ( )τ TCS is a function defining the target value from which the deviations are calculated,

and ω+∆ and ω

−∆ are nonnegative deviation variables which measure how much the cash surplus

in state Tω∈Ω differs from the target value. For example, when the target value is the mean of

the terminal wealth level, the deviation constraints are written as

( ) 0 ,T

TCS p CSω ω ω ωω

ω ω+ −′

′∈Ω

′− − ∆ + ∆ = ∀ ∈Ω∑

Using these deviation variables, some common dispersion statistics can be written as follows: AD: ( ) ( )

T

p ω ωω

ω − +

∈Ω

⋅ ∆ + ∆∑ .

LSAD: ( )T

p ωω

ω −

∈Ω

⋅∆∑ .

V: 2( ) ( )

T

p ω ωω

ω − +

∈Ω

⋅ ∆ + ∆∑

SV: 2( ) ( )T

p ωω

ω −

∈Ω

⋅ ∆∑ .

The respective fixed-target value statistics can be obtained with the deviation constraints 0 TCSω ω ωτ ω+ −− − ∆ + ∆ = ∀ ∈Ω ,

where τ is the fixed target level. EDR, for example, can then be obtained from the sum

( )T

p ωω

ω −

∈Ω

⋅∆∑ .

It is also possible to set a limit for the (critical) probability of getting an outcome below some

65

Page 74: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

12

target level. Mathematically, the critical probability is defined as ( ) ( )crit

X XP F Xτ τ= < ,

where FX denotes the cumulative distribution function of random variable X and τ is the desired

target level. This risk measure can be implemented by using the following linear constraints:

TCS Mω ωτ ξ ω− ≤ ∀ ∈Ω , and

0,1 Tωξ ω∈ ∀ ∈Ω ,

where τ is the target value from which critical probability is calculated and M is some very large

number. Critical probability can then be constrained from above using the constraint

T

p Rω ωω

ξ∈Ω

≤∑ .

3.3.4 Other Constraints

Other types of constraints can also be included in the model, such as non-negativity restraints that

prevent short selling, upper bounds for the number of shares bought, and credit limit constraints

(Markowitz 1987). In the ensuing sections, we assume, for the sake of simplicity, that no such

additional constraints have been imposed.

4 Project Valuation

4.1 Breakeven Buying and Selling Prices Because we consider projects as non-tradable investment opportunities, there is no market price

that can be used to value the project. In such a situation, it is reasonable to define the value of the

project as the amount of money at present, time 0, that is equally desirable to the project. In a

portfolio context, this can be interpreted so that the investor is indifferent between the following

two portfolios: (A1) a portfolio with the project and (B1) a portfolio without the project and cash

equal to the value of the project. However, we may alternatively define the value of a project as

the indifference between the following two portfolios: (A2) a portfolio without the project and

(B2) a portfolio with the project and a debt equal to the value of the project. The project values

obtained in these two ways will not, in general, be the same. Analogously to Luenberger (1998),

Smith and Nau (1995), and Raiffa (1968), we call the first value the “breakeven selling price”

(BSP), as the portfolio comparison can be understood as a selling process, and the second type of

value the “breakeven buying price” (BBP).

66

Page 75: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

13

A crucial element in BSP and BBP is the definition of equal desirability of two different

portfolios. In preference functional models, the investor is, by definition, indifferent between two

portfolios whenever their utility scores are equal. In a mean-risk setting, an investor is indifferent

between two portfolios if the means and risks of the two portfolios are identical. More generally,

when the risks are considered as constraints, the investor is indifferent between two portfolios if

their expectations are equal and they both satisfy the risk constraints. Hence, we obtain the pairs of optimization problems shown in Table 3. Here, the variable *az is the action variable

associated with not starting project j; thus, * 0az = indicates that investor invests in the project,

whereas * 1az = means that the project is not started. In the case that the project starting decision

is a binary “go / no go” decision, we can alternatively impose the restrictions on the variable

indicating project starting instead. Finding a BSP and BBP is an inverse optimization problem: one has to find the values for the parameters s

jv and bjv so that the optimal value of the second

optimization problem matches the optimal value of the first problem.

4.2 Inverse Optimization Procedure Inverse optimization problems have recently attracted interest among researchers, such as Ahuja

and Orlin (2001). In an inverse optimization problem, the challenge is to find the values for a set

of parameters, typically a subset of all model parameters, that yields the desired optimal solution.

Inverse optimization problems can broadly be classified into two groups: (i) finding an optimal

value for the objective function, and (ii) finding a solution vector. Ahuja and Orlin (2001) discuss

problems of the second kind, whereas the problem of finding a BSP or BBP falls within the first

class.

In principle, the task of finding a BSP is equivalent to finding a root to the function ( ) ( )s s s

j s j sf v W v W− += − , where sW + is the optimal value of the portfolio optimization problem in

the status quo and ( )ss jW v− is the corresponding optimal value in the second setting as the

function of parameter sjv . Similarly, the BBP can be obtained by finding the root to the function

( ) ( )b b bj b b jf v W W v− += − . Note that these functions are increasing with respect to their

parameters. To solve such root-finding problems, we can use any of the usual root-finding

algorithms (see e.g. Belegundu and Chandrupatla 1999) such as the bisection method, the secant

method, and the false position method, which have the advantage that they do not require the

67

Page 76: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

14

knowledge of the functions’ derivatives, which are not typically known. If the first derivatives are

known, or when approximated numerically, we can use the Newton-Raphson method to obtain

quicker convergence.

4.3 General Analytical Properties

4.3.1 Sequential Consistency

Breakeven selling and buying prices are not, in general, equal to each other. While this

discrepancy is accepted as a general property of risk preferences in expected utility theory (Raiffa

1968), it may also seem to contradict the rationality of these valuation concepts. It can be argued

that if the investor were willing to sell a project at a lower price than at which he/she would be

prepared to buy it, the investor would create an arbitrage opportunity and lose an infinite amount

of money when another investor repeatedly bought the project at its selling price and sold it back

at the buying price. In a reverse situation where the investor’s selling price for a project is greater

than the respective buying price, the investor would be irrational in the sense that he/she would

not take advantage of an arbitrage opportunity – if such an opportunity existed – where it would

be possible to repeatedly buy the project at the investor’s buying price and sell it at a slightly

higher price which is below the investor’s breakeven selling price.

However, these arguments neglect the fact that the breakeven prices are affected by the budget

and that therefore these prices may change after obtaining the project’s selling price and after

paying its buying price. Indeed, it can be shown that in a sequential setting where the investor

first sells the project, adding the selling price to the budget, and then buys the project back, the

investor’s selling price and the respective (sequential) buying price are always equal to each

other. This observation is formalized as the following proposition and it holds for any preference

model accommodated by the model in Section 3. The proof is given in the Appendix.

PROPOSITION 1. A project’s breakeven selling (buying) price and its sequential breakeven

buying (selling) price are equal to each other.

4.3.2 Consistency with Contingent Claims Analysis

Option pricing analysis, or contingent claims analysis (CCA; Luenberger 1998, Brealey and

Myers 2000, Hull 1999), can be applied to value projects whenever the cash flows of a project

68

Page 77: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

15

can be replicated using financial instruments. According to CCA, the value of project j is given

by the market price of the replicating portfolio (or trading strategy) less the investment cost of the

project: 0 0

0

nCCAj j i i

iv C S x∗

=

= − +∑ ,

where ix∗ is the amount of security i in the replicating portfolio and 0jC is the investment cost of

the project in at time 0.

It is straightforward to show that, when CCA is applicable, i.e. if there exists a replicating

portfolio, then the breakeven buying and selling prices are equal to each other and yield the same result as CCA (Smith and Nau 1995). For example, when CCA

jv is positive, we know that any

investor will invest in the project, since it is possible to make money for sure by investing in the

project and shorting the replicating portfolio. Furthermore, any investor will start the project even when he/she is forced to pay a sum b

jv less than CCAjv to gain a license to invest in the project,

because it is now possible to gain CCA bj jv v− for sure. On the other hand, if b

jv is greater than CCAjv , the investor will be better off by investing the replicating portfolio instead and hence

he/she will not start the project. A similar reasoning applies to breakeven selling prices. These

observations are formalized in Proposition 2. The proof is obvious and hence omitted. Due to the

consistency with CCA, the breakeven prices can be regarded as a generalization of CCA to

incomplete markets.

PROPOSITION 2. If there is a replicating portfolio for a project, the breakeven selling price and

breakeven buying price are equal to each other and yield the same result as CCA.

4.3.3 Sequential Additivity

The BBP and BSP for a project depend on what other assets are in the portfolio. The value

obtained from breakeven prices is, in general, an added value, which is determined relative to the

situation without the project. When there are no other projects in the portfolio, or when we

remove them from the model before determining the value of the project, we speak of the

project’s isolated value. We define the respective values for a set of projects as the joint added

value and joint value. Figure 2 illustrates the relationship between these concepts.

69

Page 78: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

16

Isolated Value

Joint Value Joint Added Value

Added Value

No other projects Additional projects

Portf

olio

of

proj

ects

Sing

le

proj

ect

Num

ber

of p

roje

cts b

eing

val

ued

Number of other projects in the portfolio Figure 2. Different types of valuations for projects.

Isolated project values are, in general, non-additive; they do not sum up to the value of the project

portfolio composed of the same projects. However, in a sequential setting where the investor buys

the projects one after the other using the prevailing buying price at each time, the obtained project

values do add up to the joint value of the project portfolio. These prices are the projects’ added

values in a sequential buying process, where the budget is reduced by the buying price after each

step. We refer to these values as sequential added values. This sequential additivity property

holds regardless of the order in which the projects are bought. Individual projects can, however,

acquire different added values depending on the sequence in which they are bought. These

observations are proven by the following proposition. The proof is in the Appendix.

PROPOSITION 3. The breakeven buying (selling) prices of sequentially bought (sold) projects add

up to the breakeven buying (selling) price of the portfolio of the projects regardless of the order

in which the projects are bought (sold).

4.4 Analytical Properties of Mean-Risk Preference Models

4.4.1 Equality of Prices and Solution through a Pair of Optimization Problems

When the investor is a mean-risk optimizer and the risk measure is independent of an addition of

a constant to the portfolio, like variance, breakeven selling and buying prices are identical,

provided that unlimited borrowing and lending are allowed. In contrast, they are typically

different under expected utility theory and under most non-expected utility models. Also, for

mean-risk optimizers the breakeven prices can be computed directly by solving the expectations

of terminal wealth levels when the investor invests and does not invest in the project and

discounting their difference back to its present value at the risk-free interest rate. Therefore, with

70

Page 79: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

17

mean-risk models, there is no need to resort to possibly laborious inverse optimization. We

formalize these claims in Proposition 4. The proof is in the Appendix.

PROPOSITION 4. Let the investor be a mean-risk optimizer with risk measure ρ that satisfies [ ] [ ]X b Xρ ρ+ = for all random variables X and constants b. When limitless borrowing and

lending are allowed, the breakeven selling price and the breakeven buying price of any given

project are identical. Moreover, the prices are equal to

( )

1

0

1 ( )T

ft

W Wvr t

+ −

=

−=

+∏,

where W + is the expectation of the terminal wealth level when the investor invests in the project,

W − is the expectation of the terminal wealth level when the investor does not invest in the

project, and rf,(t) is the risk-free interest rate from time t to t+1.

Proposition 4 is striking both in its generality and simplicity, since neither the given valuation

formula, nor the equality of breakeven prices, generally holds under expected utility theory. The

results are also intuitively appealing: the investor places only a single price on a given project,

and this price is related to the investor’s terminal wealth levels with and without the project. This

indicates that mean-risk models exhibit a very reasonable type of pricing behavior.

4.4.2 Relationship to the Capital Asset Pricing Model

According to the CAPM, the market price of any asset is given by the certainty equivalent

formula (see, e.g., Luenberger 1998):

[ ][ ]1 1

0cov ,

1 var 1j j M M fCAPM

j jf M f

E C C r E r rv C

r r r

⎡ ⎤ ⎡ ⎤ −⎣ ⎦ ⎣ ⎦= − + − ⋅+ +

,

where 0jC is the investment cost of the asset, 1

jC is the random value of the asset at time 1, and

Mr is the random return of the market portfolio. For non-market-traded assets like projects, the

outcome of the formula can be interpreted as the price that the markets would give to the asset if

it were traded.

In general, breakeven selling and breakeven buying prices are inconsistent with CAPM

valuations, because some of the CAPM assumptions do not hold in a MAPS setting. For example,

private projects are not included in the derivation of the CAPM market equilibrium, yet they may

71

Page 80: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

18

be strongly correlated with the securities. Therefore, results that hold for the CAPM do not

necessarily hold here. In particular, the optimal financial portfolio for the investor is not

necessarily a combination of the risk-free asset and the market fund consisting of all securities in

proportions according to their market capitalization.

However, there is a special case when the investor’s optimal financial portfolio is what the

CAPM predicts, namely when the project portfolio is uncorrelated with the market securities and

the investor is a mean-variance optimizer. This is proven in Proposition 5. Nevertheless, even

though the optimal financial portfolio falls on the capital market line with and without the project

being valued, the CAPM valuation is incorrect even in this case, as an extra risk premium term

appears in the breakeven selling and buying prices, as shown in Propositions 6 and 7. Proofs for

the propositions can be found in the Appendix. Note that this is a particularly interesting special

case, because project outcomes do not usually depend on the fluctuations of financial markets.

Proposition 5 is to be understood as a mixed asset portfolio extension of the usual Separation

Theorem (Tobin 1958), applicable when projects are uncorrelated with securities.

PROPOSITION 5. If the investor is a mean-variance optimizer and projects are uncorrelated with

securities, the optimal financial portfolio is a combination of the market fund and the risk-free

asset.

PROPOSITION 6. If the conditions of Proposition 5 hold and the composition of the optimal

project portfolio is the same with and without project j, the breakeven selling price of project j is

[ ]1 1

1 1 10

1

1

var var

1 11 1var

m m

k k k kk k

j k j M fsj j Mn

f fi i

i

C z C zE C E r r

v C br rS x

= =≠

=

⎛ ⎞⎡ ⎤⎜ ⎟⎡ ⎤ ⎢ ⎥−⎜ ⎟⎢ ⎥ ⎢ ⎥⎣ ⎦⎡ ⎤ ⎜ ⎟ −⎢ ⎥⎣ ⎦ ⎣ ⎦= − + − + −⎜ ⎟

+ +⎡ ⎤⎜ ⎟⎢ ⎥⎜ ⎟⎣ ⎦⎜ ⎟⎜ ⎟

⎝ ⎠

∑ ∑

∑,

where Mr is the random rate of return of the market portfolio and bM is the amount of money

spent on securities in the status quo. That is, 00

1

m

M k kk

b b x C z=

= − −∑ .

PROPOSITION 7. If the conditions of Proposition 5 hold and the composition of the optimal

72

Page 81: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

19

project portfolio is the same with and without project j, the breakeven buying price of project j is

[ ]1 1

1 1 10

1

1

var var

1 11 1var

m m

k k k kk k

j k j M fbj j Mn

f fi i

i

C z C zE C E r r

v C br rS x

= =≠

=

⎛ ⎞⎡ ⎤⎜ ⎟⎡ ⎤ ⎢ ⎥−⎜ ⎟⎢ ⎥ ⎢ ⎥⎣ ⎦⎡ ⎤ ⎜ ⎟ −⎢ ⎥⎣ ⎦ ⎣ ⎦= − + − − −⎜ ⎟

+ +⎡ ⎤⎜ ⎟⎢ ⎥⎜ ⎟⎣ ⎦⎜ ⎟⎜ ⎟

⎝ ⎠

∑ ∑

∑,

where 00

1

m

M k kkk j

b b x C z=≠

= − −∑ .

Note that bM is computed here slightly differently than in Proposition 6 due to a different situation

in the status quo. Despite the differences in the formulas in Propositions 6 and 7, we know from

Proposition 4 that the formulas will give the same result. The following proposition generalizes

Propositions 6 and 7 to a case where the optimal project portfolio changes with and without the

examined project.

PROPOSITION 8. If the conditions of Proposition 5 hold, the difference in budget required to

make two portfolios with different projects (the decision variables of the first are denoted by , 1,...,kz k m= , and those of the second by , 1,...,kz k m′ = ) equally desirable is

( )( ) [ ]

1 11

1 10 1

1 1

1

var var1 1

1 1var

m mm

k k k kk k kmM fk kk

k k k Mnk f f

i ii

C z C zE C z z E r rC z z b

r rS x

= ==

=

=

⎛ ⎞⎡ ⎤ ⎡ ⎤′⎡ ⎤ −⎜ ⎟′− ⎢ ⎥ ⎢ ⎥⎣ ⎦ −⎣ ⎦ ⎣ ⎦⎜ ⎟′∆ = − − + − + −⎜ ⎟+ +⎡ ⎤⎜ ⎟⎢ ⎥⎜ ⎟⎣ ⎦⎝ ⎠

∑ ∑∑∑

,

where 00

1

m

M k kk

b b x C z=

= − −∑ .

As the CAPM predicts that the value of an uncorrelated project is equal to its NPV discounted at

the risk-free interest rate, this suggests that breakeven buying and selling prices are, in general,

different from CAPM values, even when the optimal financial portfolio falls on the capital market

line. However, the prices approach CAPM values when the amount of securities in the portfolio

increases (i.e. when the investor’s risk tolerance increases). This can be verified by multiplying

bM inside the parentheses. The last term in Proposition 8 now becomes

73

Page 82: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

20

[ ]1 1

1 12

var var

1var

m m

k k k kM fk k

M MfM

C z C z E r rb b

rr= =

⎛ ⎞⎡ ⎤ ⎡ ⎤′−⎜ ⎟⎢ ⎥ ⎢ ⎥ −⎣ ⎦ ⎣ ⎦⎜ ⎟+ −⎜ ⎟ +⎡ ⎤⎣ ⎦⎜ ⎟⎜ ⎟⎝ ⎠

∑ ∑,

which goes to 0 as bM goes to infinity.

Convergence of the project values to CAPM prices as the investor’s risk tolerance goes to infinity

applies in general to all projects, regardless of their correlation with each other or with the

market. As long as the optimal project portfolio is the same with and without the project being

valued, every project’s BSP and BBP will converge towards the CAPM price of the project, as

shown in Proposition 9; otherwise, they will converge towards the CAPM value of a difference

portfolio ′z - z specified in Proposition 10. The proofs of the propositions are in the Appendix.

Note that these limit results are valid regardless of whether the markets actually abide by the

CAPM. If there is a discrepancy between securities’ expected returns and the capitalization

weights that a mean-variance investor would find appropriate, or vice versa, the market portfolio

in Propositions 5–10 is to be understood as the portfolio that would be the market portfolio if all

investors in the market were mean-variance optimizers.

PROPOSITION 9. If the investor is a mean-variance optimizer and the composition of the optimal

project portfolio is the same with and without project j, the project’s breakeven buying and

selling prices converge towards the project’s CAPM price

[ ]

[ ]1 10

cov ,

1 var 1j j M M fb s CAPM

j j j jf M f

E C C r E r rv v v C

r r r

⎡ ⎤ ⎡ ⎤ −⎣ ⎦ ⎣ ⎦= = = − + − ⋅+ +

as the investor’s risk tolerance goes to infinity.

PROPOSITION 10. If the investor is a mean-variance optimizer, a project’s breakeven buying and

selling prices converge towards

( )( ) ( )

[ ][ ]

11

10 1

1

cov ,

1 var 1

mm

k k k Mk k kmM fkb s CAPM k

k k kk f M f

C z z rE C z z E r rv v v C z z

r r r==

=

⎡ ⎤′⎡ ⎤ −′− ⎢ ⎥⎣ ⎦ −⎣ ⎦′= = = − − + − ⋅+ +

∑∑∑

as the investor’s risk tolerance goes to infinity. Here, , 1,...,kz k m= , denote the decision

variables for the optimal project portfolio with the project and , 1,...,kz k m′ = those without the

74

Page 83: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

21

project.

Apart from this limit behavior, another case where breakeven buying and selling prices coincide

with CAPM recommendations is when a replicating portfolio exists for the project. This is a

direct result of the fact that each of the three valuation methods is consistent with CCA.

4.5 Valuation of Opportunities and Real Options When valuing a project, we can either value an already started project or an opportunity to start a

project. The difference is that, although the value of a started project can be negative, that of an

opportunity to start a project is always non-negative, because a rational investor does not start a

project with a negative value. While BSP and BBP are appropriate for valuing started projects,

new valuation concepts are needed for valuing opportunities.

Since an opportunity entails the right but not the obligation to take an action, we need selling and

buying prices that rely on comparing the situations where the investor can and cannot invest in

the project, instead of does and does not. The lowest price at which the investor would be willing

to sell an opportunity to start a project can be obtained from the definition of the breakeven

selling price by removing the requirement to start the project in the status quo, i.e. by removing the constraint * 0az = in the top-left quadrant of Table 3. We define this price as the opportunity

selling price (OSP) of the project. Likewise, the opportunity buying price (OBP) of a project can be obtained be removing the starting requirement in the second setting, i.e. the equation * 0az =

in the bottom-right quadrant of Table 3. It is the highest price that the investor is willing to pay

for a license to start the project. Opportunity selling and buying prices have a lower bound of

zero; it is also straightforward to show that the opportunity prices can be computed by taking a

maximum of 0 and the respective breakeven price. Table 4 gives a summary of breakeven and

opportunity selling and buying prices.

Opportunity buying and selling prices can also be used to value real options (Trigeorgis 1996)

that may be contained within the project portfolio. These options result from management’s

flexibility to adapt later decisions to unexpected future developments. Typical examples include

possibilities to expand production when markets are up, to abandon a project under bad market

conditions, and to switch operations to alternative production facilities. Real options can be

valued much in the same way as opportunities to start projects. However, instead of comparing

75

Page 84: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

22

portfolio selection problems with and without a possibility to start a project, we will compare

portfolio selection problems with and without the real option. This can typically be implemented

by disallowing the investor from taking a particular action (e.g. expanding production) in the

setting where the real option is not present. Since breakeven prices are consistent with CCA, also

opportunity prices have this property, and hence they can be regarded as a generalization of the

standard CCA real option valuation procedure to incomplete markets.

Table 4. A summary of an investor’s buying and selling prices.

Valuation concept Idea Difference from breakeven prices

Properties

Breakeven buying and selling prices

Comparison of settings where the investor does and does not invest in the project.

- Breakeven buying price and selling price are in general different from each other.

Opportunity buying and selling prices

Comparison of settings where the investor can and cannot invest in the project.

The investor is not obliged to invest in the project.

Lower bound of 0. Equal to maximum of 0 and the respective breakeven price.

5 Numerical Experiments In this section, we demonstrate the use of a MAPS model in project valuation through a series of

numerical experiments. We employ the multi-period MAPS model but construct only a single-

period model, because this allows creating several insights, is easier to follow and replicate, and

because we can contrast the results with the CAPM, also a single-period model. We have also

conducted further experiments featuring real options and a multi-period MAPS model in two

working papers (De Reyck et al. 2004, and Gustafsson and Salo 2004). The present experiments

confirm some of the theoretical results obtained in the previous sections and also cast light on the

following issues:

1 How are project values affected by the presence of other projects in the portfolio?

2 How are project values affected by the opportunity to invest in securities?

3 How are project values affected by the investor’s risk tolerance (maximum level for

risk)?

4 How are project values related to the CAPM?

5 How does the presence of twin securities affect the value of a project?

5.1 Experimental Set-up The experimental set-up includes 8 equally likely states of nature, four projects, A, B, C, and D

(Table 5), and two securities, 1 and 2, which together constitute the market portfolio (Table 6).

76

Page 85: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

23

The setting can be extended to include more securities, but for the sake of simplicity we limit our

market portfolio to two assets only. This does not influence the generality of our results. Note that

projects C and D and security 2 are the same that Smith and Nau (1995) use in their examples.

Table 7 shows the correlation between the assets’ cash flows. Numbers in italic in Tables 5 and 6

are computed values. The risk-free interest rate is 8%.

Table 5. Projects.

Project A B C D Investment cost $80 $100 $104 $0.00 State 1 $150 $140 $180 $67.68 State 2 $150 $140 $180 $67.68 State 3 $150 $150 $60 $0.00 State 4 $150 $110 $60 $0.00 State 5 $50 $170 $180 $67.68 State 6 $50 $100 $180 $67.68 State 7 $50 $90 $60 $0.00 State 8 $50 $90 $60 $0.00 Expected outcome $100 $123.75 $120 $33.84 St. dev. of outcome $50 $28.26 $60 $33.84 Beta 0.000 0.431 1.637 3.683 Market price (NPV) $12.59 $11.33 -$4.00 $25.07

The market prices in Tables 5 and 6 are obtained by using the CAPM, where the expected rate of

return of the market portfolio has been chosen so that the price of security 2 is $20, the price used

by Smith and Nau (1995). Technically, this can be accomplished by including all of the assets

into the market portfolio, with projects having zero issued shares, and by finding the market

prices, excluding projects’ investment costs, that minimize the sum of squared errors (SSE)

between the rate of return given by the CAPM formula and the real expected rate of return, as

computed from the market price. The prices converge, resulting in SSE equal to zero. The desired

expected rate of return is 15.33%. The standard deviation of the market portfolio is then 35.32%.

In Table 6, the security capitalization weights represent the ratio between the market

capitalization of the security and that of the entire market. These weights are of interest, because

the CAPM predicts that an MV investor will always invest in a combination of the risk-free asset

and the market fund, where securities are present according to their capitalization weights.

77

Page 86: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

24

Table 6. Securities.

Security 1 2 Market price $39.56 $20.00 Shares issued 15,000,000 10,000,000 Capitalization weight 74.79% 25.21% State 1 $60 $36 State 2 $50 $36 State 3 $40 $12 State 4 $30 $12 State 5 $60 $36 State 6 $50 $36 State 7 $40 $12 State 8 $30 $12 Beta 0.79 1.64 Expected return 13.76% 20.00% St. dev. of return 28.26% 60.00%

Table 7. Correlation matrix.

A B C D 1 2 A 1 0.398 0 0 0 0 B 0.398 1 0.487 0.487 0.653 0.487 C 0 0.487 1 1 0.894 1 D 0 0.487 1 1 0.894 1 1 0 0.653 0.894 0.894 1 0.894 2 0 0.487 1 1 0.894 1

The experiment comprises several steps. We start with a setting where the investor can invest

only in the project being valued, which is then extended to cover a portfolio of all projects. We

then add securities 1 and 2. Apart from the very first case, we assume that limitless borrowing at

the risk-free rate and shorting of securities are allowed. As predicted by Proposition 4, the

breakeven selling and buying prices are identical in this case and hence we display both of the

prices in one entry. Unless otherwise noted, in each of the steps we use a budget of $500. The

maximum risk level is defined in terms of the rate of return of the portfolio. For example, a 50%

risk level implies that the maximum standard deviation for the terminal wealth level is $250.

5.2 Project Values without Securities In general, the value of a project depends on what would happen to the invested funds if the

project were not started. In the absence of other projects and securities, three cases described in

Table 8 are possible. These cases both provide a benchmark for project valuations obtained later,

78

Page 87: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

25

and show that even apparently small differences in available investment opportunities can make a

significant difference in the values of projects. In the first case, where unused funds are lost, the

net present value of the project is undefined, and hence we give the future value (FV) of the

project instead.

Table 8. Projects values when only the project being valued is available.

Project What happens to unused funds: A B C D Funds are lost (FV) $100.00 $123.75 $120.00 $33.84 No interest gained $20.00 $23.75 $16.00 $33.84 Deposited at risk-free interest rate $12.59 $14.58 $7.11 $31.33

Let us next assume that the investor is able to invest in all of the projects and in the risk-free

asset. The values of the projects are described in Table 9 as a function of the investor’s risk

tolerance. In the optimal policy, the investor starts the projects with positive value.

Table 9. Projects values when all of the projects are available.

Mean-Variance Risk level A B C D 15% -$1.99 $1.99 -$38.81 $18.74 20% $12.59 $14.58 -$20.06 $24.22 25% $5.48 $7.47 -$5.48 $24.22 30% and up $12.59 $14.58 $7.11 $31.33

-$60.00

-$40.00

-$20.00

$0.00

$20.00

$40.00

12% 15% 18% 21% 24% 27% 30%

Risk level

Proj

ect v

alue A

B

C

D

Figure 2. Project values for MV investor.

Table 9 and Figure 2 show that the project values behave rather erratically when the risk tolerance

79

Page 88: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

26

is varied. For example, at a 20% risk level, the MV values for projects A and B are identical to

the values they have in isolation ($12.59 and $14.58), but then they drop to $5.48 and $7.47 at

25% and rise back to the values in isolation for higher risk tolerance levels. Intuitively, one might

expect project values to rise as the risk constraint is relaxed, because more projects can be

included into the portfolio so that fewer projects impose an opportunity cost. At the limit, this is

correct: when the risk constraint is relaxed enough to allow the inclusion of all profitable projects,

the project values coincide with the prices on the last row of Table 8. For intermediate risk

values, however, another effect confounds this result: the value of a project depends on the

projects that fit into the portfolio when the project is started and when it is not. For example, the

decrease of the price of projects A and B when the risk constraint is increased from 20% to 25%

is due to the fact that, when the risk constraint is 25%, project C can be started if either of the

projects is not included into the portfolio, but this is not the case at 20%. So at a 25% risk

constraint, project C imposes an opportunity cost on these projects, but not at 20%, since the

project does not fit into the portfolio regardless the decision on projects A and B. This

counterintuive result is typical for project values in a MAPS setting.

5.3 Project Values with Securities We will now investigate what happens to the project values when securities are also available. In

this case, the risk constraint will always be binding as long as the rate of return of the optimal

security portfolio for the investor is higher than the risk-free interest rate: expectation of the

whole portfolio will be maximized by purchasing as much of the optimal security portfolio as

possible and by borrowing the necessary funds at the risk-free interest rate. When the investor

does not invest in the projects, i.e. in a pure CAPM setting, the optimal security portfolio for a

MV investor is to buy 75% of security 1 and 25% of security 2 at all risk levels. Note that these

weights are the capitalization weights of the securities in Table 6.

Columns 2–5 of Table 10 describe the project values assuming that the investor can only invest in

the project being valued and in securities. The presence of securities changes the project values

for two reasons. On the one hand, securities impose an additional opportunity cost, which lowers

project values. On the other hand, it is possible to hedge against project risks by buying

negatively correlated or shorting positively correlated securities, which increases the project

values.

80

Page 89: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

27

Table 10. Project values when securities 1 and 2 are available.

Single project available All projects available Risk level A B C D A B C D 15% $8.92 $10.78 -$4.00 $25.07 $6.36 $8.22 -$4.00 $25.07 20% $10.02 $10.92 -$4.00 $25.07 $8.61 $9.51 -$4.00 $25.07 30% $10.94 $11.06 -$4.00 $25.07 $10.14 $10.26 -$4.00 $25.07 40% $11.37 $11.13 -$4.00 $25.07 $10.80 $10.55 -$4.00 $25.07 50% $11.62 $11.17 -$4.00 $25.07 $11.17 $10.72 -$4.00 $25.07 60% $11.79 $11.19 -$4.00 $25.07 $11.42 $10.82 -$4.00 $25.07 70% $11.90 $11.21 -$4.00 $25.07 $11.59 $10.90 -$4.00 $25.07 80% $11.99 $11.23 -$4.00 $25.07 $11.72 $10.95 -$4.00 $25.07 90% $12.06 $11.24 -$4.00 $25.07 $11.81 $11.00 -$4.00 $25.07 100% $12.11 $11.25 -$4.00 $25.07 $11.89 $11.03 -$4.00 $25.07 200% $12.35 $11.29 -$4.00 $25.07 $12.24 $11.18 -$4.00 $25.07 500% $12.50 $11.31 -$4.00 $25.07 $12.45 $11.27 -$4.00 $25.07 2000% $12.57 $11.32 -$4.00 $25.07 $12.56 $11.31 -$4.00 $25.07 10000% $12.59 $11.33 -$4.00 $25.07 $12.59 $11.33 -$4.00 $25.07

Since project A is uncorrelated with the market portfolio, we can use its values to verify

Propositions 6, 7, and 8. For example, let us calculate the breakeven buying price for project A at

50% risk level. The variance of the project portfolio with the project (including only project A) is

502 = 2500 and 0 without the project. The variance of the optimal security portfolio in the status

quo (no projects) is 2502 = 62500 and the amount of money spent on securities is $708.3. As the

market rate of return is 15.33%, we get from Proposition 7 that

100 2500 0.073380 1 1 708.3 12.59 0.0202 48.07 11.621.08 62500 1.08

bAv

⎛ ⎞= − + − − − = − ⋅ =⎜ ⎟⎜ ⎟

⎝ ⎠,

which coincides with the value in Table 10. Also, as predicted by Proposition 9, when the risk

tolerance approaches infinity, the project values converge to their CAPM prices, as shown in the

row “Market price” in Table 6. Intuitively, one may expect that the values would approach risk-

neutral values as the risk-constraint is being relaxed more and more, but this is not the case.

Notice that security 2 is a twin security for projects C and D. This explains why these projects are

priced constantly to their CCA value across all risk levels. To replicate the cash flows of project

C, one needs 5 shares of security 2, costing $100. Since the investment cost of project C is $104,

we obtain the value of –$4 for the project, regardless of the investor’s risk tolerance. The cash

flows of project D can be replicated by constructing a portfolio of 2.82 shares of security 2 and

borrowing of $31.33 at the risk-free interest rate. Therefore, the value of this project is

$0 2.82 $20 31.33 $1 $25.07− + ⋅ − ⋅ = .

81

Page 90: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

28

Let us next investigate how the presence of other projects, as well as that of securities, affects

project values. Columns 6–9 of Table 10 give the results. First, we observe that the values of

projects A and B still increase monotonically with increasing risk tolerance without any erratic

behavior. We see also that the values still converge to CAPM prices. Additionally, the values of

projects A and B decrease in comparison with the single-project case, because the optimal

security portfolio with these projects is less profitable than what it was with a single project only.

5.4 Summary of Experimental Results Table 11 summarizes our findings with respect to the research questions posed at the beginning of

this section. Further, we have results on the valuation of real options and multi-period projects in

two working papers, De Reyck et al. (2004) and Gustafsson and Salo (2004). The results indicate

that (i) the present approach can readily be extended to the valuation of real options in incomplete

markets, and that (ii) it will also be practically feasible in a multi-period setting. In addition, when

valuing a single project, the present approach will yield the same results as Smith and Nau’s

(1995) method, provided that the assumptions of Smith and Nau’s method are satisfied and the

investor’s preference models in both approaches are identical.

82

Page 91: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

29

Table 11. Summary of research questions.

Research question Answer Influence of other projects on project values Alternative projects can lower project values by imposing an opportunity cost.

Influence of securities on project values

Securities can both lower and raise project values. On the one hand, they impose an opportunity cost. On the other hand, they enable better diversification of risk.

Influence of risk tolerance on project values

When securities are not available, project values may rise non-monotonically as the risk tolerance rises. When securities are available, project values for MV investor may rise (or decrease, see De Reyck et al. 2004) monotonically by risk tolerance.

Relationship to CAPM prices For an MV investor, project values converge towards CAPM prices as the risk tolerance goes to infinity.

Influence of twin securities on project values

A twin security makes the MAPS project value consistent with the project’s CCA value at all risk levels, provided that limitless borrowing is allowed. As ordinary securities for other projects, twin securities also influence values of other projects.

6 Summary and Conclusions In this paper, we analyzed the valuation of private projects in a mixed asset portfolio selection

(MAPS) setting, where an investor can invest in a portfolio of projects as well as securities traded

in financial markets, but where the replication of project cash flows with financial securities is not

necessarily possible. We developed a valuation procedure based on the concepts of breakeven

selling and buying prices. This inverse optimization procedure requires the solution of portfolio

selection problems with and without the project being valued and finding a lump sum that makes

the investor indifferent between the two situations. To make the solution of these portfolio

selection problems possible in a multi-period setting, we developed a multi-period MAPS model

using the Contingent Portfolio Programming framework (Gustafsson and Salo 2005). Finally, we

produced several theoretical results relating to the analytical properties of breakeven prices, and

studied the pricing behavior of mean-variance (MV) investors through a set of numerical

experiments.

Our theoretical results indicate that the breakeven prices are, in general, consistent valuation

measures, exhibiting sequential consistency, consistency with contingent claims analysis, and

sequential additivity. The results also show that MV investors have several notable pricing

properties. First, when limitless borrowing and lending are allowed, breakeven buying and selling

prices are identical, which is not the case, in general, under expected utility theory (Smith and

Nau 1995, Raiffa 1968). Also, the prices can be computed by solving the investor’s terminal

83

Page 92: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

30

wealth level when he/she invests and he/she does not invest in the project and by discounting the

difference back to its present value at the risk-free interest rate. In addition, we derived analytical

formulas to calculate the breakeven prices for an MV investor when the optimal portfolio at

present is known and projects are uncorrelated with securities. Finally, we showed that the project

values given by an MV investor converge towards the projects’ CAPM prices as the investor’s

risk tolerance goes to infinity.

Overall, our results suggest that alternative investment opportunities have a significant impact on

the value of a project. Therefore, valuation of projects in isolation may potentially lead to biased

estimates of the values of projects. This emphasizes the argument, which appears for example in

the real options literature, that it is crucial to also consider financial instruments in project

valuation. However, it is also important to recognize other projects in the portfolio.

Managers can draw several conclusions from our analysis. First, it is of central importance to

clearly recognize the real investment alternatives to the projects. The presence of the possibility

to borrow or short, or in general to invest in financial markets, may significantly influence the

value of a project. Second, project values are non-additive. To obtain the value of a portfolio of

two or more projects, it is necessary to calculate the terminal wealth levels when the investor

starts and does not start the projects. This can be accomplished by solving the appropriate

portfolio selection problems. Third, projects influence, in general, the optimal financial portfolio

for the investor, so that when the projects are started, the optimal financial portfolio does not

typically fall on the capital market line. Therefore, the allocation of funds to projects and

securities separately may potentially lead to suboptimally diversified portfolios.

References Ahuja, R. K., J. B. Orlin. 2001. Inverse optimization. Operations Research 49 771-783.

Belegundu, A. D., T. R. Chandrupatla. 1999. Optimization Concepts and Applications in

Engineering. Prentice Hall.

Brealey, R., S. Myers. 2000. Principles of Corporate Finance. McGraw-Hill, New York, NY.

Clemen, R. T. 1996. Making Hard Decisions – An Introduction to Decision Analysis. Duxbury

Press, Pacific Grove.

De Reyck, B., Z. Degraeve, J. Gustafsson. 2004. Valuing Real Options in Incomplete Markets.

Working Paper. London Business School.

84

Page 93: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

31

Dixit, A. K., R. S. Pindyck. 1994. Investment Under Uncertainty. Princeton University Press,

Princeton.

Eppen G. D., Martin R. K., Schrage L. 1989. A Scenario Based Approach to Capacity Planning.

Oper. Res. 37 517-527.

Fishburn, P. C. 1977. Mean-Risk Analysis with Risk Associated with Below-Target Returns.

Amer. Econ. Rev. 67(2) 116–126.

French, S. 1986. Decision Theory – An Introduction to the Mathematics of Rationality. Ellis

Horwood Limited, Chichester.

Gustafsson, J., A. Salo. 2004. Valuing Risky Projects with Contingent Portfolio Programming.

Working Paper. Helsinki University of Technology.

––––––. 2005. Contingent Portfolio Programming for the Management of Risky Projects.

Operations Research. Forthcoming.

Konno, H., H. Yamazaki. 1991. Mean-Absolute Deviation Portfolio Optimization and Its

Applications to the Tokyo Stock Market. Management Sci. 37(5) 519–531.

Lintner, J. 1965. The Valuation of Risk Assets and the Selection of Risky Investments in Stock

Portfolios and Capital Budgets. Rev. Econ. Statist. 47(1) 13–37.

Luenberger, D. G. 1998. Investment Science. Oxford University Press, New York, NY.

Markowitz, H. M. 1952. Portfolio Selection. J. Finance 7(1) 77–91.

–––––. 1959. Portfolio Selection: Efficient Diversification of Investments. Cowles Foundation,

Yale.

–––––. 1987. Mean-Variance Analysis in Portfolio Choice and Capital Markets. Frank J. Fabozzi

Associates, New Hope, PA.

Microsoft. 2004. Annual Report 2004. Available at http//www.microsoft.com.

Mulvey, J. M., G. Gould, C. Morgan. 2000. An Asset and Liability Management Model for

Towers Perrin-Tillinghast. Interfaces 30(1) 96–114.

Ogryczak, W., A. Ruszczynski. 1999. From stochastic dominance to mean-risk models:

Semideviations as risk measures. European J. Oper. Res. 116(1) 33–50.

Raiffa, H. 1968. Decision Analysis – Introductory Lectures on Choices under Uncertainty.

Addison-Wesley Publishing Company, Reading, MA.

Rubinstein, M. E. 1973. A Mean-Variance Synthesis of Corporate Financial Theory. J. Finance

28(1) 167–181.

Sharpe, W. F. 1964. Capital Asset Prices: A Theory of Market Equilibrium under Conditions of

Risk. J. Finance 19(3) 425–442.

–––––. 1970. Portfolio Theory and Capital Markets. McGraw-Hill, New York, NY.

85

Page 94: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

32

Smith, J. E., R. F. Nau. 1995. Valuing Risky Projects: Option Pricing Theory and Decision

Analysis. Management Sci. 41(5) 795–816.

Tobin, J. 1958. Liquidity Preference as Behavior Towards Risk. Rev. Econ. Studies 25 124-131.

Trigeorgis, L. 1996. Real Options: Managerial Flexibility and Strategy in Resource Allocation.

MIT Press, Massachussets.

Yu, P.-L. 1985. Multiple-Criteria Decision Making: Concepts, Techniques, and Extensions.

Plenum Press, New York, NY.

Appendix PROOF OF PROPOSITION 1: Let the breakeven selling price for the project be s

jv . The time-0

budget used to determine the sequential buying price is then sjb v+ in the status quo and

s sbj jb v v+ − in the second setting, where sb

jv is the sequential buying price. Let s sW W− += be the

optimal value for the optimization problems with the breakeven selling price and sb sbW W− += be

the optimal values for the optimization problems with the sequential buying price. By definition,

sbW − = sW − and therefore also sbW + = sW + . Thus, it follows that s sbj jb v v b+ − = , or s sb

j jv v= . A

similar reasoning shows the equality between the breakeven buying price and the respective

sequential selling price. Q.E.D.

PROOF OF PROPOSITION 3: Let us begin with the breakeven buying price and a setting where the

portfolio does not include projects and the budget is b. Suppose that there are n projects that the

investor buys sequentially. Let us denote the optimal value for this MAPS problem by W0.

Suppose then that the investor buys a project, indexed by 1, at his or her breakeven buying price,

1bv . Let us denote the resulting optimal value for the MAPS problem by W1. By definition of the

breakeven buying price, W0 = W1. Suppose then that the investor buys another project, indexed by 2, at his or her breakeven buying price, 2

bv . The initial budget is now 1bb v− , and after the second

project is bought, it is 1 2b bb v v− − . By definition, the optimal value for the resulting MAPS

problem W12 is equal to W1. Add then the rest of the projects in the same manner. The resulting budget in the last optimization problem, which includes all the projects, is 1 2 ...b b b

nb v v v− − − − .

We also know that W12…n = W12…n–1 = … = W0, and therefore, by definition, the breakeven buying price of the portfolio including all the projects is 12... 1 2 ...b b b b

n nv v v v= + + + . By re-indexing the

projects and using the above procedure, we can change the order in which the projects are added

to the portfolio. In doing so, the projects can obtain different values, but they still sum up to the

same joint value of the portfolio. Similar logic proves the proposition for breakeven selling

86

Page 95: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

33

prices. Q.E.D.

PROOF OF PROPOSITION 4: Let us first examine the breakeven selling price. Let us denote the optimal expectation for the given risk constraint in the status quo with s

SQµ and the one in the

second setting when 0sv = with s s sSS SQµ µ= + ∆ . The expectations can be made to match by

increasing the time-0 budget by sδ , which reduces the amount of money borrowed by the same

amount, leaving the effective budget unchanged. This increases the expectation of the portfolio

by ( )1

0

1 ( )T

sf

t

r tδ−

=

+∏ , where rf(t) is the risk-free interest rate from time t to t+1. This chain of

logic is valid on the condition that increasing the budget does not influence the composition of the

optimal risky portfolio, which is ensured when the risk measure is independent of an addition of a constant, i.e. it satisfies [ ] [ ]X b Xρ ρ+ = for all random variables X and constants b. Since we

desire to make the expectations in the status quo and in the second setting under a modified budget (denoted by s

SSµ ∗ ) equal, we have the equation

( )1

0

1 ( )T

s s s s s sSQ SS SQ SQ f

t

r tµ µ µ µ δ−

=

= ⇔ = + ∆ + +∏ , from which it follows that

( )1

0

/ 1 ( )T

s sf

t

r tδ−

=

= −∆ +∏ , which is also the breakeven selling price of the project.

Let us then examine the breakeven buying price and denote the optimal expectation in the status quo by b

SQµ and the one in the second setting when 0bv = by b b bSS SQµ µ= + ∆ . As before, the

expectations can be matched by lowering the budget by bδ and increasing the amount of money

borrowed by the same amount. The expectation in the second setting is now

( )1

0

1 ( )T

b b b bSS SQ f

t

r tµ µ δ−

=

= + ∆ − +∏ . By requiring that

(1 )b b b b b bSQ SS SQ SQ frµ µ µ µ δ∗= ⇔ = + ∆ − + , we get ( )

1

0

/ 1 ( )T

b bf

t

r tδ−

=

= ∆ +∏ , which is also the

breakeven buying price of the project.

Finally, we observe that s b

SS SQµ µ= and b sSS SQµ µ= , from which it follows that

s b s s b b s bSS SQ SQ SSµ µ µ µ= ⇔ +∆ = −∆ ⇔ ∆ = −∆ . Hence, .s bδ δ= To obtain the valuation

formula, let us denote bSS Wµ += and b

SQ Wµ −= , and we get the formula directly from the

equation for the breakeven buying price. Q.E.D.

87

Page 96: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

34

PROOF OF PROPOSITION 5: Since projects are uncorrelated with securities, we know that

1 1 1 1

0 1 0 1var var var

n m n m

i i k k i i k ki k i k

S x C z S x C z= = = =

⎡ ⎤ ⎡ ⎤ ⎡ ⎤+ = +⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦ ⎣ ⎦∑ ∑ ∑ ∑ and hence the risk constraint can be

rewritten as 1 1

0 1var var

n m

i i k ki k

S x R C z= =

⎡ ⎤ ⎡ ⎤≤ −⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦∑ ∑ . Similarly, the objective can rewritten as

1 1 1 1

0 1 0 1max max

n m n m

i i k k i i k ki k i k

E S x C z E S x E C z= = = =

⎡ ⎤ ⎡ ⎤ ⎡ ⎤+ ⇔ +⎢ ⎥ ⎢ ⎥ ⎢ ⎥⎣ ⎦ ⎣ ⎦ ⎣ ⎦∑ ∑ ∑ ∑ . For any project portfolio with

fixed , 1,...,kz k m∗ = , the optimal financial portfolio will be acquired by solving an optimization

problem 1

0max

n

i ii

E S x a=

⎡ ⎤ +⎢ ⎥⎣ ⎦∑ subject to 0 0

0 1

n m

i i k ki k

S x C z b∗

= =

+ =∑ ∑ and 1

0var

n

i ii

S x R∗

=

⎡ ⎤ ≤⎢ ⎥⎣ ⎦∑ ,

where 1

1

m

k kk

a E C z∗=

⎡ ⎤= ⎢ ⎥⎣ ⎦∑ is a constant and 1

1var

m

k kk

R R C z∗ ∗

=

⎡ ⎤= − ⎢ ⎥⎣ ⎦∑ is a constant. The parameter

b does not influence the xi’s that yield the maximum to the problem, and R* is just an adjusted

maximum risk level for the investor. Thus, for any given project portfolio, the optimal financial

portfolio is obtained by the solution of a usual mean-variance problem. We know from the

Separation Theorem (Tobin 1958) that a mean-variance investor will invest in the combination of

the market fund and the risk-free asset. Q.E.D.

PROOF OF PROPOSITIONS 6, 7, AND 8: Let us prove first Proposition 8. Let the project portfolio

in the status quo be 1

1

m

k kk

P C z=

= ∑ and the security portfolio 1

1

n

i ii

M S x=

= ∑ . The respective

portfolios in the second setting are 1

1

m

k kk

P C z=

′ ′= ∑ and 1

1

n

i ii

M S x=

′ ′= ∑ . The amount of money

lent in the status quo is denoted by 0x and the respective amount in the second setting by 0x′ . By

Proposition 4 we know that the portfolio weights in the security portfolios M and M’ are identical, i.e., that / , 1,...,i ix x a i n′ = = , where a is the positive ratio of funds invested in risky securities in

the second setting and in the status quo. That is, 0

01

00

1

m

k kkm

k kk

b x C za

b x C z

=

=

′ ′ ′− −=

− −

∑, (A1)

where b′ is the budget in the second setting. This implies that M aM′ = . Let us require next

that

88

Page 97: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

35

0 0

0 0

(1 ) (1 )

var (1 ) var (1 )

f f

f f

E r x M P E r x M P

r x M P r x M P

′ ′ ′⎡ ⎤ ⎡ ⎤+ + + = + + +⎣ ⎦ ⎣ ⎦′ ′ ′⎡ ⎤ ⎡ ⎤+ + + = + + +⎣ ⎦ ⎣ ⎦

By using the relationship M aM′ = , the equivalence of expectations yields us [ ] [ ] [ ] [ ]0 0(1 ) (1 )f fr x E M E P r x aE M E P′ ′+ + + = + + +

[ ] [ ] [ ]0 0

(1 )1 f

a E M E P E Px x

r′− + −

′⇒ = ++

(A2)

Similarly, from the equivalence of variances we obtain [ ] [ ] [ ] [ ]2var var var varM P a M P′+ = +

[ ] [ ][ ]

var var1

varP P

aM

′−⇒ = ± + (A3)

The minus sign alternative can be dropped, because we know that a has to be positive. Let us

define b b′∆ = − . By the definition of a in (A1), we get

0 00 0

1 1

m m

k k k kk k

b b a b x C z x C z b= =

⎛ ⎞′ ′ ′∆ = − = − − + + −⎜ ⎟⎝ ⎠

∑ ∑ .

By substituting (A2) into this, we obtain [ ] [ ] [ ]0 0

0 01 1

(1 )1

m m

k k k kk kf

a E M E P E Pb b a b x C z x C z b

r= =

′− + −⎛ ⎞′ ′∆ = − = − − + + + −⎜ ⎟ +⎝ ⎠∑ ∑ .

By denoting 00

1

m

M k kk

b b x C z=

= − −∑ and (1 )M MM r b= + , we get

[ ]( ) [ ] [ ] 00

1

(1 ) 11

mM M

M k kkf

a E r B E P E Pab x C z b

r =

′− + + −′∆ = + + + −

+ ∑ .

Let us then add and subtract 0

1

m

k kk

C z=∑ from the right-hand side of the equation to obtain

00

1

m

M k kk

b b x C z=

= − −∑ on the right-hand side. Thus we get

[ ]( ) [ ] [ ] 0 0

1 1

(1 ) 1( 1)

1

m mM M

M k k k kk kf

a E r b E P E Pa b C z C z

r = =

′− + + −′∆ = − + + −

+ ∑ ∑ .

By adding and subtracting rf inside [ ]( )1 ME r+ we are able to eliminate the term (a – 1)bM from

the beginning and obtain [ ]( ) [ ] [ ] 0 0

1 1

(1 )1

m mM f M

k k k kk kf

a E r r b E P E PC z C z

r = =

′− − + −′∆ = + −

+ ∑ ∑ .

By denoting [ ] 1

1

m

k kk

E P E C z=

⎡ ⎤= ⋅⎣ ⎦∑ and [ ] 1

1

m

k kk

E P E C z=

⎡ ⎤′ ′= ⋅⎣ ⎦∑ and by rearranging the terms,

89

Page 98: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

36

we get

[ ]

1

0 1

1

( )( ) ( 1)

1 1

m

k k kmM fk

k k k Mk f f

E C z z E r rC z z a b

r r=

=

⎡ ⎤ ′⋅ −⎣ ⎦ −′∆ = − − + − −

+ +

∑∑ . (A4)

Finally, by substituting (A3) into (A4), we obtain the desired formula

[ ]1 11

1 10 1

1 1

1

var var( )( ) 1 1

1 1var

m mm

k k k kk k kmM fk kk

k k k Mnk f f

i ii

C z C zE C z z E r rC z z b

r rS x

= ==

=

=

⎛ ⎞⎡ ⎤ ⎡ ⎤′⎡ ⎤ −⎜ ⎟′⋅ − ⎢ ⎥ ⎢ ⎥⎣ ⎦ −⎣ ⎦ ⎣ ⎦⎜ ⎟′∆ = − − + − + −⎜ ⎟+ +⎡ ⎤⎜ ⎟⎢ ⎥⎜ ⎟⎣ ⎦⎝ ⎠

∑ ∑∑∑

∑.

Proposition 6 is obtained by assuming that k kz z k j′= ∀ ≠ and 1jz = and 0jz′ = . We get now

[ ]1 1

1 1 10

1

1

var var

1 11 1var

m m

k k k kk k

j k j M fjs j Mn

f fi i

i

C z C zE C E r r

v C br rS x

= =≠

=

⎛ ⎞⎡ ⎤⎜ ⎟⎡ ⎤ ⎢ ⎥−⎜ ⎟⎢ ⎥ ⎢ ⎥⎣ ⎦⎡ ⎤ ⎜ ⎟ −⎢ ⎥⎣ ⎦ ⎣ ⎦= ∆ = − + − + −⎜ ⎟

+ +⎡ ⎤⎜ ⎟⎢ ⎥⎜ ⎟⎣ ⎦⎜ ⎟⎜ ⎟

⎝ ⎠

∑ ∑

∑.

Similarly, Proposition 7 is obtained by assuming that k kz z k j′= ∀ ≠ and 0jz = and 1jz′ = . Then,

[ ]1 1

1 1 10

1

1

var var

1 11 1var

m m

k k k kk k

j k j M fjb j Mn

f fi i

i

C z C zE C E r r

v C br rS x

= =≠

=

⎛ ⎞⎡ ⎤⎜ ⎟⎡ ⎤ ⎢ ⎥−⎜ ⎟⎢ ⎥ ⎢ ⎥⎣ ⎦⎡ ⎤ ⎜ ⎟ −⎢ ⎥⎣ ⎦ ⎣ ⎦= −∆ = − + − − −⎜ ⎟

+ +⎡ ⎤⎜ ⎟⎢ ⎥⎜ ⎟⎣ ⎦⎜ ⎟⎜ ⎟

⎝ ⎠

∑ ∑

∑.

Q.E.D.

PROOF OF PROPOSITIONS 9 AND 10: Let us prove first Proposition 10. The notation is as in the

proof of Propositions 6–8. Observe first that when the investor’s risk tolerance increases, she

invests more and more in securities and the effect of projects on the total mixed asset portfolio

diminishes. At the limit, the investor invests in an infinitely large security portfolio whose

composition matches that of the market portfolio regardless of what projects there are in the

portfolio. Therefore, the portfolio weights in the security portfolios M and M’ are identical, so

that M aM′ = , where a is some constant. Now, we have by definition

90

Page 99: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

37

0 0var (1 ) var (1 )f fr x M P r x M P′ ′ ′⎡ ⎤ ⎡ ⎤+ + + = + + +⎣ ⎦ ⎣ ⎦ ,

from which we obtain [ ] [ ] [ ] [ ] [ ] [ ]2var var 2cov , var var 2 cov ,M P M P a M P a M P′ ′+ + = + +

[ ] [ ] [ ] [ ] [ ] [ ]2 var 2 cov , var var var 2cov , 0a M a M P M P P M P′ ′⇒ + − + − − = .

Solving this for a gives

[ ] [ ] [ ] [ ] [ ] [ ] [ ]( )[ ]

[ ][ ]

[ ][ ]

[ ] [ ] [ ] [ ]( ) [ ] [ ][ ]

2

22

2 2

2cov , 4cov , 4 var var var var 2cov ,

2 var

var var var var 2 var cov ,cov , cov ,var var var

M P M P M M P P M Pa

M

M M P P M M PM P M PM M M

′ ′ ′− ± + − + +⇒ =

′+ − +′ ′−= ± +

[ ][ ]

[ ][ ]

[ ] [ ][ ]

[ ][ ]

[ ][ ]

[ ][ ]

[ ][ ]

[ ][ ]

[ ][ ]

[ ] [ ][ ]

[ ][ ]

[ ][ ]

2, ,

2 2 2, , , ,

2

,

varcov , var var1 2

var var var

var var varcov , var var2 1

var var var var var

(cov ,1

var

M P M P

M P M P M P M P

M P

P stdev PM P P PM M M stdev M

P stdev P P PM P P PM M stdev M M M M

stdev PM PM stdev M

ρ ρ

ρ ρ ρ ρ

ρ ρ

′′ ′− −= ± + + +

′′ ′− −= ± + + − + +

⎛ ⎞′−= ± + +⎜ ⎟⎜ ⎟

⎝ ⎠

[ ][ ]

[ ][ ]

[ ]

[ ][ ]

[ ][ ]

[ ][ ]

2 2, ,

var

1) var ( 1) varvar var

cov , cov , cov ,1 1

var var var

M P M P

M

P PM M

M P M P M P PM M M

ρ′

→∞

′− −−

′ ′− −→ + + = +

The minus sign alternative can be dropped, because we know that a has to be positive. Now let us

substitute this into Equation (A4) in the proof of Propositions 6–8 to obtain

[ ]

[ ]11

10 1

1

cov , ( )( )( )

1 var 1

mm

k k kk k kmM fkk

k k k Mk f f

M C z zE C z z E r rC z z b

r M r==

=

⎡ ⎤′⎡ ⎤ ⋅ −′⋅ − ⎢ ⎥⎣ ⎦ −⎣ ⎦′∆ = − − + −+ +

∑∑∑ . (A5)

For the BSP of project j, we have 1jz = and 0jz′ = . Since ∆ is the budget increment required to

make two portfolios equally desirable, ∆ is, by definition, the BSP of the project. Formula for the

BBP, which is a budget reduction, is obtained by multiplying both sides of Equation (A5) by –1.

Finally, since (1 )M MM r b= + , we obtain the CAPM formula

[ ][ ]

11

10 1

1

cov , ( )( )( )

1 var 1

mm

M k k kk k kmM fkCAPM k

k k kk f M f

r C z zE C z z E r rv C z z

r r r==

=

⎡ ⎤′⎡ ⎤ ⋅ −′⋅ − ⎢ ⎥⎣ ⎦ −⎣ ⎦′= ∆ = − − + −+ +

∑∑∑ . (A6)

91

Page 100: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

38

Note that the definition of iz ’s and iz′ ’s in Proposition 10 is different from the definition in

Proposition 8. Here, Proposition 10 for the BSP case can be obtained directly using Equation

(A6). The BBP case is obtained by changing the notation so that iz′ indicates the status quo (no

project) and iz indicates the second setting (project present). Proposition 9 is obtained for the

BSP case from (A6) by assuming that k kz z k j′= ∀ ≠ and 1jz = and 0jz′ = . Proposition 9

for the BBP case is again obtained by changing the meaning of iz′ (to no project) and iz (to

project present) and assuming that k kz z k j′= ∀ ≠ and 1jz = and 0jz′ = . Q.E.D.

Acknowledgements We would like to thank an anonymous referee and an anonymous associate editor for the

comments on an earlier version of this paper. Grants from the Finnish Cultural Foundation, the

Foundation of Jenny and Antti Wihuri, and the Foundation of Emil Aaltonen are gratefully

acknowledged.

92

Page 101: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

Helsinki University of Technology Systems Analysis Laboratory Research Reports E17, June 2005

PROJECT VALUATION UNDER AMBIGUITY

Janne Gustafsson Ahti Salo

AB TEKNILLINEN KORKEAKOULUTEKNISKA HÖGSKOLANHELSINKI UNIVERSITY OF TECHNOLOGYTECHNISCHE UNIVERSITÄT HELSINKIUNIVERSITE DE TECHNOLOGIE D’HELSINKI

93

Page 102: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

Title: Project Valuation under Ambiguity Authors: Janne Gustafsson Cheyne Capital Management Limited Stornoway House, 13 Cleveland Row, London SW1A 1DH, U.K. [email protected] Ahti Salo Systems Analysis Laboratory Helsinki University of Technology P.O. Box 1100, 02015 HUT, FINLAND [email protected] www.sal.hut.fi/Personnel/Homepages/AhtiS.html Date: June, 2005 Status: Systems Analysis Laboratory Research Reports E17 June 2005 Abstract: This paper examines the valuation of risky projects in a setting where the

investor’s probability estimates for future states of nature are ambiguous and where he or she can invest both in a portfolio of projects and securities in financial markets. This setting is relevant to investors who allocate resources to industrial research and development projects or make venture capital investments whose success probabilities may be hard to estimate. Here, we employ the Choquet-Expected Utility (CEU) model to capture the ambiguity in probability estimates and the investor’s attitude towards ambiguity. Projects are valued using breakeven selling and buying prices, which are obtained by solving several mixed asset portfolio selection (MAPS) models. Specifically, we formulate the MAPS model for CEU investors, and show that a project’s breakeven prices for (i) investors exhibiting constant absolute risk aversion and for (ii) investors using Wald’s maximin criterion can be obtained by solving two MAPS problems. We also show that breakeven prices are consistent with options pricing analysis when the investor is a non-expected utility maximizer. The valuation procedure is demonstrated through numerical experiments.

Keywords: project valuation, capital budgeting, portfolio selection, ambiguity,

mathematical programming, decision analysis

94

Page 103: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

1

1 Introduction Recently, Gustafsson et al. (2004) proposed a framework for valuing risky projects in a setting

where the investor can invest in a portfolio of private projects and market-traded securities.

This framework assumes that the investor knows – by subjective estimation, quantitative

analysis, or otherwise – the probability distribution of returns for each asset. Yet, in practice, it

can be difficult to obtain well-founded probability estimates for real-world events, such as for

the success of a research and development (R&D) project or the increase in the stock price of a

newly listed company.

Several approaches for dealing with incomplete information about probabilities (or, ambiguity)

have been developed over the past few decades. Prominent approaches include (i) Wald’s

(1950) maximin analysis, which can be applied when probabilities are unknown, (ii) Dempster-

Shafer theory of evidence (Dempster 1967, Shafer 1976), where beliefs can be assigned to sets of

states, (iii) second-order probabilities (Marschak 1975, Howard 1992), where a (second-order)

probability distribution is associated with each probability estimate, and (iv) Choquet-Expected

Utility (CEU) theory, where the expectation is taken with respect to a nonadditive probability

measure (Schmeidler 1982, 1989; Gilboa 1987). Among these, the CEU approach has attracted

substantial attention among researchers and found several applications in economic analysis,

not least due to its robust theoretic foundation (see, e.g., Camerer and Weber 1992, Sarin and

Wakker 1992, Wakker and Tversky 1993, Salo and Weber 1995).

In this paper, we examine the valuation of risky projects when the investor’s probability

estimates for future states of nature are ambiguous. This setting is relevant to investors who

allocate resources to industrial R&D projects or make venture capital investments in new

technologies. We use the CEU approach to capture the ambiguity of probability estimates as

well as the investor’s attitude towards ambiguity. We also assume that the investor is

consistent with first-degree stochastic dominance, which is widely accepted as a plausible

requirement of rational decision making. In this case, the general CEU model reduces to a more

amenable form, which is effectively identical to the Rank Dependent Expected Utility (RDEU)

model (Quiggin 1982, 1993; see Wakker 1990).

The CEU/RDEU model, hereafter referred to as the CEU model for short, has several

intuitively appealing implications. For each portfolio, an ambiguity averse investor adjusts the

initial probability estimates, if such are given, so that the probabilities of the states in which

95

Page 104: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

2

the portfolio performs poorly are adjusted upwards, while the probabilities of the states with

high outcomes are adjusted downwards. Thus, each portfolio is evaluated by using a dedicated

set of state probabilities, obtained by adjusting each state’s initial probability estimate

according to the investment portfolio’s performance in that state.

The main contributions of this paper are (i) the explicit formulation of a mixed asset portfolio

selection (MAPS) model for CEU investors and (ii) the development of a computational

framework for applying this model in project valuation. We consider three special cases of the

general CEU model: (i) the quadratic ambiguity aversion model, which leads to linear

probability distortion, (ii) the exponential ambiguity aversion model, and (iii) Wald’s (1950)

maximin model, which can be obtained as a limit case of the exponential ambiguity aversion

model. We demonstrate these models through numerical experiments and contrast the results to

expected utility models with exponential utility functions.

For a practitioner, this paper offers (a) a model for selecting an investment portfolio that is

robust with respect to probability estimation errors and (b) a framework for that uses this

model in project valuation. The framework can be particularly beneficial in the valuation of

high technology investments in large corporations, because in these firms resources are often

diversified across financial instruments and several technology projects, whereby the related

probabilities entail ambiguities.

The remainder of this paper is structured as follows. Section 2 reviews CEU theory and its link

to RDEU theory. Section 3 presents the MAPS models based on CEU theory. Section 4

describes how projects can be valued in a MAPS setting when the investor is a CEU maximizer.

Section 5 uses the framework in a series of numerical experiments. Section 6 summarizes our

findings and discusses their managerial implications.

2 Choquet Expected Utility Since the publication of the Ellsberg paradox (Ellsberg 1961), which questioned the empirical

validity of Savage’s (1954) subjective expected utility (SEU) theory, several ways of dealing

with incomplete information about probabilities have been developed. The CEU approach is

based on the hypothesis that the investor’s response to ambiguity can be captured through

nonadditive probabilities. An analogous idea of probability distortion was utilized already in

prospect theory, where Kahneman and Tversky (1979) introduced a probability distortion

function (called the weighting function) expressing the investor’s chance attitude. While

96

Page 105: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

3

prospect theory was motivated by empirical evidence about human behavior under risk, several

axiomatic models featuring distorted probabilities, or decision weights, were developed later on.

These approaches include weighted utility theory by Chew and MacCrimmon (1979), rank

dependent utility theory by Quiggin (1982, 1993), subjectively weighted linear utility of Hazen

(1987), dual theory of Yaari (1987), and cumulative prospect theory axiomatized by Wakker

and Tversky (1993), among others.

The key difference between the early decision weight models and the CEU theory is that most

of the earlier models deal with decision making under risk (see Starmer 2000), where

probabilities are either subjectively or objectively known. In contrast, the CEU approach is

concerned with decision making under uncertainty, where probabilities are not precisely known

(see Camerer and Weber 1992). The approach is largely based on the works of Schmeidler

(1982, 1989) and Gilboa (1987) who show that a nonadditive capacity measure (“nonadditive

probability measure”; Choquet 1955) represents the investor’s preferences when certain axioms

hold. Further axiomatizations of nonadditive probabilities can be found in Wakker (1989) and

Sarin and Wakker (1992).

A capacity measure is a function [ ]Ω →: 2 0,1c which satisfies ∅ =( ) 0c , Ω =( ) 1c , and

⊂ ⇒ ≤( ) ( )A B c A c B , where Ω is the set of states of nature and Ω2 is the power set of Ω . If

c is also additive, i.e. ∪ = +( ) ( ) ( )c A B c A c B for all disjoint A and B, then it is also a

probability measure. For an act Ω → R:X , the Choquet-integral with respect to c, called the

Choquet-expectation of X, is

[ ] ( )( ) ( )∞

−∞

= > − + >∫ ∫0

0

1cE X c X x dx c X x dx . (1)

Although a capacity measure is not, in general, linked to a probability measure, it can be

shown (Wakker 1990) that if the investor is consistent with first-degree stochastic dominance

with respect to probability measure P, capacity is of the form

( ) ( )( ) ( )ϕ ϕ> = > = ( )Xc X x P X x G x , where = −( ) 1 ( )X XG x F x is the decumulative

distribution function of X and ϕ is nondecreasing with ϕ =(0) 0 and ϕ =(1) 1 . For example,

such a probability measure may reflect the investor’s best but ambiguous probability estimate

or, in the case of total uncertainty, the maximum entropy (uniform) distribution. In the

following, we assume that such a probability measure exists, and that ϕ is differentiable.

Using the transformation function and integrating by parts, the Choquet-expection in (1) can

97

Page 106: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

4

be expressed as the Lebesgue-Stieltjes integral

[ ] ( )( ) ( )

( )( ) ( )

( ) ( )

( )

0

00

0

00

( ) 1 ( )

( ) 1 ( ) ( )

( ) ( ) ( )

( ) ( ) (1 ( )) ( )

c X X

X X X

X X X

X X X X

E X G x dx G x dx

x G x x G x dG x

x G x x G x dG x

x G x dG x x F x dF x

ϕ ϕ

ϕ ϕ

ϕ ϕ

ϕ ϕ

−∞

−∞−∞

∞∞

∞ ∞

−∞ −∞

= − +

′= − − +

′−

′ ′= − = −

∫ ∫

∫ ∫

This shows that the Choquet-expectation with respect to c is effectively the same as the

expectation with regard to P where probabilities are distorted by the factor ϕ′ −(1 ( ))XF x .

Note that when ϕ( )x is increasing and convex, ϕ′ −(1 )x is nonnegative and decreasing,

expressing ambiguity aversion. Thus, the Choquet-expected utility can be written as

[ ] [ ] ϕ∞

−∞

′= = −∫ ( ) ( )( ) (1 ( )) ( )c u X u XCEU X E u X s F s dF s .

With a strictly increasing u, we obtain ( ) ( )1( )( ) ( ) ( )u XF s P u X s P X u s−= ≤ = ≤

1( ( ))XF u s−= . A change of variables → ( )s u x gives

[ ] [ ]( ) ( ) (1 ( )) ( )c X XCEU X E u X u x F x dF xϕ∞

−∞

′= = −∫ . (2)

This is essentially the same preference model as in the RDEU theory (Quiggin 1982, 1993; see

especially Wakker 1990). Note that since RDEU applies to decision making under risk, the

transformation function ϕ can alternatively be interpreted to be an additional component of

the investor’s risk preferences, rather than to describe aversion to ambiguous probabilities.

However, regardless of the interpretation of ϕ , a convex ϕ always influences the investor’s

behavior by increasing the probabilities of states with low utility outcomes and decreasing those

with high utility outcomes.

3 Mixed Asset Portfolio Models Following Gustafsson et al. (2004), we consider a two-period setting with discrete states. Let

there be n securities, m projects, and l states of nature at time 1. The price of security

= 0,...,i n at time 0 is denoted by 0iS and the respective random price at time 1 by 1

iS . The

outcome of 1iS in state ω = 1,...,l is ω1( )iS . The 0-th asset is the risk-free asset with =0

0 1S

and 10( ) 1 fS rω ≡ + , where rf is the risk-free interest rate. The amounts of securities in the

investor’s portfolio are indicated by continuous decision variables =, 0,...,ix i n . The

investment cost of project = 1,...,k m at time 0 is 0kC , and the respective random cash flow at

time 1 is 1kC . The outcome of 1

kC in state ω = 1,...,l is 1( )kC ω . Binary variable zk indicates

98

Page 107: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

5

whether project k is started or not.

An investor’s preferences over risky mixed asset portfolios can be captured with a preference

functional U, which gives the utility of each risky portfolio. For example, under expected utility

theory, the investor’s preference functional is [ ] [ ]= ( )U X E u X , where u is the investor’s von

Neumann-Morgenstern (1947) utility function. A MAPS model using a preference functional U

can thus be formulated as follows:

= =

⎡ ⎤+⎢ ⎥⎣ ⎦∑ ∑1 1

,0 1

maxn m

i i k ki k

U S x C zx z

(3)

subject to

= =

+ =∑ ∑0 0

0 1

n m

i i k ki k

S x C z B (4)

∈ =0,1 1,...,kz k m (5)

= free 0,...,ix i n , (6)

where B is the budget. The expression inside U in (3) is the investor’s terminal wealth level.

The budget constraint (4) is written as equality, because we know that in the presence of a

risk-free asset a rational investor always spends her entire budget in the optimum.

3.1 Binary Variable Model

In general, the MAPS model for the maximization of CEU can be expressed by adapting (2) as

the preference functional in (3). Let Fx,z be the cumulative distribution function for the

terminal wealth level of mixed asset portfolio ( , )x z , i.e., for the random variable

= =

= +∑ ∑1 1

0 1

n m

i i k ki k

W S x C zx,z . Letting 1 1

0 1

( ) ( ) ( )n m

i i k ki k

W S x C zω ω ω= =

= +∑ ∑x,z denote the investor’s

terminal wealth level in state ω , the objective function maximizing CEU can be written as

1 1 1 1,,

1 0 1 0 1

max 1 ( ) ( ) ( ) ( )l n m n m

i i k k i i k ki k i k

p F S x C z u S x C zωω

ϕ ω ω ω ω= = = = =

⎛ ⎞⎛ ⎞ ⎛ ⎞′ − + +⎜ ⎟⎜ ⎟ ⎜ ⎟⎝ ⎠ ⎝ ⎠⎝ ⎠

∑ ∑ ∑ ∑ ∑x zx z. (7)

A challenge in this formulation is to calculate the values for the cumulative distribution

function, because the function Fx,z depends on the portfolio ( , )x z . For the given target level t,

we can calculate 1

( )l

F t pω ωω

ξ=

= ∑x,z for each portfolio ( , )x z using the following linear

constraints:

1 1

0 1

( ) ( ) 1,...,n m

i i k ki k

t S x C z M lωε ω ω ξ ω= =

+ − − ≤ =∑ ∑ , (8)

ωω ω ε ξ ω= =

+ − − ≤ − =∑ ∑1 1

0 1

( ) ( ) (1 ) 1,...,n m

i i k ki k

S x C z t M l , and (9)

99

Page 108: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

6

ωξ ω∈ =0,1 1,...,l , (10)

where M is some large number and ε some small positive number. The variable ωξ indicates

whether the value of the portfolio is less than or equal to t in state ω ; (8) ensures that ωξ is

equal to 1 when the terminal wealth level is less than t ε+ ; (9) ascertains that ωξ equals 0

when the terminal wealth level is greater than t ε+ . Note that the use of a non-zero ε is

necessary, because (8) and (9) do not define the value of ωξ at t ε+ .

In the CEU model, we need to calculate the values of Fx,y for each state, using the state-

specific portfolio outcome 1 1

0 1

( ) ( ) ( )n m

i i k ki k

W S x C zω ω ω= =

= +∑ ∑x,z as the target value. Extending

our earlier notation, let ωωξ ′ , ω′ = 1,...,l , be a binary variable indicating whether the portfolio

outcome in state ω′ is less than or equal to the target level ( )W ωx,z . The associated

constraints can now be expressed with (8)–(10) by replacing t with the appropriate target

value.

With the help of binary variables ωωξ ′ , the CEU model can be written as the following mixed

integer non-linear programming (MINLP) model:

ωω ω ω

ω ω

ϕ ξ ω ω′ ′′= = = =

⎛ ⎞ ⎛ ⎞′ − +⎜ ⎟ ⎜ ⎟⎝ ⎠⎝ ⎠

∑ ∑ ∑ ∑ξ

1 1

, ,1 1 0 1

max 1 ( ) ( )l l n m

i i k ki k

p p u S x C zx z

(11)

subject to

= =

+ =∑ ∑0 0

0 1

n m

i i k ki k

S x C z B (12)

ωωω ω ω ω ε ξ

ω ω

′= = = =

′ ′+ − − + ≤

′= =

∑ ∑ ∑ ∑1 1 1 1

0 1 0 1

( ) ( ) ( ) ( )

1,..., , 1,...,

n m n m

i i k k i i k ki k i k

S x C z S x C z M

l l (13)

ωωω ω ω ω ε ξ

ω ω

′= = = =

′ ′+ − − − ≤ −

′= =

∑ ∑ ∑ ∑1 1 1 1

0 1 0 1

( ) ( ) ( ) ( ) (1 )

1,..., , 1,...,

n m n m

i i k k i i k ki k i k

S x C z S x C z M

l l (14)

∈ =0,1 1,...,kz k m (15)

ωωξ ω ω′ ′∈ = =0,1 1,..., , 1,..., ,l l (16)

= free 0,...,ix i n (17)

Equivalently, the objective function can be expressed in a form similar to the certainty

equivalent formula in expected utility theory. This form has some useful properties, which are

discussed later in this paper. This “certainty equivalent” form is given by

100

Page 109: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

7

ωω ω ω

ω ω

ϕ ξ ω ω−′ ′

′= = = =

⎛ ⎞⎛ ⎞ ⎛ ⎞′ − +⎜ ⎟⎜ ⎟ ⎜ ⎟⎝ ⎠⎝ ⎠⎝ ⎠

∑ ∑ ∑ ∑ξ

1 1 1

, ,1 1 0 1

max 1 ( ) ( )l l n m

i i k ki k

u p p u S x C zx z

.

The model (11)–(17) has n continuous decision variables, + 2m l binary variables, and + 21 2l

linear constraints. Such a model can usually be solved in a reasonable time for models with less

than 30 states.

3.2 Rank-Constrained Model

Even when the MINLP model (11)–(17) has a reasonable number of binary variables, it can be

relatively hard to solve numerically due to the complex relationships defining the variables ωωξ ′

and the normalization problems implied by the large values with M. Our experiments with the

GAMS software package (see www.gams.com) indicate that standard MINLP algorithms do not

necessarily find the optimal solution for models of this type. Therefore, we propose an

alternative method for solving the CEU MAPS model. The method relies on solving a separate

MINLP model for each possible rank order of states (i.e., ranking is based on the outcome of

the portfolio in each state). In each model, portfolios must satisfy the given rank-order.

Consequently, distorted probabilities (decision weights) in each model are constant and can be

used as parameters in the optimization problem. The decision weight of state ω is

( )( )1 ( )p p F Wω ωϕ ω∗ ′= − x,z x,z , where ( )( )F W ωx,z x,z is the sum of the probabilities of the states

where the portfolio outcome is less than or equal to the portfolio outcome in state ω ,

1 1

0 1

( ) ( ) ( )n m

i i k ki k

W S x C zω ω ω= =

= +∑ ∑x,z .

Assuming that hω denotes the state with h:th rank, the rank-constrained model can be

formulated as

ωω

ω ω∗

= = =

⎛ ⎞+⎜ ⎟⎝ ⎠

∑ ∑ ∑1 1

,1 0 1

max ( ) ( )l n m

i i k ki k

p u S x C zx z

(18)

subject to

= =

+ =∑ ∑0 0

0 1

n m

i i k ki k

S x C z B (19)

1 1 1 11 1

0 1 0 1

( ) ( ) ( ) ( ) 0 1,..., 1n m n m

i h i k h k i h i k h ki k i k

S x C z S x C z h lω ω ω ω+ += = = =

+ − − ≤ = −∑ ∑ ∑ ∑ (20)

∈ =0,1 1,...,kz k m (21)

= free 0,...,ix i n , (22)

where ( )( )1 ( )p p F Wω ωϕ ω∗ ′= − x,z x,z . This model is solved for each possible rank-order of states,

which leads to a total of !l models. The rank-order with the highest optimum gives the optimal

101

Page 110: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

8

solution to the problem (11)–(17).

An advantage of this approach is that its complexity does not depend on the transformation

function ϕ . However, because the calculation of the optimal solution requires the solution of !l

rank-constrained models, this approach is computationally attractive only for problems with

some 10 states or less.

3.3 Specific CEU Models

In this section, we consider three interesting special cases of the general CEU model, (i) the

quadratic ambiguity aversion model, (ii) the exponential ambiguity aversion model, and (iii)

Wald’s (1950) maximin model. The pricing behavior of these models is studied in Sections 4

and 5.

3.3.1 Quadratic Ambiguity Aversion

An investor with quadratic ambiguity aversion is characterized by a convex quadratic

transformation function ϕ = + +2( )x ax bx c , a > 0. The normalization ϕ =(0) 0 , ϕ =(1) 1

imply that c = 0 and b = 1 – a, and hence ϕ = + −2( ) (1 )x ax a x and ϕ′ = + −( ) 2 1x ax a . To

ensure that ϕ is nondecreasing over the interval [0,1], the constant a must be less than or

equal to 1. This also ensures that distorted probabilities (decision weights) will be nonnegative.

The appeal of the quadratic transformation function comes from the linearity of the probability

distortion, which potentially facilitates the solution of the resulting models. Figure 1 illustrates

the decision weights obtained by applying a quadratic transformation function in a setting with

six equally likely states of nature. For the sake of comparison, the decision weights are

normalized so that they to sum up to 1.

Together with a linear utility function =( )u x x , a quadratic transformation function leads to a

mixed integer quadratic programming model, where the objective function is

1 1 1 1

,1 0 1 0 1

max 1 1 2 ( ) ( ) ( ) ( )l n m n m

i i k k i i k ki k i k

p a F S x C z S x C zωω

ω ω ω ω= = = = =

⎡ ⎤⎛ ⎞⎛ ⎞ ⎛ ⎞+ − ⋅ + +⎢ ⎥⎜ ⎟⎜ ⎟ ⎜ ⎟⎝ ⎠ ⎝ ⎠⎝ ⎠⎣ ⎦

∑ ∑ ∑ ∑ ∑x,yx z.

This model is interesting, because (i) it can be computationally appealing and (ii) it falls within

the scope of Yaari’s (1987) dual theory, expressing the properties observed under this theory,

such as linearity in the payments.

102

Page 111: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

9

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

Rank 6 Rank 5 Rank 4 Rank 3 Rank 2 Rank 1

State

Nor

mal

ized

Dec

ision

Wei

ght 0

0.2

0.4

0.6

0.8

1

Figure 1. Quadratic ambiguity aversion: Normalized decision weights for different levels of a

when there are 6 equally likely states. Rank i indicates the i:th best state.

3.3.2 Exponential Ambiguity Aversion

When the investor’s transformation function is ( ) xx a beγϕ = + , 0γ > , the investor is said to

express exponential ambiguity aversion. The condition ϕ =(0) 0 , ϕ =(1) 1 imply that

1/(1 )a b eγ= − = − so that ( ) ( )( ) 1 / 1xx e eγ γϕ = − − and ( )( ) / 1xx e eγ γϕ β′ = − . Figure 2

illustrates the decision weights implied by applying exponential ambiguity aversion in a setting

with six equally likely states of nature. Again, the decision weights are normalized to sum up to

one.

In numerical experiments (Section 5), we also consider investors who exhibit both constant

absolute risk aversion and exponential ambiguity aversion. In this case, the investor’s

preferences are captured by a concave exponential utility function ( ) xu x e α−= − , α > 0 , and a

convex exponential transformation function. This implies a CEU objective function

1 1 1 1

0 1 0 1

( ) ( ) ( ) ( )

,1

max1

n m n m

i i k k i i k ki k i k

F S x C z S x C zl

p e ee

γ ω ω α ω ω

ω γω

γ= = = =

⎛ ⎞ ⎛ ⎞− + − +⎜ ⎟ ⎜ ⎟⎜ ⎟ ⎜ ⎟

⎝ ⎠ ⎝ ⎠−

=

⎛ ⎞ ⎛ ⎞∑ ∑ ∑ ∑⎜ ⎟ ⎜ ⎟−⎜ ⎟ ⎜ ⎟−⎝ ⎠ ⎝ ⎠

∑x,y

x z. (23)

103

Page 112: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

10

0

0.1

0.2

0.3

0.40.5

0.6

0.7

0.8

0.9

1

Rank 6 Rank 5 Rank 4 Rank 3 Rank 2 Rank 1

State

Nor

mal

ized

Dec

isio

n W

eigh

t

0.001

0.01

0.1

0.5

1

2

3

5

10

Figure 2. Exponential ambiguity aversion: Normalized decision weights at different γ -levels

when there are 6 equally likely states. Rank i indicates the i:th best state.

3.3.3 Wald’s Maximin Criterion

With Wald’s (1950) maximin criterion, the investor maximizes the utility of the least preferred

state. Wald’s (1950) model is a special case of the general CEU model, as it is obtained with a

capacity measure that gives 0 to all proper subsets of Ω ; it is also an extreme case of the

exponential CEU model when γ goes to infinity.

A MAPS model using the maximin criterion can be formulated as the following mixed integer

linear programming (MILP) problem:

ππ

, ,maxx z

(24)

subject to

= =

+ =∑ ∑0 0

0 1

n m

i i k ki k

S x C z B (25)

π ω ω ω= =

− − ≤ =∑ ∑1 1

0 1

( ) ( ) 0 1,...,n m

i i k ki k

S x C z l (26)

∈ =0,1 1,...,kz k m (27)

= free 0,...,ix i n (28)

π free (29)

104

Page 113: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

11

In contrast to nonlinear CEU models, this model does not require additional binary variables or

the solution of several MINLP models. Hence, maximin models can be solved relatively rapidly

even with a large number of states.

4 Valuing Projects: CEU Investors In a MAPS setting, an individual project can be valued by comparing the values of the optimal

portfolios when the investor does and does not invest in the project. Formally, the comparison

is carried out according to the concepts of breakeven selling and buying prices (De Reyck,

Degraeve and Gustafsson 2003, Luenberger 1998, Smith and Nau 1995). Specifically, breakeven

selling price is the lowest price at which a rational investor would be willing to sell the project

if he or she had it, while the breakeven buying price is the highest price at which the investor

would buy the project if he or she did not have it. Hence, when determining the breakeven

selling price, we assume that the investor invests in the project at the beginning (referred to as

the status quo) and compare the result to the situation where he or she does not have the

project (referred to as the second setting). The breakeven selling price is the budget increment

which makes the portfolio in the second setting equally preferred to the portfolio in the status

quo. Similarly, when calculating the breakeven buying price, we assume that the investor does

not have the project in the status quo and that he or she has it in the second setting. The

breakeven buying price is the budget reduction that makes the portfolio with the project

equally desirable to the portfolio without the project.

Optimization problems for calculating breakeven selling and buying price of project j are given

in Table 1. For CEU maximizers, we can use the explicit portfolio optimization model (11)–(17)

or (18)–(22), adding the relevant constraint on project j to the model in each setting (i.e.,

= 1jz or = 0jz ). The breakeven prices are optimized iteratively by changing the parameter sjv or b

jv , as appropriate, until the optimal values of the portfolio selection problems in the

status quo and in the second setting are identical. In general, the breakeven selling price and

the breakeven buying price are not equal.

When the investor can borrow funds beyond any limit and no risk constraint is imposed on the

portfolio, the optimization problems in Table 1 may yield unbounded solutions; if so, the

breakeven prices are undefined. For example, this may happen when the utility function of a

CEU maximizer is linear. For example, the investors abiding by Yaari’s (1987) dual theory are

CEU maximizers with a linear utility function. There are also other classes of CEU models

which may lead to unbounded solutions, such as models using logarithmic utility functions.

105

Page 114: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

12

Indeed, if limitless borrowing is possible, it may be necessary to use a risk or borrowing

constraint with several types of preference models in order to avoid unbounded solutions.

Table 1. Definitions of the value of project j. For CEU investors U is as in Section 3.

Breakeven selling price Breakeven buying price

Definition sjv such that − +=W W b

jv such that − +=W W

Status quo +

= =

⎡ ⎤= +⎢ ⎥⎣ ⎦∑ ∑1 1

,0 1

maxn m

i i k ki k

W U S x C zx z

subject to

= =

+ =∑ ∑0 0

0 1

n m

i i k ki k

S x C z B

= 1jz

∈ =0,1 1,...,kz k m

= free 0,...,ix i n

= =

⎡ ⎤= +⎢ ⎥⎣ ⎦∑ ∑1 1

,0 1

maxn m

i i k ki k

W U S x C zx z

subject to

= =

+ =∑ ∑0 0

0 1

n m

i i k ki k

S x C z B

= 0jz

∈ =0,1 1,...,kz k m

= free 0,...,ix i n

Second

setting

Project sold

= =

⎡ ⎤= +⎢ ⎥⎣ ⎦∑ ∑1 1

,0 1

maxn m

i i k ki k

W U S x C zx z

subject to

= =

+ = +∑ ∑0 0

0 1

n ms

i i k k ji k

S x C z B v

= 0jz

∈ =0,1 1,...,kz k m

= free 0,...,ix i n

Project bought

+

= =

⎡ ⎤= +⎢ ⎥⎣ ⎦∑ ∑1 1

,0 1

maxn m

i i k ki k

W U S x C zx z

subject to

= =

+ = −∑ ∑0 0

0 1

n mb

i i k k ji k

S x C z B v

= 1jz

∈ =0,1 1,...,kz k m

= free 0,...,ix i n

On the other hand, portfolio optimization problems typically remain bounded when a CEU

investor has a concave exponential utility function. Moreover, for such an investor, there exists

a pricing formula analogous to the one introduced in Gustafsson et al. (2004) for mean-risk

investors, provided that the preference functional is expressed in expected utility theory’s

“certainty equivalent” form. The proof of the following proposition is in Appendix.

PROPOSITION 1. Let the investor be a CEU maximizer with α−= −( ) xu x e , α > 0 , and the

preference functional be [ ] ( )[ ]( )−= 1cU X u E u X . Then, breakeven selling and buying prices

are identical and can be computed using the formula + −−

=+1 f

V Vvr

, where V + and V − are the

optimal preference functional values with and without the project, respectively.

106

Page 115: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

13

It can also be shown that, in arbitrage-free markets, an investor who applies Wald’s (1950)

maximin criterion never invests in securities, if he or she does not invest in projects

(Proposition 2). However, if he or she does invest in projects, it may be optimal to invest in

securities as well. The breakeven selling and buying prices for a maximin investor can be

computed by calculating the value of the portfolio when the investor does and does not invest

in the project and by discounting the difference to its present value at the risk-free interest rate

(Proposition 3). The difference from analogous formulas in Gustafsson et al. (2004) is that the

value of the portfolio is here defined as the value obtained in the worst possible state, as given

by decision variable π in (24). The proofs of the propositions are given in Appendix.

PROPOSITION 2. In arbitrage-free markets, if a maximin investor does not invest in projects, she

does not invest in securities, unless the resulting portfolio replicates the risk-free asset. If a

maximin investor invests in projects, she may also invest in a risky security portfolio.

PROPOSITION 3. If the investor abides by Wald’s maximin criterion, breakeven selling and

buying prices for any project are identical and can be computed by the formula + −−

=+1 f

V Vvr

,

where V + and V − are the values of the portfolio in the worst state (given by decision variable

π ) with and without the project, respectively.

Breakeven selling and buying prices for a CEU investor (including a maximin investor) exhibit

an important consistency property. Namely, a project’s breakeven selling and buying price are

consistent with options pricing analysis, or contingent claims analysis (CCA), which is used in

the theory of finance (see, e.g., Brealey and Myers 2000, Hull 1999, Luenberger 1998). An

analogous property was proven by Smith and Nau (1995) for expected utility maximizers; the

following proposition generalizes their result to non-expected utility maximizers. The

consistency property holds in the sense that CCA and the breakeven prices yield the same

result whenever CCA is applicable, i.e. when there is a replicating portfolio for the project. The

CCA price of project j is given by

=

= − + ∑0 0

0

nCCAj j i i

i

v C x S ,

where ix ∗ is the amount of security i in the replicating portfolio. Due to this property,

breakeven buying and selling prices can be regarded as generalizations of CCA for incomplete

markets. The proof of the proposition is in Appendix.

107

Page 116: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

14

PROPOSITION 4. If there is a replicating portfolio for a project, breakeven selling and buying

prices for the project are identical and equal to the value given by options pricing analysis.

5 Numerical Experiments In this section, we examine CEU-based project valuation through numerical experiments. We

first offer comparative results with expected utility and expected value maximizers and then

analyze investors with varying degrees of ambiguity aversion. In particular, we consider the

following research questions:

Q1. When an expected utility maximizer becomes less risk averse, do the project values

converge to the values given by an expected value maximizer?

Q2. To which values do project values converge with increasing risk aversion?

Q3. When a CEU maximizer becomes less ambiguity averse, do the project values

converge to the values given by the respective expected utility maximizer?

Q4. To which values do project values converge with increasing ambiguity aversion?

5.1 Experimental Setup

The experimental setup consists of 6 states of nature, 4 projects (Table 3), and 2 securities

(Table 2) which constitute the market. Initial state probabilities are generated by assuming

maximum entropy, implying a probability of 1/6 for each state. The security prices are

obtained from the Capital Asset Pricing Model (CAPM; Sharpe 1964, Lintner 1965; see also

Luenberger 1998), where the expected rate of return of the market portfolio has been set to

13.62%, so that the price of security 2 is $20. This price is used by Smith and Nau (1995),

Trigeorgis (1996), and Gustafsson et al. (2004) for this security. The standard deviation of the

market portfolio is then 19.68%. Betas are calculated with respect to the resulting market

portfolio. The market prices in Table 3 are the prices that the CAPM would give to the

projects if they were traded in financial markets, assuming infinite divisibility and a negligible

market capitalization. Italicized text indicates a computed value. The risk-free interest rate is

8%.

108

Page 117: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

15

Table 2. Securities.

Security

1 2

Shares issued 15,000,000 10,000,000

State 1 $60 $36

State 2 $50 $36

State 3 $40 $36

State 4 $60 $12

State 5 $50 $12

State 6 $40 $12

Beta 0.66 2.13

Market price $44.75 $20.00

Capitalization weight 77.05% 22.95%

Expected rate of return 11.72% 20.00%

St. dev. of rate of return 18.24% 60.00%

Table 3. Projects.

Project

A B C D

Investment cost $80 $100 $104 $40

State 1 $150 $140 $180 $20

State 2 $50 $140 $180 $40

State 3 $150 $150 $180 $40

State 4 $150 $110 $60 $20

State 5 $50 $170 $60 $60

State 6 $150 $100 $60 $60

Beta 0.000 0.240 2.13 -1.679

Market price $28.02 $23.46 -$4.00 $0.58

Expected outcome $116.67 $135 $120 $40

St. dev. of outcome $47.14 $23.63 $60 $16.33

We consider investors who exhibit (i) constant absolute risk aversion and (ii) exponential,

quadratic, or maximin ambiguity aversion. By Propositions 2 and 4, breakeven selling and

109

Page 118: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

16

buying prices are identical, and hence both prices are displayed in a single entry. Unless

otherwise indicated, we assume that the investor’s budget is $500. Optimizations were carried

out with the GAMS software package and the rank-constrained formulation in (18)–(22), except

for the maximin model, for which the model (24)–(29) was employed.

5.2 Expected Utility, Expected Value, and Maximin Investors

Before considering the impacts of ambiguity aversion, we first examine three cases without

ambiguity aversion, namely, (i) expected utility (EU) maximizers, (ii) expected value (EV)

maximizers (risk-neutral investors), and (iii) maximin investors. Table 4 gives the project

values for an expected utility maximizer as a function of the risk-aversion coefficient α , as well

as the optimal mixed asset portfolios at each level of risk aversion. At each level, the investor

starts projects A, B, and D. Weights of securities 1 and 2 in the security portfolio are denoted

by 1w and 2w , respectively. The amount of funds invested in the risk-free asset is given in the

column “ 0x ”.

Table 4. Breakeven prices and optimal portfolios for an EU maximizer.

A B C D Expectation St.dev.

0.000001 $28.67 $23.06 -$4.00 $0.64 $83,218.7 $289,177.2 76.99% 23.01% -$1,468,382.6

0.00001 $28.67 $23.07 -$4.00 $0.64 $8,858.8 $28,919.6 77.01% 22.99% -$146,659.7

0.0001 $28.68 $23.10 -$4.00 $0.65 $1,422.8 $2,894.1 77.14% 22.86% -$14,487.5

0.001 $28.72 $23.40 -$4.00 $0.79 $679.3 $293.5 78.38% 21.62% -$1,271.5

0.005 $28.65 $24.78 -$4.00 $1.44 $613.7 $69.8 82.08% 17.92% -$102.1

0.010 $27.77 $26.54 -$4.00 $2.28 $606.1 $47.7 83.89% 16.11% $37.7

0.015 $26.77 $28.26 -$4.00 $3.09 $603.9 $42.5 83.96% 16.04% $80.0

0.020 $25.74 $29.85 -$4.00 $3.81 $603.1 $40.8 83.18% 16.82% $98.4

0.030 $23.90 $32.42 -$4.00 $4.96 $602.6 $40.1 80.79% 19.21% $114.0

0.040 $22.50 $34.02 -$4.00 $5.77 $602.6 $40.4 78.43% 21.57% $121.1

Project value Optimal mixed asset portfolio

α 1w 2w 0x

In Table 4, we observe that when α approaches zero, the values of projects A–D converge to

$28.67, $23.06, -$4.00, and $0.64, which are close to the projects’ CAPM market prices (Table

3). This is surprising, because one would expect that the project values would converge to the

values given by an EV maximizer, i.e., to $17.22, $12.50, –$4.00, and –$6.67. These values are

obtained by discounting the project cash flows at a 20% discount rate, the expected rate of

return of the security with the highest expected return (security 2). Also, the financial portfolio

converges close (but not exactly) to the CAPM market portfolio, containing 76.99% of security

1 and 23.01% of security 2 when α −= 610 . A risk-neutral portfolio would contain 100% of

110

Page 119: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

17

security 2. Even though the amount of funds borrowed and the monetary worth of the mixed

asset portfolio approach infinity with decreasing risk aversion, project values and the weights of

the investor’s financial portfolio converge to specific values. Our experiments also show that,

when the investor cannot invest in projects, he or she invests in a security portfolio with

weights (76.99%, 23.01%) at all levels of risk aversion. Taken together, these observations

indicate that the pricing behavior of EU maximizers with constant absolute risk aversion is

similar to mean-variance investors whose project values converge to CAPM market prices with

increasing risk tolerance, and who always invest in the CAPM market portfolio when there are

no projects (De Reyck, Degraeve and Gustafsson 2003).

Because the optimal security portfolio for the EU investor is near to the CAPM market

portfolio, the expected rate of return of the mixed asset portfolio converges to a value that is

close to the excess rate of return of the CAPM market portfolio. Likewise, the volatility of the

mixed asset portfolio approaches a value near to the volatility of the market portfolio. Notice

that volatility does not decrease monotonically with increasing risk aversion. This is because

standard deviation does not exactly measure the risk perceived by an EU maximizer. Hence, it

may happen that, means being equal, a portfolio with more standard deviation is preferred to

the portfolio with less standard deviation.

On the other hand, it can be shown that, when the risk aversion parameter α goes to infinity,

project values converge towards the values for a maximin investor, who prices the projects at

$17.69, $25.37, -$4.00, and $8.15. In the optimum, a maximin investor invests in a security

portfolio with weights (76.32%, 23.68%) and lends $104.3. Table 4 gives project values and

mixed asset portfolio statistics up to the α -level of 0.04. At higher α -levels optimization

problems become computationally unstable due to the very large negative values in the

exponent. Notice that the expected convergence behavior is not clear from Table 4. This

suggests that some solutions at α -levels below 0.04 may be distorted by computational issues

or convergence to a local optimum, for example.

Between the two extremes, the value of project A decreases with increasing risk aversion. In

contrast, the values of projects B and D grow when risk aversion is increased. The behavior

with project D can be explained by the negative correlation of project D with the rest of the

portfolio; the negative correlation is partly highlighted by the project’s negative beta (Table 3).

As indicated by the results from Gustafsson et al. (2004), the value of a nearly-zero-correlation

111

Page 120: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

18

project may decrease by growing risk aversion, explaining the convergence behavior of project

A. However, the value of project B increases with α , even though the beta of the project is

positive. This may be explained by the project’s negative correlation with project A (–0.599).

Note that the value must start to decrease at some point, because the maximin value of project

B is $25.37, which is less than the price of project B when 0.04α = , $34.02. Hence, the

convergence to maximin values need not be monotonic. Project C is priced at a constant –$4

level, because there exists a replicating portfolio: it is possible to replicate the cash flows of

project C by buying 5 shares of security 2. Since the replicating portfolio costs $100, it implies a

CCA price of –$4 for project C.

In summary, when the risk aversion parameter α approaches zero, the project values given by

an EU maximizer with constant absolute risk aversion converge to values near to CAPM

market prices, rather than to the values for an EV maximizer [Q1]. Also the optimal financial

portfolio converges close to the CAPM market portfolio. With increasing risk aversion project

values approach those of a maximin investor [Q2]. The results also indicate that project values

may either decrease or increase with growing risk aversion, and that the convergence behavior

is not necessarily monotonic. Finally, a replicating portfolio implies a constant pricing behavior.

5.3 Choquet-Expected Utility Maximizers

In this section, we examine investors with exponential and quadratic ambiguity aversion. The

risk aversion parameter α is 0.005 in each experiment. Table 5 describes project values and

mixed asset portfolio statistics for exponentially ambiguity averse investors as a function of

ambiguity aversion parameter γ . At each value of γ , the investor invests in projects A, B, and

D. When γ approaches zero, the values converge, as expected, towards project values given by

an expected utility maximizer. The prices for such an investor are given on the row where

0γ = . At the other extreme, when γ approaches infinity, project values converge towards the

values given by a maximin investor. These values are described on the row with γ = ∞ .

112

Page 121: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

19

Table 5. Breakeven prices and optimal portfolios for a CEU investor with exponential ambiguity aversion.

A B C D Expectation St.dev.

0 $28.65 $24.78 -$4.00 $1.44 $613.7 $69.8 82.08% 17.92% -$102.1

0.1 $28.51 $24.96 -$4.00 $1.53 $612.4 $65.9 82.75% 17.25% -$81.1

0.5 $27.98 $25.89 -$4.00 $1.84 $607.7 $52.1 85.12% 14.88% $0.0

1 $27.35 $27.41 -$4.00 $2.14 $604.2 $43.4 81.08% 18.92% $84.0

2 $25.25 $31.05 -$4.00 $3.67 $603.8 $43.5 76.32% 23.68% $104.1

3 $22.86 $34.03 -$4.00 $4.96 $603.8 $43.5 76.32% 23.68% $104.1

5 $19.78 $29.90 -$4.00 $8.05 $603.8 $43.5 76.32% 23.68% $104.1

$17.69 $25.37 -$4.00 $8.15 $603.8 $43.5 76.32% 23.68% $104.1

Project value Optimal mixed asset portfolio

γ

2w1w 0x

Similarly to increasing risk aversion, the value of project A decreases and the value of project D

increases with increasing ambiguity aversion. This is because project A has a near-zero

correlation with the rest of the portfolio, whereas project D has a negative correlation, which

reduces the aggregate risk of the portfolio. Also, because it is possible to construct a replicating

portfolio, project C has a constant price which is equal to its CCA value. Note also that that

the value of project B increases at low ambiguity aversion levels and decreases at higher levels.

This suggests that a similar behavior is likely to happen with growing risk aversion, as

conjectured in the previous section.

For values of γ higher than and equal to 2, the optimal portfolio is almost identical with a

maximin investor’s portfolio, even though project values at these γ -levels are different. The

reason for this is that, even though the optimal portfolios without restrictions on what projects

can be started are the same at each level, optimal portfolios when the investor must or must

not start a project (the settings in Table 1) are different, which implies different project values.

Table 6 presents project values and optimal mixed asset portfolios under quadratic ambiguity

aversion. As before, the investor invests in projects A, B, and D at all levels of ambiguity

aversion considered. Project values converge towards those for an expected utility maximizer

when a goes to 0. At the other extreme, however, project values do not anymore reach maximin

values, because a is bounded from above by 1. When a = 1, the level of ambiguity aversion

roughly corresponds to exponential ambiguity aversion when γ is in the range [ ]1.5,4 . Overall,

quadratic ambiguity aversion gives lower values to projects A and B and a higher value to

project D than exponential ambiguity aversion.

113

Page 122: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

20

Table 6. Breakeven prices and optimal portfolios for a CEU investor with quadratic ambiguity aversion.

A B C D Expectation St.dev.

0 $28.65 $24.78 -$4.00 $1.44 $613.7 $69.8 82.08% 17.92% -$102.1

0.2 $27.95 $25.64 -$4.00 $1.79 $608.8 $54.4 85.15% 14.85% -$16.7

0.4 $26.97 $26.70 -$4.00 $1.99 $605.6 $45.9 83.45% 16.54% $53.6

0.6 $25.84 $28.11 -$4.00 $2.57 $604.4 $43.3 80.14% 19.86% $88.4

0.8 $24.19 $29.05 -$4.00 $3.31 $604.1 $43.5 76.32% 23.68% $104.1

1 $22.03 $29.33 -$4.00 $4.10 $603.8 $43.5 76.32% 23.68% $104.1

Project value Optimal mixed asset portfolio

a 1w 2w 0x

In summary, when ambiguity aversion approaches zero, the project values and the optimal

portfolios are approach those of the expected utility maximizer [Q3]. When exponential

ambiguity aversion goes to infinity, the project values approach the values given by a maximin

investor [Q4]. Under quadratic ambiguity aversion, however, this occurs only to a certain

extent, because the ambiguity aversion parameter a is limited by 1 from above. Also, the

convergence to the extreme values is not necessarily monotonic.

6 Summary and Conclusions In this paper, we have examined project valuation in a setting where the investor’s probability

estimates are ambiguous and where she can construct an investment portfolio that includes

both projects and securities. We have used the CEU theory to capture the investor’s

preferences for portfolios characterized by ambiguity. In this framework, projects are valued

using the concepts of breakeven selling and buying prices, which require the solution of MAPS

models with and without the analyzed project.

We formulated two alternative MINLP models for solving MAPS problems for CEU investors,

(i) the binary variable model, where the values of the portfolio’s cumulative distribution

function are captured by dedicated binary variables, and (ii) the rank-constrained model, which

is solved separately for each possible rank-order of states. Although the former model is suitable

for a larger number of states than the latter, our experience with standard MINLP algorithms

suggests that models of the former type may be difficult to solve. The latter model is

computationally more tractable in this respect.

We showed that a project’s breakeven selling and buying prices for (i) CEU maximizers who

exhibit constant absolute risk aversion and for (ii) maximin investors can be computed by

solving MAPS models with and without the project and by discounting the difference of the

114

Page 123: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

21

obtained portfolio values back to its present value at the risk-free interest rate. This makes it

possible to calculate breakeven selling and buying prices from two optimization problems only.

We also showed that breakeven selling and buying prices give consistent results with CCA for

non-expected utility maximizers whenever the method is applicable, i.e., when a replicating

portfolio exists for the project. Therefore, these pricing concepts can be regarded as

generalizations of CCA. In our numerical experiments, we examined investors who exhibit (i) constant absolute risk

aversion and (ii) either quadratic, exponential, or maximin ambiguity aversion. The

experiments show that when an expected utility maximizer becomes less risk averse, project

values approach values that are close to the projects’ CAPM prices, rather than the values

given by a risk-neutral investor. Hence, it appears that expected utility maximizers with

constant absolute risk aversion behave similarly to risk-constrained mean-risk investors. When

the investor becomes increasingly averse to risk or ambiguity, project values approach the

values given by a maximin investor. Under quadratic ambiguity aversion, project values do not

reach maximin values, because the ambiguity aversion parameter is bounded from above. Our analysis has several managerial implications. The present framework makes it possible to

(i) select an investment portfolio that is relatively insensitive to changes in probability

estimates and to (ii) calculate defensible values for risky projects whose success probabilities are

ambiguous. Our results also indicate that when ambiguity is taken into account, assets that

perform well when the rest of the portfolio performs poorly often gain higher values than when

ambiguity is not accounted for. Therefore, hedging through relevant derivative securities is

particularly useful under ambiguity, because many hedging instruments are costless zero-value

contracts that make the portfolio outcomes more uniformly distributed and thereby lead to

higher values. Also, diversification possibilities can be expanded by using – in addition to usual

market-traded securities – over-the-counter (OTC) instruments, such as commodity and credit

derivatives, that many investment banks provide for their corporate clients. This work suggests several avenues for further research. In particular, efficient MINLP

algorithms for solving the binary variable model (11)–(17) are called for. Computational issues

could also be relaxed if the CEU model were changed to the weighted utility model, which may

be computationally more amenable than the CEU model. Also, since many real-world project

valuation settings involve multiple time periods, the present model should be extended to a

multi-period setting, for example, in the spirit of Gustafsson and Salo (2005).

115

Page 124: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

22

References

Brealey, R., S. Myers. 2000. Principles of Corporate Finance. McGraw-Hill, New York, NY.

Camerer, C., M. Weber. 1992. Recent Developments in Modeling Preferences: Uncertainty and

Ambiguity. Journal of Risk and Uncertainty 5 325–370.

Chew, S. H., K. R. MacCrimmon. 1979. Alpha-Nu Choice Theory: An Axiomatization of

Expected Utility. Working Paper #669. University of British Columbia Faculty of

Commerce.

Choquet, G. 1955. Theory of Capacities. Annales de l’Institut Fourier 5 131–295.

Dempster, A. P. 1967. Upper and Lower Probabilities Induced by a Multivalued Mapping.

Annals of Mathematical Statistics 38 325–339.

Ellsberg, D. 1961. Risk, Ambiguity, and the Savage Axioms. Quarterly Journal of Economics 75

643–669.

Gilboa, I. 1987. Expected Utility without Purely Subjective Non-Additive Probabilities. Journal

of Mathematical Economics 16 65–88.

Gustafsson, J. De Reyck, B., Z. Degraeve, A. Salo. 2004. Project Valuation in Mixed Asset

Portfolio Selection. Working Paper. Helsinki University of Technology and London Business

School. Downloadable at http://www.sal.hut.fi/Publications/pdf-files/mgus04a.pdf.

Gustafsson, J., A. Salo. 2005. Contingent Portfolio Programming for the Management of Risky

Projects. Operations Research 53(5), forthcoming.

Hazen, G. B. 1987. Subjectively Weighted Linear Utility. Theory and Decision 23 261–282.

Howard, R. A. 1992. The Cogency of Decision Analysis. In W. Edwards (ed.), Utility: Theories,

Measurement, Applications. Kluwer Academic Publications, Dordrecht, Holland.

Hull, J. C. 1999. Options, Futures, and Other Derivatives. Prentice Hall, New York, NY.

Kahneman, D., A. Tversky. 1979. Prospect Theory: An Analysis of Decision under Risk.

Econometrica 47(2) 263–291.

Lintner, J. 1965. The Valuation of Risk Assets and the Selection of Risky Investments in Stock

Portfolios and Capital Budgets. Review of Economics and Statistics 47(1) 13–37.

Luenberger, D. G. 1998. Investment Science. Oxford University Press, New York, NY.

Marschak, J. 1975. Personal Probabilities of Probabilities. Theory and Decision 6 121–153.

Quiggin, J. 1982. A Theory of Anticipated Utility. Journal of Economic Behavior and

Organization 3 323–343.

–––––– 1993. Generalized Expected Utility Theory: The Rank-Dependent Model. Kluwer

Academic Publishers, Dordrecht, Holland.

116

Page 125: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

23

Salo, A. A., M. Weber. 1995. Ambiguity Aversion in First-Price Sealed-Bid Auctions. Journal

of Risk and Uncertainty 11 123–137.

Sarin, R., P. Wakker. 1992. Simple Axiomatization of Nonadditive Expected Utility.

Econometrica 60 1255–1272.

Savage, L. J. 1954. The Foundations of Statistics. John Wiley & Sons, New York.

Schmeidler, D. 1982. Subjective Probability without Additivity. Proceedings of the American

Mathematical Society 97 255-261.

–––––– 1989. Subjective Probability and Expected Utility without Additivity. Econometrica 57

571–587.

Shafer, G. 1976. A Mathematical Theory of Evidence. Princeton University Press, Princeton.

Sharpe, W. F. 1964. Capital Asset Prices: A Theory of Market Equilibrium under Conditions of

Risk. Journal of Finance 19(3) 425–442.

Smith, J. E., R. F. Nau. 1995. Valuing Risky Projects: Option Pricing Theory and Decision

Analysis. Management Science 41(5) 795–816.

Starmer, C. 2000. Developments in Non-Expected Utility Theory: The Hunt for a Descriptive

Theory of Choice under Risk. Journal of Economic Literature 38 332–382.

Trigeorgis, L. 1996. Real Options: Managerial Flexibility and Strategy in Resource Allocation.

MIT Press, Massachussets.

von Neumann, J., O. Morgenstern. 1947. Theory of Games and Economic Behavior, Second

edition. Princeton University Press, Princeton.

Wakker P. 1989. Additive Representation of Preferences: A New Foundation for Decision

Analysis. Kluwer Academic Publishers, Dordrecht, Holland.

–––––– 1990. Under Stochastic Dominance Choquet-Expected Utility and Anticipated Utility

Are Identical. Theory and Decision 29 119–132.

Wakker, P., A. Tversky. 1993. An Axiomatization of Cumulative Prospect Theory. Journal of

Risk and Uncertainty 7 147–176.

Wald, A. 1950. Statistical Decision Functions. John Wiley & Sons, New York.

Yaari, M. 1987. The Dual Theory of Choice under Risk. Econometrica 55 95–115.

117

Page 126: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

24

Appendix PROOF OF PROPOSITION 1: We know that an expected utility maximizer with an exponential

utility function satisfies the delta property, i.e. ( )[ ]( ) ( )[ ]( )− −+ = +1 1u E u X b u E u X b . Since

Choquet-expectation satisfies [ ] [ ]=c cE aX aE X , we also have [ ] ( )[ ]( )1cU X b u E u X b−+ = + =

( )[ ]( ) [ ]1cu E u X b U X b− + = + . Let us next consider the breakeven buying price. Let µb

SQ be the

optimal value of the objective function in the status quo, and µ µ= + ∆b b bSS SQ be the optimal value

in the second setting, when the budget is B. To obtain an optimal value in the second setting

which is equal to µbSQ , we can lower the budget by δ b , whereby the amount borrowed effectively

increases by the same amount, lowering the optimal value by δ +(1 )bfr . When

[ ] ( )( )∗ − ∗= ⎡ ⎤⎣ ⎦1

cU X u E u X is the optimum to the portfolio selection problem with the budget B,

( )( )δ δ∗ − ∗⎡ ⎤− + = − +⎡ ⎤⎣ ⎦⎣ ⎦1(1 ) (1 )b b

f c fU X r u E u X r is the optimum to the portfolio selection

problem with budget δ− bB . By requiring that δ + = ∆(1 )b br we obtain δ = ∆ +/(1 )b bfr . By

denoting that µ −=bSQ V and µ +=b

SS V , we immediately obtain the desired formula. With

breakeven selling price, we have the optimum value µsSQ in the status quo and µ µ= − ∆s s s

SS SQ in

the second setting when the budget is B. Again, we can increase the budget by δ s to obtain an

optimal solution ( )( )δ δ∗ − ∗⎡ ⎤+ + = + +⎡ ⎤⎣ ⎦⎣ ⎦1(1 ) (1 )s s

f c fU X r u E u X r . By requiring that

δ + = ∆(1 )s sfr we obtain δ = ∆ +/(1 )s s

fr . By observing that µ +=sSQ V and µ −=s

SS V , we get

again the desired formula. Q.E.D.

PROOF OF PROPOSITION 2: In arbitrage-free markets, each risky portfolio composed of securities

must have a state in which it yields less than the risk-free interest rate, as otherwise it would be

possible, by buying the portfolio and borrowing at the risk-free interest rate, to create an arbitrage

opportunity, i.e. a portfolio with a chance of yielding a positive amount of money without an initial

capital outlay. Since a maximin investor values each portfolio by its worst state, such an investor

prefers the risk-free asset to all risky security portfolios. However, in a combined project-security

portfolio it may occur that the rate of return of the combined portfolio exceeds the risk-free

interest rate in all states, because projects are not bound by no-arbitrage conditions. For example,

consider a setting with two equally likely states, a project yielding $100 in state 1 and $0

otherwise, and a security yielding $10 in state 2 and $0 otherwise. The cost of the project is $45

and the security is priced at $4.5. The risk-free interest rate is 8% and the budget is $90. The

optimum is to start the project and buy 10 shares of the security. Q.E.D.

PROOF OF PROPOSITION 3: Let us first consider the breakeven buying price. Let π bSQ be the

118

Page 127: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

25

optimal value in the status quo and π π= + ∆b b bSS SQ the optimal value in the second setting when

the budget is B. Because the investor invests in the optimal mixed asset portfolio regardless of the

budget (as long as it yields more than the risk-free interest rate in the worst state), lessening the

budget by δ b will reduce 0x by the same amount, and consequently decrease the optimal value

(the value in the worst state) by δ +(1 )bfr . By requiring that δ + = ∆(1 )b b

fr we obtain,

δ = ∆ +/(1 )b bfr . By denoting that π −=b

SQ V and π +=bSS V , we obtain the desired formula. Let

us then consider the breakeven selling price. Here, we have the optimal value π bSQ in the status quo

and π π= − ∆s s sSS SQ in the second setting. As above, we can increase the optimal value in the

second setting by δ +(1 )sfr by increasing the budget by δ s . By setting δ + = ∆(1 )s s

fr and

denoting π +=sSQ V and π −=s

SS V we again obtain the desired formula. Q.E.D.

PROOF OF PROPOSITION 4: Let us first consider the breakeven buying price and let j indicate the

project being valued. Let ( , )SQ SQx z be the optimal portfolio in the status quo (with 0jz = ).

Notice that when there is a replicating portfolio for project j and the investor invests in the project,

it is possible to construct a shorted replicating portfolio, which nullifies the cash flows of the

project at time 1, and which jointly with project j yields money equal to ∗

=

− + ∑0 0

0

n

j i ii

C x S at time 0

(x* is the amount of security i in the replicating portfolio). Therefore, in the second setting, a

portfolio ( , ) ( , )∗= − +SS SS SQ SQjx z x x z e , where je is an m-dimensional vector with a zero value for

each element except for the j:th element which has the value of 1, will yield the same optimal value

with budget 0 0

0

( )n

j i ii

B C x S∗=

− − + ∑ as the portfolio ( , )SQ SQx z yields with budget B in the status

quo. The portfolio ( , )SS SSx z is the optimum to the problem in the second setting, because

otherwise ( , )SQ SQx z would not be the solution to the problem in the status quo. Therefore, the

breakeven buying price for project j is, by definition, equal to ∗

=

− + ∑0 0

0

n

j i ii

C x S , which is also the

CCA price of the project. A similar logic proves the theorem for breakeven selling prices. Q.E.D.

119

Page 128: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

120

Page 129: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

Helsinki University of Technology Systems Analysis Laboratory Research Reports E18, June 2005

VALUING RISKY PROJECTS WITH CONTINGENT PORTFOLIO

PROGRAMMING

Janne Gustafsson Ahti Salo

AB TEKNILLINEN KORKEAKOULUTEKNISKA HÖGSKOLANHELSINKI UNIVERSITY OF TECHNOLOGYTECHNISCHE UNIVERSITÄT HELSINKIUNIVERSITE DE TECHNOLOGIE D’HELSINKI

121

Page 130: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

Title: Valuing Risky Projects with Contingent Portfolio Programming Authors: Janne Gustafsson Cheyne Capital Management Limited Stornoway House, 13 Cleveland Row, London SW1A 1DH, U.K. [email protected] Ahti Salo Systems Analysis Laboratory Helsinki University of Technology P.O. Box 1100, 02015 HUT, FINLAND [email protected] www.sal.hut.fi/Personnel/Homepages/AhtiS.html Date: June, 2005 Status: Systems Analysis Laboratory Research Reports E18 June 2005 Abstract: This paper examines the valuation of multi-period projects in a setting

where (i) the investor maximizes her terminal wealth level, (ii) she can invest in securities and private investment opportunities, and (iii) markets are incomplete, i.e. the cash flows of private investments cannot necessarily be replicated using financial securities. Based on Gustafsson and Salo’s (2005) Contingent Portfolio Programming, we develop a multi-period mixed asset portfolio selection model, where project management decisions are captured through project-specific decision trees. This model properly captures the opportunity costs imposed by alternative investment opportunities and determines the appropriate risk-adjustment to the projects based on their effect on the investor’s aggregate portfolio risk. The project valuation procedure is based on the concepts of breakeven selling and buying prices, which require the solution of mixed asset portfolio selection models with and without the project being valued. The valuation procedure is demonstrated through numerical experiments.

Keywords: project valuation, mixed asset portfolio selection, contingent portfolio

programming

122

Page 131: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

1

1 Introduction The valuation of risky multi-period projects is one of the most fundamental topics in corporate

finance (Brealey and Myers 2000, Luenberger 1998). Among the numerous methods developed

for this purpose, the risk-adjusted net present value (NPV) method (Brealey and Myers 2000,

Chapter 2) and the decision tree technique (Raiffa 1968; Brealey and Myers 2000, Chapter 10)

are among the most popular ones. These two methods are complementary in the sense that a

decision tree can be used to structure the management decisions and uncertainties related to a

project, while the risk-adjusted NPV method gives the present value of each risky cash flow

stream that can be acquired from the decision tree, provided that the appropriate risk-adjusted

discount rate is known.

For projects of market-traded companies, the appropriate discount rate can, in principle, be

derived using the Capital Asset Pricing Model (CAPM; Sharpe 1964, Lintner 1965; see

Rubinstein 1973). However, there are no direct guidelines for determining the discount rate

when the investor is not a public company, e.g., when the investor is an individual, a

governmental agency, or a non-listed firm, which do not have a share price to maximize. In a

series of recent papers, Gustafsson et al. (2004), De Reyck et al. (2004) and Gustafsson and

Salo (2004) have studied the valuation of risky projects in this setting under the assumptions

that (i) the investor maximizes her terminal wealth level, (ii) she can invest in securities in

financial markets as well as in a portfolio of private investment opportunities, and (iii) markets

are incomplete, i.e. the cash flows of private investments cannot necessarily be replicated using

financial securities. However, due to the complexity of modeling a portfolio of multi-period

projects, the developed models have been limited to two time periods only and hence cannot be

applied to the valuation of a multi-period project whose management strategy can be altered

over the course of the project’s life cycle.

In this paper, we extend the valuation framework presented in Gustafsson et al. (2004) and

Gustafsson and Salo (2004) to multiple time periods and intermediate project management

decisions. Both the original framework and its multi-period extension are based on a mixed

asset portfolio selection (MAPS) model – a portfolio model including both projects and

securities – and the concepts of breakeven selling and buying prices, which compare the values

of optimal portfolios including and not including the analyzed project. We employ Gustafsson

and Salo’s (2005) Contingent Portfolio Programming (CPP) to develop a multi-period MAPS

model that captures the management decisions and uncertainties related to risky multi-period

123

Page 132: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

2

projects through project-specific decision trees. We extend CPP to include securities and

generalize its objective function to several types of preference models, ranging from the

expected utility model to various non-expected utility models.

The main contribution of this paper is the development of a generic framework for the

valuation of risky multi-period projects in a setting where the investor maximizes her terminal

wealth level and where the possible investment assets include both projects and securities. Also,

we discuss how the results of the earlier MAPS-based project valuation papers extend to the

multi-period setting. In particular, we show that breakeven prices remain consistent with

contingent claims analysis when the investor is a non-expected utility maximizer, and that the

resulting prices are, under certain conditions, the same as in Hillier’s (1963) method. The use of

the resulting framework is demonstrated through numerical experiments.

The remainder of this paper is structured as follows. Section 2 reviews earlier approaches to the

valuation of multi-period projects and discusses their assumptions and shortcomings. Section 3

gives an overview of the model developed in this paper. Section 4 discusses the modeling of

states, securities, and projects in the present approach, and Section 5 gives an explicit

formulation of the portfolio selection model. Section 6 describes how the developed portfolio

model can be used in project valuation, and Section 7 illustrates the approach with numerical

experiments. Finally, Section 8 draws conclusions.

2 Earlier Approaches The literature on corporate finance contains a large number of apparently rivaling methods for

the valuation of risky multi-period projects. In what follows, we briefly review some of the most

widely used approaches, namely, (i) decision trees (Hespos and Strassman 1965, Raiffa 1968),

(ii) expected utility theory (von Neumann and Morgenstern 1947, Raiffa 1968), (iii) the risk-

adjusted NPV method (Brealey and Myers 2000), (iv) real options (Dixit and Pindyck 1994,

Trigeorgis 1996), (v) Robichek and Myers’ (1966) certainty equivalent method (see also Brealey

and Myers 2000, Chapter 9), (vi) Hillier’s (1963) method, and (vii) Smith and Nau’s (1995)

method. These methods are summarized in Table 1.

124

Page 133: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

3

Table 1. Methods for the valuation of risky multi-period investments.

Use

Method CE PV ST Formula / explanation

Risk-adjusted NPV X [ ]

1 (1 )

Tt

tt adj

E cNPV I

r=

= − ++∑

Decision tree X A flow chart with decision and chance nodes

Expected utility theory X [ ] [ ]( )1 ( )CE X u E u X−=

Contingent claims analysis X 0

n

i ii

NPV I S x ∗

=

= − + ∑

Robichek and Myers (1966) X [ ]

1 (1 )

Tt

tt f

CE cNPV I

r=

= − ++∑

Hillier (1963) X 1 (1 )

Tt

tt f

cNPV I CEr=

⎡ ⎤= − + ⎢ ⎥

+⎢ ⎥⎣ ⎦∑

Smith and Nau (1995) X

NPV = breakeven selling or buying price

Preference model for cash flow streams:

( ) ( )1 2 1 2[ , ,..., ] , ,...,T TU c c c E u c c c∗= ⎡ ⎤⎣ ⎦

Gustafsson et al. (2004)

Gustafsson and Salo (2004) X

NPV = breakeven selling or buying price

Preference model for terminal wealth levels

Key: CE = Certainty equivalent for a risky alternative, PV = Present value of a risky cash flow stream,

ST = Structuring of decision opportunities and uncertainties, I = investment cost, ct = risky cash flow at

time t, radj = risk-adjusted discount rate, u = utility function, Si = price of security i, ix ∗ = amount of

security i in the replicating portfolio, rf = risk-free interest rate, u* = intertemporal (multi-attribute)

utility function.

2.1 Decision Trees and Related Approaches

In general terms, a decision tree describes the points at which decisions can be made and the

way in which these points are related to unfolding uncertainties. Conventionally, decision trees

have been utilized with expected utility theory (von Neumann and Morgenstern 1947) so that

each end node of the decision tree is associated with the utility implied by the earlier actions

and the uncertainties that have resolved earlier. This decision tree formulation does not

explicitly include the time axis or provide guidelines for accounting for the time value of money.

In corporate finance, decision trees are used to describe how project management decisions

influence the cash flows of the project (see, e.g., Brealey and Myers 2000, Chapter 10). Here,

decision trees are typically applied together with the risk-adjusted NPV method, whereby an

125

Page 134: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

4

explicitly defined time axis is also constructed. However, the selection of an appropriate

discount rate for the NPV method is often problematic, mainly because the rate is influenced

by three confounding factors, (i) the risk of the project, which depends on the project’s

correlation with other investments, (ii) the opportunity costs imposed by alternative investment

opportunities, and (iii) the investor’s risk preferences, which affect the two other factors by

determining (a) the degree of risk-adjustment for the project and (b) the optimal alternative

investment portfolio when the investor does not invest in the project.

Several methods for selecting the discount rate have been proposed in the literature. However,

most of them have problematic limitations. For example, the weighted average cost of capital

(WACC) is appropriate only for average-risk investments in a firm, whereas discount rates

based on expected utility theory do not account for the opportunity costs imposed by securities

in financial markets. The real options literature suggests the use of contingent claims analysis

(CCA) to derive the appropriate discount rate by constructing replicating portfolios using

market-traded securities. Still, it may be difficult to construct replicating portfolios for private

projects in practice. Last, the use of a CAPM discount rate is appropriate only for market-

traded companies.

2.2 Robichek and Myers’ and Hillier’s Methods

Robichek and Myer’s (1966) and Hillier’s (1963) methods are two alternative ways of

determining a risk-adjustment to a discount rate in a multi-period setting. These methods have

been widely discussed in the literature on corporate finance (see, e.g., Keeley and Westerfield

1972, Bar-Yosef and Mesznik 1977, Beedles 1978, Fuller and Kim 1980, Chen and Moore 1982,

Gallagher and Zumwalt 1991, Ariel 1998, and Brealey and Myers 2000). They both employ

expected utility theory or a similar preference model to derive a certainty equivalent (CE) for a

risky prospect.

In Robichek and Myers’ method, the investor first determines a CE for the cash flow of each

period, and then discounts it back to its present value at the risk-free interest rate. Yet,

because CEs are taken separately for each cash flow, the method does not account for the effect

of cash flows’ temporal correlation on the cash flow stream’s aggregate risk; hence, it may lead

to an unnecessarily large risk-adjustment. For example, consider a cash flow stream that yields

a random cash flow X at time 1 and (1 )fX r a− + + at time 2 (a is a constant and rf is the

risk-free interest rate). Assuming that the investor invests the funds obtained at time 1 in the

126

Page 135: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

5

risk-free asset, the investor will acquire cash equal to a for sure at time 2. To avoid arbitrage

opportunities, this cash flow stream has to be discounted at the risk-free interest rate. Still,

since cash flows at times 1 and 2 are separately risky, Robichek and Myers’ method leads to a

discount rate that is higher than the risk-free interest rate, which is incorrect. The importance

of recognizing intertemporal correlation of cash flows is further discussed in Fuller and Kim

(1980).

On the other hand, in Hillier’s (1963) method, which is in the context of certainty equivalents

also referred to as the single certainty equivalent (SCE) method (Keeley and Westerfield 1972),

we first determine the cash flow streams that can be acquired with the project in different

scenarios and then calculate the NPVs of these streams using the risk-free interest rate. The

result is a probability distribution for risk-free-discounted NPV, for which a CE is then

determined. However, the use of the risk-free interest rate essentially means that any money

received before the end of the planning horizon is invested in the risk-free asset. Yet, it might

be more advantageous to invest the funds in risky securities instead. Therefore, Hillier’s method

is, strictly speaking, applicable only in settings, where the investor cannot invest in risky

securities.

Perhaps the most restrictive assumption used in Hillier’s method as well as in Robichek and

Myers’ method is that the risk-adjustment to the discount rate is determined only using the

investor’s risk preferences without considering alternative investment opportunities. Yet, in

general, the appropriate discount rate depends also on the investment portfolio in which the

investor would invest if she did not invest in the project. For example, consider a risk-neutral

investor who can invest in one project and in risky securities in financial markets. We know

that, in financial markets, such an investor would invest all her funds in the security with the

highest expected return. Therefore, the appropriate discount rate for the NPV of the investor’s

project is equal to the expected rate of return of this security. Yet, both of the methods

discussed in this section would give a discount rate equal to the risk-free interest rate for the

project. The fact that Robichek and Myers’ and Hillier’s methods do not adequately consider

alternative investment opportunities limits their applicability to a setting where the investor

can invest only in the analyzed project and in the risk-free asset.

127

Page 136: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

6

2.3 Smith and Nau’s Method

The idea behind Smith and Nau’s (1995) method, which Smith and Nau call “full decision tree

analysis,” is to explicitly account for security trading in each decision node of a decision tree.

The main advantage of the approach is that it appropriately accounts for the effect that the

possibility to invest in securities has on the discount rate of a risky project. However, the

method does not consider alternative projects, which impose an opportunity cost on the project

being valued. Also, the method relies on the assumption that the investor maximizes the

(intertemporal) utility of the project’s cash flows rather than the utility of the investor’s

terminal wealth level. Therefore, it does not necessarily lead to an investment portfolio with

maximal NPV, when NPV is defined as the present equivalent of the investment’s future value

(Luenberger 1998).

Also, practically appealing forms of the method rely on several restrictive assumptions: in order

to develop a useful rollback procedure, Smith and Nau (1995) assume (i) additive independence

(Keeney and Raiffa 1976), (ii) constant absolute risk aversion (CARA), and (iii) partial

completeness of markets. Yet, as pointed out by Keeney and Raiffa (1976), additive

independence entails possibly unrealistic preferential restrictions. The CARA assumption seems

also questionable, because it leads to an exponential utility function with utility bounded from

above. This is known to result in an unrealistic degree of risk aversion at high levels of

outcomes (see, e.g., Rabin 2000). In practice, it may also be difficult to create a replicating

portfolio for market-related cash flows of a project, as assumed in partially complete markets.

In view of the limitations of earlier approaches, we develop a valuation method for risky multi-

period projects where project management decisions are structured as project-specific decision

trees. The method relies on fewer assumptions about the investor’s preference structure than

Smith and Nau’s integrated procedure, allowing the use of a wide range of preference models.

We also employ the objective of maximization of the investor’s terminal wealth level, which

ensures consistency with NPV maximization.

3 Model Overview In a MAPS model, there are two kinds of assets, projects and securities. Projects produce cash

flows according to the chosen project management strategy; cash flows from securities are

realized through trading decisions. Similarly to CPP, uncertainties are modeled using a state

tree, which depicts the structure of future states of nature (Figure 1). The state tree need not

128

Page 137: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

7

be binomial or symmetric; it may also take the form of a multinomial tree that has different

probability distributions in its branches. In each non-terminal state, securities can be bought

and sold in any, possibly fractional quantities. As usual in the financial literature, we assume

that there are no transaction costs or profits tax.

Using the CPP framework, projects are modeled using decision trees that span over the state

tree. Figure 2 describes how project decisions (the figure on the left), when combined with the

state tree in Figure 1, lead to project-specific decision trees (the figure on the right). The

specific feature of these decision trees is that the chance nodes – since they are generated using

the common state tree – are shared by all projects. Security trading is implemented through

state-specific trading variables, which are similar to the ones used in financial models of

stochastic programming (e.g., Mulvey et al. 2000) and in Smith and Nau’s (1995) method.

0 1 2

Stat

es

Time

ω0

50%

30%

70%

40%

60%

50%

15%

35%

20%

30%

ω1

ω2

ω12

ω21

ω22

ω11

Figure 1. A state tree.

129

Page 138: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

8

Start? Continue?Yes

No

Yes

No

0 1 2Time

0

Start?Yes

No

2

Continue?Yes

No

Yes

No

1Time

ω11

ω12ω11

ω12ω21

ω22ω21

ω22

ω1

ω2

Figure 2. A decision tree for a project.

The investor seeks to maximize the utility of her terminal wealth level. When the investor’s

preferences are captured by the preference functional U, this objective can be expressed as

max [ ]U X ,

where the random variable X represents the amount of cash at the end of the planning horizon.

When the investor is able to determine a certainty equivalent for any X, U can be expressed as

a strictly increasing transformation of the investor’s certainty equivalent operator CE. Hence,

the objective can be written as

max [ ]CE X ,

which is the general objective function used in CPP.

Alternatively, it is possible to use a risk-constrained mean-risk model (Gustafsson et al. 2004),

where the investor maximizes the mean of X and constrains the risk level for X. Let ρ be the

investor’s risk measure and R her risk tolerance, i.e. the maximum level for risk. Then, the

objective function is

max [ ]E X

and the risk constraint is

[ ]X Rρ ≤ .

A risk-constrained model is often relevant, because without a risk constraint several preference

models may yield unbounded solutions when there is a possibility to purchase securities and

borrow without limit (see Gustafsson and Salo 2004).

130

Page 139: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

9

4 Formulation of States, Assets, and Cash Flows

4.1 States Let the planning horizon be 0,...,T . The set of states in period t is denoted by tΩ , and the

set of all states is 0

T

tt=

Ω = Ω∪ . The state tree starts with base state 0ω in period 0. Each non-

terminal state is followed by at least one state. This relationship is modeled by the function

:B Ω → Ω which returns the immediate predecessor of each state, except for the base state,

for which the function gives 0 0( )B ω ω= .

The probability of state ω , when ( )B ω has occurred, is given by ( )( )Bp ω ω . Unconditional

probabilities for each state, except for the base state, can be computed recursively from the

equation ( )( ) ( ) ( ( ))Bp p p Bωω ω ω= ⋅ . The probability of the base state is 0( ) 1p ω = .

4.2 Assets

Let there be n securities available in financial markets. The amount of security i bought in

state ω is indicated by trading variable ,ix ω , 1,...,i n= , ω ∈Ω , and the price of security i in

state ω is denoted by ( )iS ω . Under the assumption that all securities are sold back in the next

period (they can immediately be re-bought), the cash flows implied by security i in state

0ω ω≠ is ( ), ( ) ,( )i i B iS x xω ωω ⋅ − . In base state 0ω , the cash flow is 00 ,( )i iS x ωω− ⋅ .

The investor can invest privately in m projects. The decision opportunities for each project are

structured as a decision tree which is formed of decision points. For each project 1,...,k m= ,

the decision tree is implemented using a set of decision points kD and the function ( )ap d that

gives the action leading to decision point 0\k kd D d∈ , where 0kd is the first decision point of

project k. Let dA be the set of actions that can be taken in decision point kd D∈ . For each

action a in dA , a binary action variable ,k az indicates whether the action is selected or not.

These variables are bound by the restriction that, at each decision point, only one of them can

be equal to one at a time. The action in decision point d is chosen in state ( )dω .

For a project k, the vector of all action variables ,k az relating to the project, denoted by kz , is

called the management strategy of project k. The vector of all action variables of all projects,

denoted by z, is termed the project portfolio management strategy. The pair ( , )x z , composed

of all trading and action variables, is the (mixed asset) portfolio management strategy.

131

Page 140: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

10

4.3 Cash Flows and Cash Surpluses

Let ( , )pkCF ωkz be the cash flow of project k in state ω when the project management strategy

is kz . When , ( )k aC ω is the cash flow in state ω implied by action a , this cash flow is given by

, ,:

( ) ( )

( , ) ( )k dB

pk k a k a

d D a Ad

CF C z

ω ω

ω ω∈ ∈∈Ω

= ⋅∑ ∑kz ,

where the restriction in the summation of decision points guarantees that actions yield cash

flows only in the prevailing state and in the future states that can be reached from the

prevailing state. The set ( )B ωΩ is defined as

( ) | 0 such that ( )B kk Bω ω ω ω′ ′Ω = ∈Ω ∃ ≥ = , where 1( ) ( ( ))n nB B Bω ω−= is the n:th

predecessor of ω ( 0( )B ω ω= ).

The cash flows from security i in state ω ∈Ω are given by

( ),

, ( ) ,

( )( , )

( )

i is

ii i B i

S xCF

S x x

ω

ω ω

ωω

ω

− ⋅⎧⎪= ⎨⋅ −⎪⎩

ix 0

0

if

if

ω ω

ω ω

=

Thus, the aggregate cash flow ( , , )CF ωx z in state ω ∈Ω , obtained by summing up the cash

flows for all projects and securities, is

( )

1 1

, , , 01 :

( ) ( )

, ( ) , , , 01 :

( ) ( )

( , , ) ( , ) ( , )

( ) ( ) , if

( ) ( ) , if

k dB

k dB

n ms p

i ki k

n

i i z a k ai d D a A

d

n

i i B i z a k ai d D a A

d

CF CF CF

S x C z

S x x C z

ω

ω ω

ω ω

ω ω

ω ω ω

ω ω ω ω

ω ω ω ω

= =

= ∈ ∈∈Ω

= ∈ ∈∈Ω

= +

⎧ − ⋅ + ⋅ =⎪⎪⎪= ⎨⎪ ⋅ − + ⋅ ≠⎪⎪⎩

∑ ∑

∑ ∑ ∑

∑ ∑ ∑

i kx z x z

Together with the initial budget of each state, cash flows define cash surpluses that would

remain in state ω ∈Ω if the investor chose portfolio management strategy ( , )x z . Assuming

that excess cash is invested in the risk-free asset, the cash surplus in state ω ∈Ω is given by

( ) ( )

( ) ( , , )

( ) ( , , ) (1 )B B

b CFCS

b CF r CS

x z

x zωω ω ω

ω ω

ω ω →

+⎧⎪= ⎨ + + + ⋅⎪⎩

0

0

if

if

ω ω

ω ω

=

≠, (1)

where ( )b ω is the initial budget in state ω ∈Ω and ( )Br ω ω→ is the short rate at which cash

accrues interest from state ( )B ω to ω . Cash surplus in a terminal state is the investor’s

terminal wealth level in that state.

132

Page 141: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

11

5 Optimization Model When using the preference functional U, the objective function for the MAPS model is written

as a function of cash surplus variables of the last time period, i.e.

( ), ,

maxU Tx z CSCS

where CST denotes the vector of cash surplus variables related to period T. Under the risk-

constrained mean-risk model, the objective is to maximize the expectation of the investor’s

terminal wealth level, viz.

, ,max ( )

T

p CSx z CS ω

ω

ω∈Ω

⋅∑ .

Three types of constraints are imposed on the model: (i) budget constraints, (ii) decision

consistency constraints, and (iii) risk constraints, which apply to risk-constrained models only.

Different versions of the MAPS model are summarized in Table 2.

5.1 Budget Constraints

Budget constraints ensure that there is a nonnegative amount of cash in each state. They can

be implemented using continuous cash surplus variables CSω , which measure the amount of

cash in state ω . Using (1), these variables lead to the following budget constraints:

00 0( , , ) ( )CF CS bωω ω− = −x z

( ) ( ) 0( , , ) (1 ) ( ) \ B BCF r CS CS bω ω ω ωω ω ω ω→+ + ⋅ − = − ∀ ∈Ωx z .

Note that if CSω is negative, the investor borrows money at the risk-free interest rate to cover

a funding shortage. Thus, CSω can also be regarded as a trading variable for the risk-free asset.

5.2 Decision Consistency Constraints

Decision consistency constraints implement the projects’ decision trees. They require that (i) at

each decision point at which the investor arrives, only one action is selected, and that (ii) at

each decision point at which the investor does not arrive, no action is taken, implying that the

point does not incur any cash flows. Decision consistency constraints can be written as

0

, 1 1,...,dk

k aa A

z k m∈

= =∑ (2)

0, , ( ) \ 1,...,d

k a k ap d k ka A

z z d D d k m∈

= ∀ ∈ =∑ , (3)

where (2) ensures that one action is selected in the first decision point, and (3) implements the

above requirements for other decision points.

133

Page 142: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

12

Table 2. Multi-period MAPS models.

Preference

functional model

General mean-risk

model

Deviation-based mean-risk model

Objective

function ( )

, ,maxU Tx z CS

CS , ,

max ( )T

p CSωω

ω∈Ω

⋅∑x y CS

Budget

constraints 00 0( , , ) ( )CF CS bωω ω− = −x y

( ) ( ) 0( , , ) (1 ) ( ) \ B BCF r CS CS bω ω ω ωω ω ω ω→+ + ⋅ − = − ∀ ∈Ωx y

Decision

consistency

constraints

0

, 1 1,...,dk

k aa A

z k m∈

= =∑

0, , ( ) \ 1,...,d

k a k ap d k ka A

z z d D d k m∈

= ∀ ∈ =∑

Risk

constraints - ( ) Rρ ≤TCS

( ), Rρ − +∆ ∆ ≤( ) 0 TCSω ω ωτ ω+ −− − ∆ + ∆ = ∀ ∈ΩTCS

Variables

, 0,1 1,...,k a d kz a A d D k m∈ ∀ ∈ ∀ ∈ =

, free 1,...,ix i nω ω∀ ∈Ω =

freeCSω ω∀ ∈Ω

, 0,1 1,...,k a d kz a A d D k m∈ ∀ ∈ ∀ ∈ =

, free 1,...,ix i nω ω∀ ∈Ω =

freeCSω ω∀ ∈Ω

0 Tω ω−∆ ≥ ∀ ∈Ω

0 Tω ω+∆ ≥ ∀ ∈Ω

5.3 Risk Constraints

A risk-constrained model includes one or more risk constraints. For the sake of analogy with

Gustafsson et al. (2004), we focus here on the single constraint case. When ρ denotes the risk

constraint and R the risk tolerance, the risk constraint can be expressed as

( ) Rρ ≤TCS .

Many common dispersion statistics such as variance (V; Markowitz 1952, 1987), semivariance

(SV; Markowitz 1959), absolute deviation (AD; Konno and Yamazaki 1991), lower semi-

absolute deviation (LSAD; Gotoh and Konno 2000), and expected downside risk (EDR; Eppen

et al. 1989) can be formulated through deviation constraints introduced in Gustafsson and Salo

(2005) (see also Gustafsson et al. 2004). In general, deviation constraints are expressed as

( ) 0 TCSω ω ωτ ω+ −− − ∆ + ∆ = ∀ ∈ΩTCS ,

where ( )τ TCS is a function defining the target value from which the deviations are calculated,

and ω+∆ and ω

−∆ are nonnegative deviation variables which measure how much the cash surplus

in state Tω ∈Ω differs from the target value. For example, when the target value is the mean

of the terminal wealth level, the deviation constraints are written as

134

Page 143: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

13

( ) 0 ,T

TCS p CSω ω ω ωω

ω ω+ −′

′∈Ω

′− − ∆ + ∆ = ∀ ∈Ω∑

Using these deviation variables, some common dispersion statistics can be written as

AD: ( ) ( )T

p ω ωω

ω − +

∈Ω

⋅ ∆ + ∆∑ .

LSAD: ( )T

p ωω

ω −

∈Ω

⋅ ∆∑ .

V: 2( ) ( )

T

p ω ωω

ω − +

∈Ω

⋅ ∆ + ∆∑

SV: 2( ) ( )T

p ωω

ω −

∈Ω

⋅ ∆∑ .

The respective fixed-target value statistics can be obtained with the deviation constraints

0 TCSω ω ωτ ω+ −− − ∆ + ∆ = ∀ ∈Ω ,

where τ is the fixed target level. EDR, for example, can then be obtained from the sum

( )T

p ωω

ω −

∈Ω

⋅ ∆∑ .

6 Project Valuation The value of a project can be defined as the lowest price at which a rational investor would be

willing to sell a project, if she had the project, and as the highest price at which a rational

investor would be willing to buy the project, if she did not have it. These prices are referred to

as the breakeven selling price and breakeven buying price of the project, respectively

(Gustafsson et al. 2004, Luenberger 1998, Smith and Nau 1995). In this section, we extend the

definitions of these prices, as presented in Gustafsson et al. (2004), to a multi-period MAPS

setting.

The breakeven prices for a project can be determined from optimization problems where the

investor invests and does not invest in the project (see Table 3). Let the project being valued

be indexed with j. The action associated with not starting the project is denoted by a*. When

calculating the breakeven selling price, we first determine the optimal investment portfolio with

the project. Then, we iteratively solve another optimization problem where the investor does

not invest in the project but where an amount sjv is added to the budget at time 0 instead. The

breakeven selling price is the amount sjv that makes the portfolios with and without the project

equally preferable. In a preference functional model, this means that the utilities of the two

portfolios are identical. In a mean-risk model, two portfolios are equally preferred if they are

equal in terms of the expected terminal wealth level and they both satisfy the risk constraint.

The breakeven buying price for a project is determined similarly as its breakeven selling price.

135

Page 144: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

14

The difference is that, in the first setting, the investor does not invest in the project, and in the

second setting, she invests in the project and deducts an amount bjv from the budget at time 0.

The breakeven selling price is the amount that makes the investor indifferent between the

portfolios with and without the project. In general, the breakeven buying price and the

breakeven selling price for a project are not equal to each other.

Table 3 describes how breakeven selling and buying prices can be determined based on the

optimization problems in Table 2. In each setting, an additional constraint is added to the

optimization model. Also, the budget at time 0 is modified in the second setting. The objective

function values of the resulting models at optimum are denoted by W + and W − depending on

whether the investor invests in the project or not.

Table 3. Definitions of the value of project j. Each setting is based on the model in Table 2.

Breakeven selling price Breakeven buying price

Definition sjv such that s sW W+ −= b

jv such that b bW W+ −=

Status quo Additional constraint:

, * 0j az = , i.e. invest in project

Budget at time 0:

0( )b ω

Optimal objective function value:

sW +

Additional constraint:

, * 1j az = , i.e. do not invest in project

Budget at time 0:

0( )b ω

Optimal objective function value:

bW −

Second

setting

Additional constraint:

, * 1j az = , i.e. do not invest in project

Budget at time 0:

0( ) sjb vω +

Optimal objective function value:

sW −

Additional constraint:

, * 0j az = , i.e. invest in project

Budget at time 0:

0( ) bjb vω −

Optimal objective function value:

bW +

Breakeven selling and buying prices exhibit a notable property: they yield the same result as

CCA whenever the method is applicable, i.e. whenever the cash flows of the project can be

replicated by trading financial securities. In general, the CCA value of the project is defined as

the value of the portfolio that is required to create a trading strategy replicating the future cash

flows of the project minus the project’s investment cost at time 0 (see Table 1). Smith and Nau

136

Page 145: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

15

(1995) show that this property holds for investors with an intertemporal utility function over

cash flow streams, whereas Gustafsson and Salo (2004) prove the same for non-expected utility

investors who maximize their terminal wealth level in a two-period setting. The following

proposition generalizes the latter result to a multi-period setting.

PROPOSITION 1. If there is a replicating trading strategy for a project, the breakeven buying

price and breakeven selling price for the project are equal and yield the same value as CCA.

PROOF: See Appendix.

Gustafsson and Salo (2004) and Gustafsson et al. (2004) also discuss valuation formulas that

facilitate the computation of project values for (i) certain types of mean-risk investors, (ii)

Choquet-expected utility maximizers exhibiting constant absolute risk aversion, and (iii)

investors using Wald’s (1950) maximin criterion. These formulas are applicable in the two-

period case and are of the form

1 f

V Vvr

+ −−=

+,

where V + and V − are the optimal objective function values with and without the project,

respectively, and rf is the risk-free interest rate. Since the propositions rely on the examination

of the last time period with no consideration for the number of intermediate periods, these

formulas can also be used in a multi-period setting with the modification that the objective

function values are discounted using the time-T spot-rate, sT, viz.

(1 )TT

V Vvs

+ −−=

+.

The formal proof of the formulas in a multi-period setting is almost identical to their proof in

the two-period setting and is hence omitted.

As shown by Proposition 2, there exists also an analogous formula for the breakeven selling

price when the investor can invest only in a single project and the risk-free asset. This

proposition holds for any rational preference model accommodated by the portfolio models in

Table 2. If we further assume that the investor exhibits constant absolute risk aversion, i.e. the

investor’s certainty equivalent operator satisfies [ ] [ ]CE X b CE X b+ = + for all random

variables X and constants b, breakeven selling and buying prices are equal to each other; they

are also independent of the budget, and they can be computed from the certainty equivalent of

the project’s future value (Proposition 3). The proofs of these two propositions are in Appendix.

137

Page 146: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

16

PROPOSITION 2. When the investor can invest in a single project and in the risk-free asset, the

breakeven selling price for the project can be computed using the formula (1 )TT

V Vvs

+ −−=

+, where

V + and V − are the investor’s certainty equivalents for her terminal wealth level with and

without the project, respectively, and sT is the time-T spot rate.

PROPOSITION 3. When the conditions of Proposition 2 hold and the investor’s certainty

equivalent operator satisfies [ ] [ ]CE X b CE X b+ = + for all random variables X and constants

b, breakeven buying and selling prices are equal and can be computed using the formula

(1 ) (1 )T

T TT T

V V CEvs s

+ −−= =

+ +, where CET is the certainty equivalent for the project’s future value

and sT is the time-T spot rate.

If all short rates are equal to rf, the formula in Proposition 3 can be expressed as

1

(1 )

(1 )

TT t

t ft

Tf

CE c rv I

r

=

⎡ ⎤+⎢ ⎥⎣ ⎦= − ++

∑,

where, following the notation in Table 1, I is the investment cost at time 0 and ct is the random

time-t cash flow implied by the project under the optimal project management strategy. Notice

that if the investor’s certainty equivalent operator further satisfies [ ] [ ]CE aX aCE X= for all

random variables X and constants a (that is, the investor exhibits also constant relative risk

aversion), the formula reduces to Hillier’s method (Table 1). This observation is formalized

with the following proposition. The proof is obvious from the above and is hence omitted.

PROPOSITION 4. When the conditions of Proposition 2 hold and the investor’s certainty

equivalent operator satisfies [ ] [ ]CE aX b aCE X b+ = + for all random variables X and

constants a and b, the breakeven buying price and breakeven selling price are equal and yield the

same value as Hillier’s method, given by the formula 1 (1 )

Tt

tt t

cNPV I CEs=

⎡ ⎤= − + ⎢ ⎥+⎣ ⎦

∑ , where I is

the investment cost of the project, ct is the cash flow at time t, and st is the time-t spot rate.

For example, Proposition 4 can be applied with risk-neutral investors, some mean-risk models

(see, e.g., Gustafsson and Salo 2005), and preference models based on Yaari’s (1987) dual

theory. However, no risk-averse expected utility maximizer exhibits both constant absolute and

138

Page 147: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

17

constant relative risk aversion, so that no risk-averse expected utility maximizer is, in general,

consistent with Hillier’s method.

While Hillier’s method can be obtained as a special case of breakeven selling and buying prices,

Robichek and Myers’ (1966) method does not comply with these prices, unless restrictive

assumptions about the analyzed cash flow streams are made. For example, when a project’s

cash flows are perfectly correlated with each other and the conditions of Proposition 4 hold,

Robichek and Myer’s method reduces to Hillier’s method, and hence gives consistent results

with the breakeven prices. On the other hand, Smith and Nau’s (1995) method coincides with

the present approach when the chosen intertemporal utility function rewards for the

maximization of the expected utility of the terminal wealth level, i.e., when the utility functions

for periods other than the last are identically zero. Indeed, maximization of intertemporal

utility and that of the utility of terminal wealth level are not equivalent objectives: in effect, it

can be shown that an intertemporal utility function can lead to an investment portfolio, which

is stochastically dominated by another investment portfolio in terms of terminal wealth levels

(Proposition 5). Since the terminal wealth level is widely regarded as the future value

equivalent of NPV, care should be taken when applying Smith and Nau’s method in a setting

where the investor aims to maximize the NPV of the investment portfolio. Table 4 summarizes

the conditions under which other project valuation methods give the same results as the present

approach.

PROPOSITION 5. Maximization of intertemporal utility may lead to a portfolio that is

stochastically dominated by another portfolio in terms of the terminal wealth level.

PROOF: See Appendix.

Table 4. The conditions when the present approach coincides with other project valuation methods.

Method Preferences Cash flows Available investments

Hillier (1963) [ ] [ ]CE aX b aCE X b+ = + – One project, risk-free asset

Robichek and Myers (1966) [ ] [ ]CE aX b aCE X b+ = + Perfectly correlated One project, risk-free asset

Smith and Nau (1995) Utility function for each period

except the last is identically 0

– One project, risk-free asset,

securities

Finally, it is worth noting that the results based on the CAPM in Gustafsson and Salo (2004)

and Gustafsson et al. (2004), such as the convergence of project values to the CAPM market

139

Page 148: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

18

prices, may not hold in a multi-period setting. The reason for this is that the mean-variance

model, on which the CAPM is based, is a dynamically inconsistent preference model (Machina

1989). Therefore, the optimal portfolio for a mean-variance investor maximizing her terminal

wealth level is, in general, different from the portfolio obtained using the two-period CAPM

consecutively in a multi-period setting.

7 Numerical Experiments In the previous sections, we developed a framework for valuing risky multi-period projects in

incomplete markets, and showed that the framework gives consistent results with CCA and,

under certain conditions, with Hillier’s (1963) method. In this section, we demonstrate the

approach through a set of numerical experiments and show that most of the phenomena

observed in a two-period MAPS setting also occur in the multi-period setting. In particular, we

cast light on the following issues:

Q1. How are the values of multi-period projects influenced by the possibility to invest

in risky securities?

Q2. How are the values of multi-period projects influenced by the possibility to invest

in other projects?

Q3. How do the values of multi-period projects depend on the investor’s risk attitude?

7.1 Experimental Setup

The experimental setup involves 3 time periods, 4 projects, and 2 securities that together

constitute the market. Uncertainties related to projects and securities are captured through a

state tree which consists of 8 equally likely time-1 states and 64 equally likely time-2 states

(Figure 3). Each state is formed of two underlying states, (i) the market state (denoted by M),

which determines the prevailing security prices, and (ii) the private state (denoted by P) that

influences project outcomes. Security prices at time 2 are given in Figure 3. Security prices at

time 1 are computed from these prices using the CAPM (Table 5), and the prices at time 0 are

obtained similarly from the time-1 security prices (Table 6).

140

Page 149: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

19

Security 1 Security 2 Project A Project B Project C Project D

State 0 State M1/P1 State M11/P11 $60 $36 $150 $150 $100 $180State M12/P11 $50 $12 $150 $100 $120 $60State M13/P11 $40 $36 $150 $50 $90 $180State M14/P11 $30 $12 $150 $0 $80 $60State M11/P12 $60 $36 $50 $250 $170 $180State M12/P12 $50 $12 $50 $200 $200 $60State M13/P12 $40 $36 $50 $150 $40 $180State M14/P12 $30 $12 $50 $100 $50 $60

State M2/P1 State M21/P11 $70 $36 $150 $200 $130 $180State M22/P11 $60 $12 $150 $150 $70 $60State M23/P11 $50 $36 $150 $100 $130 $180State M24/P11 $40 $12 $150 $50 $110 $60State M21/P12 $70 $36 $50 $300 $100 $180State M22/P12 $60 $12 $50 $250 $180 $60State M23/P12 $50 $36 $50 $200 $40 $180State M24/P12 $40 $12 $50 $150 $50 $60

State M3/P1 State M31/P11 $80 $36 $150 $250 $190 $180State M32/P11 $70 $12 $150 $200 $45 $60State M33/P11 $60 $36 $150 $150 $60 $180State M34/P11 $50 $12 $150 $100 $100 $60State M31/P12 $80 $36 $50 $350 $110 $180State M32/P12 $70 $12 $50 $300 $50 $60State M33/P12 $60 $36 $50 $250 $140 $180State M34/P12 $50 $12 $50 $200 $90 $60

State M4/P1 State M41/P11 $90 $36 $150 $300 $65 $180State M42/P11 $80 $12 $150 $250 $105 $60State M43/P11 $70 $36 $150 $200 $100 $180State M44/P11 $60 $12 $150 $150 $115 $60State M41/P12 $90 $36 $50 $400 $175 $180State M42/P12 $80 $12 $50 $350 $90 $60State M43/P12 $70 $36 $50 $300 $210 $180State M44/P12 $60 $12 $50 $250 $180 $60

State M1/P2 State M11/P21 $60 $36 $80 $220 $50 $180State M12/P21 $50 $12 $80 $170 $40 $60State M13/P21 $40 $36 $80 $120 $110 $180State M14/P21 $30 $12 $80 $70 $105 $60State M11/P22 $60 $36 $40 $260 $125 $180State M12/P22 $50 $12 $40 $210 $115 $60State M13/P22 $40 $36 $40 $160 $235 $180State M14/P22 $30 $12 $40 $110 $170 $60

State M2/P2 State M21/P21 $70 $36 $80 $270 $250 $180State M22/P21 $60 $12 $80 $220 $215 $60State M23/P21 $50 $36 $80 $170 $180 $180State M24/P21 $40 $12 $80 $120 $190 $60State M21/P22 $70 $36 $40 $310 $130 $180State M22/P22 $60 $12 $40 $260 $100 $60State M23/P22 $50 $36 $40 $210 $70 $180State M24/P22 $40 $12 $40 $160 $90 $60

State M3/P2 State M31/P21 $80 $36 $80 $320 $240 $180State M32/P21 $70 $12 $80 $270 $150 $60State M33/P21 $60 $36 $80 $220 $120 $180State M34/P21 $50 $12 $80 $170 $100 $60State M31/P22 $80 $36 $40 $360 $230 $180State M32/P22 $70 $12 $40 $310 $100 $60State M33/P22 $60 $36 $40 $260 $120 $180State M34/P22 $50 $12 $40 $210 $240 $60

State M4/P2 State M41/P21 $90 $36 $80 $370 $100 $180State M42/P21 $80 $12 $80 $320 $160 $60State M43/P21 $70 $36 $80 $270 $250 $180State M44/P21 $60 $12 $80 $220 $290 $60State M41/P22 $90 $36 $40 $410 $200 $180State M42/P22 $80 $12 $40 $360 $100 $60State M43/P22 $70 $36 $40 $310 $50 $180State M44/P22 $60 $12 $40 $260 $45 $60

Figure 3. State tree with security prices and project cash flows at time 2.

141

Page 150: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

20

Table 5. Securities in states M1–M4.

M1 M2 M3 M4

1 2 1 2 1 2 1 2

Beta 0.821 1.527 0.782 1.795 0.755 2.064 0.735 2.332

Market price $39.32 $20.00 $48.58 $20.00 $57.83 $20.00 $67.09 $20.00

Capitalization weight 74.68% 25.32% 78.46% 21.54% 81.26% 18.74% 83.42% 16.58%

Expected rate of return 14.46% 20.00% 13.23% 20.00% 12.39% 20.00% 11.78% 20.00%

St. dev. of rate of return 28.44% 60.00% 23.02% 60.00% 19.33% 60.00% 16.66% 60.00%

Market expected return 15.86% 14.68% 13.82% 13.15%

Market st. dev. 31.15% 26.49% 23.05% 20.39%

Table 6. Securities in state 0.

Security

1 2

Market price in M1 $39.32 $20.00

Market price in M2 $48.58 $20.00

Market price in M3 $57.83 $20.00

Market price in M4 $67.09 $20.00

Beta 1.264 0.000

Market price in 0 46.85 18.52

Capitalization weight 79.14% 20.86%

Expected rate of return 13.58% 8.00%

St. dev. of rate of return 22.10% 0.00%

Market expected return 12.41%

Market standard deviation 17.49%

The CAPM requires us to specify the amounts of outstanding shares and the expected rate of

return of the market portfolio in each state. We have assumed here that there are a total of

15,000,000 shares of security 1 and 10,000,000 shares of security 2. In each state, the expected

rate of return of the market portfolio is determined so that the market price of risk – the excess

rate of return of the market portfolio divided by the standard deviation of the market –

remains constant at 0.252. This is a reasonable way of selecting the expected rate of return of

the market, because hereby it is, in each state, proportional to the standard deviation of the

market (see Tables 5 and 6). Also, this implies that security 2 is priced at $20 in each time-1

market state, which is desirable, because the time-2 prices for security 2 are independent of the

time-1 market state (Figure 3). The risk-free interest rate is 8%.

142

Page 151: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

21

Table 7 presents the investments required to start projects A, B, C and D. The investments are

made in two rounds; the first financing round takes place at time 0 and the second at time 1. If

the investment is made in both rounds, the project produces a cash flow at time 3. Figure 3

gives this cash flow as a function of the time-2 state. The specific characteristics of these cash

flows are summarized in Table 7. The cash flows of project A depend only on the private states,

implying that the project is uncorrelated with the market. On the other hand, the cash flows of

project B can be separated into private and market related components; the private component

depends on the private state and the market component on the market state. Furthermore, the

markets are partially complete (see Smith and Nau 1995) with respect to project B in the sense

that the market component of project B can be replicated by buying 5 shares of security 1; the

private component is obtained by taking the negative of project A. Similarly, the outcome of

project C depends on both private and market states but it cannot be separated into market

and private components. In contrast, the cash flows of project D are entirely market-dependent

and they can be replicated by purchasing 5 shares of security 2.

Table 7. Projects.

Project

A B C D

Private state dependence Yes Yes Yes No

Market state dependence No Yes Yes Yes

Market component replication No Yes No Yes

Investment cost at time 0 $10 $20 $30 $40

Investment cost at time 1 $60 $150 $70 $40

Expected outcome $80.00 $220.00 $127.42 $120.00

St. dev. of outcome $43.01 $90.00 $62.66 $60.00

We examine two preference models, (i) the risk-constrained mean-variance (MV) model, with

which we use the deviation-based mean-risk model in Table 2, and (ii) the expected utility

(EU) model with constant absolute risk aversion (CARA). As the base case, the latter model

employs the preference functional [ ] [ ]( ) XU X E u X E e α−= = − ⎡ ⎤⎣ ⎦ with 0.003α = unless

otherwise noted, and the former the risk tolerance of $250 for the standard deviation of the

terminal wealth level. For both preference models, breakeven selling and buying prices are equal

(see Gustafsson et al. 2004, Gustafsson and Salo 2004) and therefore they are displayed in a

143

Page 152: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

22

single entry. We begin our experiments by analyzing project values when securities are not

available and then continue to the setting with securities. The optimizations were carried out

using the GAMS software package with SBB and CONOPT algorithms (www.gams.com).

7.2 Single Project without Securities

We first solve project values in a setting where the investor can invest only in a single project

and the risk-free asset. Under both preference models, each of the projects is started at time 0.

The optimal management strategies for the projects at time 1 are described in the first eight

data rows of Table 8. The mark “X” indicates that a further investment is made in the project,

while the mark “–“ means that the project is terminated. The first mark is the optimal strategy

for the MV model; the second one indicates the optimal action for the EU investor. In the MV

model, the risk constraint is not binding, wherefore it now corresponds to expected value

maximization.

Table 8. Optimal project management strategies and project values.

Project

Time-1 State A B C D

M1/P1 X/X –/– X/X X/X

M2/P1 X/X X/X X/X X/X

M3/P1 X/X X/X X/X X/X

M4/P1 X/X X/X X/X X/X

M1/P2 –/– X/– X/X X/X

M2/P2 –/– X/X X/X X/X

M3/P2 –/– X/X X/X X/X

M4/P2 –/– X/X X/X X/X

Project value, MV/EV model $5.09 $33.69 $14.43 $25.84

Project value, EU model $3.17 $25.59 $9.58 $21.24

In Table 8, observe that a further investment is made in project A only if private state P1

obtains. This is because the outcomes of project A in the states ensuing P1 are substantially

higher than in the states following P2 (Figure 3). Project B is terminated in state M1/P1 under

both preference models. In contrast, only the MV investor chooses to continue the project in

state M1/P2. The termination decisions are explained by the low outcome with the market

component of project B (i.e., security 1); furthermore, project B is more profitable in private

state P2 than in P1, explaining the continuation decision with the (risk-neutral) MV investor.

144

Page 153: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

23

Both MV and EU investors choose to continue projects C and D in all states. With project D

this is because the realized time-1 state does not convey any information about the future cash

flows of project D (Figure 3); with project C the future outcome distribution in each time-1

state is different but in each case it warrants a continuation decision.

Since we analyze here only a single project with the risk-free asset, we can use the project

values in Table 8 to numerically verify Proposition 3. Let us take project A, for example. The

certainty equivalent for an EU investor’s terminal wealth level with project A is $586.90, while

it is $583.20 without the project. Discounting the difference of these values back to its present

value yields 2($586.90 $583.20)/1.08 $3.17Av = − = , which coincides with the project’s

breakeven selling and buying prices given in Table 8.

7.3 Single Project with Securities

We next consider a setting where the investor can invest in a single project, in the risk-free

asset, and in securities 1 and 2. We first examine the investor’s trading strategy in a case where

the investor cannot invest in projects, because we can consequently examine how the starting of

a project influences the optimal trading strategy. Table 9 gives the optimal trading strategy for

the MV investor. Note that purchase of security 2 at time 0 is equivalent to investment in the

risk-free asset; for the sake of clarity, we have assumed here that the investor invests in the

risk-free asset and not in security 2. The column CS gives the amount of money lent in each

state.

Table 9. Optimal trading strategy for the MV investor without projects.

Portfolio weights Securities bought

State 1 2 1 2 CS

0 100% 0% 16.557 0 -$275.602

M1/P1 74.7% 25.3% 14.203 9.468 -$394.475

M2/P1 78.5% 21.5% 11.809 7.872 -$224.445

M3/P1 81.3% 18.7% 9.414 6.275 -$10.034

M4/P1 83.4% 16.6% 7.019 4.679 $248.724

M1/P2 74.7% 25.3% 14.203 9.468 -$394.475

M2/P2 78.5% 21.5% 11.809 7.872 -$224.445

M3/P2 81.3% 18.7% 9.414 6.275 -$10.034

M4/P2 83.4% 16.6% 7.019 4.679 $248.724

145

Page 154: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

24

In comparison, an EU investor would buy 7.634 shares of security 1 at time 0, and 5.169 shares

of security 1 and 3.478 shares of security 2 in each time-1 state. Since the price of security 1

changes between the time-1 states, the portfolio weights for an EU investor do not remain

constant across time-1 states. An EU investor lends $142.032 in state 0, implying that the EU

model with 0.003α = is a more risk averse preference model than the mean-standard

deviation model with R = $250.

When an MV investor invests in project A, the portfolio weights of the optimal security

portfolio remain the same as in Table 9, but absolute amounts of securities in the portfolio

change. The constant portfolio weights are explained by the fact that project A is uncorrelated

with the market; this was proven in a two-period setting by Gustafsson et al. (2004). For the

same reason, an EU investor preserves her investment behavior: even when she invests in

project A, the optimal trading strategy is, as before, to buy 5.169 shares of security 1 and 3.478

shares of security 2 in each time-1 state.

Recall that the cash flows of project B can be formed by buying 5 shares of security 1 and

deducting the cash flows of project A from the resulting portfolio. The weights of the optimal

security portfolio for an MV investor now change, because project B counts for 5 shares of

security 1; if these 5 shares were added to the security portfolio, the portfolio weights would

match with the weights in Table 9. With the EU investor, we clearly observe this behavior: she

invests in 0.169 shares of security 1 in each time-1 state where the project is continued (see

Table 10) and 5.169 shares otherwise.

In contrast, the cash flows of project C cannot be separated into private and market

components, wherefore the weights of the security portfolio change without a clear pattern.

Project D can be replicated using 5 shares of security 2. Therefore, in the optimum, an EU

investor shorts 5 3.478 1.522− = shares of security 2 and buys 5.169 shares of security 1.

When these 5 shares are added to the security portfolio of the MV investor, the weights of the

portfolio correspond to the ones in Table 9.

Table 10 presents the optimal project management strategy and the values of the projects for

MV and EU investors. The optimal strategies for project A, C and D are the same as in Table

8, whereas project B is now terminated in states M2/P1 and M1/P2 under both preference

models due to opportunity costs imposed by the securities, i.e. it is more profitable to invest in

146

Page 155: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

25

the securities in states M2/P1 and M1/P2 than in project B. In particular, this implies that the

opportunity costs imposed by securities can have a major effect on projects’ optimal

management strategies.

Table 10. Optimal project management strategies and project values.

Project

A B C D

M1/P1 X/X –/– X/X X/X

M2/P1 X/X –/– X/X X/X

M3/P1 X/X X/X X/X X/X

M4/P1 X/X X/X X/X X/X

M1/P2 –/– –/– X/X X/X

M2/P2 –/– X/X X/X X/X

M3/P2 –/– X/X X/X X/X

M4/P2 –/– X/X X/X X/X

Project value, MV model $4.13 $14.53 $8.54 $15.56

Project value, EU model $3.17 $14.15 $7.07 $15.56

In each case, except for project A under the EU model, the opportunity costs lower the

resulting project values. Notice also that project D is priced at $15.56 due to the existence of a

replicating portfolio: the price of 5 shares of security 2 minus the risk-free discounted NPV of

the costs is 5 $20/1.08 $40 $40/1.08 $15.56⋅ − − = .

In summary, securities typically lead to lower project values by imposing additional

opportunity costs. However, as noted by Gustafsson et al. (2004), they may also increase

project values by providing more efficient diversification possibilities. Furthermore, if a

replicating portfolio exists for a project, the project will be priced at its CCA value. Also, the

introduction of securities may, at times, change the optimal management strategies for projects.

[Q1]

7.4 Project Portfolio with Securities

In this section, we consider the case where the investor can simultaneously invest in projects A–

D as well as in the risk-free asset and in the two securities. Here, the optimal project

management strategies are the same as in Table 10, except that both MV and EU investors

147

Page 156: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

26

make here a further investment in project B in states M2/P1; this is largely due to the fact

that projects A and B are negatively correlated with each other. Note also that, apart from

project D, the values of the projects increase, because with more investment assets available,

the risks of each project can be diversified more efficiently than previously.

Table 11 presents the optimal security portfolio of the MV investor when the projects are

managed using the optimal project management strategy. This portfolio combines the effects

that were described in the previous section: in states where project B is continued (in all states

except M1/P1 and M1/P2), the investor buys five shares of security 1 less than she would

otherwise; on the other hand, the starting of project D implies that the investor buys five

shares of security 2 less in each state. With these modifications, the optimal trading strategy is

very close to the optimal trading strategy when only project C was started. The small

differences are due to correlations between projects.

Table 11. Optimal trading strategy for the MV investor with projects A–D.

Portfolio weights Securities bought

State 1 2 1 2 CS

0 100.0% 0.0% 11.136 0.000 -$121.69

M1/P1 74.7% 25.3% 9.495 6.319 -$363.27

M2/P1 76.2% 23.8% 5.167 3.912 -$239.70

M3/P1 105.3% -5.3% 4.731 -0.685 -$67.29

M4/P1 116.5% -16.6% 4.019 -1.915 $64.41

M1/P2 95.1% 4.9% 17.451 1.769 -$525.09

M2/P2 77.3% 22.7% 4.180 2.989 -$113.30

M3/P2 94.5% 5.5% 2.046 0.343 $127.43

M4/P2 114.8% -14.8% 2.488 -1.076 $210.33

Tables 12 and 13 present sensitivity analyses for project values with respect to the MV

investor’s risk tolerance R and the EU investor’s risk aversion parameter α , respectively. At

low risk tolerance levels (high degree of risk aversion), an MV investor gives high values for

projects A and B, whereas the values of the projects decrease as the investor becomes more risk

tolerant. In contrast, the price behavior of project C is opposite: project C obtains low values at

small risk tolerance levels and the values increase as the investor becomes more risk tolerant.

Project D is constantly priced at $15.56 due to the existence of a replicating portfolio. An EU

investor prices the projects similarly: values of projects A and B decrease by diminishing risk

148

Page 157: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

27

aversion, while the value of project C increases. Also in this case, the value of project D is

constant due to the existence of a replicating portfolio.

In both cases, project values converge to limit values as the investor becomes less risk averse.

However, these limit values are different, and in neither case they coincide with a risk neutral

investor’s project values ( $4.13, $9.35, $1.86, $12.83)− , obtained by using a short rate

13.58% in period 0 (the expected rate of return of security 1; Table 6) and 20.00% in period 1

(the expected rate of return of security 2; Table 5). Gustafsson et al. (2004) showed in their

numerical experiments that, in a two-period setting, the project values for an MV investor

converge towards the CAPM prices of the projects; the experiments in De Reyck et al. (2004)

showed a similar behavior for EU investors with CARA. The observed converge behavior for

both types of investors shows that a similar phenomenon can happen also in a multi-period

setting.

Table 12. Project values for an MV investor.

Risk level A B C D

75 $11.70 $16.87 $4.12 $15.56

100 $8.49 $16.21 $6.20 $15.56

125 $7.55 $15.96 $7.20 $15.56

150 $7.05 $15.82 $7.80 $15.56

175 $6.73 $15.73 $8.21 $15.56

200 $6.50 $15.67 $8.51 $15.56

225 $6.34 $15.62 $8.74 $15.56

250 $6.21 $15.58 $8.92 $15.56

300 $6.02 $15.53 $9.19 $15.56

350 $5.89 $15.49 $9.38 $15.56

400 $5.79 $15.46 $9.52 $15.56

450 $5.71 $15.44 $9.63 $15.56

500 $5.65 $15.42 $9.71 $15.56

1000 $5.38 $15.35 $10.10 $15.56

1500 $5.28 $15.32 $10.23 $15.56

2500 $5.21 $15.30 $10.33 $15.56

5000 $5.15 $15.28 $10.41 $15.56

50000 $5.10 $15.27 $10.48 $15.56

Project

149

Page 158: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

28

Table 13. Project values for an EU investor.

α A B C D

0.005 $8.03 $16.67 $5.97 $15.56

0.004 $7.43 $16.41 $6.87 $15.56

0.003 $6.83 $16.15 $7.79 $15.56

0.002 $6.24 $15.90 $8.74 $15.56

0.001 $5.67 $15.66 $9.71 $15.56

0.0001 $5.15 $15.46 $10.59 $15.56

0.00001 $5.10 $15.44 $10.68 $15.56

0.000001 $5.09 $15.44 $10.69 $15.56

Project

In summary, project values when the investor can invest simultaneously in projects A–D can be

different from the values obtained with single projects due to correlations between projects and

better diversification possibilities implied by a greater number of investment assets in the

portfolio. Also, the optimal management strategies for the projects may change. [Q2] Sensitivity

analysis reveals that the value of a project may either decrease or increase with increasing risk

aversion depending on its correlation with the rest of the investment portfolio. Furthermore, a

project with a replicating portfolio is constantly priced at its CCA value. Also, with decreasing

level of risk-aversion, project values converge towards prices that are different from the

projects’ risk-neutral values. These values are analogous to the CAPM prices and EU

convergence prices observed in a two-period setting. [Q3]

8 Summary and Conclusions In this paper, we have presented a model for selecting a multi-period mixed asset portfolio and

a framework that uses this model in the valuation of multi-period projects. In the MAPS model,

project management decisions are structured as project-specific decision trees. The project

valuation framework is based on the concepts of breakeven selling and buying prices

(Gustafsson et al. 2004, Luenberger 1998, Smith and Nau 1995); the breakeven selling price is

the lowest price at which a rational investor would sell the project if she had it, whereas the

breakeven buying price is the highest price at which the investor would buy the project if she

did not have it.

The present project valuation framework differs from Smith and Nau’s (1995) “full decision tree

analysis” in that (i) it maximizes the investor’s terminal wealth level rather than the

150

Page 159: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

29

intertemporal utility implied by the project and in that (ii) it takes into account the

opportunity costs imposed by alternative projects in the portfolio. The framework also (iii)

allows the use of a wide range of preference models, and it leads to an explicit optimization

problem, which can readily be solved using standard techniques of nonlinear programming. In

comparison with Smith and Nau’s (1995) integrated rollback procedure, which is also an

explicit project valuation method, the present framework relies on fewer assumptions about the

investor’s preference structure and the nature of project cash flows.

In addition to formulating the project valuation framework, we showed in this paper that, when

the investor uses either a preference functional or a mean-risk model, the framework gives

consistent values with CCA, wherefore it can be regarded as a generalization of CCA to

incomplete markets. In addition, we showed that the framework leads to Hillier’s (1963)

method when the investor can invest only in a single project and in the risk-free asset, and

when she exhibits both constant relative and constant absolute risk aversion.

In our numerical experiments, we examined the pricing behavior of mean-variance and expected

utility maximizers when markets were assumed to abide by the standard CAPM. We observed

that many of the phenomena occurring in a two-period setting also took place in the multi-

period setting. For example, we observed that, when compared to the situation without

securities, the presence of securities typically resulted in lower project values due to imposed

opportunity costs. Furthermore, replicating portfolios led to project values that were

independent of the investor’s risk attitude. We also observed that the value of a project, when

it did not have a replicating portfolio, could either increase or decrease with increasing risk

aversion, depending on its correlation with the rest of the investment portfolio. When the

investor became less risk averse, project values converged towards prices that were different

from the prices that a risk-neutral investor would give to the projects.

This work suggests several areas for further research. For example, it would be relevant to

examine the effect of alternative security pricing models, such as the multi-period models of

Merton (1973), Stapleton and Subrahmanyam (1978), and Levy and Samuelson (1992), on

project values and optimal trading strategies. Also, the convergence behavior of project values

demands more analysis: the results from Gustafsson et al. (2004) indicate that in a two-period

setting project values for a mean-variance investor converge towards the projects’ CAPM prices

as the investor becomes less risk averse; a similar behavior could potentially be proven in the

151

Page 160: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

30

multi-period case.

References

Ariel, R. 1998. Risk adjusted discount rates and the present value of risky costs. The Financial

Review 33 17–30.

Bar-Yosef, S., R. Metznik. 1977. On Some Definitional Problems with the Method of Certainty

Equivalents. Journal of Finance 32(5) 1729–1737.

Beedles, W. L. 1978. Evaluating Negative Benefits. Journal of Financial and Quantitative

Analysis 13(1) 173–176.

Brealey, R., S. Myers. 2000. Principles of Corporate Finance. McGraw-Hill, New York, NY.

Chen, S–N, W. T. Moore. 1982. Investment Decisions under Uncertainty: Application of

Estimation Risk in the Hillier Approach. Journal of Financial and Quantitative Analysis

17(3) 425–440.

De Reyck, B., Z. Degraeve, J. Gustafsson. 2004. Valuing Real Options in Incomplete Markets.

Working Paper. London Business School.

Dixit, A. K., R. S. Pindyck. 1994. Investment Under Uncertainty. Princeton University Press,

Princeton, NJ.

Eppen G. D., Martin R. K., Schrage L. 1989. A Scenario Based Approach to Capacity

Planning. Operations Research 37 517-527.

Fuller, R. J., S.–H. Kim. 1980. Inter-Temporal Correlation of Cash Flows and the Risk of

Multi-Period Investment Projects. Journal of Financial and Quantitative Analysis 15(5)

1149–1162.

Gallagher T. J., J. K. Zumwalt. 1998. Risk-Adjusted Discount Rates Revisited. The Financial

Review 26(1) 105–114.

Gotoh, J., H. Konno. 2000. Third Degree Stochastic Dominance and Mean-Risk Analysis.

Management Science 46(2) 289–301.

Gustafsson, J. De Reyck, B., Z. Degraeve, A. Salo. 2004. Project Valuation in Mixed Asset

Portfolio Selection. Working Paper. Helsinki University of Technology and London Business

School. Downloadable at http://www.sal.hut.fi/Publications/pdf-files/mgus04a.pdf.

Gustafsson, J., A. Salo. 2005. Contingent Portfolio Programming for the Management of Risky

Projects. Operations Research 53(5), forthcoming.

Hespos, R. F., P. A. Strassmann. 1965. Stochastic Decision Trees for the Analysis of Invest-

ment Decision. Management Science 11(10) 244–259.

152

Page 161: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

31

Hillier, F. S. 1963. The Deviation of Probabilistic Information for the Evaluation of Risky

Investments. Management Science 9(3) 443–457.

Hull, J. C. 1999. Options, Futures, and Other Derivatives. Prentice Hall, New York, NY.

Keeley, R. H., R. Westerfield. 1972. A Problem in Probability Distribution Techniques for

Capital Budgeting. Journal of Finance 27(3) 703–709.

Keeney, R. L., H. Raiffa. 1976. Decisions with Multiple Objectives: Preferences and Value

Tradeoffs. John Wiley & Sons, New York, NY.

Konno, H., H. Yamazaki. 1991. Mean-Absolute Deviation Portfolio Optimization and Its Ap-

plications to the Tokyo Stock Market. Management Science 37(5) 519–531.

Levy H., P. A. Samuelson. 1992. The Capital Asset Pricing Model with Diverse Holding

Periods. Management Science 38 1529–1542.

Lintner, J. 1965. The Valuation of Risk Assets and the Selection of Risky Investments in Stock

Portfolios and Capital Budgets. Review of Economics and Statistics 47(1) 13–37.

Luenberger, D. G. 1998. Investment Science. Oxford University Press, New York, NY.

Machina, M. 1989. Dynamic Consistency and Non-Expected Utility Models of Choice under

Risk. Journal of Economic Literature 27(4) 1622–1668.

Markowitz, H. M. 1952. Portfolio Selection. Journal of Finance 7(1) 77–91.

–––––. 1959. Portfolio Selection: Efficient Diversification of Investments. Cowles Foundation,

Yale.

–––––. 1987. Mean-Variance Analysis in Portfolio Choice and Capital Markets. Frank J.

Fabozzi Associates, New Hope, PA.

Merton, R. C. 1973. An Intertemporal Capital Asset Pricing Model. Econometrica 41 867–887.

Mulvey, J. M., G. Gould, C. Morgan. 2000. An Asset and Liability Management Model for

Towers Perrin-Tillinghast. Interfaces 30(1) 96–114.

Rabin, M. 2000. Risk Aversion and Expected-Utility Theorem: A Calibration Theorem.

Econometrica 68(5) 1281–1292.

Robichek, A. A., S. C. Myers. 1966. Conceptual Problems in the Use of Risk-Adjusted Dis-

count Rates. Journal of Finance 21(4) 727–730.

Rubinstein, A. 1973. A Mean-Variance Synthesis of Corporate Financial Theory. Journal of

Finance 28(1) 167–181.

Sharpe, W. F. 1964. Capital Asset Prices: A Theory of Market Equilibrium under Conditions of

Risk. Journal of Finance 19(3) 425–442.

Smith, J. E., R. F. Nau. 1995. Valuing Risky Projects: Option Pricing Theory and Decision

Analysis. Management Science 41(5) 795–816.

153

Page 162: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

32

Stapleton, R. C., M. G. Subrahmanyam. 1978. A Multiperiod Equilibrium Asset Pricing Model.

Econometrica 46 1077–1096.

Trigeorgis, L. 1996. Real Options: Managerial Flexibility and Strategy in Resource Allocation.

MIT Press, Massachussets.

von Neumann, J., O. Morgenstern. 1947. Theory of Games and Economic Behavior, Second

edition. Princeton University Press, Princeton.

Wald, A. 1950. Statistical Decision Functions. John Wiley & Sons, New York.

Appendix

Proofs of Propositions 1, 2, 3 and 5 Proof of Proposition 1: Let ω

*x denote the portfolio of securities in state ω that replicates the cash

flow pattern of the next period. Let ,0

( ) ( )n

j i ii

P S xωω ω ∗

=

= ∑ be the price of this replicating portfolio.

Let ω**x be the portfolio whose cash flow is equal to the value of the replicating portfolio in each

state of the next period. (For convenience, let us assume that ω =**x 0 if ω belongs to the last or

the second last period and that ω =*x 0 if ω belongs to the last period.) Then, * *ω ω ω= +** * *x x x is

the portfolio that replicates the cash flows of the next period and which can be used to construct

replicating portfolios for the period after the next period. Let us then define **ω**x so that it is

equal to *ω**x if ω belongs to the last or the second last period, and in other periods it is the

portfolio whose cash flow is equal to the value of the replicating portfolio **ω′**x in each state ω′ of

the next period. If there is a replicating trading strategy for the project, there exists portfolios **

ω**x for all states ω . Thus, the time-0 portfolio

0

**ω**x can be used to construct a replicating

strategy for the project. Its price is 00 ,

0

( )n

i ii

S xωω ∗∗∗∗

=∑ . Since a shorted replicating trading strategy,

which can be created using portfolio 0

**ω− **x , nullifies the cash flows of the project, the setting

where the investor invests in the project and shorts 0

**ω**x is identical with the situation where the

investor cannot invest in the project and receives money equal to 0

00 ,

0

( )n

j i ii

C S xωω ∗∗∗∗

=

− + ∑ , where 0jC

is the investment cost of the project. When the time-0 budget in the setting without the project is

( )0b ω , the respective budget is ( )0

00 0 ,

0

( )n

j i ii

b C S xωω ω ∗∗∗∗

=

− + ∑ with the project and portfolio 0

**ω− **x .

Hence, the breakeven buying price of the project is given by

0

00 0 , 0

0

( ) ( ) ( )n

bj i i j

i

b C S x v bωω ω ω∗∗∗∗

=

− + − =∑ , or 0

00 ,

0

( )n

bj j i i

i

v C S xωω ∗∗∗∗

=

= − + ∑ , which is also the CCA

price of the project. Conversely, in the setting with the project and budget ( )0b ω , we can nullify

the project cash flows by shorting 0

**ω**x and obtain an effective budget of

154

Page 163: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

33

( )0

00 0 ,

0

( )n

j i ii

b C S xωω ω ∗∗∗∗

=

− + ∑ . Hence, in the setting without the project, we will obtain an equally

desirable portfolio by increasing the budget by 0

00 ,

0

( )n

sj j i i

i

v C S xωω ∗∗∗∗

=

= − + ∑ , which is both the

breakeven selling price and the CCA price of the project. Q.E.D.

Proof of Proposition 2: Since the investor can invest only in a single project and in the risk-free

asset, in the setting without the project the investor invests her entire budget in the risk-free asset;

consequently, the portfolio is riskless and is its own certainty equivalent, i.e.

0(1 ) ( )TTV s b Oω− = + + , sT is the time-T spot rate and O is the future value of the budgets at

states other than 0ω . Also, changing the time-0 budget to 0( )b ω δ+ will give a portfolio value of

( )0(1 ) ( )TTV s b Oω δ∗ = + + + without the project. Setting V V+ ∗= we obtain

(1 )TTV V s δ+ −= + + , wherefrom it follows (1 )TT

V Vs

δ+ −−

=+

, which is also, by definition, the

breakeven selling price of the project. For a mean-risk investor, the proposition follows from the

observation that, when the risk-constraint is not binding, the preference model corresponds to a

preference functional model with CE[X] = E[X]. If the risk constraint prevents the starting of the

project, then 0(1 ) ( )TTV V s b Oω+ −= = + + , implying a zero value for the project. Q.E.D.

Proof of Proposition 3: When the investor can invest only in a single project and in the risk-free

asset, we have 0(1 ) ( )TTV s b Oω− = + + , where sT is the time-T spot rate and O is the future value

of the budgets at states other than 0ω . Also, we can separate the optimal terminal wealth level

with the project to random variable X, which equals to project’s future value (project’s cash flows

with the accrued interest), and to f that is the future value of the budgets. Since

[ ] [ ]CE X f CE X f+ = + , and denoting CE[X] by CET , we have TV CE f+ = + and V f− = .

Increasing the budget by δ both with and without the project will now increase the respective

certainty equivalent by (1 )TTs δ+ . Let V ∗ be the certainty equivalent of the optimal portfolio

when the investor invests in the project and the time-0 budget is 0( )b ω δ− . Then,

(1 )TTV V s δ∗ += − + . Setting V V− ∗= , we have (1 )TTV V s δ− += − + , and hence we have

(1 )TT

V Vs

δ+ −−

=+

, which is also the breakeven buying price of the project. Noticing that

TV V CE+ −− = , we immediately obtain the desired formula. Using further this observation and

Proposition 2, the proposition is also proved for the breakeven selling price. For a mean-risk

investor, the proposition follows from the observation that, when the risk-constraint is not binding,

the preference model corresponds to a preference functional model with CE[X] = E[X]. If the risk

constraint prevents the starting of the project, the project will have a zero value. Q.E.D.

155

Page 164: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

34

Proof of Proposition 5: The proposition is proven with a three-period counterexample. Consider an

investor with an intertemporal utility function 1 2 31 0.95 0.90 1 2( , , ) 5 c c cu c c c e e e∗ − − −= − − − , where we

have assumed CARA and additive independence for the sake of analogy with Smith and Nau

(1995). Note that the parameters of the utility function reflect the investor’s risk aversion and the

investor’s subjective perception about the relative importance of the cash flows received at different

periods, and hence they can be selected independently of the risk-free interest rate or available

securities. Consider then a project that yields a cash flow stream (–1, 1.1, 0) with 50% and a cash

flow stream ( 1,0,1.22)− with 50%. There is also a riskless zero-coupon bond yielding the cash flow

stream ( 1,0,1.21)− for sure. The risk-free interest rate is 10%, the time-0 budget is 1, and no other

securities are available. The terminal wealth level with the project is 1.21 with 50% and 1.22 with

50%. The bond yields a terminal wealth level of 1.21 for sure. Note that the terminal wealth level

implied by the project stochastically dominates the one implied by the bond. The expected

intertemporal utility for the project is 0.939 and for the bond 0.945. Thus, the bond is preferred to

the project even though the project’s terminal wealth level stochastically dominates that of the

bond. Q.E.D.

156

Page 165: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

MULTIPLE CRITERIA DECISION MAKING IN THE NEW MILLENNIUM, M. KÖKSALAN, S. ZIONTS (EDS.), LECTURE NOTES IN ECONOMICS AND MATHEMATICAL SYSTEMS 507, 2001, SPRINGER-VERLAG, BERLIN. REPRINTED WITH THE PERMISSION OF THE PUBLISHER.

PRIME Decisions: An Interactive Tool for Value Tree Analysis

Janne Gustafsson, Ahti Salo, and Tommi Gustafsson Helsinki University of Technology, Systems Analysis Laboratory P.O. Box 1100, FIN-02015 HUT, Finland Abstract. Several methods for the processing of incomplete preference information in additive preference models have been proposed. Due to the lack of adequate decision aiding tools, however, only a few case studies have been reported so far. In this paper, we present PRIME Decisions, a decision aiding tool which supports the analysis of incomplete preference information with the PRIME method. PRIME Decisions also offers novel features such as decision rules and guided elicitation tours. The application of PRIME Decisions is illustrated with a case study on the valuation of a high-tech company.

Keywords. Decision analysis, interval judgement, incomplete information, linear programming, decision aiding tools.

157

Page 166: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

168

Page 167: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

Systems Analysis Laboratory Research Reports, Series A A91 Electricity Derivative Markets: Investment Valuation, Production May 2005 Planning and Hedging Erkka Näsäkkälä A90 Optimal pilot decisions and flight trajectories in air combat March 2005 Kai Virtanen A89 On foresight processes and performance of innovation networks November 2004 Jukka-Pekka Salmenkaita A88 Systems Intelligence - Discovering a hidden competence in human October 2004 action and organizational life Raimo P. Hämäläinen and Esa Saarinen, eds. A87 Modelling studies on soil-mediated response to acid deposition and October 2003 climate variability Maria Holmberg A86 Life cycle impact assessment based on decision analysis June 2003 Jyri Seppälä A85 Modelling and implementation issues in circuit and network planning May 2003 tools Jukka K. Nurminen A84 On investment, uncertainty, and strategic interaction with April 2003 applications in energy markets. Pauli Murto A83 Modeling of microscale variations in methane fluxes November 2002 Anu Kettunen A82 Modelling of water circulation and the dynamics of phytoplankton October 2001 for the analysis of eutrophication Arto Inkala Systems Analysis Laboratory Research Reports, Series B B25 Systeemiäly 2005 May 2005 Raimo P. Hämäläinen ja Esa Saarinen, toim. B24 Systeemiäly - Näkökulmia vuorovaikutukseen ja kokonaisuuksien June 2004 hallintaan Raimo P. Hämäläinen ja Esa Saarinen, toim. B23 Systeemiäly! April 2003 Tom Bäckström, Ville Brummer, Terhi Kling ja Paula Siitonen, toim. B22 Loruja - Esseitä uoman ulkopuolelta April 2002 Tommi Gustafsson, Timo Kiviniemi ja Riikka Mäki-Ontto, toim.

169

Page 168: PORTFOLIO OPTIMIZATION MODELS FOR PROJECT VALUATION

Systems Analysis Laboratory Research Reports, Series E Electronic Reports: www.e-reports.sal.hut.fi E18 Valuing Risky Projects with Contingent Portfolio Programming June 2005 Janne Gustafsson and Ahti Salo E17 Project Valuation under Ambiguity June 2005 Janne Gustafsson and Ahti Salo E16 Project Valuation in Mixed Asset Portfolio Selection June 2005 Janne Gustafsson, Bert De Reyck, Zeger Degraeve and Ahti Salo E15 Hydropower production planning and hedging under inflow and May 2005 forward uncertainty Erkka Näsäkkälä and Jussi Keppo E14 Models and algorithms for network planning tools - practical May 2003 experiences Jukka K. Nurminen E13 Decision structuring dialogue April 2003 Sebastian Slotte and Raimo P. Hämäläinen E12 The threat of weighting biases in environmental decision analysis April 2003 Raimo P. Hämäläinen and Susanna Alaja E11 Timing of investment under technological and revenue related March 2003 uncertainties Pauli Murto E10 Connecting methane fluxes to vegetation cover and water table November 2002 fluctuations at microsite level: A modeling study Anu Kettunen E9 Comparison of Hydrodynamical Models of the Gulf of Finland in 1995 September 2001 - A Case Study Arto Inkala and Kai Myrberg E8 Rationales for Government Intervention in the Commercialization of September 2001 New Technologies Jukka-Pekka Salmenkaita and Ahti A. Salo E7 Effects of dispersal patterns on population dynamics and synchrony April 2000 Janica Ylikarjula, Susanna Alaja, Jouni Laakso and David Tesar Series A and E reports are downloadable at www.sal.hut.fi/Publications/ Orders for Helsinki University of Technology paper copies: Systems Analysis Laboratory P.O. Box 1100, FIN-02015 HUT, Finland [email protected]

170