groundwater games: users’ behavior in common-pool resource

177
The Pennsylvania State University The Graduate School Department of Agricultural Economics, Sociology and Education GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE ECONOMIC LABORATORY AND FIELD EXPERIMENTS A Dissertation in Agricultural, Environmental, and Regional Economics by Rodrigo Salcedo Du Bois c 2014 Rodrigo Salcedo Du Bois Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy May 2014

Upload: others

Post on 16-Mar-2022

8 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

The Pennsylvania State University

The Graduate School

Department of Agricultural Economics, Sociology and Education

GROUNDWATER GAMES:

USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

ECONOMIC LABORATORY AND FIELD EXPERIMENTS

A Dissertation in

Agricultural, Environmental, and Regional Economics

by

Rodrigo Salcedo Du Bois

c© 2014 Rodrigo Salcedo Du Bois

Submitted in Partial Fulfillmentof the Requirements

for the Degree of

Doctor of Philosophy

May 2014

Page 2: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

The dissertation of Rodrigo Salcedo Du Bois was read and approved1 by the following:

James S. ShortleDistinguished Professor of Agricultural and Environmental EconomicsDissertation Co-AdviserCo-Chair of Committee

David AblerProfessor of Agricultural, Environmental and Regional Economicsand DemographyDissertation Co-AdviserCo-Chair of Committee

Jill L. FindeisDistinguished Professor Emerita of Agricultural, Environmental andRegional Economics and Demography

Edward CoulsonProfessor of Economics

Richard ReadyProfessor of Agricultural and Environmental EconomicsCoordinator, Graduate Program in Agricultural, Environmentaland Regional Economics

1Signatures on file in the Graduate School.

Page 3: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

iii

Abstract

Groundwater currently provides 30 percent of freshwater in the world, and it is

estimated that it potentially constitutes approximately 89 percent of the world’s fresh

water. Nevertheless, extensive extraction of groundwater in some areas is leading to

aquifer depletion.

Groundwater is considered, in some situations, a common-pool resource (CPR)

with extremely high use value. Economists and political scientists have devoted a great

amount of effort to understand how common pool resources are managed, as well as the

characteristics of institutions that emerge in order to deal with the use and distribution

of CPR. The characteristics of a CPR and how agents interact in their use are important

for the success and sustainability of CPR-dependent communities.

This dissertation will analyze the behavior and decision rules of CPR users in

an experimental context. We discuss the results from laboratory and field experiments

framed as a dynamic groundwater game in which users pump water from a shared re-

source, and the actions of users in one period affect resource availability and extraction

costs of everyone in the next period. This is an innovative design, since most of previous

studies do not consider the groundwater problem in experimental settings, and do not

conduct a groundwater experiment with actual groundwater users.

The objectives of this research are twofold. In the first part of the dissertation,

I assess the presence of strategic behavior when participants have access to an efficient

technology that yields both private and public benefits to CPR users. This situation

Page 4: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

iv

might discourage users from adopting the new technology, which results in free-riding.

In this part of the dissertation, I am also concerned about the effectiveness of group

arrangements to improve water usage and to guarantee appropriation levels that yield

a socially-desirable economic outcome. Given the experimental design, I can partially

address both learning effects during the game and unobserved heterogeneity of partici-

pants, in order to obtain a precise estimation of the impact of the treatments. Finally, I

also analyze the factors that affect technology adoption and deviation from agreed water

usage.

In the second part of the dissertation, I address the degree of heterogeneity in

actions of participants, and aim to identify different types of behavior among participants

in the use of the shared resource . Using a methodology proposed by Geweke and Keane

(2000) and Houser et al. (2004), I can relax the rational expectations assumption in the

dynamic choice problem, and identify different behaviors in the population. Based on

the decisions that participants make during the experiment, I can identify and cluster

parameters that define the future function of the dynamic choice problem in different

“types” or groups of participants, with these groups endogenously created. I propose

a flexible mixed multinomial discrete choice model that allows parameters to be drawn

from a discrete mixture of distributions. In order to estimate the parameters and group

membership probabilities, Bayesian Markov Chain Monte Carlo techniques are used.

These methods allow the estimation of highly dimensional discrete choice problems and

adds flexibility to estimation and the clustering process.

Page 5: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

v

Table of Contents

List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii

List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viii

Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . x

Chapter 1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

I 8

Chapter 2. The Economics of Groundwater Management . . . . . . . . . . . . . 9

Chapter 3. The Groundwater Management Game . . . . . . . . . . . . . . . . . 173.1 Theoretical Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

3.1.1 Basic Game (Baseline) . . . . . . . . . . . . . . . . . . . . . . 183.1.2 Investment . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233.1.3 Parametrization, Myopic, Rational and Optimal Behavior . . 25

3.2 Experimental Design . . . . . . . . . . . . . . . . . . . . . . . . . . . 353.2.1 Recruitment of Participants and Framing . . . . . . . . . . . 353.2.2 Treatments and Sessions . . . . . . . . . . . . . . . . . . . . . 363.2.3 Rewards . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 393.2.4 Survey . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

Chapter 4. Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . . . 454.1 Experiment Outcomes . . . . . . . . . . . . . . . . . . . . . . . . . . 454.2 Treatment Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 634.3 Technology Adoption . . . . . . . . . . . . . . . . . . . . . . . . . . . 684.4 Agreement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 724.5 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

II 81

Chapter 5. Dynamic Decision-making in the GroundwaterExperiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82

5.1 Unbounded Rationality, Bounded Rationality and CPR . . . . . . . 825.2 Behavioral Heterogeneity . . . . . . . . . . . . . . . . . . . . . . . . . 875.3 Structural Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96

Page 6: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

vi

Chapter 6. Empirical Specification and Estimation . . . . . . . . . . . . . . . . . 1006.1 Discrete Mixture Models . . . . . . . . . . . . . . . . . . . . . . . . . 1006.2 Multinomial Probit . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1046.3 Estimation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

6.3.1 Bayesian Methods . . . . . . . . . . . . . . . . . . . . . . . . 1096.3.2 Data Likelihood and Parametrization . . . . . . . . . . . . . . 1136.3.3 Priors and Sampling Algorithm . . . . . . . . . . . . . . . . . 115

Chapter 7. Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . 1187.1 Deviance Information Criteria and Group Number . . . . . . . . . . 1187.2 Parameter Statistics and Convergence . . . . . . . . . . . . . . . . . 1217.3 Classification and Characteristics of Groups . . . . . . . . . . . . . . 1277.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133

Chapter 8. Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . 135

Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140A Functional Forms of Polynomials of a Higher Degree . . . . . . . . . 140B MCMC of Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . 151

Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159

Page 7: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

vii

List of Tables

3.1 Parameters used in experiment . . . . . . . . . . . . . . . . . . . . . . . 273.2 Revenues with less-efficient technology . . . . . . . . . . . . . . . . . . . 283.3 Revenues with highly-efficient technology . . . . . . . . . . . . . . . . . 293.4 Pumping costs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303.5 Summary of sessions: Laboratory and field experiments . . . . . . . . . 373.6 Summary statistics of survey variables, laboratory experiment . . . . . . 423.7 Summary statistics of survey variables, field experiment . . . . . . . . . 43

4.1 Average outcomes by type of session, round and period: Laboratoryexperiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

4.2 Average outcomes by type of session, round and period: Field experiments 484.3 Average end-of-round outcomes by type treatment and round: Labora-

tory experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 584.4 Average end-of-round outcomes by type treatment and round: Field ex-

periments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 584.5 Difference-in-Difference estimation of final outcomes in rounds 1 and 2:

Lab experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 654.6 Difference-in-Difference estimation of final outcomes in rounds 1 and 2:

Field experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 664.7 Cox proportional hazard model of investment: Laboratory experiments . 714.8 Cox proportional hazard model of deviation from agreement: Field ex-

periments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

7.1 Deviance Information Criterion . . . . . . . . . . . . . . . . . . . . . . . 1207.2 Distribution of groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1217.3 Descriptive statistics of posterior distributions: Field experiment . . . . 1227.4 Descriptive statistics of posterior distributions: Laboratory experiment . 1237.5 Total benefits and final well depth . . . . . . . . . . . . . . . . . . . . . 132

Page 8: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

viii

List of Figures

3.1 Equilibrium Pumping hours for Myopic, Rational/strategic and FullyCooperative behaviors . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

3.2 Benefits of pumping for Myopic, Rational/strategic and Fully Coopera-tive behaviors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

3.3 Well depth for Myopic, Rational/strategic and Fully Cooperative behaviors 333.4 Simulated benefits for Myopic, Rational/strategic and Fully Cooperative

behaviors, and fixed arrangements . . . . . . . . . . . . . . . . . . . . . 34

4.1 Average individual hours pumped by period and treatment . . . . . . . 504.2 Average individual benefits by period and treatment . . . . . . . . . . . 514.3 Average well depth by period and treatment . . . . . . . . . . . . . . . . 524.4 Histogram of pumping hours from laboratory experiment, by Round,

Period and Treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . 554.5 Histogram of pumping hours from field experiment, by Round, Period

and Treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 564.6 Average individual total benefits by round and treatment . . . . . . . . 604.7 Average individual total hours pumped by round and treatment . . . . . 614.8 Average final well depth by round and treatment . . . . . . . . . . . . . 624.9 Percentage of technology adoption by period and experiment . . . . . . 694.10 Percentage of technology adoption by period and round: Laboratory

experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 694.11 Histogram of agreed values of pumping hours in “AG” sessions . . . . . 744.12 Agreement deviation percentages in “AG” sessions . . . . . . . . . . . . 76

5.1 Average individual hours pumped by experiment and period . . . . . . . 895.2 Hours of pumping by period: Laboratory experiment . . . . . . . . . . . 915.3 Hours of pumping by period: Field experiment . . . . . . . . . . . . . . 925.4 Average individual hours pumped by well depth . . . . . . . . . . . . . . 94

6.1 Single vs. Mixed Distributions . . . . . . . . . . . . . . . . . . . . . . . 102

7.1 Estimated posterior distributions of parameters: Field experiment . . . 1247.2 Estimated posterior distributions of parameters: Laboratory experiment 1257.3 Trace of deviance, two chains . . . . . . . . . . . . . . . . . . . . . . . . 1267.4 Average individual hours pumped by period and type . . . . . . . . . . 1287.5 Average individual pumping costs by period and type . . . . . . . . . . 1297.6 Histogram of choices by cluster and period: Field experiment . . . . . . 1317.7 Histogram of choices by cluster and period: Laboratory experiment . . . 131

B.1 Trace of the MCMC of all the parameters of the future function of thefield experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152

Page 9: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

ix

B.2 Trace of the MCMC of all the parameters of the future function of thefield experiment - Group 1 . . . . . . . . . . . . . . . . . . . . . . . . . . 153

B.3 Trace of the MCMC of all the parameters of the future function of thefield experiment - Group 2 . . . . . . . . . . . . . . . . . . . . . . . . . . 154

B.4 Trace of the MCMC of all the parameters of the future function in thelaboratory experiment . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155

B.5 Trace of the MCMC of all the parameters of the future function of thelaboratory experiment - Group 1 . . . . . . . . . . . . . . . . . . . . . . 156

B.6 Trace of the MCMC of all the parameters of the future function of thelaboratory experiment - Group 2 . . . . . . . . . . . . . . . . . . . . . . 157

B.7 Trace of the MCMC of all the parameters of the future function of thelaboratory experiment - Group 3 . . . . . . . . . . . . . . . . . . . . . . 158

Page 10: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

x

Acknowledgments

This research would not have been possible to conduct without the support of

the Latin American and Caribbean Environmental Economics Program (LACEEP). I

am indebted with them for all their help. I also would like to thank my advisers James

Shortle and David Abler for their help, support and patience in all this process, especially

at the hard times. I would like to thank Jill Findeis for her encouragement and support,

especially during the first part of the process. I am also thankful to Edward Coulson

for his feedback and support. I am deeply indebted with the farmers of Aguascalientes

and the Penn State students that participated in the experiments. Without them, this

dissertation would not exist. Professors and authorities of the Universidad Autnoma de

Aguascalientes were extremely helpful in the organization of the sessions. In particular

professors Joaqun Sosa, Jess Meraz and Jos Luna. I would also like to thank Ing. Juan

Zamarripa, who was key in the recruitment of many farmers and the development of many

sessions with well owners. The work of Berenice Castillo, Viviana del Hoyo, Gabriela

Flores, AnneLiese Nachman and Charlotte Benson was crucial for the implementation

of the sessions. I also appreciate the comments received during the LACEEP workshops

and the European Summer School in Resource and Environmental Economics, especially

the comments received by Erik Ansink and Ariel Dinar. I would like to express my deep

gratitude to Miguel Angel Gutierrez and all my family in Aguascalientes, Marta, Susana,

Michel, Tey and Miguel. Their love and care made me feel as home in Aguascalientes. I

would like to thank all my friends in State College, especially David, Andrea, Roberto,

Page 11: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

xi

Eli, Carlos, Erin, Miguel, Claudia, Hernan, GG, Luis, Juancito, Wil and Sandra. State

College would not have been the nice place that was during this time without them.

Thanks to all my extended family that was always supporting and encouraging me,

and my immediate family who always cared about my physical and mental well being.

Finally, I am deeply indebted to Veronica, who always was on my side helping and giving

me suggestions and comments, but most importantly, support and strength.

Page 12: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

xii

Dedication

A mis padres, quienes con su trabajo sacrificado pudieron darme la salud y la edu-cacion para poder crecer profesional y humanamente; y a mi esposa, Veronica, quien consu amor incondicional y alegrıa pude obtener las fuerzas necesarias para poder terminaresta etapa, y ası poder continuar una vida juntos.

Page 13: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

1

Chapter 1

Introduction

Groundwater currently provides 30 percent of freshwater in the world IWMI

(2010), and it is estimated that groundwater potentially constitutes approximately 89

percent of the world’s fresh water (Koundouri 2000). Groundwater has been key on the

provision of freshwater in arid areas, and subsequent poverty reduction and food secu-

rity FAO (2003). However, extensive extraction of groundwater in some areas is leading

to aquifer depletion. Many areas in the world in which groundwater is essential, such

as Western US, Mexico and India, show important declines on the water table Sekhri

(2011). Some problems related to groundwater overexploitation are increasing pumping

costs; land subsidence and further damages to surface infrastructure; changes in ecosys-

tems that rely on groundwater, such as wetlands; and availability of water for drinking

and irrigation (FAO 2003). Moreover, water quality problems are becoming frequent

in some of these areas, since high concentrations of fluor and arsenic are deposited in

deeper levels of groundwater (IWMI 2010)

Groundwater is considered, in some situations, a common-pool resource (CPR)

with extremely high use value (FAO 2003). Economists and political scientists have de-

voted a great amount of effort to understand the use of common pool resources (McGinnis

2000). CPR are, for economic and institutional reasons, nonexcludable. (Ostrom and

Gardner 1993; McGinnis 2000). Examples of CPR include fisheries, pastures, irrigation

Page 14: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

2

systems, or even oceans and the biosphere (Ostrom and Gardner 1993). Traditional

microeconomic theory predicts that the rent-seeking behavior of individuals will result

in over-exploitation, degradation and possible extinction of such resources, the so called

“Tragedy of the Commons” (Hardin 1968).

Some scholars argue that the solution to the tragedy of the commons is privati-

zation through assignment of clear property rights. If enforceable, property rights can

ensure exclusion and provide incentives for efficient use. Others point to challenges re-

sulting from high enforcement and transaction costs, leading to market failures. Due

to these market failures, they propose that a government should be responsible for the

management of the resource and guarantee the socially optimal exploitation. Recently,

a third line argues that communities can self-regulate (Coward 1976; Ostrom 1986, 1990,

1992; Ostrom and Gardner 1993; Trawick 2003). These scholars argue that rules that

are internally created and agreed upon by the community, along with a set of tools that

ensure the enforcement of those rules, can be effective in the provision and preserva-

tion of CPR (Ostrom 1990; Ostrom and Gardner 1993; Dayton-Johnson 2000; Trawick

2003; Swallow et al. 2006). These studies have found some evidence that cooperation

among the users of a CPR can emerge, and that face-to-face communication between the

agents is important to ensure compliance of rules (Hackett et al. 1994; Cardenas 2011;

Moreno-Sanchez and Maldonado 2010).

One way to increase the availability of the resource to all its users is through

the adoption of technology that improves the efficiency of how people use the resource.

For instance, in the case of groundwater, users can adopt efficient irrigation technologies

(e.g. drip irrigation) and reduce exploitation of the resource. Nevertheless, in the case

Page 15: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

3

of CPR, private investment in efficient technologies can yield benefits to all users. Thus,

some users might have incentives to free ride, preventing the adoption of the efficient

technology unless arrangements are made within the group.

In this regard, the characteristics of a system and how agents interact within are

important for the success of the community as an institution that ensures the sustainable

and responsible extraction of the resource. Moreover, the dynamics of the system and

how CPR users make decisions, as well the study of the behavior they manifest in a

CPR setup is important. It is necessary to understand the conditions working for and

against the sustainability of local cooperation in situations of general social and economic

interdependence (Bardhan 2000).

The analysis of institutions involved in CPR’s and the behavior of users demands

an analytical framework that goes beyond the one traditionally used by economists that

assumes non-cooperative rent seeking. We usually assume that agents choose“as if”

he/she is unboundedly rational, meaning that the potential to develop rational thinking,

given the information on hand, has no limits (Conlisk 1996). However, there is strong

evidence that agents make systematic mistakes from the perspective of economic ratio-

nality (Simon 1955; Kahneman and Tversky 1979; Conlisk 1996; Camerer 1998; Ellison

2006).

Several labels, such as “bounded rationality”, “behavioral failures” or “rules of

thumb” have been used to name those types of behavior that depart from economic

rationality. Unfortunately, there is no agreement on how these behaviors can be handled

analytically, and thus have been analyzed separately in different contexts. This is the

major limitation of non-traditional models. Their inability to replace the traditional

Page 16: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

4

economically rational framework is due the lack of a unified formal conceptual framework,

although it has been proven that these types of behaviors are consistently supported by

data Harstad and Selten (2013).

According to Shogren et al. (2010), the economists that focus on the analysis

of deviations from unboundedly rational behavior (conventionally called “behavioral

economists”) have clearly shown two important facts of economic behavior not allowed

by the traditional economic perspective: i) preferences are context-dependent, and ii)

social preferences have an important role in economic choice. The challenge is to in-

corporate these facts in the analysis conducted by environmental and natural resource

economists, where an unbiased valuation of environmental non-market goods and the

design of institutions and mechanisms that promote cooperation in a CPR context are

needed.

Another challenge is to identify and classify these types of behaviors within a

heterogenous population. This is key in a CPR context, since the composition of the

group might have important effects on the final state of the resource and well being

of users. One way to identify the behavior of subjects is, in experimental setups, to

design the experiment in order to verify whether participants behave according to one

theory or another. To mention some studies, Lettau et al. (1999) consider different

types of agents and propose a theoretical model in which individuals choose the best

of several decisions rules, based on the state and comparison over past experiences.

Similarly, Suleimain and Rapoport (1997) consider three types: Those concerned with

equity, those who maximized utility and those that cannot be categorized. Moreover,

Page 17: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

5

Fischbacher et al. (2001) find that a significant number of agents that participated of a

laboratory experiment can be categorized as “conditional cooperators”.

A slightly more general approach is the one followed by El-Gamal and Grether

(1995). They propose a pre-defined set of behaviors that participants might show and

then classify participants according to their actions and how close they are to the pre-

defined theoretical behaviors. A more flexible approach is one in which researchers “let

the data speak”, and behavior types are clustered according to the actions agents make.

After the clustering process, one can categorize the different behaviors existent in the

population. This approach is followed by Houser et al. (2004).

In summary, allowing behavior to depart from economic rationality implies not

only the formulation of theories that might explain the behavior of agents in a better

way, but also entails the identification of these behaviors in a heterogenous population.

Given that relaxing the rational behavior assumption opens the door to many other

behavioral theories that, in addition, allow context-dependent preferences, the identifi-

cation of agents following one or another behavior seems to be a necessary but difficult

task to pursue.

The present work contributes to this last objective. In this dissertation, I will

analyze and discuss the results from CPR laboratory experiments, in which the results

are intended to reveal the behavior of users. I propose a dynamic groundwater game

in which users pump water from a shared well, and the actions of users in one period

affect the extraction costs of everyone for the next period. This is an innovative design,

since most of previous studies do not consider the groundwater problem in experimental

settings, and do not conduct a groundwater experiment with farm groundwater users.

Page 18: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

6

The objectives of this research are twofold. In the first part of the dissertation,

I assess the presence of strategic behavior when participants have access to an efficient

technology that yields both private and public benefits to CPR users. This situation

might discourage users from adopting the new technology, which results in free-riding.

In this part of the dissertation, I am also concerned about the effectiveness of group ar-

rangements to improve water usage and to endorse certain appropriation levels that yield

a socially-desirable economic outcome. Given the experimental design, I can partially

address both learning effects during the game and unobserved heterogeneity, in order to

obtain a precise estimation of the impact of the treatments. Finally, I also analyze the

factors that affect technology adoption and deviation from agreed water usage.

In the second part of the dissertation, I analyze the degree of heterogeneity of the

choices that participants make, and aim to identify the decision rules that participants

use during the game. Using a methodology proposed by Geweke and Keane (2000) and

Houser et al. (2004), I can relax the rational expectations assumption in the dynamic

choice problem, and identify different behaviors in the population. Based on the deci-

sions that participants make during the experiment, I can identify and cluster “types”

or groups of participants with a flexible mixed multinomial discrete choice model that

allows parameters from the structural model to be drawn from a discrete mixture of

distributions. In order to estimate the parameters and group membership probabilities,

Bayesian Markov Chain Monte Carlo techniques are used. These methods allow the esti-

mation of highly dimensional discrete choice problems and adds flexibility to estimation

and the clustering process.

Page 19: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

7

In this dissertation, I repeatedly compare the results obtained in the laboratory

experiment with those obtained by Salcedo and Gutierrez (2014) and Salcedo (2014),

henceforth LACEEP 1 (2014) and LACEEP 2 (2014), respectively. In these studies,

the authors recruited 256 farmers from the State of Aguascalientes, Mexico in order to

conduct the same experimental design presented in this dissertation but on the field.

Details about the implementation of these experiments can be found in LACEEP 1

(2014).

Page 20: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

Part I

8

Page 21: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

9

Chapter 2

The Economics of Groundwater Management

Groundwater characterizes for being replenishable but depletable. Aquifers recharge

over time due to filtration of surface water. However, this process could be very slow,

and when the total use of groundwater exceeds recharge, the resource will be depleted.

Depletion of groundwater continues until the marginal cost of extraction is prohibitive

or the stock of water in the aquifer is exhausted. This leads to the consideration of the

use value of groundwater, since water used today reduces the opportunity of obtaining

future benefits from water. Then, an efficient allocation of water entails the considera-

tion of both the marginal cost of pumping (increasing with the water table) and the use

value. The initial economic studies that analyzed the groundwater problem focused on

the analysis of the efficient allocation of the resource. In the seminal works of Burt and

Brown and Deacon (Burt 1964, 1967, 1970; Brown and Deacon 1972), the authors used

dynamic optimization models to analyze the optimal allocation of water over time.

Later, economists started to address groundwater allocation in a competitive econ-

omy, exploiting the characteristic of open access of common-property resources. Conven-

tionally, groundwater users’ behavior is assumed to be myopic, with no consideration of

the “use value” of water. In this setup, the current marginal value of water is equalized

to the current marginal cost of water extraction (Gisser and Sanchez 1980; Nieswiadomy

1985; Worthington et al. 1985, see).

Page 22: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

10

Most of these studies conclude that the optimal path in the controlled and the

competitive situations is the same, implying that there are no welfare gains from a

controlled optimal allocation of groundwater. This result is usually referred as the Gisser-

Sanchez effect, and it relies on the non-excludable property of groundwater. However, as

pointed out by Koundouri (2004a,b), there are reasons to believe that the conclusions of

Gisser and Sanchez should be taken with caution, since several restrictive assumptions

are made. When these assumptions are relaxed, the features of the Gisser-Sanchez

effect do not necessarily hold. First, some of the assumptions that Gisser and Sanchez

make for their model might drive the results (e.g. linear negatively sloped demand, linear

pumping cost). Second, there is no interaction among users, and therefore the possibility

of strategic behavior is absent. Third, the model does not consider heterogeneity among

users and/or locations. Fourth, it is assumed that the aquifer will never be depleted.

Finally, the model is deterministic, there is no uncertainty on recharge or maximum

capacity.

For instance, Worthington et al. (1985) solve a dynamic water allocation model

for a confined aquifer system in Southwest Montana. Due to the topological structure

of the confined system, the cost function of water pumping is nonlinear. They find

significant welfare loses under the perfect competition compared to the dynamically

efficient allocation, especially when land productivity heterogeneity exists. In a similar

way, Brill and Burness (1994) find that the differences between optimal and competitive

setups increase with a growing (non-stationary) demand, declining well yields (non-linear

cost), and low social discount rates.

Page 23: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

11

Later, economists began using game theory to develop groundwater analytical

models. The major focus of this group of studies was to analyze the behavior and

interactions among water users given the strategic behavior that might arise in different

situations, as well as the welfare losses due to strategic behavior. Negri (1989) argues

that little attention has been paid to the “more realistic” situation in which access to the

resource could be restricted to some users. In such situations, rents do not fully dissipate

(in opposition to unrestricted setups), and welfare loses vary inversely with the number

of resource users (Negri 1989). In this situation, game-theoretic tools are appropriate.

Studies that rely on game theory for the analysis of CPR usually focus on the

sources of inefficiency based on three types of externalities generated by the appropriation

of the resource: i) Stock (Cost) Externalities, that arise when changes in the stock of the

resource affect the cost of extraction of the resource to all users; ii) Strategic Externalities,

related to the common-property feature of the resource and the difficulty of property

rights allocation, which might encourage users to extract more than optimal level of the

resource because of fear of appropriation of the scarce resource from other users in the

future (Negri 1989); and iii) Congestion Externalities, related to the spatial distribution

of the points of extraction of the resource. Provencher and Burt (1993) also identifies, iv)

Risk Externalities, that arise when the uncertainty in the availability of surface water

is considered, which increases the optimal use value of groundwater for all firms, but

firms fail to internalize this value. Gardner et al. (1990) also considers “technological

Page 24: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

12

externalities”, that arise when the presence of a new technology adopted by one group

of users affects the extraction costs of those that did not adopt the technology. 1.

Negri distinguishes between open-loop and feedback solutions for the dynamic

program. Open-loop solutions are based on information that users hold at the begin-

ning of the game, which obligate users to commit to an efficient extraction path for the

whole time span. For the case of groundwater extraction, these solutions only consider

inefficiencies based on stock or cost externalities. On the other hand, feedback strate-

gies depend on current-period information. This type of behavior incorporates dynamic

strategic behavior, and therefore include both stock and strategic externalities. Negri

argues that commitment to an extraction path in open-loop solutions is not credible

with a lack of a property rights structure, and therefore these type of solutions are less

robust than feedback solutions. He concludes that resource extraction in feedback solu-

tions tend to increase the likelihood of overexploitation of the resource in comparison to

open-loop solutions.

The analysis developed by Negri (1989) yields to the conclusion that strategic

behavior affects the rates of water pumping, and therefore, steady-state stock levels are

lower than the open-loop equilibrium. Provencher and Burt (1993) take Negri’s argu-

ment, but relates his conclusion about strategic externalities to the fact that a finite stock

of water in the aquifer aggravates the inefficiencies that already exist (cost externality).

In their study, Provencher and Burt analyze the strategic behavior of groundwater users

where there is a limited stock of water (the name they assigned to this type of externality

1For this study, I will not consider risk and congestion externalites, given that none ofthe parameters considered in the experiments are stochastic, nor spatial or network effects areintroduced.

Page 25: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

13

was stock externality). They analytically calculate both the cost externality and strate-

gic (stock) externality and show that the aquifer steady-state water stock level in an

open-loop equilibrium is higher than is obtained in the feedback equilibrium. Moreover,

the two equilibria are bounded from above and below by the controlled and the free (un-

controlled) solutions, respectively. Rubio and Casino (2003) also prove that a feedback

solution increases the overexploitation of the aquifer. Moreover, using nonlinear policy

functions, they also prove that, if the storage capacity is large, the difference between

the controlled and uncontrolled setup is negligible.

An instructive numerical exercise is developed in Madani and Dinar (2012). The

authors define six different management institutions for groundwater extraction and cal-

culate the differences in extraction and welfare. In the six models, three different (in

parameters) water users interact through the physical conditions of the aquifer, and only

cost externalities are considered. The six setups developed in the study are: ignorant

myopic management, smart myopic management with drawdown penalty, smart myopic

management with profit penalty, fixed ignorant non-myopic management, variable igno-

rant non-myopic management, and smart non-myopic management. They define agents

as “smart” when they notice that there are interactions with other users that affect their

profits; and “myopic” as agents that do not care about the use value of the resource. They

find that agents from a “smart non-myopic” management institution yield the highest

welfare and resource conservation, whereas the ignorant myopic management institution

(free or uncontrolled access) yields the worst. It is worth noting that, although they

do not consider feedback equilibrium, agents in the “smart non-myopic” management

institution recalculate their long-term plans extraction in each period based on the new

Page 26: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

14

information they obtained from the water table. This situation is not exactly a feedback

strategy, given that conjectures about the actions of others are not computed, but it

considers adaptation to additional information.

Finally, in a recent paper, Suter et al. (2012) develop a groundwater game in

a laboratory setup and explore the effects of different spatially explicit hydrological

groundwater models on pumping rates and the behavior of water users. They find that

spatially explicit models (in which the depth to water is specific to each location but at

the same time depends on the decisions of the other users) show a less myopic behavior

in comparison to the “bathtub” model.

Another strand of research has focused on the analysis of investment decisions in

common-pool resources. In the context of groundwater, investment in efficient irrigation

technology might alleviate the depletion of the resource. However, there is no agreement

among researchers on the impacts of irrigation efficiency on water saving (Peterson and

Ding 2005; Ward and Pulido-Velazquez 1989; Pfeiffer 2009), since individual gains in

efficiency might yield public benefits in this context. Moreover, increases in efficiency

often lower the cost of consumption, which might increase water consumption through

substitution or income effect (Pfeiffer 2009). This is known as the Jevon’s Paradox

or “rebound effect” (Jevons 1865). Now, farmers might have the incentive to shift

towards improving parcels that were not irrigated, or switch to more water-demanding

crops. Lichtemberg (1989) develops a model in which the introduction of land quality-

augmenting technologies generates a substitution between the acreage of good and bad

quality land. In an empirical test of the model, Lichtemberg uses the adoption of center

pivot technology as a land quality-augmenting technology, and finds that the capital cost

Page 27: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

15

of the technology significantly affects the cropping patterns. Ward and Pulido-Velazquez

(1989) develops a basin scale hydrologic model of the Upper Rio Grande Basin of North

America and show that subsidies directed to partially support the adoption of more

efficient irrigation technologies encourage a shift to more water-efficient technologies,

but do not reduce water depletion under any of the scenarios. They find that irrigated

acreage increases with the adoption of the new technology. Moreover, the study shows

that important return flows are lost because of the increase in efficiency. The same

results are found in Peterson and Ding (2005) and Scheierling et al. (2006). This research

suggests that irrigation efficiency improving investments give farmers incentives to shift

to more water-demanding crops.

When there is interaction between users (as there is in the aquifer problem),

strategic investment behavior might arise due to technological externalities (Gardner et

al. 1990). For instance, Agaarwal and Narayan (2004) develop a dynamic two-stage game

in which agents choose the level of initial investment for the capacity of a well and subse-

quent extraction. They show that agents strategically invest in excess capacity, leading

to further depletion of the aquifer. Barham et al. (1998) also analyze the relationship

between sunk costs and strategic behavior but in a context of a non-renewable resource.

The groundwater model studied in this dissertation considers the possibility of

ameliorating the depletion of the resource through adopting efficient irrigation tech-

nologies. However, due to strategic behavior, some groundwater users could have the

incentive to delay adoption, given that they are already being benefited by those farmers

who already adopted. The problem presented in this study is similar to the one presented

Page 28: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

16

in Moretto (2000) and Dosi and Moretto (2010) without considering uncertainty in re-

turns. Moretto (2000) develops a theory where irreversibility effects and war-of-attrition

effects are compounded in the decision of producers to adopt a new technology. They

find trigger values at which producers will switch from one technology to another.

Page 29: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

17

Chapter 3

The Groundwater Management Game

The theoretical model that serves as the basis for the framed experiments follows

the model proposed by Provencher and Burt (1993). In order to facilitate the exposition,

the model of groundwater management with no investment is presented first, with the

solutions for three different types of behavior: myopic, rational and fully coordinated.

Then, the individual model for both groundwater consumption and optimal investment

is presented, followed by the description of the parameters used and the experimental

design. It is important to mention that all the theoretical models were solved numerically,

and water use paths were calculated for each type of behavior. Therefore, it is possible

to compare the theoretical behaviors obtained from the model with the results obtained

in the experiments.

3.1 Theoretical Model

The model considered in the study is a simple dynamic groundwater model with

a finite number of users that interact through the aquifer. All users are assumed to have

dug the wells beforehand. Also, it is assumed that pumping costs change with the water

table. Nevertheless, open access is not considered, so it is possible to perform a game

theoretical perspective, since rents do not fully dissipate.

Page 30: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

18

The functional forms that I have chosen for the experimental design follow Wang

and Segarra (2011). These authors consider a profit function that is linear in respect to

the demand of water. They argue that crop water-related yields tend to increase linearly

with the amount of water applied until they reach a plateau, due to the natural capacity

of plants to absorb the water. Provencher and Burt (1993) is modified to consider depth-

to-water instead of the stock of water that remains in the aquifer. This is because depth

is more meaningful to user as it determines pumping costs directly. The model considered

by Provencher and Burt (1993) is presented, and then the modifications are shown for

clarification.

3.1.1 Basic Game (Baseline)

There are N groundwater users. They pump water from a “bathtub-type” aquifer

tapped through wells and they do not have access to surface water. In the model, I

assume wells have been dug, and costs are only those related to pumping from existing

wells. 1.

Farmer profits conditioned on water is are denoted by:

Bit = αwith− witβ

St− k = wit

(αh− β

St

)− k

Where wit is the amount of water pumped by farmer i in period t, α is the marginal

value of production of irrigated land, h, 0 < h ≤ 1 is the level of efficiency in the use of

water, βSt

is the marginal cost of water pumped from the well, St is the stock of water in

1Although I believe that these decisions and costs are extremely important, I tried to simplifythe model in order to make the application with farmers easier.

Page 31: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

19

the aquifer in period t, and k is the fixed cost of production. As commonly assumed in

the groundwater literature, only production costs related to groundwater are variable. It

is assumed that farmers already made all the decisions regarding the use of other inputs,

and their cost is already considered in the marginal cost β and the fixed cost k. Note

that the marginal cost of water pumped is inversely related to the stock of water in the

aquifer. If the stock of water increases, then the water table increases and farmers have

to pump water from a higher point in the well, which requires less electricity and reduces

pumping costs. Finally, there is no heterogeneity between farmers.

To specify the model in terms of water table instead of stock, I considered the

aquifer as a cylinder with a radius of R and height of D. Thus, the maximum capacity

of the aquifer is denoted by D × πR2. However, the stock of water in the aquifer will

change over time due to exploitation. Due to this, the stock can be observed using the

difference between the maximum depth of the aquifer, D, and the depth at which water

is pumped, dt. The stock of water at time t can be represented by: St =(D − dt

)×πR2.

With this identity, I can transform the current-period profit function to:

Bit = wit

(αh− β(

D − dt)× πR2

)− k

The equation of motion for the water table is given by:

dt+1 = dt +Wt

πR2 −f

πR2

Page 32: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

20

Where Wt is the demand of water of the N farms, Wt =∑N

i=1wit, and f is the natural

recharge of the aquifer in period t. Since no water beyond D can be pumped, total water

use is limited to:

dt +Wt

πR2 ≤ D

Farmers’ decisions depend on behavioral assumptions. Types considered here are

myopic, rational and fully cooperative.

Given the functional form of the model, a myopic water user will pump the max-

imum possible amount of water in every period, as long as profits are positive, since

myopic water users do not internalize the future consequences of their water consump-

tion today. Because there are no contemporaneous externalities from other water users

in the model, the demand of a myopic user is given by:

wm =

w if dt ≤ D − β

πR2αh

0 otherwise

Where w is the maximum amount of water that can be pumped.

Alternatively, a rational water user will consider the future value of water in their

decisions. Since the decisions of others will affect the stock of water in the future, rational

Page 33: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

21

users will also consider the choices of others. The rational water users problem is:

Vi1 = max{wit}Tt=1

[T∑t=1

δt−1Bit

]s.t

dt+1 = dt +1

πR2

N∑j=1

wjt + f

dt +

1

πR2

N∑j=1

wjt ≤ D

d1 given

Where δ is the discount rate. Assuming then that agent i can predict with certainty

other users’ choices, then user i maximizes the value function given the other users’ best

response2. Recalling the optimality principle, it is possible to write the Bellman equation

for agent i and period t as:

Vit (dt) = maxwit

[wit

(αh− β(

D − dt)× πR2

)− k + δVi,t+1 (dt+1)

]s.t.

dt+1 = dt +1

πR2 [wit + (N − 1)φ (dt) + f ]

dt +1

πR2 [wit + (N − 1)φ (dt)] ≤ D

d1 given

2Nevertheless, for empirical purposes, conjectures about other users’ decisions should varybetween individuals, depending on observable and unobservable variables.

Page 34: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

22

Where φ (d) is the best decision taken by the other N − 1 firms and, as before, h < 1.

The solution of the problem will depend on the strategies that agents choose. Open-

loop strategies depend on time and not on the current state (dt), whereas feedback or

close-loop strategies will depend on both time and current state. Negri (1989) and other

authors argue that open-loop strategies are not time consistent, since agents will correct

their paths over time. Provencher and Burt (1993) show that the solution of this problem

involves two different types of externalities: strategic and stock. Strategic externalities

arise when feedback strategies are followed. These externalities negatively affect the

efficient allocation of the resource and could encourage the depletion of the resource. It

is important to mention that the myopic behavior will be rational if full open access is

guaranteed or if the user cost is very small. In other words, the full/rational behavior

will tend to to myopic if N increases (Provencher and Burt 1993)

A socially optimal (dynamically efficient) solution of the problem differs from the

individual problem in that the collective profits are maximized. Assuming homogeneity

of agents, the symmetric total profit maximization problem is:

NVit (dt) = maxwit

N

[wit

(αh− β(

D − dt)× πR2

)− k + δVi,t+1 (dt+1)

]s.t.

dt+1 = dt +1

πR2

[wit + (N − 1) φ (dt) + f

]dt +

1

πR2

[wit + (N − 1) φ (dt)

]≤ D

d1 given

Page 35: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

23

Where ∼ denotes values at the social optimum. Given that V and V are concave,

Provencher and Burt (1993) show that the individual myopic or rational demands for

water are greater than the socially optimal demand, and lead to a lower steady-state

equilibrium stock of water. With a proper parametrization, it is possible to solve both

problems numerically.

3.1.2 Investment

Next I consider consider the problem when farmers can invest in new irrigation

technology. This new technology has an efficiency h of 1, meaning less water is required

to achieve the same profits exclusive of pumping costs, compared to the previous tech-

nology. This technology is adopted only once at a cost of I. The new technology also

requires maintenance cost of m every period3. The periodic benefits with the less ef-

ficient technology are denoted as B0 and are given by equation (3.1), whereas benefits

with the more efficient technology, without considering the investment cost, are given

by:

B1it

= wit

(α− β(

D − dt)× πR2

)− k −m

With the two technologies available, the farmer not only has to decide the op-

timal path of pumped water, but also the optimal moment to switch to more efficient

technology. These two situations can be combined as follows. In any period, say τ , the

3This additional cost can be interpreted as the cost incurred in hose and filter replacementfor drip irrigation technology.

Page 36: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

24

farmer will choose whether or not to invest in the technology, in order to maximize the

present value of the utility, Viτ , taking into account the equation of motion of the stock

of water in the aquifer and the boundaries of wit:

Viτ = max{V 0iτ, V 1

iτ− I}

s.t.

dτ+1 = dτ +1

πR2 [wiτ + (N − 1)φ (dτ ) + f ]

dτ +1

πR2 [wiτ + (N − 1)φ (dτ )] ≤ D

d1 given

Where

V 0iτ

= B0iτ

+ δVi,τ+1

and

V 1iτ

=

T∑t=τ

δt−1B1it

Again, φ (·) denotes the best strategy of the other users. The present value of the utility

of not investing in τ , V 0iτ

, considers the possibility of investing in the new technology in

the next period τ + 1, whereas investment in τ is considered irreversible.

This problem can be solved using the two-step method proposed by Agaarwal

and Narayan (2004). The first stage involves the investment decision whereas the second

Page 37: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

25

stage solves the optimal extraction path {wit}Tt=1. This problem is solved backwards,

starting from the second stage. It is possible to search for the optimal extraction path

conditional on the investment timing decision t, {wit(t)}Tt=1

, t = {1, 2, ..., T}. Each t will

yield a lifetime utility V(t)i1

. Now, the optimal investment time, t∗, can be represented

as:

t∗ = argmax{V(t)i1}, ∀t = {1, 2, . . . , T} (3.1)

Myopic agents do not care about the future, so they will not invest in the new

technology. Rational/strategic agents might be willing to invest in the new technology if

every user is willing to invest. However, there is the possibility of free-riding : Farmer i

will benefit from the water saving resulting from others’ investments. Therefore, farmer

i will have no incentive to invest if others do. If this is the case, then all the agents

will have the same strategy. Thus, Nash equilibrium will be the resulting equilibrium.

So two equilibria might arise, one in which everyone invests in the first period, and the

other, in which no one invests.

3.1.3 Parametrization, Myopic, Rational and Optimal Behavior

The total number of water users for each well considered in simulations is N = 4.

To facilitate the decision-making process of participants, I discretized the number of

hours of water that they can use. Thus, for this experiment, wit = {0, 1, . . . , 10}. The

irrigation technology efficiency is h = 0.5 for the less efficient technology, and h = 1

with the more efficient one. With this specification, 10 units of water will be required to

Page 38: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

26

irrigate all the land with the less efficient technology, while only 5 units is required with

the more efficient one.

The planning horizon of participants is 5 production years. After year 5, they

retire from farming and receive their payoff (Vi1). Thus, it is not important for them

if there is water left in the aquifer4. The initial depth of extraction is set to d0 = 170

yards.

All the parameters that were used in the experiments are presented in Table

3.1. Also, tables 3.2, 3.3 and 3.4 present the revenues with the less-efficient technology,

highly-efficient technology and the costs of extraction. Note that revenues with the

highly-efficient technology increase with the number of units used up to five units. After

five units, revenues do not change. This is because farmers posses a fixed agricultural

area and, with the parameters used, all the area that they hold is irrigated with 5 units

(see Section 3.2.1). Also, note that the cost changes with each level of the depth of

extraction as well as number of water units required.

4Although I believe that it is important to give a value to that water, it would required thevaluation of the ecological benefits of that stock, and this valuation goes beyond the scope of thispaper.

Page 39: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

27

Table 3.1.Parameters used in experiment

Parameter Value

α 10

β 390π

h 1 1

h 0 0.5

N 4

H 10

D 250

R 1

C 200

I 150

f 20π

m 9.460879

k 3

d 0 170

Page 40: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

28

Table 3.2.Revenues with less-efficient technology

Hours Revenue

0 -3

1 7

2 17

3 27

4 37

5 47

6 57

7 67

8 77

9 87

10 97

Revenues by hours of water pumped

with flood irrigation

Table A

Page 41: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

29

Table 3.3.Revenues with highly-efficient technology

Hours Revenue

0 -12

1 8

2 28

3 48

4 68

5 88

6 88

7 88

8 88

9 88

10 88

Table B

Revenues by hours of water pumped

with drip irrigation

Page 42: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

30

Tab

le3.4

.P

um

pin

gcost

s

12

34

56

78

910

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

38

39

40

41

42

43

44

45

46

47

48

49

50

Wa

ter

lev

el

24

92

48

24

72

46

24

52

44

24

32

42

24

12

40

23

92

38

23

72

36

23

52

34

23

32

32

23

12

30

22

92

28

22

72

26

22

52

24

22

32

22

22

12

20

21

92

18

21

72

16

21

52

14

21

32

12

21

12

10

20

92

08

20

72

06

20

52

04

20

32

02

20

12

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

390

195

130

98

78

65

56

49

43

39

35

33

30

28

26

24

23

22

21

20

19

18

17

16

16

15

14

14

13

13

13

12

12

11

11

11

11

10

10

10

10

99

99

88

88

8

780

390

260

195

156

130

111

98

87

78

71

65

60

56

52

49

46

43

41

39

37

35

34

33

31

30

29

28

27

26

25

24

24

23

22

22

21

21

20

20

19

19

18

18

17

17

17

16

16

16

1170

585

390

293

234

195

167

146

130

117

106

98

90

84

78

73

69

65

62

59

56

53

51

49

47

45

43

42

40

39

38

37

35

34

33

33

32

31

30

29

29

28

27

27

26

25

25

24

24

23

1560

780

520

390

312

260

223

195

173

156

142

130

120

111

104

98

92

87

82

78

74

71

68

65

62

60

58

56

54

52

50

49

47

46

45

43

42

41

40

39

38

37

36

35

35

34

33

33

32

31

1950

975

650

488

390

325

279

244

217

195

177

163

150

139

130

122

115

108

103

98

93

89

85

81

78

75

72

70

67

65

63

61

59

57

56

54

53

51

50

49

48

46

45

44

43

42

41

41

40

39

2340

1170

780

585

468

390

334

293

260

234

213

195

180

167

156

146

138

130

123

117

111

106

102

98

94

90

87

84

81

78

75

73

71

69

67

65

63

62

60

59

57

56

54

53

52

51

50

49

48

47

2730

1365

910

683

546

455

390

341

303

273

248

228

210

195

182

171

161

152

144

137

130

124

119

114

109

105

101

98

94

91

88

85

83

80

78

76

74

72

70

68

67

65

63

62

61

59

58

57

56

55

3120

1560

1040

780

624

520

446

390

347

312

284

260

240

223

208

195

184

173

164

156

149

142

136

130

125

120

116

111

108

104

101

98

95

92

89

87

84

82

80

78

76

74

73

71

69

68

66

65

64

62

3510

1755

1170

878

702

585

501

439

390

351

319

293

270

251

234

219

206

195

185

176

167

160

153

146

140

135

130

125

121

117

113

110

106

103

100

98

95

92

90

88

86

84

82

80

78

76

75

73

72

70

3900

1950

1300

975

780

650

557

488

433

390

355

325

300

279

260

244

229

217

205

195

186

177

170

163

156

150

144

139

134

130

126

122

118

115

111

108

105

103

100

98

95

93

91

89

87

85

83

81

80

78

51

52

53

54

55

56

57

58

59

60

61

62

63

64

65

66

67

68

69

70

71

72

73

74

75

76

77

78

79

80

81

82

83

84

85

86

87

88

89

90

91

92

93

94

95

96

97

98

99

100

Wa

ter

lev

el

19

91

98

19

71

96

19

51

94

19

31

92

19

11

90

18

91

88

18

71

86

18

51

84

18

31

82

18

11

80

17

91

78

17

71

76

17

51

74

17

31

72

17

11

70

16

91

68

16

71

66

16

51

64

16

31

62

16

11

60

15

91

58

15

71

56

15

51

54

15

31

52

15

11

50

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

88

77

77

77

77

66

66

66

66

66

55

55

55

55

55

55

55

55

44

44

44

44

44

44

44

15

15

15

14

14

14

14

13

13

13

13

13

12

12

12

12

12

11

11

11

11

11

11

11

10

10

10

10

10

10

10

10

99

99

99

99

98

88

88

88

88

23

23

22

22

21

21

21

20

20

20

19

19

19

18

18

18

17

17

17

17

16

16

16

16

16

15

15

15

15

15

14

14

14

14

14

14

13

13

13

13

13

13

13

12

12

12

12

12

12

12

31

30

29

29

28

28

27

27

26

26

26

25

25

24

24

24

23

23

23

22

22

22

21

21

21

21

20

20

20

20

19

19

19

19

18

18

18

18

18

17

17

17

17

17

16

16

16

16

16

16

38

38

37

36

35

35

34

34

33

33

32

31

31

30

30

30

29

29

28

28

27

27

27

26

26

26

25

25

25

24

24

24

23

23

23

23

22

22

22

22

21

21

21

21

21

20

20

20

20

20

46

45

44

43

43

42

41

40

40

39

38

38

37

37

36

35

35

34

34

33

33

33

32

32

31

31

30

30

30

29

29

29

28

28

28

27

27

27

26

26

26

25

25

25

25

24

24

24

24

23

54

53

52

51

50

49

48

47

46

46

45

44

43

43

42

41

41

40

40

39

38

38

37

37

36

36

35

35

35

34

34

33

33

33

32

32

31

31

31

30

30

30

29

29

29

28

28

28

28

27

61

60

59

58

57

56

55

54

53

52

51

50

50

49

48

47

47

46

45

45

44

43

43

42

42

41

41

40

39

39

39

38

38

37

37

36

36

35

35

35

34

34

34

33

33

33

32

32

32

31

69

68

66

65

64

63

62

61

59

59

58

57

56

55

54

53

52

52

51

50

49

49

48

47

47

46

46

45

44

44

43

43

42

42

41

41

40

40

39

39

39

38

38

37

37

37

36

36

35

35

76

75

74

72

71

70

68

67

66

65

64

63

62

61

60

59

58

57

57

56

55

54

53

53

52

51

51

50

49

49

48

48

47

46

46

45

45

44

44

43

43

42

42

41

41

41

40

40

39

39

101

102

103

104

105

106

107

108

109

110

111

112

113

114

115

116

117

118

119

120

121

122

123

124

125

126

127

128

129

130

131

132

133

134

135

136

137

138

139

140

141

142

143

144

145

146

147

148

149

150

Wa

ter

lev

el

14

91

48

14

71

46

14

51

44

14

31

42

14

11

40

13

91

38

13

71

36

13

51

34

13

31

32

13

11

30

12

91

28

12

71

26

12

51

24

12

31

22

12

11

20

11

91

18

11

71

16

11

51

14

11

31

12

11

11

10

10

91

08

10

71

06

10

51

04

10

31

02

10

11

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

00

44

44

44

44

44

43

33

33

33

33

33

33

33

33

33

33

33

33

33

33

33

33

33

33

33

88

88

77

77

77

77

77

77

77

77

66

66

66

66

66

66

66

66

66

66

65

55

55

55

55

12

11

11

11

11

11

11

11

11

11

11

10

10

10

10

10

10

10

10

10

10

10

10

99

99

99

99

99

99

99

88

88

88

88

88

88

8

15

15

15

15

15

15

15

14

14

14

14

14

14

14

14

13

13

13

13

13

13

13

13

13

12

12

12

12

12

12

12

12

12

12

12

11

11

11

11

11

11

11

11

11

11

11

11

11

10

10

19

19

19

19

19

18

18

18

18

18

18

17

17

17

17

17

17

17

16

16

16

16

16

16

16

15

15

15

15

15

15

15

15

15

14

14

14

14

14

14

14

14

14

14

13

13

13

13

13

13

23

23

23

23

22

22

22

22

21

21

21

21

21

21

20

20

20

20

20

20

19

19

19

19

19

19

18

18

18

18

18

18

18

17

17

17

17

17

17

17

17

16

16

16

16

16

16

16

16

16

27

27

27

26

26

26

26

25

25

25

25

24

24

24

24

24

23

23

23

23

23

22

22

22

22

22

21

21

21

21

21

21

21

20

20

20

20

20

20

20

19

19

19

19

19

19

19

18

18

18

31

31

30

30

30

29

29

29

29

28

28

28

28

27

27

27

27

26

26

26

26

26

25

25

25

25

25

24

24

24

24

24

23

23

23

23

23

23

22

22

22

22

22

22

22

21

21

21

21

21

35

34

34

34

33

33

33

33

32

32

32

31

31

31

31

30

30

30

29

29

29

29

29

28

28

28

28

27

27

27

27

27

26

26

26

26

26

25

25

25

25

25

25

24

24

24

24

24

24

23

39

38

38

38

37

37

36

36

36

35

35

35

35

34

34

34

33

33

33

33

32

32

32

31

31

31

31

30

30

30

30

30

29

29

29

29

28

28

28

28

28

27

27

27

27

27

27

26

26

26

Ta

ble

C

Pu

mp

ing

co

sts

by

wa

ter

lev

el

an

d p

um

pin

g h

ou

rs

9 10

Pu

mp

ing

ho

urs

Pu

mp

ing

ho

urs

Pu

mp

ing

ho

urs

4 5 6 7 810 0 1 2 35 6 7 8 90 1 2 3 40 1 2 3 4 105 6 7 8 9

Page 43: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

31

The parameters were set so that it is not profitable to use water if there is less

in the aquifer than necessary to supply the maximum amount that users can require.

Thus, beyond dt = 210, it is more profitable to use zero hours of water and just cover

the fixed cost of production. With these parameters, the myopic, rational/strategic and

fully coordinated paths for the basic game were calculated.

Given that the decision variables are discrete and the functional forms are linear,

linear integer programming can be used to solve the problems numerically. A Branch-

and-Bound algorithm was used to find the optimal solutions. For the case of the social

optimum, a restriction requiring the same use of water for every farmer is imposed. For

Nash equilibrium, the steps presented below are followed:

1. Set the decisions of agent i to a random discrete number between 0 and 10

2. Set the decisions of agents j 6= i to a random discrete number between 0 and 10.These 3 agents have the same values but they are different to agent i.

3. Solve the problem for agent i.

4. Take these values and assign them to the decision variables of the other threeagents j 6= i (symmetric game).

5. Evaluate whether this solution is feasible for period t (i.e. dt + 4w∗t≤ D).

(a) If feasible ⇒ Go to next step.

(b) If not feasible ⇒ Turn all values for that period to zero and go to next step.

6. Repeat steps 3-5 until convergence of the value function

Given that the time span is short and that the action set is discrete and relatively

short, it is not difficult to find a stable solution for the Nash equilibrium. The trajectories

of units of water pumped, costs and well depth for each type of equilibrium are presented

in figures 3.1, 3.2 and 3.3. The total benefits from possible “fixed arrangements” in the

Page 44: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

32

number of water units is calculated. This exercise could represent the situation in which

there is an agreement between users. The total benefits with arrangements from 0 to 10

hours of pumping were calculated. I considered zero hours of pumping whenever the cost

of extraction was too high and no production was more profitable. I consider 10 hours

for the last period, since the final well depth does not affect participant’s profits. The

results are presented in Figure 3.4. The equilibrium total benefits without including the

initial capital for myopic agents is 76.25; 108.75 for rational/strategic agents; and 165.82

for fully coordinated agents. Also, the value function is not linear for different fixed

arrangements. Moreover, with arrangements of 3, 4, 5, 6 and 7 units for each period, it

is possible to achieve net benefits higher than the Nash equilibrium. This indicates that

it is possible to be better-off than the Nash equilibrium if water users comply with those

fixed arrangements.

Fig. 3.1.Equilibrium Pumping hours for Myopic, Rational/strategic and Fully Coop-erative behaviors

Page 45: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

33

Fig. 3.2.Benefits of pumping for Myopic, Rational/strategic and Fully Cooperativebehaviors

Fig. 3.3.Well depth for Myopic, Rational/strategic and Fully Cooperative behaviors

Page 46: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

34

Fig. 3.4.Simulated benefits for Myopic, Rational/strategic and Fully Cooperative be-haviors, and fixed arrangements

In order to solve the model when the efficient technology is available, I first solved

the same problem in the case when only the highly-efficient technology is available. I

set the parameters in a way that it is possible to achieve the same social optimum with

the two technologies. This was done in order to avoid the introduction of a bias from

the parameters chosen. Then, I solved the second phase that involves the investment

decision, allowing the model to switch between the two technologies, given the efficient

path. As mentioned before, the model shows two equilibria: one in which all users adopt

the technology in the first period, and another in which none of the users adopt the

technology.

Page 47: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

35

3.2 Experimental Design

3.2.1 Recruitment of Participants and Framing

Penn State undergraduate students were recruited through the Smeal College of

Business Laboratory for Economics Management Auctions (LEMA) recruitment system

for the laboratory experiments. A total of 200 students were recruited. Subjects could

participate in the experiment only once. The sessions were conducted between June 25

and July 15, and August 30 and September 17 of 2013.

Farmers from the State of Aguascalientes, Mexico were recruited for the framed

field experiments. Results from sessions with 256 farmers were used in LACEEP 1

(2014). For more information about the recruitment of the field experiment, please see

LACEEP 1 (2014).

Both experiments were framed as follows. Each participant holds H acres of land

devoted to agriculture. This land could be irrigated or not. The amount of land is fixed

over time throughout the game and it is the same for all participants. Each participant

shares a well among N . Participants are homogenous (i.e. have identical choice sets

and payoff) but they do not know for sure what the other users do (see below for a

description of the session).

Each period will represent one year of production. Water users decide the daily

amount of water they will use for the entire production year. Decisions on the amount

of water are expressed as hours of water/day pumped from the well, wit. Each irrigated

acre requires 12h hours of water/day from the well. For instance, to produce on the H

acres requires wit = H2h hours of water per day. The total level of water use depends on

Page 48: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

36

the number of acres that are irrigated. Thus, the total irrigated land for the period will

depend on both the water requirements that they scheduled and the level of efficiency,

wit2h.

As mentioned above, changes in the depth of the water well will depend on the

amount of water that farmers use. For the sake of consistency all participants are told

that one hour of water pumped from the well will contribute to the depth of the well in

one yard . Finally, each farmer has an initial capital of C.

3.2.2 Treatments and Sessions

Treatments

The experimental design consists of two treatments:

Investment in technology (IN): In this treatment, participants will be allowed to invest

in irrigation technology. Participants are free to choose the time at which the investment

is done. The cost of investment is I.

Internal agreement (AG): Before participants make any decision regarding water con-

sumption, they agree on a fixed amount of water that they should pump in each period.

After participants make their agreements, they individually and anonymously decide

whether they comply with or deviate from the agreement.

The “Counterfactual” treatment (CF) refers to the sessions when the two treat-

ments are deactivated. Table 3.5 presents the distribution of sessions and participants

among the treatments for both field and laboratory experiments.

Page 49: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

37

Table 3.5.Summary of sessions: Laboratory and field experiments

a. Laboratory

Treatment Sessions Groups Participants Periods

Counterfactual 14 23 92 920

Investment 7 14 56 560

Agreement 6 13 52 520

Total 27 50 200 2,000

b. Field

Treatment Sessions Groups Participants Periods

Counterfactual 7 21 84 840

Investment 8 21 84 840

Agreement 10 22 88 880

Total 25 64 256 2,560

Source: LACEEP 1 (2014)

Sessions

Games were played by groups of 4 participants. Each group represented a water

well and the four members of the group had to pump water from the same well. I

named the wells with colors: yellow, blue, orange, green and red. The number of wells

that participated in the session varied depending upon the number of attendants. Each

participant received a card of the color of the group membership. The card was randomly

assigned, but I ensured other participants with the same color did not sit nearby each

other. Each well was played independently. Wells did not compete for water and there

Page 50: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

38

was no relationship between them. Finally, all experiments (both field and laboratory)

were paper-based.

All the sessions had at most 12 participants. Participants sat in individual cubicles

and members of the same well did not sit next to each other. Participants played for 15

periods. The first 5 periods (Round 1) corresponded to the baseline situation, in which

no investment nor agreement treatments were in place. After the first 5 periods, the

following 5 periods (Round 2) corresponded to a specific treatment. 3 types of sessions

in the experiments were applied. Participants were informed of the type of session

(CF,IN,AG) and corresponding instructions after the end of Round 1. After Round 2,

all the groups were randomly reshuffled and the same game as Round 2 was conducted.5.

Before participants made their pumping decisions for the first round, all group

members were informed of the initial well depth. This information was public throughout

the game and was the same for all the water wells.

Each participant had a revenue table for the case of the less-efficient technology

and, only for the case when the IN treatment is in place, participants had the revenues

of the two technologies, as shown in tables 3.2 and 3.3. They also had the cost table

presented in Table 3.4. Costs did not change between treatments. With this information,

participants could decide how many hours they pumped. They wrote down all their

decisions for the first period of Round 1 in their account sheet. Once all participants

made their decisions, calculated their income, cost and benefits for the first period (the

facilitator and assistants helped with these tasks), the facilitator proceeded to count how

5There is a variation in application of the CF session in Round 3, but that round will not beused in this study.

Page 51: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

39

many hours were used by each participant for each well. This exercise was done in a

way that the other members of the well did not notice who their group partners were.

Then, I calculated the change in depth of each well (considering recharge) and made this

information public to proceed with the second period. This exercise was repeated four

more periods with the corresponding update well depth.

For Round 2, I started again from the initial well depth (170 yards). If the type

of session was “IN”, participants were able to invest in a highly-efficient technology.

All participants were informed of this opportunity before the round starts. However,

there was a cost of I to acquire the new technology. This payment was made only once

during the game. Thus, every period, they had to decide whether or not to invest in the

technology and the amount of water to pump. If the technology is adopted, they would

only calculate their benefits using the table of the highly-efficient technology, shown in

Table 3.3.

If the type of session is “AG”, the members of each well met for 10 minutes

before the game starts and discussed a fixed amount of water that they would request

for each period throughout the game. This agreement was written in their accounting

sheet. After the meeting, the participants proceeded to play a game similar to the

baseline. Participants could anonymously decide whether to comply with the agreement

or deviate, and no penalties were applied.

3.2.3 Rewards

As mentioned above, each participant received an initial endowment C which was

their show up fee. The payoffs at the end of each round were composed by C plus the

Page 52: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

40

total net benefits from production that the participant earned through the five periods

minus I if the participant adopted the technology in the second (and third round for lab

experiments) of the “IN” sessions . At the end of the session, each participant drew a ball

from a bag with six balls, two of them marked with “1”, “2” or “3”,. The drawn number

indicated the round that was used to determine the reward. Participants received a

show up fee of $10 and the total earnings that participants received were between $10

and $226.

3.2.4 Survey

A short survey was conducted at the end of each session. The survey asked

about age, gender, major, minor, whether participants have taken courses in Economics,

Finances, Business Management, Psychology and Hydrology, and whether participants

had a farming background. The Cognitive Reflection Test (CRT) proposed by Frederick

(2005) was also included. This test is composed by three questions and it measures the

reflection capacity of agents, since the three questions could lead participants to give an

intuitive but wrong answer. Frederick (2005) mentions that the scores from the CRT is

correlated with patience, but also with less risk aversion.

Table 3.6 presents means and proportions of the variables gathered in the survey

for each treatment group, “Counterfactual”, “Investment” and “Agreement”. Most of

the students have a major related to “Business”, which includes Finance, Accounting,

Supply Chain Management and other similar majors, and also most of students have

6Some participants earned total benefits lower than the show up fee, however, the minimumreward was to $10.

Page 53: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

41

taken a course in Economics. Also, there are no major differences between the groups,

with the exception of the percentage of male participants, participants with farming

background and participants that have taken courses in finance. The CRT survey results

show that, although on average participants obtained a low score, the sample is evenly

distributed between the “reflective” and the correct answers in Questions I and III, and

favors the “reflective” answer in Question II7.

Descriptive statistics of a survey conducted in the field experiment are presented

in Table 3.7. The survey consisted of five sections and farmers were asked about land

tenure, major crops and livestock, farmers’ experience and demographics, water access

and use, and irrigation technology available on the field.

7The CRT assigns a score of 1 to a correct answer and zero to a wrong answer. Therefore,the CRT final score has values of zero, one, two and three.

Page 54: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

42

Table 3.6.Summary statistics of survey variables, laboratory experiment

Variable Counterfactual Investment Agreement Overall

Number of participants 92 56 52 200

Number of participants (%) 46.00 28.00 26.00 100.00

Mean of age (Years) 21.98 20.84 21.02 21.41

Male (%) 36.95 69.64 46.15 48.50

Farm background (%) 7.61 8.93 1.92 6.50

Courses (%):

Economics 85.87 89.29 84.62 86.50

Business Magement 48.91 69.64 61.54 58.00

Finance 53.26 75.00 59.62 61.00

Ag. Management 4.35 1.79 3.85 3.50

Environmental Economics 6.52 5.36 7.69 6.50

Psychology 42.39 46.43 51.92 46.00

Hydraulics 5.43 0.00 1.92 3.00

Major category (%):

Business 42.39 69.64 59.62 54.50

Education/Psychology 5.43 0.00 7.69 4.50

Energy/Environment 2.17 1.79 1.92 2.00

Engineering 5.43 3.57 1.92 4.00

Fundamental Sciences 7.61 3.57 7.69 6.50

Human development/medicine 3.26 1.79 7.69 4.00

Information Sciences 5.43 3.57 1.92 4.00

Liberal arts 7.61 5.36 9.62 7.50

Media 7.61 3.57 0.00 4.50

Natural Sciences 13.04 7.14 1.92 8.50

Cognitive Reflection Test (CRT)

Average Score (min:0, max:3) 1.59 1.04 0.96 1.27

Percentile 25 0 0 0 0

Median 2 0 0 1

Percentile 75 3 2 2 3

Answers Question I (%)

Answer = 5 57.61 33.93 34.62 45.00

Answer = 10 35.87 60.71 61.54 49.5

Question II (%)

Answer = 5 36.96 21.43 13.46 26.50

Answer = 100 45.65 58.93 71.15 56.00

Question III (%)

Answer = 47 43.48 35.71 26.92 37.00

Answer = 24 26.09 33.93 42.31 32.50

Page 55: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

43

Table 3.7.Summary statistics of survey variables, field experiment

Variable Baseline Investment Agreement Overall

Mean of irrigated land (has.) 7.11 4.23 3.68 4.99

Mean of rain-fed land (has.) 5.11 4.82 3.66 4.52

Farmers with land in Ejido (%) 70.24 79.76 94.25 81.57

Owner farmers (%) 86.90 76.19 88.51 83.92

Renter farmers (%) 21.43 25.00 17.24 21.18

Farmers whose major crop is: (%)

Corn for food 27.71 69.05 41.38 46.06

Corn for Grazing 19.28 20.24 19.54 19.69

Alfalfa 12.05 8.33 22.99 14.57

Beans 9.64 2.38 6.90 6.30

Grapes 9.64 0.00 8.05 5.91

Farmers whose major livestock is: (%)

Bovine for dairy 53.70 42.11 34.92 43.10

Bovine for beef 22.22 36.84 47.62 36.21

Bovine for double purpose 5.56 7.02 0.00 4.02

Goat 9.26 8.77 7.94 8.62

Swine 3.70 3.51 3.17 3.45

Mean of age (Years) 54.42 46.19 54.98 51.91

Male Farmers (%) 92.86 90.48 95.45 92.97

Marital status (%)

Single 11.9 26.19 5.75 14.51

Married 79.76 66.67 79.31 75.29

Cohabitant 2.38 3.57 5.75 3.92

Widow 5.95 3.57 9.2 6.27

Education level (%)

Preschool/none 2.38 1.19 11.49 5.1

Incomplete Primary 34.52 22.62 27.59 28.24

Complet Primary 16.67 21.43 16.09 18.04

Incomplete Secondary 4.76 5.95 8.05 6.27

Complete Secondary 20.24 19.05 17.24 18.82

Incomplete Technician 1.19 5.95 4.6 3.92

Complete Technician 8.33 5.95 2.3 5.49

Incomplete agricultural technician 0 1.19 0 0.39

Complete agricultural technician 2.38 1.19 1.15 1.57

Incomplete upper secondary 1.19 0 0 0.39

Complete upper secondary 0 2.38 6.9 3.14

Incomplete undergraduate 2.38 4.76 1.15 2.75

Complete undergraduate 3.57 7.14 2.3 4.31

Graduate School 2.38 1.19 1.15 1.57

Source: LACEEP 1 (2014)

Page 56: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

44

Table 3.7 (cont.)Summary statistics of survey variables, field experiment

Variable Baseline Investment Agreement Overall

Working regime (%)

Only on farm 41.67 38.10 54.02 44.71

Mostly on farm 20.24 20.24 6.90 15.69

Half-time on farm, half-time off-farm 26.19 30.95 31.03 29.41

Mostly off-farm 9.52 8.33 6.90 8.24

Only off-farm 0.00 1.19 0.00 0.39

Retired 2.38 1.19 1.15 1.57

Starting year (year) 1984 1992 1983 1986

Starting land (has.) 7.41 5.86 5.57 6.28

Major source of water (%)

Bordo 2.41 0.00 2.33 1.59

Water well 85.54 62.20 54.65 67.33

Dam 8.43 35.37 41.86 28.69

Other 3.62 2.43 1.16 2.39

Water well property

Private 3.57 0.00 0.00 1.18

Shared 82.14 58.33 55.81 65.35

Ejidal 4.76 10.71 1.16 5.51

No water well 9.52 30.95 43.02 27.95

Farmer is part of a water well association 90.79 81.03 95.92 89.07

Times of crop watering in first month 2.12 4.26 3.33 3.24

Times of crop watering in second month 2.12 3.77 3.32 3.07

Times of crop watering in third month 2.01 3.79 3.36 3.05

Hours for each watering 16.53 14.97 25.16 18.96

Farmers with irrigation technology (%)

Flood 69.88 74.70 65.12 69.84

Sprinkle 7.23 2.41 0.00 3.17

Drip 15.66 21.69 34.88 24.21

Micro Sprinkle 6.02 0.00 0.00 1.98

Other 1.20 1.20 0.00 0.79

Area with irrigation technology (has.)

Flood 3.99 2.64 2.35 2.99

Sprinkle 0.56 0.15 0.00 0.24

Drip 1.63 1.27 1.24 1.38

Micro Sprinkle 0.16 0.00 0.00 0.05

Other 0.04 0.04 0.00 0.02

Source: LACEEP 1 (2014)

Page 57: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

45

Chapter 4

Results and discussion

This section summarizes the results of the laboratory experiment and those pre-

sented in ?. The experimental outcomes (pumping decisions) are presented first. They

show significant differences, on average, between the ones observed on the field and lab-

oratory experiment. Then, average treatment effects of the treatments are estimated

for each type of experiment. Effects of the treatments also differ between the two pop-

ulations, suggesting different reactions to the treatment. Finally, a deeper analysis is

conducted for each treatment: technology adoption and agreement compliance.

4.1 Experiment Outcomes

Tables 4.1 and 4.2 show the average values in each treatment, round and period

for the number of hours, benefits and well depth for both field and lab experiments.

Figures 4.1, 4.2 and 4.3 present the results graphically One noteworthy feature of the

results is that, in Round 1, the paths from the field experiment are very different from

the laboratory experiment. The average path of pumping hours in the field experiment

decreases in a smoother way than in the lab experiments, and similar differences are

observed on the average benefits and well depth. Also, important differences can also be

observed between the three treated groups of the field experiments in Round 1 when the

treatment has not been applied yet, whereas there are no significant differences between

Page 58: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

46

the three treated groups of the lab experiments for the same round. These significant

differences are also shown in tables 4.1 and 4.2. Further, the average paths in Round 1

in both field and lab experiments greatly differ from the theoretical outcomes presented

in figures 3.1, 3.2 and 3.3.

Page 59: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

47

Table 4.1.Average outcomes by type of session, round and period: Laboratory experi-ments

Period Treatment Round

Round 1 8.27 39.29 170.00

Round 2 8.83 42.18 170.00

Round 1 7.96 37.80 170.00

Round 2 7.23 *† 49.88 *† 170.00

Round 1 7.75 36.60 170.00

Round 2 5.85 *† 27.23 *† 170.00

Round 1 8.47 32.00 183.09

Round 2 8.99 32.43 185.30

Round 1 8.32 33.07 181.86 †

Round 2 6.91 *† 48.30 *† 178.93 *†

Round 1 8.02 31.58 181.00 *†

Round 2 5.06 *† 21.31 *† 173.38 *†

Round 1 8.01 18.29 196.96

Round 2 7.32 12.50 201.26

Round 1 7.82 19.71 195.14 †

Round 2 6.45 † 42.84 *† 186.57 *†

Round 1 8.13 22.37 *† 193.08 *†

Round 2 5.87 *† 24.67 *† 173.62 *†

Round 1 4.49 3.23 209.00

Round 2 3.29 1.70 210.52

Round 1 5.30 4.82 206.43

Round 2 5.66 *† 37.64 *† 192.36 *†

Round 1 6.38 *† 6.42 *† 205.62 *†

Round 2 5.83 *† 23.56 *† 177.08 *†

Round 1 6.14 7.63 206.96

Round 2 7.35 13.08 203.70

Round 1 4.84 4.16 207.64

Round 2 5.98 *† 38.82 *† 195.00 *†

Round 1 4.54 † 2.31 *† 211.15 *†

Round 2 8.38 30.44 *† 180.38 *†

Calculation of differences at 5% of error between CF and treatment for: * means (T-test)

and, † distributions (Mann-Whitney test)

3

4

5

Agreement

Counterfactual

Investment

Agreement

Counterfactual

Investment

Agreement

Counterfactual

Investment

Agreement

Counterfactual

Investment

1

2

Agreement

Hours of

pumpingBenefit Well Depth

Counterfactual

Investment

Page 60: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

48

Table 4.2.Average outcomes by type of session, round and period: Field experiments

Period Treatment Round

Round 1 6.99 32.83 170.00

Round 2 7.24 34.26 170.00

Round 1 8.05 *† 38.19 *† 170.00

Round 2 5.61 *† 48.74 *† 170.00

Round 1 7.38 34.89 170.00

Round 2 7.15 34.09 170.00

Round 1 6.51 26.20 177.95

Round 2 6.95 28.52 178.95

Round 1 7.42 *† 28.25 182.19 *†

Round 2 5.46 *† 50.58 *† 172.43 *†

Round 1 7.26 *† 29.83 *† 179.50

Round 2 6.85 27.83 178.59

Round 1 6.58 22.88 184.00

Round 2 6.89 22.68 186.76

Round 1 7.18 18.40 *† 191.86 *†

Round 2 5.25 *† 51.54 *† 174.29 *†

Round 1 7.20 22.10 188.55 *†

Round 2 6.60 20.75 186.00

Round 1 6.60 16.92 190.33

Round 2 6.80 13.76 194.33

Round 1 6.90 7.75 *† 200.57 *†

Round 2 5.00 *† 50.12 *† 175.29 *†

Round 1 6.89 10.92 *† 197.36 *†

Round 2 6.40 10.88 192.41

Round 1 5.96 10.94 196.71

Round 2 6.46 5.40 201.52

Round 1 6.02 -5.00 *† 208.19 *†

Round 2 5.04 *† 51.96 *† 175.29 *†

Round 1 6.16 -2.08 *† 204.91 *†

Round 2 6.18 -19.08 198.00

Calculation of differences at 5% of error between CF and treatment for: * means (T-test)

and, † distributions (Mann-Whitney test)

Agreement

Counterfactual

Investment

Agreement

Agreement

Counterfactual

Investment

Agreement

Counterfactual

Investment

1

2

3

4

5

Counterfactual

Investment

Agreement

Counterfactual

Investment

Hours of

pumpingBenefit Well Depth

Source: LACEEP 1 (2014)

Page 61: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

49

The results show important differences between the three groups in Round 2,

where the treatments are applied. For the field experiments, only the “IN” treatment

shows an improvement in the three variables analyzed with respect the counterfactual;

whereas for the lab experiments, both the “IN” and the “AG” treatments show significant

improvements with respect to the counterfactual in most of the periods. Finally, it is

worth noting that the average path of hours of pumping of the counterfactual group

of the lab experiments is very similar to the myopic theoretical behavior presented in

figure 3.1. This finding suggests that, after the first round, most participants notice the

behavior of their partners and decided to show this kind of behavior in opposition to a

more cooperative one. This result is not observed in the field experiments.

Page 62: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

50

Fig. 4.1.Average individual hours pumped by period and treatment

a. Laboratory

b. Field

Source: LACEEP 1 (2014)

Page 63: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

51

Fig. 4.2.Average individual benefits by period and treatment

a. Laboratory

b. Field

Source: LACEEP 1 (2014)

Page 64: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

52

Fig. 4.3.Average well depth by period and treatment

a. Laboratory

b. Field

Source: LACEEP 1 (2014)

Page 65: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

53

Turning to individual decisions, Figures 4.4 and 4.5 show histograms of pumping

hours for each period, round and treatment from the field and laboratory experiments,

respectively. The dispersion of choices from the field experiment is much higher than

from the laboratory experiment, which might reflect different degrees of game compre-

hension. However, when looking at the decisions of the field experiment in the first

period of the first round, that decisions are concentrated around 5 and 10 units, whereas

in the laboratory experiment they are mostly concentrated around 10 units, and a lesser

proportion is around 5 units. After period one, the distribution becomes flatter in the

field experiment, suggesting that farmers consider the potential cost of using high vol-

umes of the resource in early periods. In contrast, in the laboratory experiment, the

concentration around 10 units increases until period 4, where the distribution splits in

two parts, one around zero units and another around ten units. This is because it is not

profitable for some groups to use more water, given the depth of the well.

The differences between the decisions of students and farmers might reflect struc-

tural differences in behavior. Farmers appear to be more pro-social than students, or at

least they consider a higher use value of the resource. Students tend to be more myopic

and selfish, and this behavior becomes stronger on the following rounds. These results

are consistent with other findings in the Behavioral Economics literature. For instance,

Belot et al. (2010) find that students are more likely to behave in accordance to the eco-

nomic theory than non-students in games that engage other-regarding preferences (such

as Trust Game or Public Good Game). Carpenter et al. (2005), Carpenter et al. (2008)

and Anderson et al. (2013) find similar results.

Page 66: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

54

To fully account for the effects of each treatment on the use of the resource,

it is necessary to develop a deeper analysis of the different behaviors within the two

populations. This issue will be partially addressed in Part II of the dissertation.

Page 67: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

55

Fig. 4.4.Histogram of pumping hours from laboratory experiment, by Round, Periodand Treatment

a. Counterfactual

b. Investment

c. Agreement

Page 68: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

56

Fig. 4.5.Histogram of pumping hours from field experiment, by Round, Period andTreatment

a. Counterfactual

b. Investment

c. Agreement

Source: LACEEP 1 (2014)

Page 69: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

57

For the case when the investment treatment is imposed (Round 2), the distribution

collapses to 5 units in the field experiment, meaning that most farmers chose to adopt

the technology, whereas in the laboratory experiment, the distribution splits into two

parts, one concentrated in 5 units and another in 10. This evidence confirms the results

from the theory, where two attraction points were found. Finally, when the agreement

treatment is imposed, the distribution of the field experiment is initially concentrated

around 5 and 7 units, and then becomes more dispersed and the concentration around

8 and 9 units increases. In the laboratory experiments, there are not any significant

changes in the distribution until the last period, when most of the participants switched

to 10 units.

The end-of-round outcomes of the experiment are also of interest. Tables 4.3 and

4.4 show the mean values for the three end-of-round variables for both field and labora-

tory experiments. Difference tests in both mean and distribution are performed between

rounds for each treatment (CF, IN and AG) in part a. of each table. The results show

that there are significant differences for the two treatments in the laboratory experi-

ments, whereas significant differences for the IN treatment are found only in the field

experiment. Difference tests between the counterfactual and each of the two treatments

within each round (Part b.) were also conducted. The results of the field experiment

show that there are some differences between the counterfactual and the IN treatment

in the total benefits in Round 1, and in the total hours pumped and final well depth

in Round 2. The differences in Round 1 suggest that, even though the treatments were

assigned randomly, there are some unobservable differences between the participants of

the sessions that has to be considered in the analysis.

Page 70: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

58

Table 4.3.Average end-of-round outcomes by type treatment and round: Laboratoryexperiments

Total Benefits 299.27 323.73 *† 327.21 *†

Total hours

pumped34.83 32.23 *† 30.98 *†

Final Well Depth 209.31 198.93 *† 193.92 *†

Calculation of differences at 1% of error between Rounds for: * means (T-test) and, † distributions (Mann-Whitney test)

Total Benefits 299.27 323.73 *† 327.21 *†

Total hours

pumped34.83 32.23 † 30.98 *†

Final Well Depth 209.31 198.93 193.92 *†

Calculation of differences at 1% of error between CF and treatment for: * means (T-test) and, † distributions (Mann-Whitney test)

211.52

300.45

35.38

211.52

300.45

35.38

Investment Agreement Investment AgreementVariable

Round 2Round 1

Counterfactual Counterfactual

Round 2

Counterfactual Investment Agreement Counterfactual Investment Agreement

a. Differences for each treatment between Round 1 and Round 2

b. Differences between CF and other treatments within Round 1 or Round 2

207

299.57

34.25

207

299.57

34.25

213.09

301.89

35.77

213.09

301.89

35.77

VariableRound 1

Table 4.4.Average end-of-round outcomes by type treatment and round: Field experi-ments

Total Benefits 287.6 295.66 315.44 *† 274.47

Total hours

pumped35.57 34.88 26.36 *† 33.18

Final Well Depth 212.29 209.55 175.43 *† 202.73

Calculation of differences at 1% of error between Rounds for: * means (T-test) and, † distributions (Mann-Whitney test)

Total Benefits 287.6 *† 295.66 315.44 274.47

Total hours

pumped35.57 34.88 26.36 *† 33.18

Final Well Depth 212.29 209.55 175.43 *† 202.73

Calculation of differences at 1% of error between CF and treatment for: * means (T-test) and, † distributions (Mann-Whitney test)

207.38

304.63

34.35

207.38

304.63

34.35

Counterfactual

VariableRound 1 Round 2

Counterfactual Investment Agreement Counterfactual Investment Agreement

a. Differences for each treatment between Round 1 and Round 2

b. Differences between CF and other treatments within Round 1 or Round 2

200.57

309.77

32.64

200.57

309.77

32.64

Investment Agreement Investment AgreementVariable

Round 2Round 1

Counterfactual

Source: LACEEP 1 (2014)

Page 71: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

59

For the lab experiments, there are no differences in the outcomes in Round 1,

between the CF and the treatments, which suggests that the samples are similar a priori.

On the other hand, in Round 2, all outcomes from the AG sessions are significantly

different from the CF treatment in both mean and median, whereas only the Total

Benefits from the IN sessions are significantly different from the CF sessions in both

mean and median. There are no significant differences in median for the Total Hours

pumped between IN and CF, whereas for the Final Well Depth, there are no significant

difference between IN and CF in either mean or median.

Figure 4.6 shows the average total net benefit and the theoretical equilibrium

values for each treatment and round. In both the field and laboratory experiments, the

three groups overcome, on average, the value of the Myopic theoretical result (275.25),

but do not reach the Nash solution in Round 1 (308.75). In Round 2, the total net

benefits of the CF group do not reach the Nash equilibrium in the field experiments,

whereas the benefits for the IN group barely reaches the Nash level, and the average net

benefits of the AG groups do not even reach those of the myopic benchmark. In the case

of the laboratory experiments, again the CF group does not reach the Nash. However, in

this case, both the IN and AG are significantly higher than the Nash. A similar pattern

is observed with total hours of pumping and final well depth (Figures 4.7 and 4.8).

Page 72: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

60

Fig. 4.6.Average individual total benefits by round and treatment

a. Laboratory

b. Field

Source: LACEEP 1 (2014)

Page 73: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

61

Fig. 4.7.Average individual total hours pumped by round and treatment

a. Laboratory

b. Field

Source: LACEEP 1 (2014)

Page 74: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

62

Fig. 4.8.Average final well depth by round and treatment

a. Laboratory

b. Field

Source: LACEEP 1 (2014)

Page 75: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

63

4.2 Treatment Effects

A key interest of this study is the analysis of how access to efficient technology

and communication affect water consumption and welfare. Three outcomes of interest

are the three end-of-stage outcomes: individual total hours pumped, individual total

benefits and final well depth.

As mentioned before, there are some intrinsic differences between the treated

groups. If these differences are not taken into account, any estimation of the impact

of the treatment will be biased. Moreover, participants learn about the game between

rounds. Then, it is also necessary to control for learning effects. In order to obtain reliable

estimates of the impact of the treatment, I will estimate a difference-in-difference model.

In this case, I consider a reduced form approach to analyze the treatment effects.

Consider the outcome ygis

observed at the end of the round without treatment (s = 1) and

the round after treatment (s = 2) for three different treated groups: “Counterfactual”

(g = 0), “Investment” (g = 1) and “Agreement” (g = 2). Recall that for the case of

g = 0, there is no treatment imposed in both s = 1 and s = 2. The following linear

equation can be estimated:

ygis

= α+ δ + θ1D1 + θ2D2 + φ1D1 × δ + φ2D2 × δ + βXi + υi + εis

Where α is a constant, δ is a binary variable that takes the value of one if s = 1 and zero

otherwise, D1 is a binary variable that takes the value of one if g = 1 and zero otherwise,

D2 is a binary variable that takes the value of one if g = 2 and zero otherwise, Xi are

Page 76: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

64

fixed individual (group) level observable variables, υi is a specific individual (group) level

random term and εis is a individual (group) and time level error term.

The average treatment effect is denoted by φg for each treatment, g = 1, 2, given

that:

φ1 = E[y1i2− y1

i1

]− E

[y0i2− y0

i1

]and

φ2 = E[y2i2− y2

i1

]− E

[y0i2− y0

i1

]

Estimation of this model with OLS would yield inconsistent estimates because

of the presence of unobserved heterogeneity represented by υi. A random-effects model

to estimate the equation presented above in order to account for specific individual and

treated-group effects. The full model which includes some of the variables gathered

through the survey was also considered to control for observable characteristics of par-

ticipants. Results are presented in Table 4.6 for the field experiments and Table 4.5 for

the laboratory experiments.

Page 77: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

65

Table 4.5.Difference-in-Difference estimation of final outcomes in rounds 1 and 2: Labexperiments

(1) (2) (3) (4) (5)

Total pumping

hours

Total pumping

hours - FullTotal Benefits

Total Benefits -

FullFinal Well Depth

Treatment (Base = CF)

IN -1.130 -1.130 -0.874 -2.971 -4.522

AG -0.554 -0.554 -1.176 -2.682 -2.214

Round 2 0.391 0.391 1.446 1.446 1.565

IN x Round 2 -2.409 -2.409 22.72*** 22.72*** -9.637

AG x Round 2 -4.237*** -4.237*** 26.50*** 26.50*** -16.95***

Age 0.142 -0.0317

Female 1.105 1.147

Major group (Base = Business )

Education/Psychology -1.041 -6.991

Energy/Envionment 3.372 14.00

Engineering -0.761 -11.35

"Hard" Sciences -1.451 -2.517

Human Devlopment/Medicine -0.397 -7.784

Information Sciences -1.952 -7.364

Liberal Arts -0.660 -8.363

Media -3.186* -13.77*

Natural Sciecnes 0.797 -7.820

CRT answer correct

Question 1 0.823 2.092

Question 2 -1.078 -2.030

Question 3 0.257 2.974

Constant 35.38*** 32.10*** 300.4*** 303.2*** 211.5***

σu 3.187*** 2.924*** 10.23*** 8.604*** 0

σe 6.135*** 6.135*** 24.73*** 24.73*** 13.01***

Observations 400 400 400 100 100

Number of caseid 200 200 200 50 50

Equation (3) is estimated at the well (group) level

*** p<0.01, ** p<0.05, * p<0.1

Variables

F test performed for significancy of σu

Page 78: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

66

Table 4.6.Difference-in-Difference estimation of final outcomes in rounds 1 and 2: Fieldexperiments

(1) (2) (3) (4) (5)

Total pumping

hours

Total pumping

hours - FullTotal Benefits

Total Benefits -

FullFinal Well Depth

Treatment (Base = CF)

IN = 1 2.929** 3.292** -22.18* -21.19** 11.71**

AG = 1 2.244* 3.204** -14.11 -21.91** 8.974*

Round 2 1.702* 1.821* -5.143 -6.154 6.810

IN X Round 2 -10.92*** -11.36*** 32.99** 35.44*** -43.67***

AG X Round 2 -3.407** -3.858*** -16.05 6.413 -13.63**

Age -0.434** -0.455

Age2 0.00330* 0.00237

Female 0.829 2.859

Starting Year 3.748 47.02

Starting Year2 -0.000946 -0.0119

Head of household -2.644 2.92

Marital Status (Base = Single)

Married 0.262 -13

Cohabitant 1.462 24.68

Widow 1.483 3.803

Working time (Base = Only on farm)

Mostly on farm -0.554 4.729

Half on farm, half off-farm -2.741** -5.044

Mostly off-farm -4.827*** -15.14

Only off-farm -19.36*** -106.0***

Retired -1.462 -5.008

Education level (Base = Preeschool/None)

Primary school -1.745 -5.281

Secondary school -2.078 1.196

Technical school/Preparatory -1.659 8.714

Agricultural Technical school 1.692 21.13

University and Graduate school -3.586 5.804

Irrigated hectares 0.762*** 3.633***

Irrigated hecatares2 -0.00998*** -0.0416***

Land is ejido 0.761 17.62*

Source of water (Base = Dam)

Only water from well 1.246 7.183

Both water from well and dam 1.046 3.191

Primary crop alfalfa -0.0794 -4.915

Primary crop vine 3.520* 9.321

Area with drip irrigation -0.410*** -2.031***

Municipality (Base = Pabellón de Arteaga)

Aguascalientes -1.092 24.5

Asientos -0.678 -42.51***

Calvillo -4.246 10.3

Cosio -1.214 10.55

El Llano 0.383 -7.745

Rincón de Romos -3.922** -2.355

San Francisco 0.349 -45.68

San José de Gracia -4.809** 30.34***

Tepezalá -0.932 -8.894

Constant 32.64*** -3689 309.8*** -46349 200.6***

σu 5.882*** 4.373*** 25.95*** 9.117**

σe 6.411*** 6.380*** 74.32*** 48.67*** 14.93***

Observations 512 474 512 474 128

Number of caseid 256 237 256 237 64

Equation (5) estimated at the well (group) level

*** p<0.01, ** p<0.05, * p<0.1

Variables

Likelihood-ratio test performed for significancy of σu

Source: LACEEP 1 (2014)

Page 79: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

67

The coefficients of interest are “IN x Round2” for the effect of the “IN” treatment,

and “AG x Round2” for the effect of the “AG” treatment. For the field experiments, the

two treatments have a significant negative effect on the total pumping hours and final well

depth, whereas only the IN treatment has a positive effect on final total benefits. As the

authors of LACEEP1 mention, this could be explained by the fact that some participants

deviated from the agreement. Thus, even though the amount of water was reduced, it

did not increase benefits. Also, there are intrinsic differences in the total water used in

the groups, since the coefficients of “IN” and “AG” are significantly different from zero

in equation (1) and (2). This result shows the value of the DID approach.

The authors of LACEEP1 also present the full model including contextual and

individual variables. It seems that older people, people that participate more in non-

agricultural activities, and people with a larger area with drip irrigation installed tend

to use less hours of pumping within the experiment, whereas people with total larger

areas tend to use more. At the same time, people that work only off-farm and people

with drip irrigation tend to earn less benefits in the game, and people with more area

tend to earn more. These results suggest that some individual characteristics might be

related to the types of behavior that the participant show during the game.

With respect to the laboratory experiments, the two treatments only show an

effect on the total benefits earned by participants, and only the “AG” treatment shows

effects on total hours pumped and final well depth. This is because, as I will show

in the next section, several participants did not adopt the technology and “free-ride”,

Page 80: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

68

taking advantage of the benefits generated by the adoption of technology by other par-

ticipants. I also included individual characteristics in the regressions, but none of them

are significant.

4.3 Technology Adoption

The results presented in the previous section suggested that participants of the

field experiment show higher rates of adoption than the laboratory experiment, since

the “IN” treatment did not have any effect on the latter, but an important effect on the

former. As mentioned before, this might suggest the presence of strategic behavior in

adoption for some participants of the laboratory experiments.

Figure 4.9 shows the cumulative technology adoption percentages in each period

of Round 2 for both field and laboratory experiments. 81 percent of participants in the

field experiments adopted the technology in the first period of Round 2, and adoption

increased up to 90 percent in the fifth period. In contrast, roughly 40 percent of partici-

pants of the laboratory experiments adopted the new technology in the first period, and

up to 60 percent in the fifth period.

Recall that the theoretical model shows that there are two equilibria for the

investment problem, one in which all adopted the technology in the first period, and

one in which no one adopted the technology. The empirical findings suggest that the

theoretical predictions hold in the laboratory, since the sample is divided into people that

adopted immediately and people that never adopted. This is also observed on the third

round of the laboratory experiment. Although in Round 3, most participants adopted

Page 81: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

69

the new technology on the first period, the adoption rate slightly increased from 62 to

66 percent, as shown in Figure 4.10.

Fig. 4.9.Percentage of technology adoption by period and experiment

Source: LACEEP 1 (2014)

Fig. 4.10.Percentage of technology adoption by period and round: Laboratory exper-iment

Page 82: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

70

It is possible to analyze the factors that affect adoption rates, such as the current

well depth or personal characteristics of participants, using a proportional hazard model.

Define the hazard function as the instantaneous probability of leaving a state conditional

on survival on time t (Cameron and Trivedi 2005):

λ(t) = lim∆t→0

Pr [t ≤ T < t+ ∆T |T ≥ t]∆t

In this case, λ(t) represents the instantaneous probability of adopting the technology,

given that it has not been adopted by period t. The proportional hazard rate model

considers the hazard rate as a product of an intrinsic hazard rate that only depends on

time, and a factor that depends on the covariates x:

λ(t|x, β) = λ0(t)φ(x, β)

Where the most common functional form of φ(x, β) is exp(x′β).

Results from the Cox proportional hazard model of adoption for Round 2 of the

laboratory experiment are presented in Table 4.7. Well depth increases the hazard rate

by 275 percent whereas the squared of the well depth decreases the hazard ratio by

0.4 percent. This last result, although significant, is very weak. Therefore, well depth

has almost a linear effect on the hazard rate of adoption. With respect to the individual

characteristics, to be enrolled in the Information Sciences major increases the hazard rate

by 248 percent with respect to the Business major. Finally, differences in the results of

the CRT test also affects the hazard rate of adopting. The hazard rate of participants

Page 83: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

71

that obtained zero, one and two in the CRT test is lower in 65.6, 58.4 and 48 percent with

respect to the effect of having full score (three), respectively. This result is interesting,

since the CRT is correlated with both patient and less risk averse people (Frederick

2005). Thus, people with higher CRT tend to be more thoughtful about the decision of

investing, or are less risk averse and are willing to bear the risk of investing, even if they

know that other participants might free-ride.

Table 4.7.Cox proportional hazard model of investment: Laboratory experiments

Variables Hazard ratio

Well depth 3.755**

Well depth sq. 0.996**

Age 6.540

Age sq. 0.957

Female 0.855

Major group (Base = Business )

Energy/Envionment 0.000

Engineering 0.559

"Hard" Sciences 1.323

Human Devlopment/Medicine 0.792

Information Sciences 3.483***

Liberal Arts 0.350

Media 1.598

Natural Sciecnes 0.000

CRT score (Base: Score = 3)

Score = 0 0.3440***

Score = 1 0.416**

Score = 2 0.520*

Observations 56

Robust standard errors in parentheses

*** p<0.01, ** p<0.05, * p<0.1

Page 84: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

72

Turning to the field experiment, a possible explanation of the high adoption rate is

the background information that farmers hold about efficient irrigation technologies. Ac-

cording to LACEEP 1 (2014), there are several institutions in the area that have exten-

sion programs and demonstration plots with which information about efficient irrigation

technologies is given to farmers. Also, the Government of the State of Aguascalientes is

conducting several programs that aim at the improvement of water use efficiency through

the adoption of better irrigation technologies. A major constraint for adoption is the

lack of funding and budget constraints that farmers face. Only richer farmers can adopt

the technology, but most farmers are willing to adopt. In contrast, college students do

not posses any prior information about irrigation technologies, and they will only use

the information in the game to make their decisions. They will not adopt if they believe

that it is not worth to adopt the technology and, moreover, if their behavior is such

that they believe that they can benefit from not adopting. This difference between field

and laboratory experiments is important. If the experiment is meant to test behavior in

response to policy, then the test should be done with populations representative of the

affected population, since people with different backgrounds will see the experiment in

different ways.

4.4 Agreement

As mentioned above, participants from the AG sessions were asked to meet for

ten minutes before Round 2 starts in order to agree on a fixed individual level of pump-

ing throughout the five periods. However, no penalties were applied if the participant

deviated from the agreement.

Page 85: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

73

Figure 4.11 shows histograms with the agreed values of pumping hours. Panel a.

shows the agreements in the laboratory experiment in Round 2. More than 60 percent

of the groups agreed on a value of 5 hours, and 22 percent agreed on 7 hours in Round

2. Two groups are categorized as “zero” that actually agreed on a variable number of

hours that involved combinations 0 and 10 hours, or combinations of 3 and 10 hours.

These groups earned the highest profit of the round. Variable agreements can also be

observed in Round 3 of the laboratory experiments (Panel b.). In this round, after the

groups were reshuffled, three groups chose combinations of 0 and 10 hours, and 0 and 3

hours. Again, these three groups earned the highest total benefits, and the members of

one group earned $364, which is almost the value that participants would have earned

at the social optimum ($365).

Page 86: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

74

Fig. 4.11.Histogram of agreed values of pumping hours in “AG” sessions

a. Laboratory - Round 2

b. Laboratory - Round 3

c. Field

Source: LACEEP 1 (2014)

Page 87: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

75

In the field experiments (Panel c.), most of the agreements are of 5 and 7 units,

and no group chose less than 5 hours.

Figure 4.12 shows cumulative percentages of deviation from the agreement by

type of agreement. Groups with agreements of “zero” were excluded, since the agrement

involved variable values. It is clear that participants of the field experiment show higher

rates of deviation, in general, than participants of the laboratory experiment. This

explains in part why the treatment effect of the AG treatment was significantly higher

in the laboratory experiments than in the field ones. Moreover, participants of the field

experiments whose group chose values of six pumping hours deviate earlier than other

participants.

It might be possible that the well depth in each period, along with personal

characteristics might affect deviation rates. The authors of LACEEP1 test the influence

of other factors on deviation rates with a Cox proportional hazard model. Results from

the Cox proportional hazard model for the field experiment are taken from LACEEP1

and presented in Table 4.8. A Cox model was also ran with the data from the laboratory

experiment. However, the effects of all covariates was significantly different from zero

and deviations were mostly explained by the baseline hazard rate which depends only

on time.

Page 88: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

76

Fig. 4.12.Agreement deviation percentages in “AG” sessions

a. Laboratory - Round 2

b. Laboratory - Round 3

c. Field

Source: LACEEP 1 (2014)

Page 89: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

77

Table 4.8. Cox proportional hazard model of deviation from agreement:Field experiments

Variable Hazard rate

Well Depth 0.920*

Well Depth sq. 1.0002

Agrement (Base = 5)

Agreement = 6 2.953***

Agreement = 7 2.104

Agreement = 8 5.789***

Agreement = 9 1.542

Agreement = 10 1.365

Age 0.988

Education level (Base = Pre-school/None)

Primary school 0.372***

Secondary school 0.158***

Technical school/Preparatory 0.345**

Agricultural Technical school 1.401

University and Graduate school 0.207**

Irrigated hectares 0.899

Working time (Base = Only on farm)

Mostly on farm 2.915

Half on farm, half off-farm 1.234

Mostly off-farm 1.679

Retired 3.336*

Source of water (Base = Well)

Only water from dam 0.357*

Both water from well and dam 0.444

Land is ejido 0.224***

Primary crop alfalfa 0.532***

Municipality (Base = Pabellón de Arteaga)

Aguascalientes 1.005

Cosio 0.087***

El Llano 3.676*

San Francisco 3.142*

Tepezalá 1.183

Observations 83

Robust standard errors in parentheses

*** p<0.01, ** p<0.05, * p<0.1

For the field experiment, LACEEP 1 (2014) find that well depth and its square do

not influence the hazard rate of defection. This result is consistent with the results from

non-cooperative game theory, where agents will deviate from the agreement, regardless

the state of the resource. Turning to the constant variables, having an agreement of

Page 90: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

78

“6” and “8” significantly increases the probability of deviation in 195 and 579 percent,

respectively. However, these values are very high and should be taken with caution,

since the number of groups that chose those values is small. Furthermore, some individ-

ual characteristics also affect the hazard rate of defection. For instance, education has

significant effects in reducing the hazard of defection, although not in a monotonic way.

Also, being part of an ejido and having alfalfa as major crop have very strong effects on

reducing the probability of defection, reducing the hazard rate in 77.6 and 46.8 percent,

respectively. An ejido is a legally recognized agricultural community in Mexico originally

created manage agricultural land as a community. However, farmers that are part of the

ejido currently possess individual property rights over their land. Nevertheless, in many

ejidos land and water use decisions are still discussed with all their members. According

to the authors, the fact that farmers that are part of an ejido tend to deviate less from

the agreement gives some insights about the importance of reputation and social values

when face-to-face communication is part of the negotiation.

4.5 Discussion

The results of the experiments presented in this study show that there are impor-

tant differences between the behavior of the participants from the field and laboratory

experiments. Some researchers may argue that it is a matter of time and as more rounds

are played, actions will converge to the equilibrium and participants field and laboratory

experiments will behave similarly. However, several authors have found that there are

structural differences between the behavior of pools of students and non-students in labo-

ratory experiments (Anderson et al. 2013; Belot et al. 2010; Carpenter et al. 2005, 2008).

Page 91: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

79

The results obtained in the present study align with the results from those studies in

the sense that students tend to behave in accordance to economic theory and seem to be

more selfish in a game with other-regarding preferences than non-students. In addition,

the experiments have given some insights about the importance of background informa-

tion when making decisions in experiments. The difference in the results obtained in the

IN treatment are very clear. The level adoption of new technology by farmers is very

high, even though there are incentives for not adopting. Moreover, the failure of the

efficient technology to improve the benefits in the laboratory experiment suggests the

presence of strategic behavior of participants, who do not have background information

about the technology.

Background information is extremely important when analyzing the behavior of

people in experiments. This is a key issue for policy design. When local authorities are

willing to design a program that helps to preserve a common-property resource, it is

imperative to work with communities and consider local traditions and history to build

institutions. Theoretical results are important as a benchmark, but local norms and the

way they evolve with the provision of the natural resource are fundamental.

Turning to the laboratory experiment, an interesting result from the IN treatment

was that the probability of adoption in the laboratory setting was highly correlated with

the results of the Cognitive Reflection Test (CRT), which is usually correlated with

patient and/or less risk-averse participants. This result suggests that participants that

adopt the technology are willing to bear risk of obtaining lower benefits due to potential

free-riding from other users.

Page 92: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

80

The LACEEP1 study also shows that participants of the field experiment are less

likely to comply with the agreement, in comparison to those of the laboratory experiment.

This is surprising, since one would expect that social norms are more important in the

field than in the lab. The results show that farmers that are part of an ejido are more

likely to comply than other farmers. This result partially confirms that importance of

social norms in these settings.

Finally, it is necessary to mention that unobserved heterogeneity has not been

considered in the analysis . As shown in several figures, there is a high degree of hetero-

geneity in decision making among participants of the field experiment, and in a lesser

extent in the laboratory experiment. The presence of heterogeneity might reflect differ-

ent attitudes on the use of the resource. As mentioned before, some individuals might

behave more cooperatively and other more competitively. To identify those attitudes

and analyze whether they are correlated with some individual characteristics is funda-

mental for policy design based on economic experiments. The success of institutions

in common-pool resources management relies on the matching between its design and

user’s behavior. To identify these different behaviors and decision rules is the aim of the

second part of the dissertation.

Page 93: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

Part II

81

Page 94: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

82

Chapter 5

Dynamic Decision-making in the Groundwater

Experiment

5.1 Unbounded Rationality, Bounded Rationality and CPR

Economists traditionally assume that economic agents make choices “as if” they

are unboundedly rational, meaning that the potential to develop rational thinking, given

the information on hand, has no limits (Conlisk 1996). These assumptions have be-

come the keystone of scientific and policy-driven economics, and its power relies on the

formality with which behavior can be represented by formal models.

Nevertheless, since the late 1950’s, several economists have started to question the

rationality assumption. Many concepts such as “bounded rationality”, “heuristics” and

“rules of thumb” started to appear in economic studies as alternative behaviors (Simon

1955). Most of the critiques of rational behavior rely on the idea that the process of

choosing economically rationally (i.e. through profit or utility optimization) is complex,

and individuals have to rely on other criteria to make decisions (Simon 1955; Conlisk

1996). Moreover, some authors emphasize that the cost of acting in such a way is too

high. People incur a deliberation cost to process all the information available in order

to maximize the subjects utility and choose the optimal solution, and this deliberation

cost may be very high (Simon 1955; Conlisk 1996). Indeed, there is evidence that our

capability of making fully rational choices is context-dependent, since individuals show

Page 95: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

83

a diminished cognitive performance when facing temporal adverse situations (Mani et

al. 2013).

Conlisk (1996) reviews several studies that make evident the presence of bounded

rationality in decision making. He discusses the arguments that traditional economists

make to defend the unbounded rationality assumption, and concludes that there is

enough evidence to switch to alternative theories that do not assume unbounded ra-

tionality. Camerer (1998) also emphasizes the necessity of exploring new theories of

individual decision. He focuses on the expected utility theory, and considers that the

prospect theory proposed by Kahneman and Tversky (1979) is a superior theory be-

cause it includes the traditional theory of expected utility and, at the same time, allows

other behaviors, such as loss adversity. Ellison (2006) gives an extensive exploration

of theory and evidence that implies bounded rationality specifically for industrial or-

ganization. Given the evidence that he found, he considers that a boundedly rational

behavior could be more realistic in some setups, and theory based on bounded ratio-

nality could be more flexible, which allows the incorporation of additional features to

models. In a recent article, Harstad and Selten (2013) discuss the reasons why neoclas-

sical economics is currently the prominent way of thinking in economics, and encourage

behavioral economists to pursue a formal and unified model with which it is possible to

compete with the traditional path. They identify some approaches as potential competi-

tors of unbounded rationality if a coherent set of tools that allow comparative statics

and stationary behavior is developed (Harstad and Selten 2013)

Most of the studies that have focused on alternative decision-making theories

have conducted economic laboratory experiments. (Houser and Winter 2004) estimate

Page 96: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

84

discounting rates in an experimental setup using two different estimations strategies:

The first one is based on the usual rationality assumptions, whereas the second one uses

weaker behavioral assumptions. They define several decision rules based on heuristics and

estimate discount rates using these decision rules. They found that the estimation based

on decision rules gives a better prediction of the outcomes than rational expectations.

Goree and Holt (1999) discuss the results of three laboratory experiments and compare

the traditional predictions with those obtained using three different estimation strategies:

learning dynamics, logit equilibrium and iterated noisy introspection. By relaxing the

assumptions of perfect (unbounded) rationality and perfect foresight in the estimation,

they obtain a significant improvement in predicting the outcomes of their experiments.

Other studies have focused on the modeling and identification of classes of subjects.

Lettau et al. (1999) consider different types of agents and propose a theoretical model

in which individuals choose the best of several decisions rules, based on the state and

comparison over past experiences. Similarly, Suleimain and Rapoport (1997) consider

three types: Those concerned with equity, those who maximized utility and those that

cannot be categorized. Moreover, Fischbacher et al. (2001) find that a significant number

of agents that participated of a laboratory experiment can be categorized as “conditional

cooperators”.

A slightly more general approach is the one followed by El-Gamal and Grether

(1995). They propose a pre-defined set of behaviors that participants might show and

then classify participants according to their actions and how close they are to the out-

comes of the pre-defined theoretical behaviors. A more flexible approach is one in which

researchers “let the data speak”, and types of behavior are clustered according to the

Page 97: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

85

actions agents make. After the clustering process, one can categorize the different be-

haviors existent in the population. This approach is followed by Houser et al. (2004).

For the case of natural resource and environmental economics, the question of

whether agents behave unboundedly rational or not is relevant and very important.

Shogren and Taylor (2008) discuss several issues regarding research on “imperfect ra-

tional behavior”, or “behavioral failures” as the authors call deviations from the eco-

nomically rational behavior. They discuss the relevance of these types of behaviors on

environmentally-related economic issues and the relation between behavioral and market

failures, which are common in environmental setups. They bring the question that, given

the specialities related to economics in environmental setups, is the lack of environmental

markets the cause of the existence of behavioral failures? How is it possible to “correct”

for environmental failures in order to have markets that work properly?

Nevertheless, some authors recognize that markets emerge from the interactions

of individuals with different types of behavior that try to exchange a specific good/bad

(Conlisk 1996). These authors propose that markets are self-organizing, and their func-

tionality rely on population heterogeneity. If self-organizing markets do not emerge,

it should be because the behavior of individuals does not allow their creation, or the

goods/bads in mind have some peculiarities that does not allow its exchange (e.g. trans-

action costs), or both. Thus, it is not clear whether behavior “correction” is possible

in order to obtain well-functioning markets, and therefore other types of institutions

have to emerge. Market-oriented policies will not necessarily work if the agents that

are involved are not “economically rational”, as pointed out by Gsottbauer and van den

Bergh (2011). The authors review models that depart from “unbounded” rationality and

Page 98: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

86

analyze the possibilities of applying those models to environmental problems and policy

design. In their review, Gsottbauer and van den Bergh (2011) consider previous studies

that analyze departures from economic rationality and propose that environmental pol-

icy should go beyond price-based regulation or market-based instruments. For instance,

they mention that policy strategies can include social reward or punishment, which could

affect reputation and a consequent change in behavior. Gsottbauer and van den Bergh

(2011) also emphasize the lack of research on behavioral models applied to adaptation to

climate change. In addition, Shogren et al. (2010) note that behavioral economists have

clearly shown two important facts of economic behavior not allowed by the traditional

economic perspective: i) preferences are context-dependent, and ii) social preferences

have an important role in economic choice. The challenge is to incorporate these facts

on the analysis conducted by environmental and natural resource economists, where an

unbiased valuation of environmental non-market goods and the design of institutions

and mechanisms that promote cooperation in a CPR context are needed.

Literature, in general, recognizes the presence of agents whose behavior depart

from homo economicus. However, there is no agreement on whether these types of

behaviors have impacts on market outcomes, if existent. Nevertheless, to identify these

types of behavior in the population is crucial, since the presence of different types of

behaviors in the population might open the door to ways of avoiding the tragedy of the

commons, and facilitates the design of policy and institutions that would help to preserve

the resource, especially when markets are non-existent. This is key in a CPR context,

since the composition of the group might have important effects on the final state of the

resource and well being of users.

Page 99: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

87

In summary, allowing behavior to depart from economic rationality implies not

only the need for theories that might explain the behavior of agents in a better way,

but also the need to identify these behaviors in a heterogenous population. Given that

relaxing the rational behavior assumption opens the door to many other behavioral the-

ories that, in addition, allow context-dependent preferences, the identification of agents

following one or another behavior seems to be a necessary but difficult task to pursue.

This part of the dissertation aims to identify different types of behavior in the

experiment. Data from both field and laboratory economic framed experiments was used

on the context of groundwater management to analyze users’ choices and endogenously

cluster the population on different groups, based on their choices during the game. For

that purpose, a Bayesian classification procedure based on the one proposed by Houser

et al. (2004) was used. The field experiment is presented in LACEEP 2 (2014) and

results from that study are presented here.

5.2 Behavioral Heterogeneity

According to the theoretical model presented in Chapter 3 , an optimum forward-

looking behavior fully internalizes the use value of the resource when a decision about

the demand for water is made. Then, at the beginning of the game, on the one hand,

fully-coordinated agents define the optimum path, which will depend only on time, given

that there is no uncertainty in the game. On the other extreme, fully-myopic agents will

disregard the use value of water remaining in the aquifer, and will decide according to

the current marginal benefits and costs.

Page 100: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

88

Nevertheless, many people might base their decisions on heuristics, or follow “rules

of thumb”, especially when the consequences of their actions are not fully clear. Then,

it might be possible to identify some “types” of participants that lie in the range of

fully-myopic and forward-looking. This section aims at the identification of those types

of participants. In other words, I will try to identify clusters of participants based on

the decisions they make in the experiment.

Figure 5.1 shows the average paths of rounds 1 and 2 of both the field and lab-

oratory experiments. In Round 1, the resulting paths of the field experiment are very

different to those of the laboratory experiment. The average path of pumping hours

in the field experiment decreases in a smoother way than in the lab experiment, which

decreases abruptly at period 4. In Round 2, participants of the field experiment, on

average, used more water than in Round 1, but the paths are very similar. On the other

hand, the average path of the laboratory experiment changed drastically, showing now a

steep drop in period 9. In addition, the average paths obtained in the field experiments

significantly differ from the theoretical benchmarks. However, the average path of the

laboratory experiment in Round 2 resembles the “myopic” path.

Page 101: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

89

Fig. 5.1. Average individual hours pumped by experiment and period

Source: LACEEP 1 (2014)

Although average values per period give us a general picture of the behavior

over time, a different scenario is observed when the distribution of choices is analyzed.

Figure 5.2 shows the histograms of the pumping hours chosen during rounds 1 and 2 for

the laboratory experiment, and Figure 5.3 for the field experiment. In the laboratory

experiment, the majority of choices are concentrated around “10” hours in periods 1

through 3 in Round 1, and then split between “0” and “10”, but there are several

participants that chose values between “3” and “9”. In Round 2, the concentration of

choices increased at “10” and, overall, there are less participants choosing intermediate

values. This change in the distribution might be a response to the learning process that

participants face during the game between rounds 1 and 2. With respect to the field

Page 102: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

90

experiment, the choices are more dispersed and, in Round 1, the value of “5” concentrates

an important proportion of choices, along with the value of “10” in the initial periods,

and then dissipates towards intermediate values. In Round 2, a similar behavior but

with a higher concentration at “10” is observed, suggesting, as in the lab experiment,

that participants are adopting a competitive behavior during the game, but in this case,

the learning process is much slower.

Page 103: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

91

Fig. 5.2. Hours of pumping by period: Laboratory experiment

a. Round 1

b. Round 2

Page 104: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

92

Fig. 5.3. Hours of pumping by period: Field experiment

a. Round 1

b. Round 2

Source: LACEEP 1 (2014)

Page 105: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

93

Another way to show heterogeneity in the data is mapping the choices against

another variable, such as well depth. Figure 5.4 shows the average number of pumping

hours for each depth level for each period of Round 2. A kernel-weighted polynomial

function estimated with the actual data is also shown. In the case of the laboratory

experiment, there is no clear pattern between choices and well depth. Initially, choices

are positively correlated with well depth, but then the relationship becomes negative

until the fifth period, where the sample is divided in two parts, one at the top and one

at the bottom. For the case of the field experiment, a positive relationship between

pumping hours and well depth is shown, but this relationship gets weaker over time and

choices become more dispersed.

Page 106: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

94

Fig. 5.4. Average individual hours pumped by well depth

a. Laboratory

b. Field

Source: LACEEP 1 (2014)

Page 107: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

95

Figure 5.4 also shows the choices made by four members of a randomly-selected

well from both the laboratory and field experiment. With respect to the laboratory exper-

iment, some initial differences in the behavior of participants is observed. For instance,

participant 12076 starts low and then increases the number of hours. On the other hand,

participant 12073 starts high and in period three abruptly decreases to zero, even when

it is still profitable to pump water. This behavior resembles the “rational/strategic” be-

havior of the theoretical model. In the field experiment, participant 16199 consistently

reduces its water demand until period four, and in period five its demand increases. At

the same time, the demand of water of participant 16200 consistently increases, again

until period four, and in period five it is reduced.

These differences in individual behavior could be identified if a more general

framework than the one proposed by the theoretical model is used. For instance, it might

be possible that not all agents behave as if they were unboundedly rational, suggesting

that some agents might use heuristics or “rules of thumb” to make decisions. These

differences are important since most of the theoretical models related to groundwater

management presume that all agents are either myopic or rational.

Houser et al. (2004) develop a methodology in which several “types” of agents

can be identified in a dynamic decision problem. They combine the flexible dynamic

model proposed by Geweke and Keane (2000) with a binary choice finite mixture model

estimated using Bayesian methods. The approach is based on Houser et al.’s application,

but some further steps were taken. First, a common-pool game was considered, in which

participants interact through the environment, and beliefs about other’s actions have to

be included in the model. Second, a discrete choice model as in Houser et al. (2004)

Page 108: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

96

was used, the model is extended to a multinomial probit, since participants can make

choices on the number of hours between 0 and 10. The multinomial probit is a complex

model to estimate with traditional methods, but it is possible to estimate with Bayesian

methods.

5.3 Structural Model

Consider the Bellman equation of farmer i in the groundwater game presented

above:

Vi(dt) = B (wit, dt) + V(dt+1

∣∣∣dt, Wit, wit, t)

Where B (wit, dt) = wit

(α− β

D−dt

)− k are the current-period benefits defined

in the theoretical model, which depend on both the choice of hours of pumping and the

current well depth, and Wjt represents the beliefs of agent i of the total pumping hours

that the other members of the well will use.

As mentioned above, agents could have different behaviors. A rational agent

would solve a maximization problem and choose the level of wit that maximizes the

Bellman equation, considering his/her beliefs about the behavior of the other agents.

However, a myopic agent will use the maximum amount of water as long as Bit is positive,

whereas a cooperative agent will solve a maximization problem in which all agents are

considered altogether in the solution. Moreover, other agents might not make decisions

according to the behaviors presented in the theory. As mentioned above, researchers have

increasingly paid attention to decisions rules based on “rules of thumb” or heuristics.

Page 109: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

97

In order to define a general equation with which all these behaviors could be

represented, Geweke and Keane (2000) consider a general functional form for the “future

function” Vi,t+1:

Vi(dt) = B (wit, dt) + F(dt+1, t+ 1

∣∣∣dt, Wit, wit, t)

Where F (·) is a general function that can be approximated by a N−order poly-

nomial. Note that the equation of motion of the state variable dt+1 and time t + 1 are

considered in the equation. Geweke and Keane (2000) show that, given a proper polyno-

mial approximation, this specification fits data generated with a dynamic programming

algorithm very well. Therefore, according to the authors, it is not necessary to solve

the dynamic programming problem with this specification. Moreover, this specification

gives us the necessary flexibility to analyze other types of behavior that depart from the

forward-looking rational one, since it is not necessary to make any assumption about

how agents form expectations.

Suppose now that there areK types of behavior or decision rules in the population.

These types affect the way how agents form expectations about the future. Then, the

function F for each type of decision rule k will be defined by a a set of parameters βik:

Vi(dt) = B(wit, dt

)+ F

k(dt+1, t+ 1

∣∣∣dt, Wit, wit;βik

)

Page 110: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

98

Where the subscript ik denotes that individual i is classified in type k. For

instance, the third-order polynomial for the approximation of Fk

is:

Fk(wit = h) = β0ik

+ β1ik

(dt+1

∣∣dt, Wit, t, h)

+ β2ik

(d

2

t+1

∣∣dt, Wit, t, h)

+ β3ik

(dt+1

∣∣dt, Wit, t, h)

(t+ 1)

+ β4ik

(dt+1

∣∣dt, Wit, t, h)

(t+ 1)2

+ β5ik

(d

2

t+1

∣∣dt, , Wit, t, h)

(t+ 1)

+ β6ik

(d

3

t+1

∣∣dt, Wit, t, h)

+ β7ik(t+ 1) + β8ik

(t+ 1)2

+ β9ik(t+ 1)

3

The expression of Fk

will depend on the assumptions made about the beliefs of the

actions of other agents, Wjt. For now, it is assumed that agents have symmetric beliefs,

then Wit = 3h1. After inserting the equation of motion of dt+1, the future function for

1It is possible to consider a model of expectations, in which expectations vary with alternativesand states. This is another model that can be simultaneously estimated. Also, it is possible toinclude people beliefs in the equation. This exercise will not be conducted in this study

Page 111: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

99

a 3rd-order polynomial is represented by:

Fk(wit = h) = β0ik

+ β1ik

[dt + 4h− r

]+ β2ik

[d

2

t+ 16h

2+ r

2+ 8dth− 2dtr − 8hr

]+ β3ik

[dt + h+ Wit − r

](t+ 1)

+ β4ik

[dt + h+ Wit − r

](t+ 1)

2

+ β5ik

[d

2

t+ 16h

2+ r

2+ 8dth− 2dtr − 8hr

](t+ 1)

+ β6ik

[d

3

t+ 64h

3 − r3+ 12d

2

th− 3d

2

tr + 48h

2dt − 48h

2r

+3r2dt + 12hr

2 − 24dthr]

+ β7ik(t+ 1) + β8ik

(t+ 1)2

+ β9ik(t+ 1)

3

Where h ∈ {0, 1, ..., 10} are the alternatives the agents face. Polynomials of order

3, 4, 5 and 6 are also specified in this study, and the one that best fits that data according

to a model selection criteria will be selected. In appendix A, the functional forms of the

polynomials of orders 4, 5 and 6 are presented.

Page 112: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

100

Chapter 6

Empirical Specification and Estimation

6.1 Discrete Mixture Models

Unobserved heterogeneity in statistical models has been treated in different ways.

One way to overcome problems of parameters identification due to unobserved hetero-

geneity is through using discrete mixture models. Discrete mixture models consider that

the parameters of the behavioral model, which determine the data likelihood, are drawn

from a combination of a discrete number of distributions. For instance, instead of di-

rectly modeling a multi-modal distribution, flexibility is introduced if the distribution is

modeled as a combination of distributions.

Discrete mixture models are considered a model-based clustering algorithm. Tra-

ditional clustering algorithms (e.g. k-means or hierarchical clustering) are based on the

clustering criterion, usually the minimization of the distances to a centering measure

(mean or median). In contrast, model-based cluster algorithms assume a probabilistic

model from which multivariate observations of some variable are generated by a multi-

variate probability distribution.

Discrete mixture models can consider a point process (fixed coefficients drawn

from a base distribution), and its generalization, a mixture of distributions, from which

random parameters are drawn, and the parameters that define each distribution are

fixed or obtained from a baseline distribution (if allowed by the data). In contrast,

Page 113: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

101

the continuous mixed model, which is commonly use with multinomial logit models,

considers a single continuous distribution from which the parameters are drawn. Figure

6.1 illustrates the approaches usually taken in the literature. On one extreme, the fixed

coefficient model considers a single coefficient for the entire population. Then, the model

can be generalized to several types of coefficients, a single family from which random

coefficients are drawn, and several families from which random coefficients are drawn.

Page 114: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

102

Fig. 6.1. Single vs. Mixed Distributions

a. True distribution

b. Fixed effect c. Point process

d. Random effects e. Mixed distributions

Page 115: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

103

More specifically, a discrete mixture distribution can be represented by:

p(yi|θ,π

)=

K∑k=1

πkf(yi|θk)

Where K is the number of latent classes, clusters or types considered in the model, f(·)

are the distributions to be mixed, θ are the parameters that define each distribution and

π denote the proportions of the population that come from each class. It is possible

to introduce an unobserved indicator that denotes the class membership. Consider the

indicator ζi ∈ {1, 2, ...,K}, and:

I(ζik

)=

1 if ζi = k

0 otherwise

Then, the model in equation 6.1 is equivalent to:

p(yi, ζik

|θ,π)

= p(ζik

)p(yi|θ,π

)=

K∏k=1

(πkf

(yi|θk

))I(ζi)

And the complete data likelihood is represented by:

p (y, ζ|θ,π) =

N∏i=1

K∏k=1

(πkf

(yi|θk

))I(ζik)

Finite mixture models are used in several disciplines (see Congdon (2010); Fruwirth-

Schantter (2006); Gelman et al. (2012)). In economics, one of the most prominent articles

that discussed finite mixture models theoretically is Heckman and Singer (1984). They

Page 116: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

104

estimate a proportional hazard model with a mixture model for unobserved heterogene-

ity. Other studies that have used finite mixture models in economics are Greene and

Hensher (2003); Hess et al. (2011); Howard and Roe (2013) and Scherenberg et al. (2014)

Estimation of finite mixture models is difficult due to the multi-modal shape of the

likelihood. Traditional methods such as maximum likelihood are not suitable for these

types of distributions (Fruwirth-Schantter 2006). Over the last twenty years, researchers

have taken advantage of the computational improvements, and other methods are being

used, such as Expectation-Maximization (EM) and Markov Chain Monte Carlo (MCMC)

Bayesian methods. In this study, Bayesian methods for estimation will be used. The

method is presented in section 6.3.

6.2 Multinomial Probit

In order to identify the classes or “types” of agents in the experiments, it is

necessary to estimate the set of parameters βikthat define the utility function of each

agent. The empirical strategy used by Houser et al. (2004) is based on a random utility

mixture model, where both a point process for coefficients and normal distributions

for the error term are assumed. In other words, Houser et al. (2004) assume that the

coefficients of the future function are drawn from a normal distribution but fixed for

each class, whereas the error term varies between agents, alternatives and classes. The

error term represents any type of information that is unobserved by the econometrician

and that agents might use to make their decisions.

Allowing the error term vary among classes implies that their variances also vary

with classes. Some types will therefore show more dispersion in their error. This is a

Page 117: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

105

strong assumption that might heavily affect the estimates if wrong. Thus, a more con-

servative approach is considered. It is assumed that the error terms are not drawn from

a mixture of distributions. Instead, the error terms of the choice model are considered

iid distributed with mean zero and fixed covariance matrix, and only the coefficients

follow a mixed point process.

Consider the iid normally distributed error term νit(h), with mean zero and co-

variance matrix Σ, and the empirical specification of the value function for alternative h

as:

Vit(dt+1|h) = Bi(h, dt

)+ F

k(dt+1

∣∣∣dt, Wit, h, t;βik

)+ νit(h)

The utility levels are not observed directly. However, the choices that agents make

are. Thus, it is assumed that those decisions that are chosen yield the highest utility

among all the alternatives:

wit = h ⇐⇒ Vit(h) = max {V (j),∀j = {0, 1, . . . , 10}}

Given the empirical specification used, a multinomial probit is the adequate choice

for the estimation of the structural parameters.

It is well known that the multinomial probit model is not identified for levels

(Albert and Chib 1993; Geweke et al. 1994; McCulloch and Rossi 1994; McCulloch et al.

2000; Nobile 1998), so it is necessary to specify the model in relative terms. The utility

Page 118: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

106

of 10 pumping hours (h = 10) is used as the base category. Then, for every h 6= 10:

Vit(h) = Vit(h)− Vit(10)

= Bit(h, dt

)−Bit

(10, dt

)+F

k(dt+1

∣∣∣dt, Wit, h, t;βik

)− F k

(dt+1

∣∣∣dt, Wit, 10, t;βik

)+νit(h)− νit(10)

= Bit(h, dt

)+ F

k(dt+1

∣∣∣dt, Wit, h;βik

)+ ηit(h)

Where η(h) = ν(h)− ν(10) are normally distributed.

Page 119: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

107

Assuming symmetric beliefs and considering a polynomial of order 3, the relative

value function V k is represented by:

Vit (h) = Bi(h, dt

)+ β1ik

(h− 10)

+ β2ik(h− 10)

[2 (h+ 10) + dt − r

]+ β3ik

(h− 10) (t+ 1)

+ β4ik(h− 10) (t+ 1)

2

+ β5ik(h− 10) (t+ 1)

[2 (h+ 10) + dt − r

]+ β6ik

(h− 10)[3(dt − r)

2+ 12(h+ 10)(dt − r) + 16

(h

2+ 10h+ 100

)]+ η(h)

Where β are multiples of the original β. Note that several terms of the polynomial cancel

out when taking the difference. More specifically, any term that is not multiplied by the

the alternative h will be eliminated. The functional form of polynomials of degrees 4, 5

and 6 are presented in appendix A.

In addition to the location constraint, it is necessary to impose another constraint

on the covariance matrix Σ for scale identification (Albert and Chib 1993; Geweke et

al. 1994; McCulloch and Rossi 1994; McCulloch et al. 2000; Nobile 1998). The first

term of the covariance matrix has to be fixed at 1. This is problematic, since it is not

straightforward to sample random matrices with this restriction. In order to be able to

Page 120: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

108

identify the covariance parameters, a method proposed by Chib et al. (1998) is used (see

Section 6.3).

The traditional estimation of the multinomal probit model relies on the probability

of choosing an option as the alternative that yields the highest utility. This requires the

calculation of several integrals, which makes the estimation extremely difficult. For this

study, data augmentation (Albert and Chib 1993; McCulloch and Rossi 1994; McCulloch

et al. 2000; Nobile 1998) will be used to avoid the complexity of the estimation of the

multinomial probit. With this method, the problem reduces to a truncated multivariate

normal model:

yit ∼ TN(µy,Σ

)µy = Bi

(h, dt

)+ F

k(dt+1

∣∣∣dt, Wit, h;βik

)

Where yit is the latent relative utility. A truncated normal distribution is considered

because only values that meet the data constraints can be sampled from the normal

distribution. In this case, the data constraints are determined by the chosen alternative.

In other words, the sampled value of y(h) has to be consistent with the condition that, if

alternative h is chosen, then y(h) has to meet the condition that y(h) = max (y(j), ∀j).

The procedure proposed by Chib et al. (1998) is used to sample from the truncated

normal distribution. This method is presented in Section 6.3.

Finally, clusters of participants are formed assuming that βik= bk. In other

words, only heterogeneity between clusters is consider, leading to a point process model.

Page 121: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

109

Group membership is denoted by the latent index variable ζik, and underlying group

proportions are denoted by λk.

6.3 Estimation

The empirical model that will be estimated is a discrete choice model with hetero-

geneity considered as a discrete point process. Traditional methods of estimation, such

as Maximum Likelihood, have shown problems when estimating mixture models. More-

over, it is well known that the multinomial probit is not easily estimated via Maximum

Likelihood, and that it can only be estimated up to 3 categories. Beyond this number,

it is necessary to use simulation methods, such as Simulated Maximum Likelihood. Re-

cently, researchers started to use two methods for these type of models. EM algorithm

and Bayesian. In this study, Bayesian methods to estimate the parameters of the model

are used. A brief explanation of the nature of Bayesian methods and the idea behind

the use of Monte Carlo Markov Chains is presented below. Then, the specific algorithm

that will be used in the study is explained.

6.3.1 Bayesian Methods

Bayesian inference relies on Bayes’ rule to set up a full probability model for the

data and parameters, and then, after conditioning on the data, calculate a posterior

distribution of the parameters (Gelman et al. 2012). The main idea behind Bayes’ rule

is that the joint probability density of two random variables (y and θ) can be written as

Page 122: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

110

the product of two densities:

p (y, θ) = p (y|θ) p (θ)

Researchers often refer to p (y|θ) as the sampling or data distribution, and p (θ)

as the prior distribution. Then, conditioning on the data y, the posterior distribution

of θ is:

p (θ|y) =p (y, θ)

p (y)=p (y|θ) p (θ)

p (y)

Where p (y) =∫p (y|θ) p (θ) dθ. This term is fixed given the parameter θ, thus, the

literature usually refers to the unnormalized posterior density for Bayesian inference:

p (θ|y) ∝ p (y|θ) p (θ)

This rule indicates that the data y affects the posterior inference of the parameter

θ only through p (y|θ) which, when combined with a probability model, is called the data

likelihood. Traditional methods rely only on the data likelihood for inference. Maximum

Likelihood methods find the parameter θ that maximizes the function p (y|θ). Bayesian

inference uses the data likelihood to update prior information about the parameter. The

prior reflects the state of knowledge about the parameter, which can be informative

or non-informative. The degree of information carried by the prior will determine the

dependence of the posterior on the prior. In other words, a highly informative prior will

yield a very similar posterior since the data does not have an effect on it. Researchers

Page 123: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

111

usually consider non-informative priors for estimation, in order to let the data speak.

However, in some applications, it is necessary to add some information in order to obtain

reliable posterior distributions.

The process of updating information from the prior with the observed data is

usually called Bayesian updating. If this process is done iteratively, given some properties

of the prior and likelihood and a sufficiently high number of iterations, it is possible to

obtain estimates of posterior density characteristics, such as moments and quantiles.

This iterative process is done using Monte Carlo Markov Chains (MCMC). To construct

the MCMC, the parameters have to be sampled from the estimated posterior distribution

in each iteration. The key to MCMC simulation is to create a Markov process whose

stationary distribution is the posterior p (θ|y) (Gelman et al. 2012).

Ideally, one would build the MCMC with direct samples from the posterior distri-

bution. However, this is computationally difficult in most cases and sampling algorithms

that approximate the posterior have to be used (Gelman et al. 2012). The baseline for

MCMC sampling schemes is the Metropolis-Hastings (M-H) algorithm. Intuitively, the

M-H algorithm “builds” the posterior distribution using draws from a proposal distribu-

tion (chosen a priori) and comparing the posterior density from the last updated draw

with that from the previous draw. These draws compose the MCMC, and the higher

the number of draws, the more similar the resulting distribution it is to the target dis-

tribution, the posterior. To determine whether the obtained distribution in each draw is

Page 124: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

112

getting closer to the target distribution, the algorithm uses an acceptance rate:

α(θ∗|θ(t)

)= min

1,p(θ∗|y)f(θ

(t)|θ∗)

p(θ(t)|y

)f(θ∗|θ(t)

)

Where f(·) is the proposal distribution, p(·) is the target distribution, and θ∗

is the

proposal parameter sampled from f(·). Then, the MCMC updates to the next value:

θ(t+1)

=

θ∗

with probability min (r, 1)

θ(t)

otherwise

Details on how to choose f(·) and θ∗

can be found in Gelman et al. (2012) and Congdon

(2010).

The other commonly used sampling algorithm is the Gibbs sampler. This algo-

rithm is considered a special case of the M-H. The advantage of the Gibbs sampler is

that it allows sampling independent blocks of parameters. In other words, it is possible

to sample each parameter separately, conditional on the sampled values of the other pa-

rameters. Thus, each iteration of the MCMC with the Gibbs sampler consists on several

sampling steps, one for each parameter. In contrast, M-H samples the joint distribution

of the parameters simultaneously.

With the M-H, it is possible to sample from any posterior distribution. In contrast,

the Gibbs sampler can only be used whenever the prior and the posterior are conjugate

distributions, meaning that the posterior distribution is in the same family as the prior,

given the specification of the likelihood function. It is very common to combine Gibbs

Page 125: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

113

sampling with M-H for those parameters that do not posses the conjugacy property

(Gelman et al. 2012).

There are several statistics and methods to analyze convergence of MCMC sam-

ples. The most commonly used method is the Brooks-Gelman-Rubin statistic, which

compares several interval lengths of different sampled MCMC for one parameter. Con-

vergence is achieved when the ratio of the mean of the chosen length of the chains is

equal to 1. Another method to assess convergence is the Geweke chi-square tests. This

procedure uses different portions of the MCMC and evaluates whether the two portions

can come from the same distribution. Graphical assessment is also possible through the

analysis of the MCMC trace and MCMC quintiles.

6.3.2 Data Likelihood and Parametrization

The first step to apply Bayesian inference to this case is to to define the data

likelihood. In the model, both the utility levels y and the class indicator ζ are not

observed in the data, but could be inferred (indeed, the major purpose of the study is

to estimate the class indicator ζ). Traditional methods rely on the “observed” likeli-

hood to estimate the parameters. This likelihood is incomplete, since y and ζ are not

observed. Nevertheless, since Bayesian methods rely on a complete probability model

for both parameters and data, it is possible to specify the “complete” data likelihood

as if the latent variables are observed, and then estimate those parameters from the

MCMC simulations. This procedure is called data augmentation (Albert and Chib 1993;

McCulloch and Rossi 1994; McCulloch et al. 2000), and is commonly used with Bayesian

methods.

Page 126: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

114

Considering the probit model presented in Section 6.2, the complete data likeli-

hood of the model is represented by:

L (y, ζ|Σ, b, λ) = p (ζ|Σ, b, λ) p (y|ζ, λ,Σ, b)

=

N∏i=1

T∏t=1

p(ζi|Σ, b, λ

)p(yit|ζ, λ,Σ, b

)=

N∏i=1

T∏t=1

K∏k=1

[λkp

(yit|Σ, bk

)]I(ζi)=

K∏k=1

∏ik

T∏t=1

p(yit|Σ, bk

)( K∏k=1

λNk(ζ)

k

)(6.1)

Where Nk(ζ) is the number of individuals in category k and p(yit|Σ, bk

)is the

truncated multivariate normal:

p(yit|Σ, bk

)= (2π)

−H2 |Σ|−

12 exp

[−1

2

(yit −Bit − F

)′Σ−1 (

yit −Bit − F)]× I

(Sit)(6.2)

In equation 6.2, Sit denotes the support of the truncated normal distribution,

which is defined by the data (the chosen alternatives). Taking advantage of the Gibbs

sampler, Chib et al. (1998) and Geweke and Keane (2000) propose sampling the latent

utilities yit(h) from a truncated univariate normal distribution, given the values of the

latent utilities of the other alternatives. Thus, the support of the truncated normal

distribution is defined by:

Sit(h) =

(max{0,max{yit(−h)}},∞

)if wit = h, h = 1, 2, ...,H(

−∞,max{yit(−h)})

if wit 6= h, h = 1, 2, ...,H

(−∞, 0) if wit = H

Page 127: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

115

Before describing the priors and the sampling algorithm, it is necessary to describe

the parametrization of the covariance matrix Σ. Recall that, to ensure identification, the

multinomial probit requires one of the diagonal elements of the covariance matrix to be

one. Usually, the first element of the matrix is chosen. As mentioned before, to obtain

random matrices with this restriction is a difficult task. Chib et al. (1998) propose a

Choleski decomposition to represent and sample free elements of Σ. Let Σ = LL′, where

L is a (J − 1) × (J − 1) lower triangular matrix with the first element equal to one

(l11 = 1). Chib et al. (1998) prove that the remaining elements can be sampled from

unrestricted univariate normal distributions with the following structure:

θ =(l21, log(l22), l31, l32, log(l33), ..., log(lH−1,H−1)

)

In other words, all the elements can be sampled from normal distributions, but those

elements that compose the diagonal of L are exponentiated. This method ensures a

positive-definite covariance matrix.

6.3.3 Priors and Sampling Algorithm

The set of parameters to estimate is composed by ψ = {b, λ, θ, {ζ}, {y}}, where

{} denote a latent variable.

Page 128: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

116

The following priors for the parameters were considered:

b ∼ N(µb,Σb)

θ ∼ N(0, 0.001)

λ ∼ Dirichlet(α), αk = 2

And the latent variables are sampled, as mentioned before, from:

yit ∼ N(Bit + F ,Σ)I(Sit(h))

ζi ∼ Categorical(λ)

Note that the hyperparameters that define the prior of the parameters (b), µb and

Σb are allowed to be random. Thus, the baseline distribution of the coefficients of the

future function will be endogenously identify . The hyperpriors considered are:

µb ∼ N(µ0,Σ0)

Σb ∼ InverseWishart(M,ρ)

Where µ0 is a V x1 vector of zeros, where V is the number of terms in the future

function, Σ0 is the identity matrix, M is also the identity matrix, and ρ = V . With

the exception of θ, all the priors chosen are the natural conjugate priors. Also, with the

exception of α, relatively non-informative priors are considered.

Page 129: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

117

The Gibbs sampling algorithm is based on the procedures presented in Fruwirth-

Schantter (2006) and Houser et al. (2004) combined with the sampling procedure for the

covariance matrix proposed by Chib et al. (1998):

1. Draw the latent utilities yit(h) from p(yit(h)|yit(−h), b,Σ, λ, ζ)I(Sit(h))

2. Draw the coefficients of the future function bk, for all k = 1, ..,K from the marginalposterior p(bk|y,Σ, λ, ζ), which is proportional to the likelihood in equation (6.1)times the prior of bk.

3. Draw the parameters θ that compose the covariance matrix Σ from p(θ|y, b, λ, ζ)

4. Draw the type proportions λ from the marginal posterior p(λ|ζ) = Dirichlet(α0 +Nk(ζk),∀k)

5. Draw the type membership indicator ζi from p(ζi|y, b,Σ, λ, ζ−i) ∝ p(yi|ζi, ζ−i)p(ζi|ζ−i)

Two steps have to be added in order to make inference about the hyperparameters

µb and Σb:

6. Draw the hyperparameter µb from the marginal posterior p(µb|b,Σb)

7. Draw the hyperparameter Σb from the marginal posterior p(Σb|b, µb)

With the exception of b, the remaining parameters are sampled directly from

the resulting conjugate posteriors. For the case of b, it is necessary to include a M-H

sampling within the Gibbs sampling, as shown in Section 6.3.1.

All the MCMC were run using JAGS on the lionfx computer core at Penn State.

Page 130: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

118

Chapter 7

Results and Discussion

MCMC simulations were performed to approximate the posterior distributions of

the parameters and obtain the classification index ζ in order to cluster participants of the

laboratory experiment into groups. Two different chains for each model were simulated,

and used 20,000 iterations from those chains after considering an adaptation period of

20,000 iterations in which the samplers gain efficiency, and 60,000 iterations of burn-in.

We also discuss the results obtained in the field experiment presented in LACEEP 2

(2014). The results from the two experiments are compared.

7.1 Deviance Information Criteria and Group Number

One key issue that has not been discussed is the way the number of groups is

determined. Houser et al. (2004) performed simulations with several number of groups

and different model specifications (based on the degree of the polynomial of the future

function). For each model, they calculated the marginal likelihood in order to assess the

performance of the model. Once all the models are evaluated, they chose the model with

the best performance in order to obtain the classification.

Even though there are methods with which it is possible to identify the optimal

number of classes, a similar path to Houser et al. (2004) is pursued in this study, with the

difference that, instead of the calculation of the marginal likelihood, model assessment is

Page 131: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

119

based on the Deviance Information Criterion (DIC) (Spiegelhalter et al. 2002). Once the

best specification is identified, identification of the typology of participants is possible.

The DIC is the parallel of the Akaike Information Criteria applied to Bayesian

inference. Consider the deviance defined as D(y, θ) = −2log[p(y|θ)]. The expected value

of D(·), D is a measure of how well the model fits the data, with a better performance the

lower D is. It is always possible to increase the number of parameters and improve model

fitting. The DIC penalizes the model performance adding a measure of complexity. This

measure of complexity, represented by pD = D −D(θ), measures the effective number

of parameters in the model. Thus, pD measures the gap between the average deviance

and the “reference” deviance (Congdon 2010). The reference deviance as the deviance

evaluated with the “true” parameters, θ, is taken. Then, the DIC is calculated as:

DIC = D + pD

MCMC’s with models of 3,4,5 and 6 groups were simulated, and polynomial de-

grees of 3,4, 5 and 6. Table 7.1 presents the DIC for each model. The model with the best

performance in the laboratory experiment has 6 groups with a polynomial of degree 3,

whereas the model with the best performance in the field experiment has 6 groups with

a polynomial of degree 4, as presented by LACEEP 2 (2014). Henceforth, the analysis

will be based on those models. Table 7.2 presents the number of members in each group

for each experiment. Although the number of categories allowed was 6 and 4 clusters in

the laboratory and field experiments, respectively, not all the categories have members

Page 132: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

120

in it. In both the laboratory and field experiments, only three groups have members in

each experiment.

Table 7.1. Deviance Information Criterion

1 2 1 2 1 2 1 23 17,413 17,415 - - 19,503 20,614 18,344 18,3494 17,321 17,451 19,956 17,663 20,084 19,802 18,401 18,3415 17,615 17,440 17,738 19,291 - - 18,253 18,3786 16,823 16,678 19,723 17,389 20,149 20,474 18,313 18,366

6Groups

Degree

Chain3 4 5

a. Laboratory

1 2 1 2 1 2 1 23 21,350 23,303 25,824 24,951 89,234 65,331 66,969 68,3774 27,512 22,638 25,688 21,864 65,398 69,083 61,272 76,8185 - - 21,351 20,902 63,676 94,259 56,128 62,4896 25,117 20,746 25,026 20,202 69,057 68,714 63,233 82,353

Groups3 4 5 6

Chain

Degree

b. Field

Source: LACEEP 1 (2014)

Page 133: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

121

Table 7.2. Distribution of groups

N % N %

Group 1 69 82.1 31 33.7

Group 2 15 17.9 28 30.4

Group 3 - - 33 35.9

Total 84 100.0 92 100.0

Field LaboratoryGroup

Source: LACEEP 1 (2014)

7.2 Parameter Statistics and Convergence

Tables 7.3 and 7.4 show the mean, standard deviations, minimum and maximum

values of the estimated posterior distributions of the parameters. Although very close,

in most cases the parameters are significantly different from zero, since the range of the

distribution does not cross the origin. The size of the parameters is very small because

the variables that enter into the future function are powers of the well depth and time,

which could take very high values if the choices taken are among the highest.

Page 134: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

122

Table 7.3. Descriptive statistics of posterior distributions: Field experiment

Parameter N Mean SD Min Max Gewekebeta 1 20001 -0.000045 0.000002 -0.000050 -0.000038beta 2 20001 -0.000004 0.000004 -0.000014 0.000003 ***beta 3 20001 -0.000002 0.000007 -0.000015 0.000012 ***beta 4 20001 0.000029 0.000003 0.000022 0.000037 ***beta 5 20001 0.000010 0.000003 0.000002 0.000018beta 6 20001 -0.000329 0.000005 -0.000341 -0.000319 ***beta 7 20001 0.000029 0.000000 0.000028 0.000030 ***beta 8 20001 0.000113 0.000003 0.000108 0.000121beta 9 20001 0.000152 0.000013 0.000127 0.000174 ***beta 10 20001 -0.000013 0.000002 -0.000017 -0.000007

Parameter N Mean SD Min Max Gewekebeta 1 20001 -0.000135 0.000007 -0.000161 -0.000119 ***beta 2 20001 0.000062 0.000004 0.000050 0.000074beta 3 20001 0.000028 0.000018 0.000001 0.000071beta 4 20001 0.000131 0.000011 0.000110 0.000153beta 5 20001 0.000143 0.000009 0.000125 0.000169 **beta 6 20001 -0.000586 0.000019 -0.000628 -0.000557 ***beta 7 20001 0.000047 0.000001 0.000044 0.000050beta 8 20001 0.000314 0.000012 0.000283 0.000337 ***beta 9 20001 0.000534 0.000011 0.000510 0.000554 ***beta 10 20001 0.001100 0.000008 0.001082 0.001115 ***

Group 2

Group 1

***=0.01, **=0.05, *=0.1

Source: LACEEP 1 (2014)

Page 135: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

123

Table 7.4. Descriptive statistics of posterior distributions: Laboratory ex-periment

Parameter N Mean SD Min Max Gewekebeta 1 20001 -0.00053 0.00020 -0.00095 -0.00020 ***beta 2 20001 -0.00037 0.00019 -0.00074 -0.00005beta 3 20001 0.00430 0.00017 0.00399 0.00465 ***beta 4 20001 0.00536 0.00022 0.00495 0.00582beta 5 20001 0.01775 0.00043 0.01686 0.01853 **beta 6 20001 -0.00067 0.00001 -0.00069 -0.00065 ***

Parameter N Mean SD Min Max Gewekebeta 1 20001 0.00138 0.00011 0.00108 0.00165beta 2 20001 0.00549 0.00011 0.00518 0.00577beta 3 20001 0.00694 0.00026 0.00654 0.00749beta 4 20001 0.00962 0.00021 0.00906 0.01005beta 5 20001 0.02808 0.00054 0.02706 0.02911 ***beta 6 20001 -0.00023 0.00004 -0.00034 -0.00017 **

Parameter N Mean SD Min Max Gewekebeta 1 20001 -0.00054 0.00019 -0.00104 -0.00027 ***beta 2 20001 -0.00218 0.00013 -0.00241 -0.00183 ***beta 3 20001 0.00317 0.00016 0.00269 0.00342beta 4 20001 -0.00686 0.00008 -0.00708 -0.00665beta 5 20001 0.01992 0.00031 0.01938 0.02050beta 6 20001 -0.00066 0.00001 -0.00067 -0.00064 ***

Group 1

Group 2

Group 3

***=0.01, **=0.05, *=0.1

Figures 7.1 and 7.2 show the estimated posterior distributions. Several parameters

show multimodal distributions, which is a sign of non-convergence. Also, some of the

estimated posterior distributions overlap, which is a sign that discrimination of groups

is not fully achieved.

Page 136: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

124

Fig. 7.1. Estimated posterior distributions of parameters: Field experiment

Source: LACEEP 1 (2014)

Page 137: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

125

Fig. 7.2. Estimated posterior distributions of parameters: Laboratory ex-periment

Convergence of chains is important in order to obtain reliable estimates of the

posterior distributions. One way to assess whether the chains have converged or not is

through tracing the Deviance. Figure 7.3 shows the estimated deviance of the two chains

of each experiment. All the chains are still decreasing but close to achieve a lower bound.

This means that the models would have converged if more iterations would have been

ran.

Page 138: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

126

Fig. 7.3. Trace of deviance, two chains

a. Field

b. Laboratory

Source: LACEEP 1 (2014)

Another way to assess convergence is using Geweke’s convergence test. This test

compares the mean values of the initial and ending parts of the chain. If these averages

are significantly different, then the chains have not converged. Tables 7.3 and 7.4 show

Page 139: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

127

the level of significance of the Geweke’s test. Several parameters have not converged.

However, given the small values of the parameters, it is very unlikely that the stationary

values of the parameters are too different from the current estimates. Traces of the

parameters for each experiment are presented in Appendix B.

7.3 Classification and Characteristics of Groups

Although convergence have not been fully achieved, it is still possible to use the

classification algorithm to cluster participants according to their choices, since several

parameters already converged and correctly discriminated groups. Figure 7.4 shows

the average paths of the hours of pumping for the second round of the experiment for

each type and experiment. In the field experiment, Group 1 starts pumping water at

around 7 units and its consumption does not change significantly over the game. On the

other hand, Group 2 shows a significantly higher use of water during the game, with an

important reduction by the end.

Page 140: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

128

Fig. 7.4. Average individual hours pumped by period and type

a. Field

b. Laboratory

Source: LACEEP 1 (2014)

Page 141: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

129

Fig. 7.5. Average individual pumping costs by period and type

a. Field

b. Laboratory

Source: LACEEP 1 (2014)

Page 142: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

130

With respect to the laboratory experiment, Group 1 could be labeled as “Quasi

myopic”, since they follow a similar pattern to the one followed by the myopic prediction

of the theoretical model. This group keeps a high level of consumption until period

9 when they decrease their consumption to very low levels. Group 2 also follows a

similar patter to the myopic behavior from the theoretical model, as Group 1. It will be

shown that, although clustered in different groups, there are not significant differences

in the final outcomes of these two group. Finally, with respecto to Group 3, this group

follows a similar patter to the rational/strategic theoretical prediction: they keep a high

consumption until period 8 when they drastically reduce their pumping hours without a

severe rise on costs. Then, they raise their number of pumping hours again for the rest

of the game.

It is worth mentioning the differences between figures 5.1 and 7.4, especially for

the field experiment. Once the population is clustered, the nice and smooth path pre-

sented in Figure 5.1 disappears, yielding a more realistic path with more variance. Thus,

researchers should be careful of making inference about the behavior of participants

based on averages, since many unobserved structures could be hidden.

Looking at individual decisions, figures 7.6 and 7.7 show histograms of the choices

taken by participants by period and cluster. As mentioned before, choices of Group 1

in the field experiment show a bimodal distribution in the initial periods, becoming

flatter in latter periods, whereas Group 2 shows values concentrated around the highest

values. With regard the laboratory experiment, both groups 1 and 2, the “myopic”

groups mostly chose 10 units, with the exception of period 4, whereas Group 3 shows

more diversity.

Page 143: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

131

Fig. 7.6. Histogram of choices by cluster and period: Field experiment

Source: LACEEP 1 (2014)

Fig. 7.7. Histogram of choices by cluster and period: Laboratory experiment

Page 144: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

132

Moving towards the end-of-stage variables (total hours, total benefits and final

well depth), Table 7.5 shows important differences between groups in both the field and

laboratory experiments. In the field experiment, Group 2, the “greedy group”, ended

up earning significantly less than Group 1, since its pumping costs increased. With

respect to the laboratory experiment, groups 1 and 2, which used a relatively high

amount of water on average, are the groups that earned the most, whereas Group 3

earned significantly less than the other two groups, even though they used less water.

As mentioned before, there are no significant differences between Group 1 and Group 2

in the laboratory experiment.

Table 7.5. Total benefits and final well depth

Group 1 33.17 315.10 202.55 38.35 309.74 216.77

Group 2 39.73 ** 256.47 *** 229.60 *** 38.75 303.50 219.29

Group 3 - - - 30.82 *** 293.15 *** 204.36 ***

Mean difference tests (T-tests) with respecto to Group 1

GroupHours Benefit Well depth Hours Benefit Well depth

LaboratoryField

Source: LACEEP 1 (2014)

Given these results, it is necessary to note that the composition of the groups is

an important determinant of the the final outcome. In the laboratory experiment, the

dominant behavior is the one from groups 1 and 2, with 64% of participants. This group

shows high extraction rates and, at the end, will also affect the benefits from the other

groups. On the other hand, the dominant behavior in the field experiment is the one

Page 145: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

133

from Group 1, which shows low extraction rates. In this case, total benefits of most of

the participants are improved. They do not achieve social efficiency (365), but surpass

the level of the Nash equilibrium (308).

7.4 Discussion

This part of the dissertation tries to reveal the “hidden behaviors” of the partici-

pants of the experiment. This has been done in way that “the data is allow to speak”, in

opposition to more directed ways of testing behavior, as in experimental economics. In

order to identify and classify different behaviors in the population, a Bayesian classifica-

tion algorithm is used, based on the one proposed by Houser et al. (2004), but adapted

to this case.

The results suggest that it is possible to identify different groups with the method.

Three clear groups in the field experiment were identified: one “greedy” group that

consumes most of the resource at the beginning of the game, a “patient” group that

starts using a relatively low amount of the resource and then increased the quantity

consumed, and a third group that starts with the lower amount of water and remains

low on the following periods. For the case of the laboratory experiment, we found a group

that resembles the prediction of the myopic behavior in the theory. This group starts

with a very high consumption and abruptly reduces it to very low levels, switching again

to higher levels at the last two periods. We also found another group that resembles the

rational/strategic behavior predicted by the theory. Besides the water pumping path,

we also found significant differences in both the total earnings and final well depth of

Page 146: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

134

the groups. Moreover, some groups consumed very similar amounts of the resource at

the end of the game, but the total earnings are significantly different.

As before, the results align with what is found in the literature: non-student par-

ticipants tend to show a more pro-social behavior than the student population. However,

in the field experiment, we were able to identify a small group of participants that did

not behave in this manner. Identifying these different behaviors is crucial for the design

and success of institutions.

It is worth mentioning some comments about the classification method. First, to

achieving convergence is difficult. Some strategies that might help achieving convergence

include a reparametrization of the model and considering more informative priors. The

first strategy is difficult to perform since the empirical model is based on a structural

model that is already very flexible. A reparametrization of the model might change

the structural form from a random utility-based choice model to something else. The

strength of the model relies on the fact that it is random-utility based. The second

strategy is also difficult to apply since it is not possible to obtain more prior information.

Nevertheless, the results suggest that convergence might be achievable, and that reliable

estimates of the posteriors can be obtained, given that some chains did converge.

Page 147: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

135

Chapter 8

Concluding Remarks

This dissertation summarizes the results obtained from economic framed experi-

ments conducted with both farm water users from Mexico and undergraduate students.

Few studies have compared the results of artefactual experiments with the two popu-

lations. As discussed in previous chapters, there are several differences in the behavior

of the two populations, which confirms the results obtained by other researchers. Thus,

it is imperative for experimental economists to replicate experiments with several types

of populations in order to ensure the external validity of the study, especially if the

experiment is framed, as in this case.

The first part of this dissertation presents a dynamic CPR experiment framed as

a groundwater game. This frame allows the analysis of dynamic decisions in CPR, and

does not rely on the analysis of steady state equilibrium outcomes, as most dynamic

experiments do. This is an innovative design, since previous studies do not consider

the groundwater problem in experimental settings, with the exception of Suter et al.

(2012). Moreover, this is the first study that conducts a groundwater experiment with

farm groundwater users.

The results obtained in Part I are very suggestive on the idea that the kinds

of participants are an important issue usually overlooked by researchers. These results

confirm that external validity of experiments is difficult to achieve, and it is only possible

Page 148: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

136

to derive conclusions about the target population. External validity is an important

property that must be sought if the intention is to use the results of experiments as

benchmarks for policy design in a broader population.

In addition, these results suggest that farmers bring to experiments prior infor-

mation that affect their choice behavior. This goes in line with what is proposed by

Cardenas and Ostrom (2004). Even though students tend to behave more in line with

what traditional economic theory proposes, exceptions are consistently found in exper-

imental setups and researchers are trying to explain, with alternative theories, these

departures from the traditional economic benchmark. The setup becomes more compli-

cated on the field, not only because of comprehension issues (which could be overcome

with more practice rounds, simpler and concise instructions, etc.), but also because of

the information that farmers bring to the session.

As Cardenas and Ostrom (2004) mention, there are three layers of information

that participants in a social dilemma experiment might take into account to make their

choices: i) Material payoff game layer, ii) Group-context layer and, iii) Identity layer.

The results obtained in the investment treatment in the field experiment give insights

about the importance of the identity layer when compared with those obtained in the

laboratory experiment, since farmers have more experience and are more familiar with

irrigation technologies than students.

Another challenge is identifying and classifying different types of behaviors within

a heterogenous population. This is key in a CPR context, since the composition of groups

might have important effects on the final state of the resource and well being of users.

Moreover, since usually CPR users interact in a dynamic setup, and preferences are

Page 149: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

137

context-dependent, agents might switch their behavior based on the partial results of the

“game”. Then, users with a “dominant” behavior might exert their influence on users

with other behaviors, which might result in the outcomes predicted by the dominant

behavior even though the population was originally heterogenous.

It is possible, in experimental contexts, to design the experiment and apply the

treatments in order to verify whether participants behave according to a given theory or

not. Nevertheless, this strategy does not exclude the possibility that other behavioral

theories are at stake, and that these theories could offer better explanations than the ones

considered in the study. Departing from the traditional economic rational framework

opens a whole new set of behavioral motives that should be investigated. The problem

is that there are no limits on the size of the set of alternatives.

In this study, we clearly found two groups. In the field experiment, a small group

of participants try to consume most of the resource at the beginning of the game, whereas

the majority of participants are characterized for being more cautious. In contrast, in

the laboratory experiment, most of participants are categorized as myopic, and a small

group resembles the behavior of the rationa/strategic agent from the theory.

The ultimate purpose of the classification exercise is to analyze the effects of

policies on the different groups that compose the populations. Groups with some type

of behavior will react in different ways to a certain policy than other groups. Moreover,

the analysis of the interaction of these groups and the resulting wellbeing is a major task

to do.

Further research should incorporate the analysis of switching behaviors in the

population. As mentioned before, some groups with a given behavior could be more

Page 150: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

138

dominant than others. In a dynamic setup, this might lead to changes in the behavior of

others, since agents learn about the process and the attitude of the other agents. Some

researchers are starting to analyze economic complex systems under this evolutionary

approach. Some tools that, traditionally, were used by biologists or ecologists, such as

Agent-based models and system dynamics, are becoming more popular among environ-

mental and natural resource economists. The analysis of the properties of the steady

state equilibrium as a general characterization of the economic system is becoming less

useful when a heterogenous population constantly interacts in a changing environment.

Nevertheless, in order to conduct research on complex and evolutionary systems, it is

necessary to establish analytical boundaries that allow identification of causal effects.

This is a hard task with complex systems.

Finally, it is important to mention the potential benefits of using economic framed

field experiments as a pedagogical tool. Some researchers are starting to conduct projects

in order to assess the effectiveness of these research methods for strengthening collective

action in communities in developing countries (IFPRI 2013). Although it was not pos-

sible to perform a formal evaluation in this project, participants of the field experiment

manifested their “gratitude” for showing them in “simple terms” the groundwater prob-

lems that they are facing, and for making them think about how severe these problems

are. As mentioned in LACEEP 2 (2014), farmers at the end of all the sessions started

a discussion about the groundwater problems in Aguascalientes without any encourage-

ment from the facilitator. Although I do not expect that these sessions generated any

significant change in the behavior of farmers and the way they use water, it was ex-

tremely rewarding to see that my work helped at least a small bit in the comprehension

Page 151: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

139

of the commons problem and how to avoid the commons tragedy in the study region.

Ultimately, this is the aim of economic research: to generate knowledge to aid the critical

thinking of decision makers.

Page 152: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

140

A Functional Forms of Polynomials of a Higher Degree

In this appendix, we will present the functional forms of the polynomials of order4, 5 and 6.Degree = 4The future function of the polynomial of degree 4 is represented by:

Fk(wit = h) = β0ik

+ β1ik

(dt+1

∣∣dt, Wit, t, h)

+ β2ik

(d

2

t+1

∣∣dt, Wit, t, h)

+ β3ik

(dt+1

∣∣dt, Wit, t, h)

(t+ 1) + β4ik

(dt+1

∣∣dt, Wit, t, h)

(t+ 1)2

+ β5ik

(d

2

t+1

∣∣dt, Wit, t, h)

(t+ 1) + β6ik

(d

3

t+1

∣∣dt, Wit, t, h)

+ β7ik(t+ 1) + β8ik

(t+ 1)2

+ β9ik(t+ 1)

3+ β10ik

(d

4

t+1

∣∣dt, Wit, t, h)

+ β11ik

(dt+1

∣∣dt, Wit, t, h)

(t+ 1)3

+ β12ik

(d

2

t+1

∣∣dt, Wit, t, h)

(t+ 1)2

+ β13ik

(d

3

t+1

∣∣dt, Wit, t, h)

(t+ 1)

+ β14ik(t+ 1)

4

Page 153: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

141

After considering the assumption of Wit = 3h and plugging in the equation ofmotion of well depth, we get:

Fk(wit = h) = β0ik

+ β1ik

[dt + 4h− r

]+ β2ik

[d

2

t+ 16h

2+ r

2+ 8dth− 2dtr − 8hr

]+ β3ik

[dt + 4h− r

](t+ 1)

+ β4ik

[dt + 4h− r

](t+ 1)

2

+ β5ik

[d

2

t+ 16h

2+ r

2+ 8dth− 2dtr − 8hr

](t+ 1)

+ β6ik

[d

3

t+ 64h

3 − r3+ 12d

2

th− 3d

2

tr + 48h

2dt − 48h

2r + 3r

2dt

+12hr2 − 24dthr

]+ β7ik

(t+ 1) + β8ik(t+ 1)

2+ β9ik

(t+ 1)3

+ β10ik

[d

4

t+ 16hd

3

t− 4rd

3

t+ 96h

2d

2

t− 48hrd

2

t+ 6r

2d

2

t+ 256h

3dt

−192h2rdt + 48hr

2dt − 4r

3dt + 256h

4 − 256h3r + 96h

2r

2

−16hr3

+ r4]

+ β11ik

[dt + 4h− r

](t+ 1)

3

+ β12ik

[d

2

t+ 16h

2+ r

2+ 8dth− 2dtr − 8hr

](t+ 1)

2

+ β13ik

[d

3

t+ 64h

3 − r3+ 12d

2

th− 3d

2

tr + 48h

2dt − 48h

2r + 3r

2dt

+12hr2 − 24dthr

](t+ 1)

+ β14ik(t+ 1)

4

Page 154: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

142

And the relative future function, Fk

i represented by:

Fk

(h) = β1ik(h− 10)

+ β2ik(h− 10)

[2 (h+ 10) + dt − r

]+ β3ik

(h− 10) (t+ 1)

+ β4ik(h− 10) (t+ 1)

2

+ β5ik(h− 10) (t+ 1)

[2 (h+ 10) + dt − r

]+ β6ik

(h− 10)[3(dt − r)

2+ 12(h+ 10)(dt − r) + 16

(h

2+ 10h+ 100

)]+ β7ik

(h− 10)[(dt − r)

3+ 6(h+ 10)(dt − r)

2

+16(h

2+ 10h+ 100

)(dt − r)− 16

(h

3+ 10h

2+ 100h+ 1000

)]+ β8ik

(h− 10) (t+ 1)3

+ β9ik(h− 10) (t+ 1)

2 [2 (h+ 10) + dt − r

]+ β10ik

(h− 10) (t+ 1)[3(dt − r)

2+ 12(h+ 10)(dt − r) + 16

(h

2+ 10h+ 100

)]

Page 155: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

143

Degree = 5The future function of the polynomial of degree 5 is represented by:

Fk(wit = h) = β0ik

+ β1ik

(dt+1

∣∣dt, Wit, t, h)

+ β2ik

(d

2

t+1

∣∣dt, Wit, t, h)

+ β3ik

(dt+1

∣∣dt, Wit, t, h)

(t+ 1) + β4ik

(dt+1

∣∣dt, Wit, t, h)

(t+ 1)2

+ β5ik

(d

2

t+1

∣∣dt, Wit, t, h)

(t+ 1) + β6ik

(d

3

t+1

∣∣dt, Wit, t, h)

+ β7ik(t+ 1) + β8ik

(t+ 1)2

+ β9ik(t+ 1)

3+ β10ik

(d

4

t+1

∣∣dt, Wit, t, h)

+ β11ik

(dt+1

∣∣dt, Wit, t, h)

(t+ 1)3

+ β12ik

(d

2

t+1

∣∣dt, Wit, t, h)

(t+ 1)2

+ β13ik

(d

3

t+1

∣∣dt, Wit, t, h)

(t+ 1)

+ β14ik(t+ 1)

4

+ β15ik

(d

5

t+1

∣∣dt, Wit, t, h)

+ β16ik

(dt+1

∣∣dt, Wit, t, h)

(t+ 1)4

+ β17ik

(d

2

t+1

∣∣dt, Wit, t, h)

(t+ 1)3

+ β18ik

(d

3

t+1

∣∣dt, Wit, t, h)

(t+ 1)2

+ β19ik

(d

4

t+1

∣∣dt, Wit, t, h)

(t+ 1)

+ β20ik(t+ 1)

5

Page 156: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

144

After considering the assumption of Wit = 3h and plugging in the equation ofmotion of well depth, we get:

Fk(wit = h) = β0ik

+ β1ik

[dt + 4h− r

]+ β2ik

[d

2

t+ 16h

2+ r

2+ 8dth− 2dtr − 8hr

]+ β3ik

[dt + 4h− r

](t+ 1)

+ β4ik

[dt + 4h− r

](t+ 1)

2

+ β5ik

[d

2

t+ 16h

2+ r

2+ 8dth− 2dtr − 8hr

](t+ 1)

+ β6ik

[d

3

t+ 64h

3 − r3+ 12d

2

th− 3d

2

tr + 48h

2dt − 48h

2r + 3r

2dt

+12hr2 − 24dthr

]+ β7ik

(t+ 1) + β8ik(t+ 1)

2+ β9ik

(t+ 1)3

+ β10ik

[d

4

t+ 16hd

3

t− 4rd

3

t+ 96h

2d

2

t− 48hrd

2

t+ 6r

2d

2

t+ 256h

3dt

−192h2rdt + 48hr

2dt − 4r

3dt + 256h

4 − 256h3r + 96h

2r

2

−16hr3

+ r4]

+ β11ik

[dt + 4h− r

](t+ 1)

3

+ β12ik

[d

2

t+ 16h

2+ r

2+ 8dth− 2dtr − 8hr

](t+ 1)

2

+ β13ik

[d

3

t+ 64h

3 − r3+ 12d

2

th− 3d

2

tr + 48h

2dt − 48h

2r + 3r

2dt

+12hr2 − 24dthr

](t+ 1)

+ β14ik(t+ 1)

4

+ β15ik

[d

5

t+ 20hd

4

t− 5rd

4

t+ 160h

2d

3

t− 80hrd

3

t+ 10r

2d

3

t+ 640h

3d

2

t

−480h2rd

2

t+ 120hr

2d

2

t− 10r

3d

2

t+ 1280h

4dt − 1280h

3rdt

+480h2r

2dt − 80hr

3dt + 5r

4dt + 1024h

5 − 1280h4r

+640h3r

2 − 160h2r

3+ 20hr

4 − r5]

+ β16ik

[dt + 4h− r

](t+ 1)

4

+ β17ik

[d

2

t+ 16h

2+ r

2+ 8dth− 2dtr − 8hr

](t+ 1)

3

Page 157: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

145

+ β18ik

[d

3

t+ 64h

3 − r3+ 12d

2

th− 3d

2

tr + 48h

2dt − 48h

2r + 3r

2dt

+12hr2 − 24dthr

](t+ 1)

2

+ β19ik

[d

4

t+ 16hd

3

t− 4rd

3

t+ 96h

2d

2

t− 48hrd

2

t+ 6r

2d

2

t+ 256h

3dt

−192h2rdt + 48hr

2dt − 4r

3dt + 256h

4 − 256h3r + 96h

2r

2

−16hr3

+ r4]

(t+ 1)

+ β20ik(t+ 1)

5

Page 158: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

146

And the relative future function, Fk

i represented by:

Fk

(h) = β1ik(h− 10)

+ β2ik(h− 10)

[2 (h+ 10) + dt − r

]+ β3ik

(h− 10) (t+ 1)

+ β4ik(h− 10) (t+ 1)

2

+ β5ik(h− 10) (t+ 1)

[2 (h+ 10) + dt − r

]+ β6ik

(h− 10)[3(dt − r)

2+ 12(h+ 10)(dt − r) + 16

(h

2+ 10h+ 100

)]+ β7ik

(h− 10)[(dt − r)

3+ 6(h+ 10)(dt − r)

2

+16(h

2+ 10h+ 100

)(dt − r)

−16(h

3+ 10h

2+ 100h+ 1000

)]+ β8ik

(h− 10) (t+ 1)3

+ β9ik(h− 10) (t+ 1)

2 [2 (h+ 10) + dt − r

]+ β10ik

(h− 10) (t+ 1)[3(dt − r)

2+ 12(h+ 10)(dt − r) + 16

(h

2+ 10h+ 100

)]+ β11ik

(h− 10)[5(dt − r)

4 − 40(h+ 10)(dt − r)3

+160(h2 − 10h+ 100)(dt − r)

2

+320(h3

+ 10h2

+ 100h+ 1000)(dt − r)

+256(h

4+ 10h

3+ 100h

2+ 1000h+ 10000

)]+ β12ik

(h− 10) (t+ 1)4

+ β13ik(h− 10) (t+ 1)

3 [2 (h+ 10) + dt − r

]+ β14ik

(h− 10) (t+ 1)2[3(dt − r)

2+ 12(h+ 10)(dt − r) + 16

(h

2+ 10h+ 100

)]+ β15ik

(h− 10) (t+ 1)[(dt − r)

3+ 6(h+ 10)(dt − r)

2

+16(h

2+ 10h+ 100

)(dt − r)

−16(h

3+ 10h

2+ 100h+ 1000

)]

Page 159: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

147

Degree = 6The future function of the polynomial of degree 5 is represented by:

Fk(wit = h) = β0ik

+ β1ik

(dt+1

∣∣dt, Wit, t, h)

+ β2ik

(d

2

t+1

∣∣dt, Wit, t, h)

+ β3ik

(dt+1

∣∣dt, Wit, t, h)

(t+ 1) + β4ik

(dt+1

∣∣dt, Wit, t, h)

(t+ 1)2

+ β5ik

(d

2

t+1

∣∣dt, Wit, t, h)

(t+ 1) + β6ik

(d

3

t+1

∣∣dt, Wit, t, h)

+ β7ik(t+ 1) + β8ik

(t+ 1)2

+ β9ik(t+ 1)

3+ β10ik

(d

4

t+1

∣∣dt, Wit, t, h)

+ β11ik

(dt+1

∣∣dt, Wit, t, h)

(t+ 1)3

+ β12ik

(d

2

t+1

∣∣dt, Wit, t, h)

(t+ 1)2

+ β13ik

(d

3

t+1

∣∣dt, Wit, t, h)

(t+ 1)

+ β14ik(t+ 1)

4

+ β15ik

(d

5

t+1

∣∣dt, Wit, t, h)

+ β16ik

(dt+1

∣∣dt, Wit, t, h)

(t+ 1)4

+ β17ik

(d

2

t+1

∣∣dt, Wit, t, h)

(t+ 1)3

+ β18ik

(d

3

t+1

∣∣dt, Wit, t, h)

(t+ 1)2

+ β19ik

(d

4

t+1

∣∣dt, Wit, t, h)

(t+ 1)

+ β20ik(t+ 1)

5

+ β21ik

(d

6

t+1

∣∣dt, Wit, t, h)

+ β22ik

(dt+1

∣∣dt, Wit, t, h)

(t+ 1)5

+ β23ik

(d

2

t+1

∣∣dt, Wit, t, h)

(t+ 1)4

+ β24ik

(d

3

t+1

∣∣dt, Wit, t, h)

(t+ 1)3

+ β25ik

(d

4

t+1

∣∣dt, Wit, t, h)

(t+ 1)2

+ β26ik

(d

5

t+1

∣∣dt, Wit, t, h)

(t+ 1)

+ β27ik(t+ 1)

6

Page 160: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

148

After considering the assumption of Wit = 3h and plugging in the equation ofmotion of well depth, we get:

Fk(wit = h) = β0ik

+ β1ik

[dt + 4h− r

]+ β2ik

[d

2

t+ 16h

2+ r

2+ 8dth− 2dtr − 8hr

]+ β3ik

[dt + 4h− r

](t+ 1)

+ β4ik

[dt + 4h− r

](t+ 1)

2

+ β5ik

[d

2

t+ 16h

2+ r

2+ 8dth− 2dtr − 8hr

](t+ 1)

+ β6ik

[d

3

t+ 64h

3 − r3+ 12d

2

th− 3d

2

tr + 48h

2dt − 48h

2r + 3r

2dt

+12hr2 − 24dthr

]+ β7ik

(t+ 1) + β8ik(t+ 1)

2+ β9ik

(t+ 1)3

+ β10ik

[d

4

t+ 16hd

3

t− 4rd

3

t+ 96h

2d

2

t− 48hrd

2

t+ 6r

2d

2

t+ 256h

3dt

−192h2rdt + 48hr

2dt − 4r

3dt + 256h

4 − 256h3r + 96h

2r

2

−16hr3

+ r4]

+ β11ik

[dt + 4h− r

](t+ 1)

3

+ β12ik

[d

2

t+ 16h

2+ r

2+ 8dth− 2dtr − 8hr

](t+ 1)

2

+ β13ik

[d

3

t+ 64h

3 − r3+ 12d

2

th− 3d

2

tr + 48h

2dt − 48h

2r + 3r

2dt

+12hr2 − 24dthr

](t+ 1)

+ β14ik(t+ 1)

4

+ β15ik

[d

5

t+ 20hd

4

t− 5rd

4

t+ 160h

2d

3

t− 80hrd

3

t+ 10r

2d

3

t+ 640h

3d

2

t− 480h

2rd

2

t+

120hr2d

2

t− 10r

3d

2

t+ 1280h

4dt − 1280h

3rdt + 480h

2r

2dt − 80hr

3dt + 5r

4dt

+1024h5 − 1280h

4r + 640h

3r

2 − 160h2r

3+ 20hr

4 − r5]

+ β16ik

[dt + 4h− r

](t+ 1)

4

+ β17ik

[d

2

t+ 16h

2+ r

2+ 8dth− 2dtr − 8hr

](t+ 1)

3

+ β18ik

[d

3

t+ 64h

3 − r3+ 12d

2

th− 3d

2

tr + 48h

2dt − 48h

2r + 3r

2dt

+12hr2 − 24dthr

](t+ 1)

2

+ β19ik

[d

4

t+ 16hd

3

t− 4rd

3

t+ 96h

2d

2

t− 48hrd

2

t+ 6r

2d

2

t+ 256h

3dt

−192h2rdt + 48hr

2dt − 4r

3dt + 256h

4 − 256h3r + 96h

2r

2

−16hr3

+ r4]

(t+ 1)

Page 161: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

149

+ β20ik(t+ 1)

5

+ β21ik

[d

6

t+ 24hd

5

t− 6rd

5

t+ 240h

2d

4

t− 120hrd

4

t+ 15r

2d

4

t+ 1280h

3d

3

t− 960h

2rd

3

t

+240hr2d

3

t− 20r

3d

3

t+ 3840h

4d

2

t− 3840h

3rd

2

t+ 1440h

2r

2d

2

t− 240hr

3d

2

t

+15r4d

2

t+ 6144h

5dt − 7680h

4rdt + 3840h

3r

2dt − 960h

2r

3dt + 120hr

4dt

−6r5dt + 4096h

6 − 6144h5r + 3840h

4r

2 − 1280h3r

3+ 240h

2r

4 − 24hr5

+ r6]

+ β22ik

[dt + 4h− r

](t+ 1)

5

+ β23ik

[d

2

t+ 16h

2+ r

2+ 8dth− 2dtr − 8hr

](t+ 1)

2

+ β24ik

[d

3

t+ 64h

3 − r3+ 12d

2

th− 3d

2

tr + 48h

2dt − 48h

2r + 3r

2dt

+12hr2 − 24dthr

](t+ 1)

3

+ β25ik

[d

4

t+ 16hd

3

t− 4rd

3

t+ 96h

2d

2

t− 48hrd

2

t+ 6r

2d

2

t+ 256h

3dt − 192h

2rdt

+48hr2dt − 4r

3dt + 256h

4 − 256h3r + 96h

2r

2 − 16hr3

+ r4]

(t+ 1)2

+ β26ik

[d

5

t+ 20hd

4

t− 5rd

4

t+ 160h

2d

3

t− 80hrd

3

t+ 10r

2d

3

t+ 640h

3d

2

t− 480h

2rd

2

t+

120hr2d

2

t− 10r

3d

2

t+ 1280h

4dt − 1280h

3rdt + 480h

2r

2dt − 80hr

3dt + 5r

4dt

+1024h5 − 1280h

4r + 640h

3r

2 − 160h2r

3+ 20hr

4 − r5]

+ β27ik(t+ 1)

6

Page 162: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

150

And the relative future function, Fk

i represented by:

Fk

(h) = β1ik(h− 10)

+ β2ik(h− 10)

[2 (h+ 10) + dt − r

]+ β3ik

(h− 10) (t+ 1)

+ β4ik(h− 10) (t+ 1)

2

+ β5ik(h− 10) (t+ 1)

[2 (h+ 10) + dt − r

]+ β6ik

(h− 10)[3(dt − r)

2+ 12(h+ 10)(dt − r) + 16

(h

2+ 10h+ 100

)]+ β7ik

(h− 10)[(dt − r)

3+ 6(h+ 10)(dt − r)

2

+16(h

2+ 10h+ 100

)(dt − r)

−16(h

3+ 10h

2+ 100h+ 1000

)]+ β8ik

(h− 10) (t+ 1)3

+ β9ik(h− 10) (t+ 1)

2 [2 (h+ 10) + dt − r

]+ β10ik

(h− 10) (t+ 1)[3(dt − r)

2+ 12(h+ 10)(dt − r) + 16

(h

2+ 10h+ 100

)]+ β11ik

(h− 10)[5(dt − r)

4 − 40(h+ 10)(dt − r)3

+160(h2 − 10h+ 100)(dt − r)

2

+320(h3

+ 10h2

+ 100h+ 1000)(dt − r)

+256(h

4+ 10h

3+ 100h

2+ 1000h+ 10000

)]+ β12ik

(h− 10) (t+ 1)4

+ β13ik(h− 10) (t+ 1)

3 [2 (h+ 10) + dt − r

]+ β14ik

(h− 10) (t+ 1)2[3(dt − r)

2+ 12(h+ 10)(dt − r) + 16

(h

2+ 10h+ 100

)]+ β15ik

(h− 10) (t+ 1)[(dt − r)

3+ 6(h+ 10)(dt − r)

2

+16(h

2+ 10h+ 100

)(dt − r)

−16(h

3+ 10h

2+ 100h+ 1000

)]+ β16ik

(h− 10)[3(dt − r)

5+ 30(h+ 10)(dt − r)

4

+160(h2

+ 10h+ 100)(dt − r)3

+480(h3

+ 10h2

+ 100h3 − 1000)(dt − r)

2

+768(h4

+ 10h3

+ 100h2

+ 1000h+ 10000)(dt − r)

+512(h5

+ 10h4

+ 100h3

+ 1000h2

+ 10000h− 100000)]

Page 163: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

151

+ β17ik(h− 10) (t+ 1)

5

+ β18ik(h− 10) (t+ 1)

4 [2 (h+ 10) + dt − r

]+ β19ik

(h− 10) (t+ 1)3[3(dt − r)

2+ 12(h+ 10)(dt − r) + 16

(h

2+ 10h+ 100

)]+ β20ik

(h− 10) (t+ 1)2[(dt − r)

3+ 6(h+ 10)(dt − r)

2

+16(h

2+ 10h+ 100

)(dt − r)

−16(h

3+ 10h

2+ 100h+ 1000

)]+ β21ik

(h− 10) (t+ 1)[5(dt − r)

4 − 40(h+ 10)(dt − r)3

+160(h2 − 10h+ 100)(dt − r)

2

+320(h3

+ 10h2

+ 100h+ 1000)(dt − r)

+256(h

4+ 10h

3+ 100h

2+ 1000h+ 10000

)]B MCMC of Parameters

In this appendix, we will present the MCMC of the parameters of the futurefunction.

Page 164: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

152

Fig. B.1. Trace of the MCMC of all the parameters of the future functionof the field experiment

Group 1

Group 2

Page 165: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

153

Fig. B.2. Trace of the MCMC of all the parameters of the future functionof the field experiment - Group 1

Page 166: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

154

Fig. B.3. Trace of the MCMC of all the parameters of the future functionof the field experiment - Group 2

Page 167: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

155

Fig. B.4. Trace of the MCMC of all the parameters of the future functionin the laboratory experiment

Group 1

Group 2

Group 3

Page 168: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

156

Fig. B.5. Trace of the MCMC of all the parameters of the future functionof the laboratory experiment - Group 1

Page 169: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

157

Fig. B.6. Trace of the MCMC of all the parameters of the future functionof the laboratory experiment - Group 2

Page 170: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

158

Fig. B.7. Trace of the MCMC of all the parameters of the future functionof the laboratory experiment - Group 3

Page 171: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

159

Bibliography

Agaarwal, Rimjhim M. and Tulika A. Narayan (2004), Does inequality lead to greater efficiencyin the use of local commons? The role of strategic investment in capacity, Journal of Environ-mental Economics and Management, Vol. 47: 163-182.

Albert, James H. and Siddhartha Chib (1993), Bayesian analysis of binary and polychotomusdata, Journal of the American Statistical Association, Vol. 88: 669-679

Anderson, Jon, Stepehen V. Burks, Jeffrey Carpenter, Lorenz Gotte, Karsten Maurer, DanieleNosenzo, Ruth Potter, Kim Rocha and Aldo Rustichini (2013), Self-selection and variationsin the laboratory measurement of other-regarding preferences across subject pools: evidencefrom college students and two adult samples, Experimental Economics, Vol. 16: 170-189.

Bardhan, Pranab (2000), Irrigation and Cooperation: An Empirical Analysis of 48 IrrigationCommunities in South India, Economic Development and Cultural Change, Vol 48 (4): 847-865.

Barham, Bradford L., Jean-Paul Chavas and Oliver T. Coomes (1998), Sunk Costs and theNatural Resource Extraction Sector: Analytical Models and Historical Examples of Hysteresisand Strategic Behavior in the Americas, Land Economics, Vol. 74 (4): 429-448.

Belot, Michele, Luis Miller and Raymond Duch (2010), Who Should be Called to the Lab?: Acomprehensive comparison of students and non-students in classic experimental games, NuffieldCentre for Experimental Social Sciences Discussion Paper Series 2010-001.

Brill, Thomas C. and H. Stuart Burness (1994), Planning versus competitive rates of groundwaterpumping, Water Resources Research, Vol. 30(6): 1873-1880

Brown, Gardner Jr. and Robert Deacon (1972), Economic Optimization of Single-Cell Aquifer,Water Resources Research, Vol. 8(3): 557-564

Burt, Oscar R. (1964), Optimal Resource Use Over time with and Application to Ground Water,Management Science, Vol. 11(1) Series A: 80-93

Burt, Oscar R. (1967), Temporal Allocation of Groundwater, Water Resources Research, Vol.3(1): 45-56

Burt, Oscar R. (1970), Groundwater Storage Control undr Institutional Restrictions, WaterResources Research, Vol. 6(6): 1540-1548

Caldera Ortega, Alex Ricardo (2009), Gobernanza y sustenibilidad: Desarrollo institucional yprocesos polıticos en torno al agua subterranea en Mxico. Los casos del valle de Leon y delvalle de Aguascalientes, Ph.D. Dissertation in Political Sciences, Facultad Latinoamericanade Ciencias Sociales-FLACSO, Mexico City.

Camerer, Colin (1998), Bounded Rationality in Individual Decision Making, Experimental Eco-nomics, Vol. 1: 163-183

Cameron, A. Colin and Pravin K. Trivedi (2005), Microeconometrics: Methods and Applications,New York: Cambridge University Press.

Cardenas, Juan Camilo, Luz Angela Rodriguez and Nancy Johnson (2011), Collective Actionfor Watershed Management: Field Experiments in Colombia and Kenya, Environment andDevelopment Economics, Vol.16: 275-303.

Cardenas, Juan Camilo and Elinor Ostrom (2004), What do people bring into the game: Exper-iments in the field about cooperation in the commons, CAPRi Working Paper No. 32.

Page 172: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

160

Carey, Janis M. and David Zilberman (2002), A Model of Investment under Uncertainty: Mod-ern Irrigation Technology and Emerging Markets in Water, American Journal of AgriculturalEconomics, Vol. 84 (1): 171-183.

Carpenter, Jeffrey, Stephen Burks and Eric Verhoogen (2005), Workers: The Effects of SocialFraming on Behavior in Distribution Games, Research in Experimental Economics, Vol 10:261-290.

Carpenter, Jeffrey, Cristina Connolly and Caitlin Knowles Myers (2008), Altruistic Behavior ina Representative Dictator Experiment, Experimental Economics, Vol. 11:282-298.

Chib, Siddhartha, Edward Greenberg and Yuxin Chen (1998), MCMC Methods for Fitting andComparing Multinomial Response Models, Working paper series, Washington University atSt. Lois

Congdon, Peter D. (2010), Applied Bayesian Hierarchical Methods, New York: Chapman andHall/CRC

Conlisk, John (1996), Why Bounded rationality, Journal of Economic Literature, Vol. 34(2), pp.669-700

Comite Tecnico de Aguas Subterrneas/Ojocaliente-Aguascalientes-Encarnacion (COTAS-OAE)(2006), Escenarios del Agua 2015 y 2030 en el Acuıfero Interestatal Ojocaliente-Aguascalientes-Encarnacion: acciones para un desarrollo con sostenibilidad ambiental, dowloaded fromhttp://www.cotas.org.mx/.

Coward, E.W. (1976), Principles in social organization in an indigenous irrigation system, Humanorganization, Vol. 38(1): 28-36.

Dayton-Johnson, Jeff (2000), Determinants of collective action on the local commons: a modelwith evidence from Mexico, Journal of Development Economics, Vol. 62, pp.181-208.

Dosi, Cesare and Michele Moretto (2010), Environmental Innovation, War of Attrition and In-vestment Grants, International Game Theory Review, Vol. 12 (1): 37-59.

El-Gamal, Mahmoud A. and David M. Grether (1995), Are people Bayesian? Uncovering behav-ioral strategies, Journal of the American Statistical Association, Vol. 90(432):1137-1145.

Ellison, Glenn (2006), Bounded Rationality in Industrial Organization, MIMEO

Food and Agriculture Organization (2003), Groundwater Management: The Search for PracticalApproachers, Water Reports, 25

Fischbacher, Urs, Simon Gachter and Ernst Fehr (2001), Are People Conditionally Cooperative?Evidence from a Public Good Experiment, Economic Letters, Vol. 71: 397-404

Frederick, Shane (2005), Cognitive Reflection and Decision Making, Journal of Economic Per-spectives, Vol. 19(4),pp:25-42

Fruwirth-Schantter, Sylvia (2006), Finite Mixture and Markov Switching Models, New York:Springer

Gelman, Andrew, John B. Carlin, Hal S. Stern and Donald B. Rubin (2009), Bayesian DataAnalaysis, Second Edition, New York: Chapman and Hall/CRC

Geweke, John, Michael Keane and David Runkle (1994), Alternative computational approachesto inference in the multinomial probit model, Review of Economics and Statistics, Vol. 76:609-632

Geweke, John and Michael Keane (2000), Bayesian Inference for Dynamic Discrete Choice ModelsWithout the Need for Dynamic Programming, in eds. Roberto Mariano, Til Schuermann andMelvyn J. Weeks, Simulation-Based Inference in Econometrics: Methods and Applications,Cambridge University Press

Page 173: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

161

Gardner, Roy, Elinor Ostrom and James M. Walker (1990), The Nature of Common-Pool Re-source Problems, Rationality and Society, Vol. 2(3):335-358.

Geweke, John and Michael Keane (2000), Bayesian Inference for Dynamic Discrete Choice ModelsWithout the Need for Dynamic Programming, in eds. Roberto Mariano, Til Schuermann andMelvyn J. Weeks, Simulation-Based Inference in Econometrics: Methods and Applications,Cambridge University Press

Gisser, Micha and David A. Sanchez (1980), Comptetion Versus Optimal Control in GroundwaterPumping, Water Resources Research, Vol. 16(4):638-642.

Gobierno del Estado de Aguascalientes (2009), Sistema de Indicadores Ambientales del Estado deAguascalientes, Aguascalientes, Mexico: Gobierno del Estado de Aguascalientes and Institutodel Medio Ambiente.

Goree, Jacon K and Charles Holt (1999), Stochastic game theory: For playing games, not forjust doing theory, Procedures of the National Academy of Sciences, Vol. 96, pp. 10564-10567

Greene, William H. and David A. Hensher (2003), A latent class model for discrete choice analysis:contrast with mixed logit, Transportation Research Part B, Vol.37:681-698

Gosttbauer, Elisabeth and Jeroen C.J.M. van den Bergh (2011), Environmental Policy TheoryGiven Bounded Rationality and Other-ragarding Preferences, Environmental and ResourceEconomics, Vol. 49(2): 263-304

Hackett, Steven, Edella Schlager and James Walker (1994), The Role of Communication in Re-solving Common Dilemmas: Experimental Evidence with Heterogenous Appropriators, Jour-nal of Environmental Economics and Management, Vol. 27: 99-126.

Hardin, Garrett (1968), The Tragedy of the Commons, Science, 162: 1243-1248.

Harstad, M. Ronald and Reinhard Selten (2013), Bounded-Rationality Models: Tasks to BecomeIntellectually Competitive, Journal of Economic Literature, Vol. 51(2): 496-511.

Heckman, J. and B. Singer (1984), A Method for Minimizing the Impact of Distributional As-sumptions in Econometric Models for Duration Data, Econometrica, Vol. 52(2):271-320.

Hess, Stephane, Moshe Ben-Akiva, Dinesh Gopinath and Joan Walker (2011), Advantages oflatent class over continuos mixture of Logit models

Houser, Daniel (2003), Bayesian analysis of a dynamic stochastic model of labor supply andsaving, Journal of Econometrics, Vol. 113, pp: 289-335.

Houser, Daniel, Michael Keane and Kevin McCabe (2004), Behavior in a dynamic decision prob-lem: An analysis of experimental evidence using Bayesian type classification algorithm, Econo-metrica, Vol. 72 (3): 781-822

Houser, Daniel and Joachim Winter (2004), How Do Behavioral Assumptions Affect StructuralInference? Evidence From a Laboratory Experiment, Journal of Business & Statistics, Vol.22(1), pp:64-79

Howard, Gregory and Brian E. Roe (2014), Stripping Because You Want to Versus StrippingBecause the Money is Good: A Latent Class Analysis of Farmers Preferences Regarding Fil-ter Strip Programs, Paper presented at the 2013 AAEA and CAES Joint Annual Meeting,Washington, DC August 4-6, 2013

International Food Policy Research Institute, Experimental Games for Strength-ening Collective Action, http://www.ifpri.org/book-6330/node/9632 andhttp://wle.cgiar.org/blog/2013/10/14/ifpris-water-games-to-strengthen-collective-action/

Page 174: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

162

Instituto Nacional de Estadıstica y Geografıa (2009), Estados Unidos Mexicanos, CensoAgropecuario 2007, VII Censo Agrıcola, Ganadero y Forestal, Mexico D.F.: INEGI.

International Water Management Institute (2010), Banking on Groundwater in Times of Change,Water Policy Brief, Issue 32

Janssen, Marco, John M. Anderies, J.C. Cardenas (2011), Head-enders as stationary banditsin asymmetric commons: Comparing irrigation experiments in the laboratory and the field,Ecological Economics, Vol. 70 (9): 1590-1598.

Jevons, W. S. (1865), The Coal Question, London: Macmillan and Co.

Kahneman, Daniel and Amos Tversky (1979), Prospect Theory: An Analysis of Decision underRisk, Ecnometrica, Vol. 47(2): 263-292

Koundouri, Phoebe (2000), Three Approaches to Measuring Natural Resource Scarcity: Theoryand Application to Groundwater, Doctoral Dissertation, University of Cambridge

Koundouri, Phoebe (2004), Current Issues in the Economics of Groundwater Resource Manage-ment, Journal of Economic Surveys, Vol 18(5): 703-740

Koundouri, Phoebe (2004), Potential for groundwater management: Gisser-Sanchez effect recon-sidered, Water Resources Research, Vol. 40, W06S16, doi:10.1029/2003WR002164.

Lee, Myoung-Jae (2005), Micro-Econometrics for Policy, Program and Treatment Effects, NewYork: Oxford University Press.

Lettau, Martin and Harald Uhlig (1999), Rules of Thumb versus Dynamic Programming, TheAmerican Economic Review, Vol. 89(1), pp:148-174

Lichtemberg, Erik (1989), Land Quality, Irrigation Development and Cropping Patterns on theNorthern High Planes, American Journal of Agricultural Economics, Vol. 71(1): 187-194.

Madani, Kaveh and Ariel Dinar (2012), Non-cooperative institutions for sustainable common poolresource management: Applications to groundwater, Ecological Economics, Vol. 74: 34-45.

Mani Anandi, Sendhill Mullainathan, Eldar Shafir and Jiaying Zhao (2013), Poverty ImpedesCognitive Function, Science, Vol. 341:976-980

McCulloch, Robert and Peter Rossi (1994), An exact likelihood analysis of the multinomial probitmodel, Journal of Econometrics, Vol. 64: 207-240.

McCulloch, Robert, Nicholas Polson and Peter Rossi (2000), A Bayesian Analysis of the multi-nomial probit model with fully identified parameters, Journal of Econometrics, Vol.99, pp:173-193

McGinnis, Michael D., ed. (2000), Polycentric Games and Institutions: Readings from the Work-shop in Political Theory and Policy Analysis, Ann Arbor, MI: The University of MichiganPress.

Moreno-Sanchez, Rocıo del Pilar and Jorge Higinio Maldonado (2010), Evaluating the role of co-management in improving governance of marine protected areas: An experimental approachin the Colombian Caribbean, Ecological Economics, Vol. 69: 2577-2567.

Moretto, Michele (2000), Irreversible investment with uncertainty and strategic behavior, Eco-nomic Modelling, Vol. 17: 589-617.

Nieswiadomy, Michael (1985), The Demand for Irrigation Water in the High Plains of Texas,1957-1980, American Journal of Agricultural Economics, Vol. 67(3): 619-626

Negri, Donald H. (1989), The Common Property Aquifer as a Differential Game, Water ResourcesResearch, Vol. 25 (1): 9-15.

Page 175: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

163

Nobile, Agostino (1998), A hybrid Markov chain for the Bayesian analysis of the multinomialprobit model, Statistics and Computing, Vol.8: 229-242

Ostrom, Elinor (1986), Issues of definition and theory: some conclusions and hypothesis, Proceed-ings of the conference on common property management, National Research Council, Wash-ington D.C.: National Academy Press.

Ostrom, Elinor (1990), Governing the Commons: the evolution of institutions for collective ac-tion, New York: Cambridge University Press.

Ostrom, Elinor (1992), Crafting institutions for self-governing irrigation systems, San Francisco,CA: Institute for Contemporary Studies University Press.

Ostrom, Elinor and Roy Gardner (1993), Coping with Asymmetries in the Commons: Self-Governing Irrigation Systems Can Work, The Journal of Economic Perspectives, Vol. 7(4):93-112.

Ostrom, Elinor (2002), Chapter 24: Common-Pool Resources and Institutions: Toward a Re-vised Theory, in Bruce Gardner L. and Gordon C. Rausser (eds.), Handbook of AgriculturalEconomics, Vol. 2, Amsterdam: Elsevier Science B.V.

Peterson, Jeffrey M. and Ya Ding (2005), Economic Adjustments to Groundwater Depletionin the High Planes: Do Water-Saving Irrigation Systems Save Water?, American Journal ofAgricultural Economics, Vol. 87(1): 147-159.

Pfeiffer, Lisa (2009), Three Essays on the Economics of Groundwater Extraction for Agriculture:Property Rights, Externalities and Policy, Doctoral Dissertation, University of California -Davis.

Provencher, Bill and Oscar Burt (1993), The Externalities Associated with the Common PropertyExploitation of Groundwater, Journal of Environmental Economics and Management, Vol. 24:139-158.

Rodrguez-Haros, Benito and Jacinta Palerm-Viqueira (2007), Before transference: Handing overwater districts, Agricultura, Sociedad y Desarrollo, Vol. 4 (2): 105-125.

Rubio, Santiago J. and Begona Casino (2003), Strategic Behavior and Efficiency in the CommonProperty Extraction of Groundwater, Environmental and Resource Economics, Vol.26(1): 73-87

Salcedo, Rodrigo (2014), Dynamic decision making in common-pool resource economic experi-ments: Behavioral heterogeneity in the field and the lab, LACEEP Working paper

Salcedo, Rodrigo and Miguel Angel Gutierrez (2014), Cooperation makes it happen? A ground-water economic framed experiment, LACEEP Working paper

Scheierling, Susanne M., Robert A. Young and Grant E. Cardon (2006), Public subsidies forwater-conserving irrigation investments: Hydrologic, agronomic, and economic assessment,Water Resources Research, Vol. 42(3), W03428.

Sekhri, Sheetal (2011), Public Provision and Protection of Natural Resources: GroundwaterIrrigation in Rural India, American Economic Journal: Applied Economics, Vol. 3: 29-55

Shen, Junyi (2009), Latent class model over mixed logit? A comparison by transport choice data,Applied Economics, Vol. 41: 2915-2924

Scherenberg Farret, Ericka, Patrick T. Brandt and Catherine Eckel (2014), A finite mixturemodel of behavior in a Three-Person Trust Game, MIMEO

Shogren, Jason F., Gregory M. Parkhurst and Prasenjit Banerjee (2010), Two Cheers and aQualm for Behavioral Economics, Environmental and Resource Economics, Vol. 46: 235-247

Page 176: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

164

Shogren, Jason F. and Laura O. Taylor (2008), On Behavioral-Environmental Economics, Reviewof Environmental Economics and Policy, Vol. 2(1): 26-44

Simon, Herbert A. (1955), A Behavioral Model of Rational Choice, Quarterly Journal of Eco-nomics, Vol. 69(1): 99-118

Spiegelhalter, David J., Nicola G. Best, Bradley P. Carlin and Angelika Van Der Linde (2002),Bayesian measures of model complexity and fit, Statistical Methodology, Vol. 64(4):583-639

Suleimain, R. and A. Rapoport (1997), Provision of step-level goods with continuous contribu-tions, Journal of Behavioral Decision Making, Vol 5: 133-153

Suter, Jordan F., Joshua M. Duke, Kent D. Messer and Holly A. Michael (2012), Behavior ina Spatially Explicit Groundwater Resource: Evidence from the Lab, American Journal ofAgricultural Economics, Vol. 94(5):1094-1112.

Swallow, Brent, Nancy Johnson, Rurth Meinzen-Dick and Anna Knox (2006), The challenges ofinclusive cross-scale collective action in watersheds, Water International, Vol. 31(3): 361-376.

Trawick, Paul (2003), Against the Privatization of Water: An Indigenous Model for ImprovingExisting Laws and Successfully Governing the Commons, World Development, Vol. 31(6):977-966.

Wang, Chenggang and Eduardo Segarra (2011), The Economics of Commonly Owned Groundwa-ter When User Demand Is Perfectly Inelastic, Journal of Agricultural and Resource Economics,Vol 36(1): 95-120

Ward, Frank A. and Manuel Pulido-Velazquez (2008), Water conservation in irrigation can in-crease water use, Proceedings of the National Academy of Science, Vol. 105(47): 18215-18220.

Worthington, Virginia E., Oscar R. Burt and Richard L. Brustkern (1985), Optimal Managementof a Confined Groundwater System, Journal of Environmental Econmics and Management,Vol. 12(3): 229-245

Page 177: GROUNDWATER GAMES: USERS’ BEHAVIOR IN COMMON-POOL RESOURCE

Vita

Rodrigo Salcedo Du Bois

Education

The Pennsylvania State University State College, Pennsylvania 2006–Present

Ph.D. in Agricultural, Environmental and Regional Economics

Universidad del Pacıfico Lima, Peru 1997–2002

B.S. in Economics

Awards and Honors

Latin American and Caribbean Environmental Economics Program (LACEEP) 2012Research Grant

Seminario Permanent de Investigacion Agraria 2005Research Fellow

European Association of Environmental and Resource Economists (EAERE),Fondazione Eni Enrico Mattei (FEEM) andVenice international University (VIU) 2012Grant to attend the European Summer School in Resource and EnvironmentalEconomics

Research Experience

Dissertation Research The Pennsylvania State University 2012–PresentResearch Advisor: Prof. James S. Shortle and Prof. David Abler

Graduate Research The Pennsylvania State University 2006–2012Supervisor: Prof. Jill L. Findeis

Undergraduate Research Universidad del Pacıfico 199?–199?

Working Experience

Research Assistant The Pennsylvania State University 2006–2012

Assistant Researcher Grupo de Analiss para el Desarrollo, Lima, Peru 2004-2006

Project Analyst Instituto Nacional de Investigacion Agraria, Lima, Peru 2003–2004

Project Analyst Ministerio de Agricultura, Lima, Peru 2001–2002