Iowa State University Capstones, Theses and Retrospective Theses and Dissertations Dissertations

2008 Exploring host- relationships through computer simulations of intracellular infection Garrett aM rc Dancik Iowa State University

Follow this and additional works at: https://lib.dr.iastate.edu/rtd Part of the Bioinformatics Commons

Recommended Citation Dancik, Garrett aM rc, "Exploring host-pathogen relationships through computer simulations of intracellular infection" (2008). Retrospective Theses and Dissertations. 15834. https://lib.dr.iastate.edu/rtd/15834

This Dissertation is brought to you for free and open access by the Iowa State University Capstones, Theses and Dissertations at Iowa State University Digital Repository. It has been accepted for inclusion in Retrospective Theses and Dissertations by an authorized administrator of Iowa State University Digital Repository. For more information, please contact [email protected]. Exploring host-pathogen relationships through computer simulations of

intracellular infection

by

Garrett Marc Dancik

A dissertation submitted to the graduate faculty

in partial fulfillment of the requirements for the degree of

DOCTOR OF PHILOSOPHY

Major: Bioinformatics and Computational Biology

Program of Study Committee: Karin Dorman, Co-major Professor Douglas Jones, Co-major Professor Alicia Carriquiry David Fern´andez-Baca Drena Dobbs

Iowa State University

Ames, Iowa

2008

Copyright c Garrett Marc Dancik, 2008. All rights reserved. 3316191

Copyright 2008 by Dancik, Garrett Marc

All rights reserved

3316191 2008 ii

DEDICATION

For Potamus iii

TABLE OF CONTENTS

LISTOFTABLES ...... vii

LISTOFFIGURES ...... viii

ACKNOWLEDGEMENTS...... x

ABSTRACT...... xi

CHAPTER1. Generalintroduction ...... 1

1.1 Thesisorganization...... 1

1.2 The immune response to intracellular pathogen ...... 2

1.2.1 Innateimmunity ...... 2

1.2.2 processing and presentation ...... 3

1.2.3 The adaptive immune response ...... 3

1.3 Leishmania infection ...... 5

1.3.1 Clinicaldiseases ...... 5

1.3.2 Transmission ...... 6

1.3.3 Immunology of Leishmania major ...... 7

1.3.4 Treatmentandvaccination ...... 8

1.4 Systems biology and computer modeling ...... 9

1.4.1 The systems biology approach ...... 9

1.4.2 Cellular automata and agent-based models ...... 10 iv

1.4.3 Simulationenvironments...... 12

1.5 Gaussian processes and computer model prediction ...... 14

1.5.1 Overview ...... 14

1.5.2 The Gaussian process as a computer model approximation...... 15

1.5.3 ExperimentalDesign...... 19

1.6 Gaussian process prediction and simulation-based based research ...... 21

CHAPTER 2. Parameter estimation and sensitivity analysis in an agent-

based model of Leishmania major infection...... 32

2.1 Introduction...... 33

2.2 An agent-based model of Leishmania major infection...... 36

2.2.1 Themodelenvironment ...... 37

2.2.2 Stagesofinfection ...... 37

2.3 Statisticalmethods...... 42

2.3.1 The Gaussian process as a computer model surrogate ...... 43

2.3.2 Gaussian process diagnostics ...... 46

2.3.3 Sensivity analysis using the Gaussian process ...... 46

2.3.4 Computer model calibration using field observations ...... 48

2.3.5 MarkovChainMonteCarlo ...... 50

2.3.6 Bayesianmodelcomparison ...... 51

2.4 Results...... 53

2.4.1 Experimental design and computer model output ...... 53

2.4.2 Fitting the Gaussian processes ...... 54

2.4.3 Sensitivityanalysis...... 55

2.4.4 Calibration using simulated data ...... 57

2.4.5 Calibrationusingfielddata ...... 58 v

2.4.6 Comparison of alternative models of behavior...... 59

2.5 Discussion...... 60

2.5.1 Gaussian process computer model approximation ...... 61

2.5.2 Sensitivityanalysis...... 62

2.5.3 Calibration using simulated data ...... 65

2.5.4 Calibrationusingfielddata ...... 66

2.5.5 Model comparison and macrophage emigration ...... 70

2.6 Conclusionsandfuturework ...... 72

2.7 Appendix: Agent-based model parameters ...... 73

CHAPTER 3. Identifying correlates of optimal immune escape strategies

and immune escape success in an in silico model of viral infection . . . . . 93

3.1 Introduction...... 94

3.2 MaterialsandMethods...... 96

3.2.1 Acomputermodelofviralinfection ...... 96

3.2.2 Computermodelanalysis ...... 99

3.3 Results...... 102

3.3.1 Effect of computer model parameters on simulation output ...... 102

3.3.2 Correlates of successful immune escape ...... 104

3.3.3 Simultaneous cellular and humoral escape ...... 107

3.3.4 Viral fitness and immune escape success ...... 108

3.4 Discussion...... 109

CHAPTER 4. mlegp: statistical analysis for computer models of biological

systemsusingR ...... 129

4.1 Introduction...... 130 vi

4.2 Statisticalmethods...... 131

4.2.1 TheGaussianprocess ...... 131

4.2.2 Sensitivityanalysis...... 132

4.3 Software...... 133

4.4 Example application: analysis of a computer model of disease ...... 133

4.5 SupplementaryMaterials ...... 136

4.5.1 Methods...... 136

4.5.2 ResultsandDiscussion...... 136

CHAPTER 5. General conclusions and future work ...... 142

5.1 Host-pathogen interactions and emergent behavior in intracellular infections . . 142

5.2 Futurework...... 143 vii

LIST OF TABLES

2.6 ABM parameters and values used in initial calibration ...... 74

2.1 Gaussianprocessdiagnostics...... 78

2.2 FANOVAdecomposition ...... 78

2.3 Posterior summary of computer model parameters following calibra-

tionusingsimulatedfielddata...... 78

2.4 Posterior summary of computer model parameters following calibra-

tionusingfielddata...... 82

2.5 Posterior summary of computer model parameters following calibra-

tionusingfielddata...... 82

2.7 Updated ranges for ABM parameters following initial calibration . . . 87

3.1 Subset of computer model parameters ...... 115

3.2 Proportion of escape mutants that fail to establish secondary

infection ...... 116

3.3 Consistency of optimal escape strategies for each pair of inj-

ectiontimes ...... 118

3.4 Effect of viral fitness on immune escape success ...... 119

4.1 FANOVA Decomposition of Gaussian process predictors for pathogen

loadatfiveand18dpi ...... 139 viii

LIST OF FIGURES

2.1 Simulated output over time and the field data used in calibration . . . 75

2.2 Main effects of computer model parameters on simulated pathogen

loadovertime...... 76

2.3 Gaussianprocessdiagnostics...... 77

2.4 Main effects of the most influential parameters ...... 79

2.5 Two-wayinteractionplots ...... 80

2.6 Posterior distributions of αI , kI , βMrecr, and Tthreshold following cali- brationusingsimulatedfielddata ...... 81

2.7 Posterior distributions of αI , kI , and Tthreshold following calibration usingfielddata ...... 83

2.8 Posterior predictive distribution of macrophage counts and parasite load 84

2.9 Posterior distribution of Anecr following calibration using field data . . 85 2.10 Main effects of computer model parameters on simulated macrophage

countsovertime...... 86

3.1 Subset of C-IMMSIM rules and parameters ...... 114

3.2 Simulation output clustered according to plasma viral load ...... 116

3.3 Classification tree for simulated plasma load profiles ...... 117 ix

3.4 Regression tree for overall relative humoral importance against com-

putermodelparameters ...... 118

3.5 Correlation of various immune response characteristics with peak plasma

humoral and cellular escape mutants over time ...... 119

3.6 Correlation of various immune response characteristics with dpeak = peak plasma viral load (CEM) - peak plasma viral load (HEM) . . . . 120

3.7 Classification tree for optimal immune escape strategy using overall

relative importance of the humoral response ...... 121

3.8 Peak plasma viral load for cellular, humoral and dual escape mutants

when injected at tpeak ...... 122 3.9 Cellular and humoral escape mutants have synergistic effects...... 123

3.10 Humoral escape mutants that are less infectious are more successful

than escape mutants with the same infectivity as wild type virus. . . . 124

4.1 Sensitivity analysis of a computer model of parasitic infection using

the R package mlegp ...... 132

4.2 Gaussian process diagnostic plots ...... 138

4.3 Main effects of all parameters using the Guassian process predictors

for pathogen load at five and 18 dpi...... 138

4.4 Interaction effects of the most important interactions for pathogen load

atfiveand18dpi...... 140

4.5 Gaussian process (GP) diagnostic plots for pathogen for a GP with

constant nugget term and GP with user-specified diagonal nugget

matrix ...... 141 x

ACKNOWLEDGEMENTS

I would like to take this opportunity to express my thanks to those who helped me with various aspects of conducting research and the writing of this thesis. I am particularly grateful for the advice and encouragement of my co-major professors, Dr. Karin Dorman and Dr. Doug

Jones. Karin, your strong work ethic, attention to detail, and level of patience are admirable qualities that will not be forgotten. Doug, your interest in the Leishmania computer model and your willingness to discuss many modeling details was much appreciated. I would also like to thank my committee members: Dr Alicia Carriquiry, Dr. David Fern´andez-Baca, and Dr.

Drena Dobbs. I owe a special thanks to Dr. Max Morris for his advice regarding experimental design and the use of Gaussian processes to analyze computer models.

From Ramapo college, I am grateful to Dr. Paramjeet Bagga and Dr. Amruth Kumar.

Their wisdom, encouragement, and guidance were vital in preparing me for my graduate studies.

Although moving to Iowa resulted in family and many good friends being left behind,

I never felt too far from home. To my family, I greatly appreciate all of your love and encouragement. It was nice knowing that every week, at least someone would be calling in to make sure I was still alive. To my friends, your friendship and support does not go unnoticed, even (no, especially) when it comes from a thousand miles away. Finally, I owe a special debt of gratitude to my fianc´ee, Katie Skowronski, who despite putting up with me for the better part of the last 3 years, still wants to marry me in July. xi

ABSTRACT

Computer simulations of infectious disease allow for the identification and estimation of important pathogen and immune parameters, the validation of theoretical biological models with experimental data, and the characterization of the host-pathogen interactions that lead to emergent and sometimes counterintuitive behavior. This thesis describes the development, analysis, and calibration of a computer model of Leishmania major infection, the identification of correlates of escape mutant success and optimal escape strategies in a computer model of a viral infection, and statistical software to aid in computer model analysis and calibration.

In an agent-based model of L. major infection, sensitivity analysis reveals that increasing growth rates can favor or suppress parasite load, depending on the stage of the infection and the ability of the pathogen to avoid detection. Calibration of the computer model suggests that the pathogen has a relatively slow growth rate and can grow for an extended time before damaging the host cell.

In a computer model of viral infection, we find that the relative overall importance of the cellular (or humoral) response consistently correlates with both the success of immune escape and the optimal escape strategy, and that correlation is relatively robust to the time the escape mutant arises. Mutants that simultaneously escape both responses perform substan- tially better than humoral or cellular escape mutants alone, highlighting the importance of both responses in controlling infection. Interestingly, loss of infectiousness of humoral escape mutants favors the virus, likely because decreasing infectivity weakens the cellular response. xii

Finally, Gaussian processes (GP) are commonly used as fast predictors of computer model output and are an essential part of the calibration and analyis of time-consuming computer models. We describe the R package mlegp, which fits GPs to scalar or multivariate computer model output and performs sensitivity analysis to identify and characterize the effects of important model parameters. 1

CHAPTER 1. General introduction

1.1 Thesis organization

This thesis describes the development, analysis, and calibration of a computer model of Leishmania major infection, the identification of correlates of escape mutant success and optimal escape strategies in a computer model of a viral infection, and statistical software to aid in the analysis and calibration of computer models. The thesis is presented in journal paper format.

Chapter 1 is an introductory chapter that provides background in immunology, systems bi- ology and computer modeling, and Gaussian processes which are commonly used for computer model calibration. The purpose of the introduction is to present the biological, computational, and statistical concepts in a broader context than what is presented in the individual jour- nal articles. Chapter 2 describes the analysis and calibration of an agent-based model of

L. major infection. Analysis of the model indicates that increasing pathogen growth rate can favor or hinder pathogen survival, depending on the stage of the infection and the ability of the pathogen to avoid detection. Calibration of the model indicates that L. major can repli- cate extensively before destroying the host cell. This chapter will be submitted for publication in the Journal of Theoretical Biology. Chapter 3 describes in silico experiments that identify the overall importance of the cellular (or humoral) arm of the immune response as a consis- tent correlate of optimal immune escape strategy and escape mutant success. We also find 2 a synergestic relationship between cellular and humoral escape, and that counterintuitively, humoral escape mutants benefit from a loss in infectivity. This work will also be submitted to the Journal of Theoretical Biology. Chapter 4 describes publicly available statistical software for the analysis of computer models. The chapter has been accepted for publication, subject to minor revisions to the software, in Bioinformatics as an Application Note. Finally, in chap- ter 5 we end with some general conclusions about understanding host-pathogen interactions through computer modeling and briefly describe future work.

1.2 The immune response to intracellular pathogen

Adaptive immunity involves the proliferation and differentiation of antigen-specific lym- phocytes (T cells and B cells) into effector cells that (hopefully) eliminate the pathogen from the host. The cellular arm, which consists of cytotoxic T lymphocytes, recognizes and elimi- nates infected cells, while the humoral arm, which consists of B cells and , recognizes and eliminates extracellular pathogen (Janeway et al., 2005). In this section we provide a brief description of the immune response as it relates to the current work.

1.2.1 Innate immunity

The innate immune response is what first recognizes that pathogen is present and plays an important role in initiating the adaptive immune response. Pattern-recognition receptors, which exist as free proteins and on phagocytic cells, recognize features that are common to and alert the host that pathogen is present. For example, Toll-like receptor 4

(TLR-4), in association with the receptor CD14, recognize bacterial lipopolysaccharide (LPS), while TLR-5 recognizes flagellin, a protein necessary for bacterial motility (Du et al., 1999;

Ivison et al., 2007). Pathogen recognition initiates an inflammatory response that attracts 3 additional immune cells to the infected area, and causes phagocytes to express co-stimulatory molecules which are required for the initiation of adaptive immunity.

1.2.2 Antigen processing and presentation

For certain pathogens, such as , pathogen replication requires the host’s machinery and protein synthesis occurs in the cytoplasm of an infected cell. Other pathogens, such as Leishmania major and Mycobacterium tuberculosis, survive and reproduce in intracellular vesicles in . In both cases, antigen (processed pathogen) is presented on major histocompatibility complex (MHC) molecules on the surface of the cell. Proteins degraded in the cytosol are presented on MHC class I molecules, signalling to cytotoxic T lymphocytes

(CTLs) that a cell is infected. Proteins degraded in intracellular vesicles are presented on

MHC class II molecules, signifying that a pathogen is present. Dendritic cells and macrophages take up pathogen in a non-specific manner through phagocytosis and macropinocytosis, and present antigen on MHC class I or class II complexes. B cells are antigen-specific, recognize specific on the pathogen’s surface, and internalize and present antigen on MHC class II molecules. Dendritic cells, macrophages and B cells are collectively known as antigen presenting cells (APCs). After taking up pathogen, APCs migrate to the draining lymph node where they present antigen to na¨ıve T cells. Antigen presentation in the presence of requisite co-stimulatory signals then initiates the adaptive immune response.

1.2.3 The adaptive immune response

A na¨ıve that recognizes a MHC:peptide complex on an APC in the presence of the required co-stimulatory molecules undergoes clonal expansion and differentiation into an effector cell type. The expanding T cell population has the same antigen specificity as the na¨ıve T cell; the immune response that develops is therefore specific for a particular antigen. 4

1.2.3.1 Cellular immunity

Cytotoxic T cells (CTLs), or CD8+ T cells, target cells presenting antigen on MHC class

I molecules. Binding between a CTL receptor and a MHCI:peptide complex triggers lysis of

(Lieberman, 2003) or (Barry & Bleackley, 2002) in the infected cell.

CD4+ T cells, which primarily differentiate into Th1 or Th2 cells, target cells present- ing antigen on MHC class II molecules. Binding between a CD4+ T cell receptor with a

MHCII:peptide complex triggers the release of effector molecules whose function depends on the phenotype of the CD4+ T cell. Th1 and Th2 cells are characterized by the they produce (Mossmann & Coffman, 1989). Th1 cells, which are primarily associated with cell- mediated immunity and the production of opsonizing antibdodies that promote phagocytosis of -antgen complexes, activate infected macrophages through production of IFN-γ.

Once activated, macrophages efficiently fuse their lysosomes with their phagosomes, pro- duce TNF-α, which synergizes with IFN-γ and induces production of nitric oxide (NO), making the macrophage efficient at eliminating intracellular pathogen. However, the nox- ious chemicals produced by activated macrophages damage nearby tissue and extracellular structures, and the process must be carefully regulated (Duffield, 2003).

1.2.3.2 Humoral immunity

Th2 cells, which are primarily associated with humoral or antibody-mediated immunity, release cytokines such as IL-4 that activate B cells. Activated B cells proliferate and differ- entiate into antibody-secreting plasma B cells. Antibodies recognize extracellular pathogens of the same specificity as the B cells they derive from. Antibodies help clear the pathogen in a variety of ways. For example, neutralizing antibodies prevent viruses from binding to host cell receptors and infecting new cells (Shibata et al., 1999) and opsonizing antibodies 5 promote phagocytosis by macrophages or (Gounni et al., 1994). Antibodies can also trigger the classical pathway of complement activation which results in a biochemical cascade ending with the lysis of the antibody-coated pathogen (Cooper, 1984).

1.3 Leishmania infection

Leishmania are obligate, intracellular protozoan parasites that are transmitted by the bites of infected female sand flies of the genus Phlebotomus in the Old World (i.e., Europe, Asia, and

Africa). In the New World (i.e., the Americas and Australia), the sand fly vector is primarily of the genus Lutzomyia. Over 20 species of Leishmania can cause disease, which is usually either cutaneous (reviewed in Reithinger et al., 2008) or visceral (reviewed in Chappuis et al.,

2007).

1.3.1 Clinical diseases

Cutaneous disease manifests as localized and spontaneously-healing or diffuse skin lesions.

Visceral disease is systemic and is almost always fatal if left untreated. The World Health

Organization (WHO) estimates that there are 1.5 million new cases of cutaneous Leishmania each year and 0.5 million new cases of visceral disease 1. Localized cutaneous disease is caused by L. major and L. tropica in the Old World and by L. braziliensis and L. mexicana in the

New World. Diffuse cutaneous disease is caused by L. aethiopica and L. amazonensis. Vis- ceral Leishmania, or Kala-azar, is caused by L. donovani or L. infantum. Additional clinical manifestations include leishmaniasis recidivans, or recurrent leishmaniasis, which is primarily caused by L. topica; post-kala-azar dermal leishmaniasis, caused by L. donovani; mucocu- taneous leishmaniasis, caused primarily by L. braziliensis; and viscerotropic leishmaniasis, caused by L. tropica (Choi & Lerner, 2002).

1http://www.who.int/tdr/disease/leish/diseaseinfo.htm 6

Since the mid-1990’s, leishmaniasis has emerged as an increasingly frequent opportunistic infection in patients infected with human immunodeficiency virus (HIV) and suffering from acquired immune deficiency syndrome (AIDS). Leishmania and HIV co-infections typically involve L. infantum, which causes visceral disease. However, co-infection with species normally associated with cutaneous disease often results in visceral disease in an immunocompromised host. Species involved in these co-infections include L. braziliensis (Hern´andez et al., 1995),

L. major (Gillis et al., 1995), L. tropica (Magill et al., 1993), and L. aethiopica (Berhe et al.,

1995). According to the WHO, up to 70% of visceral disease in adults are associated with

HIV infection in southern Europe (Desjeux et al., 1999). In regions in Spain and Portugal, visceral leishmaniasis is the most common opportunistic infection after Toxoplasma gondii and Cryptosporidium parvum (Alvar et al., 1997).

1.3.2 Transmission

While inside an infected sand fly, Leishmania exist as extracellular flagellated promastigote forms which are transmitted via saliva to a mammalian host during a bloodmeal. Once inside the mammalian host, promastigotes are taken up by macrophages and quickly change into a non-motile amastigote form (Awasthi et al., 2004) that survives and proliferates in the phagolysosome of the macrophage. The transmission cycle is completed when a sand fly ingests infected macrophages from a mammalian host during a blood meal. Amastigotes transform into non-infectious promastigotes and attach to the sand fly midgut, where they differentiate into free-swimming, infectious promastigotes, ready to be transmitted to a new host (Kamhawi et al., 2004).

A variety of mammalian hosts, including humans, rodents, mustelids, hedgehogs, and rabbits, support Leishmania infection and transmission. These mammalian reservoirs are necessary for the parasite’s survival. Reservoirs include canids for L. tropica; the forest 7 rodent for L. mexicana; and rodents and marsupials for L. amazonensis (Awasthi et al.,

2004). Dogs are an important reservoir for the visceral disease-causing L. donovani and L. infantum (Vanloubbeeck & Jones, 2004).

1.3.3 Immunology of Leishmania major

In murine models of L. major infection, host resistance or susceptibility depends on the genetic background of the host. Infection in the majority of inbred strains, including C3H/He,

CBA, and C57BL/6, results in cutaneous lesions that spontaneously resolve, and the host acquires immunity to reinfection. Inbred strains such as BALB/c, however, fail to control infection and develop chronic lesions and systemic disease. Differences between strains is generally attributed to the expansion of CD4+ T helper 1 (Th1) cells in resistant mice and

CD4+ T helper 2 (Th2) cells in susceptible mice (reviewed in von Stebut & Udey, 2004; Gumy et al., 2004).

Dendritic cells (DCs) are the primary antigen-presenting cells in L. major infection (Moll et al., 1993). The fact that they can phagocytize amastigotes, but not promastigotes (the form transmitted from the sand fly), may explain the delay before the adaptive immune response develops (von Stebut et al., 1998). Antigen-processing and presentation of L. major are similar in susceptible and resistance mice (von Stebut et al., 2000). However, the ability of DCs to drive polarization toward a Th1 or Th2 phenotype is strain specific (Filippi et al.,

2003). DCs release the interleukin-12 (IL12; von Stebut et al., 1998) and are believed to be the source of IL-12 early in infection. IL-12 induces the Th1 response and is required to initiate protective immunity (Filippi et al., 2003). In susceptible mice with Th2 phenotypes, exogenous IL-12 has been shown to promote a Th1 response which leads to resistance under certain conditions (Heinzel et al., 1993; Sypek et al., 1993).

In resistant mice, CD4+ Th1 cells release interferon-γ (IFN-γ) which activates macrophages 8 via the inducible nitric oxide synthase pathway, killing intracellular parasite (Bodgan et al.,

2000). A functional Fas-FasL pathway, which enables Th1 CD4+ and CD8+ cytotoxic cells to trigger apoptosis in infected cells, is also required for healing (Con¸ceicao-Silva et al., 1998;

Huang et al., 1998). CD8+ T cells are not believed to have a primary role in healing typical high dose experimental infections, but are required for maintaining host immunity to rein- fection (Muller et al., 1993) and for controlling low dose intradermal challenge that mimics natural transmission (Belkaid et al., 2002).

In susceptible mice, the development of a Th2 response is dependent on IL-4, as mice treated with anti-IL-4 antibodies develop a Th1 response and are resistant to infection (Sadick et al., 1990). However, IL-4 is not sufficient for suppressing the Th1 response, as otherwise resistant mice that constitutively express IL-4 develop a relatively strong Th1 response but become susceptible to disease (Erb et al., 1996).

In resistant mice, low levels of pathogen persist following resolution of the infection

(Belkaid et al., 2000), and the host maintains protective immunity to reinfection. How- ever, reactivation of disease becomes possible (Mendez et al., 2004). Pathogen persistence is maintained by a pool of natural regulatory T cells; when this pool is depleted, sterile cure

(i.e., complete pathogen elimination) is achieved, and immunity to reinfection is lost. In the presence of L. major infected-macrophages, regulatory T cells suppress IFN-γ production by CD4+ Th1 cells, and promote parasite survival through both IL-10-dependent and inde- pendent mechanisms (Belkaid et al., 2003). Protective immunity also requires IL-12, whose persistent source is not clear (Scott, 2003).

1.3.4 Treatment and

Pentavalent antimonials are commonly used to treat all forms of Leishmania, although drug resistant forms of Leishmania are becoming increasingly common (Croft & Coombs, 2003; Lira 9 et al., 1999). Other treatments include amphotericin B and pentamidine. Immunotherapies, such as transfer of antigen-pulsed DCs, have also been used successfully in mice (von Stebut et al., 2000).

Although no are available against any parasitic infections, it is generally believed that successful Leishmania vaccination is feasible. Namely, because self-healing cu- taneous forms of Leishmania confer immunity to reinfection, vaccination with live, virulent

Leishmania is possible. The practice, known as ‘leishmanisation’, has been performed for over a century (Greenblatt, 1988) and has been used extensively in the former Soviet Union, Israel, and Iran (Higashi, 1988). Leishmanisation is the only proven effective vaccination strategy in humans (Tabbara et al., 2005). However, because leishmanisation can result in chronic lesions that do not heal without medical treatment, the practice has declined in recent years.

Various vaccination strategies are being examined, including killed vaccines, live-attenuated vaccines, molecularly defined subunit vaccines, vaccinates with recombinant DNA-derived pro- teins, DNA vaccines, poly-protein vaccines, and vaccines against sand fly components. For a recent review of Leishmania vaccines, see Kedzierski et al. (2006). It is generally believed that a successful vaccine against cutaneous Leishmania will pave the way for a vaccine against visceral disease. However, several scientific issues will need to be considered, including dif- ferences between murine and human immune responses to Leishmania infection, and the fact that infection outcome and immunopathology is strain-specific and may depend on the site of infection. Also problematic are the political challenges of introducing vaccines into devel- oping countries (Clemens & Jodar, 2005) and the fact that there is little financial incentive to develop vaccines for use in the Third World (Plotkin, 2005). 10

1.4 Systems biology and computer modeling

1.4.1 The systems biology approach

A common expression is that ‘the whole (of a system) is greater than the sum of its parts’.

More formally, the emergent behavior of a system cannot be understood by examining its components in isolation; it is also necessary to consider how the components of a system in- teract. Systems biology involves the study of biological processes at the holistic (i.e., systems) level by explicitly considering interactions between system components (e.g., genes, proteins, cells).

There are two broad areas within the field of systems biology: knowledge discovery, or data mining, involves inferring relationships and structure within large experimental data sets; simulation-based research involves identifying important system parameters and vali- dating complex theoretical models with experimental data using mathematical or computer models. An in silico model whose output is not consistent with biological observations indi- cates that there is incomplete understanding of one or more of the underlying components of the system, or that important elements are being excluded. In general, systems biology is used to generate predictions that can be tested through bench experiments (Kitano, 2002).

Knowledge discovery has been used to infer regulatory networks using protein-protein inter- action data, DNA-protein interaction data, and mRNA expression data (Ideker et al., 2002).

An example of simulation-based research includes mathematical modeling of HIV infection that led to the finding that HIV replicates at a rapid rate (Ho et al., 1995; Wei et al., 1995).

Computer models of disease help facilitate understanding of complex host-pathogen in- teractions, adverse and inefficient immune responses, and are being used as tools in the drug discovery process (Butcher et al., 2004). Recent computer models have been used to study granuloma formation in Mycobacterium tuberculosis, (Segovia-Juarez et al., 2004), systemic in- 11

flammatory response and multiple organ failure (An, 2001), influenza A infection (Beauchemin et al., 2005), and antigenic escape mutants in HIV infection (Bernaschi and Castiglione, 2002).

1.4.2 Cellular automata and agent-based models

A cellular automata (CA) model consist of a grid of cells. Each cell is in one of a finite number of states. The states of all cells are updated at discrete time steps based on the states of neighboring cells (Codd, 1968). The most well-known CA is John Conway’s game of life.

In this CA, a cell is either alive or dead; if a living cell has more than 3 or less than two living neighbors, the cell dies (due to overcrowding or loneliness); and a dead cell with three living neighbors comes alive (Gardner, 1970). Despite the simplicity of the rules, Conway’s game of life can be used to construct a universal turing machine (Rendell, 2001) and can produce emergent patterns including still-lifes (i.e., stable patterns that do not change over time) and oscillations. CA models have been used to study a variety of biological processes

(see Ermentrout & Edelstein-Keshet (1993) for an early review), including DNA sequence evolution (Sirakoulisa et al., 2003), and cell, tissue, and tumor growth (Deutsch & Dormann,

2004).

An agent-based model contains distinct entities, or agents, that inhabit a spatial environ- ment. A simulation visualizes agents as they move and interact according to update rules that are executed at discrete time steps (Wooldridge, 2002). ABMs inherently capture the dynamics of complex systems whose properties depend on the collective behavior of the sys- tem’s interacting components. The ABM modeling choice is particularly appropriate when interactions are non-linear (e.g. are threshold-dependent), behavior is stochastic, agents are heterogenous, or learning and adaptation occurs (Bonabeau, 2002). An example of an ABM is given in Segovia-Juarez et al. (2004), who use an ABM to study Mycobacterium tuberculosis infection. The environment consists of a square grid of microcompartments, agents consist of 12 macrophages and CD4+ Th1 cells, and each CD4+ Th1 cell that contacts (i.e., neighbors) an infected macrophage independently activates it with a fixed probability.

ABMs have been used in a variety of scientific disciplines, including ecology, where they are referred to as individual-based models (Grimm & Railsback, 2005), in the social sciences

(Gilbert, 2007), and in computational economics (Tesfatsion, 2006). More specifically, ABMs of biological systems have been used to identify important system parameters in a model of angiogenic sprouting (Bentley et al., 2008) and to validate a model of the intracellular NF-κB pathway (Pogson et al., 2002).

ABMs have also been used to study the and various diseases. In their model of Mycobacterium tuberculosis infection, Segovia-Juarez et al. (2004) identify chemokine diffusion rates and the arrival time, location, and macrophage activation efficiency of T cells as important factors in granuloma formation. In a model of systemic inflammatory response and multiple organ failure, An (2002) reproduces outcomes of unsuccessful clinical trials involving blockage of proinflammatory mediators. In a model of influenza A infection, Beauchemin

(2006) show how infection dynamics depend on the spatial structure of initially infected cells.

Finally, in a model of antigen escape in HIV infection, Bernaschi & Castiglione (2002) find that escape mutants with low transcription rate can explain the long-term asymptomatic phase of disease. For a recent review of computational immune system models, see Forrest &

Beauchemin (2007).

The difference between a CA model and an ABM are in some ways semantic. If each complex arrangement of agents in an ABM can be regarded as a state in a CA, and the total number of possible states is finite, then there is a computational equivalence between the two types of models. Therefore, although the immune system simulator IMMSIM (see below) is often described as a modified CA model, it is probably more natural to consider IMMSIM within the ABM framework. 13

1.4.3 Simulation environments

1.4.3.1 General overview

Although most computational models are developed to be application (e.g., disease) spe- cific, some effort has been made to create generalized modeling environments which are easily extendable. In social science and economics, common ABM environments include Swarm

(Minar et al., 1996), Repast (North & Collier, 2006), Ascape (Inchiosa & Parker, 2002), and

NetLogo (Tisue & Wilensky, 2004). ABM modeling environments tailored for research in systems biology also exist. CyCells is a three-dimensional ABM designed for studying in- tercellular interactions (Warrender, 2004; Forrest & Beauchemin, 2007) and has been used to compare two hypothesis regarding macrophage proliferation and migration into the lung

(Warrender et al., 2004). Simulation, analysis and modeling (SAAM) software, described in

Barrett et al. (1998), provides an environment for pharmokinetic simulations. SAAM has been used to study the clearance kinetics of intrathecal opiod injections (Ummenhofer et al.,

2000) and to identify disturbed metabolism kinetics in men with visceral obesity (Chan et al.,

2002).

1.4.3.2 Immune system simulators

There are several modeling environments available for in silico immunological research.

Most notable is the immune system simulator (IMMSIM; Seiden & Celada, 1992), which is the basis of the model we use in Chapter 3. In IMMSIM, lymphoctye receptors and their ligands (i.e., MHC:peptide complexes) are represented by binary bit strings. Interactions (i.e., binding events) occur probabilistically based on the number of complementary bits between receptor and ligand. Clonal expansion occurs following binding of a lymhocyte receptor and

MHC:peptide complex. Agents in the original version of IMMSIM include B cells, antibodies, 14 helper T cells, and general antigen-presenting cells. Immunological memory in B cells is also modeled. Through a thymus selection process, helper T cells undergo positive and negative selection based on their affinity for self-peptides, which are specified as bit strings. Seiden &

Celada (1992) validate their model by carrying out several in silico experiments. Specifically,

IMMSIM reproduces several known features of the immune response: after initial injection with antigen, the immune response efficiently controls a secondary challenge; anergy of B cells mitigates autoimmune responses when large amounts of self-antigen are present, and autoimmune responsiveness decreases as thymus selection efficiency increases.

IMMSIM has undergone several updates since the initial version, starting with the addition of CTLs to model the cellular response (Puzone et al., 2002). Although the first version to con- tain CTLs is coded in the relatively uncommon IBM APL2 language, various and more flexible versions are now available (http://www.cs.princeton.edu/immsim/software.html). These in- clude a C-language version (IMMSIM-C), an educational version (IMMSIM++) that includes a graphical interface, a second C-language version (C-IMMSIM) that is optimized for larger scale simulations, and a parallel version of C-IMMSIM (PARIMM) (Bernaschi & Castiglione,

2001). More recent extensions to the C-IMMSIM version, which is the version we use in chapter 3, allow for modeling HIV infection and tumor growth. Using C-IMMSIM to study

HIV, Bernaschi & Castiglione (2002) propose that viral escape mutants with low transcription rate explains the asymptomatic phase of disease; Castiglione et al. (2004) identify cytotoxic

T lymphoctye escape as a predictive marker of disease progression.

Additional immune system modeling environments include the synthetic immune system

(SIS; Mata & Cohn, 2007), the Basic Immune Simulator (BIS; Folcik et al., 2007), and SIM-

MUNE (Meier-Schellersheim, 1999). SIS is suited for theoretical research in Protecton The- ory (Cohn & Langman, 1990), where emphasis is on discrimination between self and non-self antigen and specificity is an all-or-none phenomena. BIS includes both innate and adaptive 15 immune responses, and is designed for research into how the innate immune response directs the development of the adaptive immune response. SIMMUNE appears similar to IMMSIM but is not publicly available (Mata & Cohn, 2007).

1.5 Gaussian processes and computer model prediction

1.5.1 Overview

Computer model calibration is the process of estimating computer model parameters so that output from the calibrated computer model is consistent with field data (Kennedy &

O’Hagan, 2001). The parameters to be estimated are referred to as calibration (or tuning) parameters. The computer model may also contain parameters controlled by an experimenter in the physical process being simulated; these parameters are referred to as input parameters.

Because models are often high-dimensional, time-consuming, and non-linear, ad hoc methods of computer model calibration are traditionally used. However, statistically precise model calibration is a fundamental and necessary step for making meaningful inference on model parameters, for assessing the validity of a model, and for performing model comparisons

Bayarri et al. (2002). This process generally entails 1) approximation of the computer code through a surrogate statistical model (i.e., a Gaussian process), 2) definition of a relationship between the computer model output and data that is observed in the field, and 3) estimation of computer model parameters in a Bayesian fashion.

A computer model approximation is needed because computer models are usually expen- sive (i.e., time consuming), and the parameter estimation step requires that the computer model be evaluated at a large number of parameter values as the parameter space is explored.

The Gaussian process quickly predicts computer model output by interpolating and smooth- ing a known data set. The Gaussian process approximation is described in section 1.5.2. An 16 important consideration in Gaussian process prediction is the choice of computer model pa- rameters (i.e., experimental design) used to generate the known data set, and this is discussed in section 1.5.3. The book by Santner et al. (2003) is an excellent source for both Gaussian process models and experimental design. More details on computer model calibration are described in section 2.3.4. For a recent review of computer model calibration and validation, see Bayarri et al. (2007).

1.5.2 The Gaussian process as a computer model approximation

Gaussian processes (GPs) are flexible and useful tools for approximating computer models when the computer model response surface of interest is a smooth function of the parameter space (Sacks et al., 1989; Kennedy & O’Hagan, 2001; Bayarri et al., 2002). The computer model, denoted as z(θ), is considered a random function that produces scalar output when

2 evaluated at the computer model parameter vector θ = (θ1, . . . , θp) . In the GP framework, a collection of computer model outputs is assumed to be multivariate normal with a correlation structure that depends on θ. A GP is specified by a mean and covariance function.

The mean function is commonly taken to be a linear regression model of the form

k E [z(θ)] = fj(θ)λj, (1.1) jX=1

′ where f1(·),...,fk(·) are known regression functions and λ = (λ1,...,λk) is a vector of regression coefficients. The variance-covariance function is usually specified in terms of a correlation function. In a stationary GP, the correlation between two computer model outputs depends only on the distance between their corresponding parameter vectors, independent of

2We choose to denote the computer model parameter vector by θ to emphasize the fact that we are exclu- sively dealing with calibration parameters. However, the GP model is not sensitive to the type of parameters represented by θ. 17 where the vectors are in the parameter space. Mathematically, stationary GPs have correlation functions that satisfy

cor z (θ) ,z θ′ = R θ, θ′ = R θ − θ′ .     The correlation function R must be positive definite with R(θ, θ) = 1. GPs that also have stationary mean functions (i.e., with E[z(·)] = λ0) are considered second-order stationary. Stationary correlation functions are commonly assumed. A notable exception is found in the work of Gramacy (2005), who explores non-stationary models by fitting separate GPs to each region of a partitioned parameter space.

A common family of correlation functions is the power exponential correlation function which has the form p R z(θ),z(θ′) = exp − β |θ − θ′ |αj . (1.2)  j j j  j=1  X   th With respect to the j component of the parameter vector, j = 1,...,p,βj > 0 determines how sharply correlation declines based on a measure of distance in the parameter space, and the power parameter αj ∈ (0, 2] controls the smoothness of the GP. When αj = 2 ∀j, sample paths (i.e., realizations) from the GP are infinitely differentiable, while sample paths are theoretically nondifferentiable when αj < 2 ∀j. A comparison between infinitely differentiable sample paths and theoretically nondifferentiable sample paths is given in Santner et al. (2003).

Power exponential correlation functions are convenient because they are separable, i.e., they are products of one-dimensional marginal correlation functions. The special case where αj = 2 ∀j is known as the Gaussian correlation function, and is the most popular correlation function for GP predictors of computer model output (Santner et al., 2003). Other families of correlation functions include the Mat´ern (Mat´ern, 1960, 1986) and cubic (Currin et al., 1989) correlation functions. 18

For deterministic computer models, the covariance function has the form

′ 2 ′ cov z (θ) ,z θ = σGPR z (θ) ,z θ ,   

2 where σGP is the unconditional variance var[z(θ)] and R is a correlation function. Let zknown = z θ(1),...,z(θ(m) be a collection of m computer model outputs. We will also use R to   denote the correlation matrix with element (i, j) equal to cor z(θ(i)),z(θ(j)) . For stochastic

2   computer models, a nugget term σe is added to the diagonal of the variance-covariance matrix, so that the variance-covariance matrix has the form

2 2 σGPR + σe I,

2 The matrix R is traditionally still referred to as a correlation matrix, even though for σe > 0 elements of R are no longer correlations.

Inferences of computer model output z(t) at parameter vector t, conditioned on zknown, follow from standard multivariate normal distribution theory (Rencher, 2002). We describe

GP prediction using a fitted GP that uses maximum likelihood estimates of GP parameters.

However, a Bayesian treatment of GP parameters is also possible, in which case the posterior predictive distribution of z(t) will depend on the GP parameter priors - for example, see

Santner et al. (2003). We present GP prediction and parameter estimation for the case where

2 there is no nugget term (i.e., σe = 0). (i) Define F to be an m × k matrix with element (i, j) equal to fj(θ ), i = 1,...,m,j =

1, . . . , k. Then F λ is the unconditional mean matrix of zknown, for mean function of Equation

2 (1.1). Furthermore, zknown ∼ Nm Fλ,σGPR . Also define ft = (f1(t),...,fk(t))) so that  E[z(t)] = ftλ, and define r(t) to be an m × 1 vector of correlations with element i equal to 19 cor z (t) ,z θ(i) .    Conditional on zknown, computer model output at parameter vector t has a predictive distribution that is normal with mean

′ ′ −1 zˆ(t)=E[z(t)|zknown]= ftλ + r (t)R (zknown − F λ) (1.3) and variance

2 ′ −1 var[z(t)|zknown]= σGP 1 − r (t)R r(t) , (1.4) h i 2 where we currently assume that the GP parameters (α,β,λ,σGP) are known. With no dis- tributional assumptions on zknown, the predictorz ˆ(t) minimizes the expected mean squared prediction error (MSPE) of z(t).

In practice, none of the GP parameters are known and therefore must be estimated.

However, if the correlation parameters (α,β) are known, then closed form maximum likelihood

2 estimates (MLEs) exist for λ and σGP and are

′ −1 −1 ′ −1 λ = (F R F ) F R zknown (1.5)

b and 1 σ2 = (z − F λ)′R−1(z − F λ) (1.6) GP m known known d b b Equation 1.5 is the generalized least squares estimate of λ, and replacing λ with λ in Equation

1.3 yields the best linear unbiased predictor of z(t). This is true regardless of theb distribution of zknown. In combination with Eqs (1.5) and (1.6), numerical methods (e.g., simplex and gradient-based methods) are used to explore the GP correlation parameters in order to find 20

MLEs of all GP parameters. The empirical best linear unbiased predictor (EBLUP) of z(t) is

′ ′ −1 zˆ(t)= ftλ + r (t)R (zknown − F λ),

b b b b where r(t) and R are obtained using the MLEs (α, β) for the GP correlation parameters (α,β).

Note thatb the termb EBLUP is a misnomer becauseb bα and β are not linear estimates in zknown.

b b 1.5.3 Experimental Design

An important consideration in GP prediction is the choice of the parameter vectors

(1) (m) θ , . . . , θ (i.e., the experimental design) used to generate zknown. An overview of de-   signs for computer models is given by Santner et al. (2003) and Koehler & Owen (1996).

In this section we focus our attention on the maximin Latin hypercube design, which we use in chapters 2 and 4. Other designs include uniform designs (Fang et al., 2000) and designs optimized according to an integrated mean square error criterion (Sacks et al., 1992) or a minimax criterion (Johnson et al., 1990). Additional specialized Latin hypercube designs include orthogonal array-based Latin hypercube designs (Tang, 1993) and symmetric Latin hypercube designs (Ye et al., 2000).

1.5.3.1 Maximum entropy and maximin designs

Let D be the experimental design, and let zD correspond to all computer model outputs evaluated at parameter vectors not in D. Intuitively, we want to select D so that the uncer- tainty in zD|zknown is minimized (Currin et al., 1989). Formally, information (the negative measure of entropy) is used to quantify the change in information between zD and zD|zknown (Lindley, 1956). The goal is to maximize the expected change in information, which is equiv- alent to maximizing the entropy of zknown(Shewry & Wynn, 1987). The entropy of zknown is 21

maximized by maximizing det(V ), where V is the variance-covariance matrix of zknown. A geometrical interpretation of maximum entropy is described by Johnson et al. (1990), who show that for weak Gaussian correlation functions, a maximum entropy design maximizes

min θ(i) − θ(j) , (1.7) i

where θ(i) − θ(j) is the Euclidian distance between θ(i) and θ(j). If more than one design

maximizes Equation 1.7, the optimal design is the design with the fewest pairs of points with maximum minimum distance. Designs satisfying Equation (1.7) are called maximin designs, and designs based on entropy or maximin criteria are said to be space filling. Space-filling designs are intuitively appealing because uncertainty in a predictionz ˆ(t) (Equation 1.4) is a function of correlations between z(t) and all observations in zknown, and the correlation between two observations increases with the distance between the corresponding parameter vectors. A space filling design prevents t from being far away from points in D, which generally leads to smaller predictive variances.

1.5.3.2 Latin hypercube designs

If two design points θ and θ′ differ in only one parameter (i.e., a single component of θ), and that parameter has a trivial effect on the response, then θ′ cannot provide any meaningful information that would not be provided by θ. Therefore, it would not be desirable to evaluate the computer model at both θ and θ′, and the two design points are said to ‘collapse’ onto one another. A second desirable property in an experimental design for GP prediction is that the design should be non-collapsing. A common class of designs that have this non-collapsing property are Latin hypercube designs (LHDs) (McKay et al., 1979).

Generation of an LHD of size n on the parameter space Θ = (Θ1,..., Θp), where each 22

component Θj is independent, involves partitioning each component of the parameter space into m distinct regions, Θj = (Θj1,..., Θjm). Let [D]ij ,i = 1,...m, j = 1,...,p, be the th design matrix, where Dij is the value of the j component of the parameter vector used in

th ∗ ∗ the i evaluation of the computer model. First, let the scaled design matrix D = [D ]ij, where each column of D∗ is a permutation of {1,...,m}. To generate the unscaled design

∗ th matrix, set Dij to be the value of a uniform random draw from the Dij region of Θj, i = 1,...m,j = 1,...,p (Saltelli et al., 2000). In practice, the fixed midpoint of each region

Θj is often used instead of a uniform random draw from each region. Geometrically, the parameter space is divided into mp cells. The LHD selects m cells in such a way that the projection of the midpoint of each cell onto a single dimension j is marginally uniformly distributed over Θj ∀j.

1.5.3.3 Maximin Latin hypercube designs

Maximin designs tend to push design points toward the edges of the parameter space

(e.g., toward the edges of a rectangular parameter space when p = 2), and are therefore not non-collapsing in general. Latin hypercube designs have desirable projection properties but are generally not space-filling (e.g., for any value of n, and with p = 2, a LHD could put design points along the diagonal of a rectangular parameter space). A maximin LHD is a maximin design that is constrained to be a LHD (Morris & Mitchell, 1995), and therefore is both space-filling and non-collapsing.

1.6 Gaussian process prediction and simulation-based based research

Gaussian process (GP) prediction is the foundation for sensitivity analysis and calibration of time-consuming computer models. For simulation-based research in systems biology, sen- sitivity analysis identifies the relative importance of biological (e.g., pathogen and immune) 23 parameters and characterizes their effects, and calibration yields estimates of biological pa- rameters based on experimental data. Model comparison methods quantify the evidence in favor of one biological model (e.g., model of immune behavior) over another. In chapter 2 we use GP predictors in the analysis and calibration of a computer model of L. major infection and in the comparison of alternative models of macrophage emigration. 24

Bibliography

Alvar, J., Canavate,˜ C., Gutierrez-Solar,´ B., Jimenez,´ M., Laguna, F., Lopez-´ Velez,´ R., Molina, R. & Moreno, J. (1997). Leishmania and human immunodeficiency virus coinfection: the first 10 years. Clin. Microbiol. Rev. 10, 298–319.

An, G. (2002). Agent-based computer simulation and SIRS: building a bridge between basic science and clinical trials. Shock 16, 266–273.

Awasthi, A., Kumar, R. & Saha, B. (2004). Immune response to Leishmania infection. Indian J. Med. Res. 119, 238–258.

Barrett, P. H. R., Bell, B. M., Cobelli, C., Golde, H., Schumitzky, A., Vicini, P. & Foster, D. M. (1998). SAAM II : Simulation, analysis, and modeling software for tracer and pharmacokinetic studies. Metab. Clin. Exp. 47, 484–492.

Barry, M. & Bleackley, R. C. (2002). Cytotoxic T lymphocytes: all roads lead to death. Nat. Rev. Immunol. 2, 401–409.

Bayarri, M., Berger, J. O., Higdon, D., Kennedy, M., Kottas, A., Paulo, R., Sacks, J., Cafeo, J., Cavendish, J. & Tu, J. (2002). A framework for the validation of computer models. In: Proceedings of the Workshop on Foundations for V&V in the 21st Century (Pace, D. & Stevenson, S., eds.). San Diego, CA: Society for Modeling and Simulation International.

Bayarri, M., Berger, J. O., Paulo, R., Sacks, J., Cafeo, J. A., Cavendish, J., Lin, C.-H. & Tu, J. (2007). A framework for validation of computer models. Technometrics 49, 138–154.

Beauchemin, C. (2006). Probing the effects of the well-mixed assumption on viral infection dynamics. J. Theor. Bio. 242, 464–477.

Belkaid, Y., Mendez, S., Lira, R., Kadambi, N., Milon, G. & Sacks, D. (2000). A natural model of Leishmania major infection reveals a prolonged “silent” phase of parasite amplification in the skin before onset of lesion formation and immunity. J. Immunol. 165, 969–977. 25

Belkaid, Y., Piccirillo, C. A., Me´ndez, S., Shevach, E. M. & Sacks, D. L. (2003). CD4+CD25+ immunoregulatory T cells control Leishmania major persistence and the de- velopment of concomitant immunity. Nature 420, 502–507.

Belkaid, Y., Von Stebut, E., Mendez, S., Lira, R., Caler, E., Bertholet, S., Udey, M. C. & Sacks, D. (2002). CD8+ T cells are required for primary immunity in C57BL/6 mice following low-dose, intradermal challenge with Leishmania major. J. Immunol. 168, 3992–4000.

Bentley, K., Gerhardt, H. & Bates, P. A. (2008). Agent-based simulation of notch- mediated tip cell selection in angiogenic sprout initialisation. J. Theor. Bio. 250(1), 25–36.

Berhe, N., Hailu, A. & Gemetchu, T. (1995). Human immunodeficiency virus and recurrence of cutaneous leishmaniasis long after healed localized cutaneous leishmaniasis due to L. aethiopica. Trans. R. Soc. Trop. Med. Hyg. 89, 205–207.

Bernaschi, M. & Castiglione, F. (2001). Design and implementation of an immune system simulator. Comput. Biol. Med. 31(5), 303–331.

Bernaschi, M. & Castiglione, F. (2002). Selection of escape mutants from immune recognition during HIV infection. Immunol. Cell. Bio. 80, 307–313.

Bodgan, C., Rollinghoff,¨ M. & Diefenbach, A. (2000). The role of nitric oxide in innate immunity. Immunol. Rev. 173, 17–26.

Bonabeau, E. (2002). Agent-based modeling: Methods and techniques for simulating human systems. PNAS 99(3), 7280–7287.

Castiglione, F., Poccia, F., D’Offizi, G. & Bernaschi, M. (2004). Mutation, fitness, viral diversity, and predictive markers of disease progression in a computational model of HIV type 1 infection. AIDS Res. Hum. Retroviruses 20(12), 1314–1323.

Chan, D. C., Watts, G. F., Redgrave, T. G., Mori, T. A., Hugh, M. P. & Barrett, H. R. (2002). Apolipoprotein B-100 kinetics in visceral obesity: Associations with plasma apolipoprotein C-III concentration. Metabolism 51(8), 1041–1046.

Chappuis, F., Sundar, S., Hailu, A., Ghalib, H., Rijal, S., Peeling, R. W., Alvar, J. & Boelaert, M. (2007). Visceral leishmaniasis: what are the needs for diagnosis, treatment, and control. Nat. Rev. Microbiol. 5(11), 873–882.

Choi, C. M. & Lerner, E. A. (2002). Leishmaniasis: recognition and management with a focus on the immunocompromised patient. Am. J. Clin. Dermatol 3(2), 91–105.

Clemens, J. & Jodar, L. (2005). Introducing new vaccines into developing countries: obstacles, opportunities, and complexities. Nat. Med. 11, S12–S15. 26

Codd, E. F. (1968). Cellular Automata. New York: Academic Press. Cohn, M. & Langman, R. E. (1990). The Protecton: the evolutionarily selected unit of humoral immunity. Immunol. Rev. 115, 107–147. Conc¸eicao-Silva, F., Hahne, M., Schroter,¨ M., Louis, J. & Tschopp, J. (1998). The resolution of lesions induced by Leishmania major in mice requires a functional Fas (APO-1, CD95) pathway of cytoxicity. Eur. J. Immunol. 28, 237–245. Cooper, N. R. (1984). The classical complement pathway. Activation and regulation of the first complement component. Adv. Immunol. 37, 151–215. Croft, S. L. & Coombs, G. H. (2003). Leishmaniasis -current chemotherapy and recent advances in the search for novel drugs. Trends Parasitol. 19, 502–508. Currin, C., Mitchell, T., Morris, M. & Ylvisaker, D. (1989). Bayesian prediction of deterministic functions, with applications to the design and analysis of computer exper- iments. JASA 86, 953–963. Desjeux, P., Meert, J. P., Piot, B., Alvar, J., Medrano, F. J., Portus, M., Munoz, C., Laguna, F., Velez, R. L., Salas, A., Sirera, G., Cisterna, R., Montalban, C., Quero, H., Gradoni, L., Gramiccia, M., Russo, R., Dedet, J. P., Prat- long, F., Dereure, J., Deniau, M., Izri, A., Matheron, S., Farault, F., Marty, P., Rosenthal, E., Antunes, F., Abranches, P. & Pradinaud, R. (1999). Leish- mania/HIV co-infection in southwestern Europe 1990-1998: retrospective analysis of 965 cases. URL http://whqlibdoc.who.int/hq/2000/WHO LEISH 2000.42.pdf. Deutsch, A. & Dormann, S. (2004). Cellular Automotaon Modeling of Biological Pattern Formation. Boston: Birkh¨auser. Du, X., Poltorak, A., Silva, M. & Beutler, B. (1999). Analysis of Tlr4-mediated LPS signal transduction in macrophages by mutational modification of the receptor. Blood Cells Mol. Dis. 25(5-6), 328–338. Duffield, J. S. (2003). The inflammatory macrophage: a story of Jekyll and Hyde. Clin. Sci. (Lond.) 104, 27–38. Erb, K. J., Blank, C. & Moll, H. (1996). Susceptibility to Leishmania major in IL-4- transgenic mice is not correlated with the lack of a Th1 immune response. Immunol. Cell. Biol. 74, 239–244. Ermentrout, G. B. & Edelstein-Keshet, L. (1993). Cellular automata approaches to biological modeling. J. Theor. Bio. 160(1), 97–133. Fang, K. T., Lin, D., Winker, P. & Zhang, Y. (2000). Uniform design: theory and applications. Technometrics 42, 327–248. 27

Filippi, C., Hugues, S., Cazareth, J., Julia, V., Glaichenhaus, N. & Ugolini, S. (2003). CD4+ T cell polarization in mice is modulated by strain-specific major histo- compatibility complex-independent differences within dendritic cells. J. Exp. Med. 198, 201–209.

Folcik, V. A., An, G. C. & Orosz, C. G. (2007). The Basic Immune Simulator: an agent- based model to study the interactions between innate and adaptive immunity. Theor. Biol. Med. Model 4, 29.

Forrest, S. & Beauchemin, C. (2007). Computer immunology. Immunol. Rev. 216, 176–197.

Gardner, M. (1970). Mathematical games: the fantastic combinations of John Conway’s new solitaire game ‘life’. Sci. Am. 223, 120–123.

Gilbert, N. (2007). Agent-Based Models (Quantitative Applications in the Social Sciences). Thousand Oaks, CA: Sage Publications, Inc.

Gillis, D., Klaus, S., Schnur, L. F., Piscopos, P., Maayan, S., Okon, E. & Engel- hard, D. (1995). Diffusely disseminated cutaneous Leishmania major infection in a child with acquired immunodeficiency syndrome. Pediatr. Infect. Dis. J. 14, 247–249.

Gounni, A. S., Lamkhioued, B., Ochiai, K., Tanaka, Y., Delaporte, E., Capron, A., Kinet, J. P. & Capron, M. (1994). High-affinity IgE receptor on eosinophils is involved in defence against parasites. Nature 367, 83–86.

Gramacy, R. B. (2005). Bayesian treed Gaussian process models. Ph.D. thesis, University of California.

Greenblatt, C. L. (1988). Cutaneous leishmaniasis: the prospects for a killed vaccine. Parasitol. Today 4(2), 53–54.

Grimm, V. & Railsback, S. F. (2005). Individual-based Modeling and Ecology. Princeton, NJ: Princeton University Press.

Gumy, A., Louis, J. A. & Launois, P. (2004). The murine model of infection with Leishmania major and its importance for deciphering of mechanisms underlying differences in Th cell differentiation in mice from different genetic backgrounds. Int. J. Parasitol. 34(4), 433–444.

Heinzel, F. P., Schoenhaut, D. S., Rerko, R. M., Rosser, L. E. & Gately, M. K. (1993). Recombinant interleukin-12 cures mice infected with Leishmania major. J. Exp. Med. 177, 1505–1509. 28

Hernandez,´ D. E., Oliver, M. & Planas, G. (1995). Visceral leishmaniasis with cuta- neous and rectal dissemination due to Leishmania braziliensis in acquired immunodeficiency syndrome (AIDS). Int J Dermatol 34, 114–115.

Higashi, G. I. (1988). Vaccines for parasitic diseases. Ann. Rev. Public Health 9, 483–501.

Ho, D. D., Neumann, A. U., Perelson, A. S., Chen, W., Leonard, J. M. & Markowitz, M. (1995). Rapid turnover of plasma virions and CD4 lymphocytes in HIV-1 infection. Nature 373, 123–126.

Huang, F. P., Xu, D., Esfandiari, E. O., Sands, W., Wei, X. Q. & Liew, F. Y. (1998). Mice defective in Fas are highly susceptible to Leishmania major infection despite elevated IL-12 synthesis, strong Th1 responses and enhanced nitric oxide production. J. Immunol. 160, 4143–4147.

Ideker, T., Ozier, O., Schwikowski, B. & Siegel, A. F. (2002). Discovering regulatory and signalling circuits in molecular interaction networks. Boinformatics 18(Suppl. 1), S233– S240.

Inchiosa, M. E. & Parker, M. T. (2002). Overcoming design and development challenges in agent-based modeling using ASCAPE. PNAS 99(3), 7304–7308.

Ivison, S. M., Khan, M. A., Graham, N. R., Q, B. C., Kaleem, A., Tirling, C. O., Cherkasov, A. & Steiner, T. S. (2007). A phosphorylation site in the Toll-like receptor 5 TIR domain is required for inflammatory signalling in response to flagellin. Biochem. Biophys. Res. Commun. 352(4), 936–941.

Janeway, C. A., Travers, P., Walport, M. & Shlomchik, M. J. (2005). Immuno- biology: The Immune System in Health and Disease, 6th ed. New York: Garland Science Publishing.

Johnson, M. E., Moore, L. M. & Ylvisaker, D. (1990). Minimax and maximin distance designs. J. Statist. Plann. Inf. 13, 455–492.

Kamhawi, S., Ramalho-Ortigao, M., Kumar, S., Lawyer, P., Turco, S., Barillas- Mury, C., Sacks, D. & Valenzuela, J. (2004). A role for insect galectins in parasite survival. Cell 119, 329–341.

Kedzierski, L., Zhu, Y. & Handman, E. (2006). Leishmania vaccines: progress and problems. Parasitology 133, S87–S112.

Kennedy, M. C. & O’Hagan, A. (2001). Bayesian calibration of computer models. J.R. Statist. Soc. B 63, 425–464.

Kitano, H. (2002). Computational systems biology. Nature 420, 206–210. 29

Koehler, J. R. & Owen, A. B. (1996). Computer experiments. In: Handbook of Statistics (Ghosh, S. & Rao, C. R., eds.), vol. 13. New York: Elsevier Science, pp. 261–308.

Lieberman, J. (2003). The ABCs of granule-mediated cytotoxicity: new weapons in the arsenal. Nat. Rev. Immunol. 3, 361–370.

Lindley, D. V. (1956). On a measure of information provided by an experiment. Ann. Math. Stat. 27, 986–1005.

Lira, R., Sundar, S., Makharia, A., Kenney, R., Gam, A., Saraiva, E. & Sacks, D. (1999). Evidence that the high incidence of treatment failures in Indian kala-azar is due to the emergence of antimony-resistant strains of Leishmaniasis donovani. J. Infect. Dis. 180, 564–567.

Magill, A. L., Grogl, M., Gasser, R. A., Sun, W. & Oster, C. N. (1993). Visceral infection caused by L. tropica in veterns of Operation Desert Storm. N. Engl. J. Med. 387, 1383–1387.

Mata, J. & Cohn, M. (2007). Cellular automata-based modeling program: synthetic im- mune system. Immunol. Rev. 216(1), 198–212.

Matern,´ B. (1960). Spatial Variation. Ph.D. thesis, Meddelanden fran Statens Skogsforskn- ingsinstitut.

Matern,´ B. (1986). Spatial Variation (Second Edition). New York: Springer-Verlag.

McKay, M., Beckman, R. & Conover, W. (1979). A comparison of three methods for selecting values of input variables in the analysis of output from a computer code. Technometrics , 239–246.

Meier-Schellersheim, M. (1999). SIMMUNE, a tool for simulating and analyzing immune system behavior. Ph.D. thesis, University of Hamburg.

Mendez, S., Reckling, S. K., Picciriloo, C. A., Sacks, D. & Belkaid, Y. (2004). Role for CD4+CD25+ regulatory T cells in reactivation of persistent leishmaniasis and control of concomitant immunity. J. Exp. Med. 200, 201–210.

Minar, N., Burkhart, R., Langton, C. & Askenazi, M. (1996). The Swarm Simulation System, A Toolkit for Building Multi-Agent Simulations. URL http://www.santafe.edu/projects/swarm/overview/overview.html.

Moll, H., Fuchs, H., Blank, C. & Rollinghoff, M. (1993). Langerhans cells transport Leishmania major from the infected skin to the draining lymph node for presentation to antigen-specific T cells. Eur. J. Immunol. 23, 1595–1601. 30

Morris, M. D. & Mitchell, T. J. (1995). Exploratory designs for computational experi- ments. J. Stat. Plan. Infer. 43, 381–402.

Mossmann, T. R. & Coffman, R. L. (1989). Th1 and Th2 cells: different patterns of lymphokine secretion lead to different functional properties. Annu. Rev. Immunol. 7, 145– 173.

Muller, I. P., Kropf, R. J. & Etges, J. A. L. (1993). γ Interferon response in secondary Leishmania major infection: role of CD8+ T cells. Infect. Immun. 61, 3730–3738.

North, M. J. & Collier, N. T. (2006). Experiences creating three implementations of the repast agent modeling toolkit. TOMACS 16(1), 1–25.

Plotkin, S. A. (2005). Vaccines: past. present, and future. Nat. Med. 11, S5–S11.

Pogson, M., Smallwood, R., Qwarnstrom, E. & Holcombe, M. (2002). Formal agent-based modelling of intracellular chemical interactions. Biosystems 85(1), 37–45.

Puzone, R., Kohler, B., Seiden, P. & Celada, F. (2002). IMMSIM, a flexible model for in machina experiments on immune system responses. Future Gener. Comp. Sy. 18(7), 961–972.

Reithinger, R., Dujardin, J. C., Louzir, H., Pirmez, C., Alexander, B. & Brooker, S. (2008). Cutaneous leishmaniasis. Lancet Infect Dis. 8(1), 581–596.

Rencher, A. C. (2002). Methods of Multivariate Analysis, 2nd Edition. New York: Wiley.

Rendell, P. (2001). Turing universality of the game of life. In: Collision-based computing (Adamatzky, A., ed.). London: Springer-Verlag, pp. 513–539.

Sacks, J., Schiller, S. B. & Welch, W. J. (1992). Design for computer experiments. Technometrics 31, 41–47.

Sacks, J., Welch, W. J., Mitchell, T. J. & Wynn, H. P. (1989). Design and analysis of computer experiments. Stat. Sci. 4(4), 409–435.

Sadick, M. D., Heinzel, F. P., Holaday, B. J., Pu, R. T., Dawkins, R. S. & Lock- sley, R. M. (1990). Cure of murine leishmaniasis with anti-interleukin 4 monoclonal antibody. Evidence for a T cell-dependent, interferon-γ independent mechanism. J. Exp. Med. 171, 115–127.

Saltelli, A., Chan, K. & Scott, E. M. (2000). Sensitivity Analysis. New York: Wiley.

Santner, T. J., Williams, B. J. & Notz, W. (2003). The Design and Analysis of Computer Experiments. New York: Springer. 31

Scott, P. (2003). Development and regulation of cell-mediated immunity in experimental leishmaniasis. Immunol. Res. 27, 489–498. Segovia-Juarez, J. L., Ganguli, S. & Kirschner, D. (2004). Identifying control mech- anisms of granuloma formation in M. tuberculosis infection using an agent-based model. J. Theor. Biol. 231, 357–376. Seiden, P. & Celada, R. (1992). A model for simulating cognate recognition and response in the immune system. J. Theor. Biol. 158, 329–357. Shewry, M. C. & Wynn, H. P. (1987). Maximum entropy sampling. J. Appl. Statist. 14, 165–170. Shibata, R., Igarashi, T., Haigwood, N., Buckler-White, A., Ogert, R., Ross, W., Willey, R., Cho, M. W. & Martin, M. A. (1999). Neutralizing antibody di- rected against HIV-1 envelope glycoprotein can completely block HIV-1/SIV chimeric virus infections of macaque monkeys. Nat. Med. 5, 204–210. Sirakoulisa, G. C., Karafyllidis, I., Mizasa, C., Mardirisa, V., Thanailakisb, A. & Tsalidesb, P. (2003). A cellular automaton model for the study of DNA sequence evolution. Comput. Biol. Med. 5, 439–453. Sypek, J. P., Chung, C. L., Mayor, S. E., Subramanyam, J. M., Goldman, S. J., Sieburth, D. S., Wolf, S. F. & Schaub, R. G. (1993). Resolution of cutaneous leishmaniasis: interleukin 12 initiates a protective T helper type 1 immune response. J. Exp. Med. 177, 1797–1792. Tabbara, K. S., Peters, N. C., Afrin, F., Mendez, S., Bertholet, S., Belkaid, Y. & Sacks, D. L. (2005). Conditions influencing the efficacy of vaccination with live organisms against Leishmania major infection. Infect. Immun. 73(8), 4714–4722. Tang, B. (1993). Orthogonal array-based Latin hypercubes. JASA 88, 1392–1397. Tesfatsion, L. (2006). Agent-based computational economics: a constructive approach to economic theory. In: Handbook of Computational Economics (Tesfatsion, L. & Judd, K. L., eds.), vol. 2. North Holland, Oxford, UK: Elsevier/North-Holland, 1 ed., pp. 831–880. Tisue, S. & Wilensky, U. (2004). NetLogo: Design and implementation of a multi-agent modeling environment. In: SwarmFest. Ann Arbor, MI: Swarm Development Group. Ummenhofer, W. C., Arends, R. H., Shen, D. D. & Bernards, C. M. (2000). Compar- ative Spinal Distribution and Clearance Kinetics of Intrathecally Administered Morphine, Fentanyl, Alfentanil, and Sufentanil. Anesthesiology 92(3), 739–753. Vanloubbeeck, Y. & Jones, D. E. (2004). The immunology of Leishmania infection and the implications for vaccine development. Ann. N.Y. Acad. Sci. 1026, 267–272. 32 von Stebut, E., Belkaid, Y., Jakob, T., Sacks, D. L. & Udey, M. C. (1998). Uptake of Leishmania major amastigotes results in activation and interleukin 12 release from murine skin-derived dendritic cells: implications for the initiation of anti-Leishmania immunity. J. Exp. Med. 61(188), 1547–1552. von Stebut, E., Belkaid, Y., Nguyen, V. B., Cushing, M., Sacks, D. L. & Udey, M. C. (2000). Leishmania major-infected murine Langerhans cells-like dendritic cells from susceptible mice release IL-12 after infection and vaccinate against experimental cutaneous leishmaniasis. Eur. J. Immunol. 30, 3800–3811. von Stebut, E. & Udey, M. C. (2004). Requirements for Th1-dependent immunity against infection with Leishmania major. Microbes Infect. 6, 1102–1109.

Warrender, C., Forrest, S. & Segel, L. (2004). Homeostasis of peripheral immune effectors. Bull. Math. Biol. 66, 1493–1514.

Warrender, C. E. (2004). Modeling intercellular interactions in the peripheral immune system. Ph.D. thesis, University of New Mexico, Computer Science Department.

Wei, X. Q., Charles, I. G., Smith, A., Feng, G. H., Huang, F. P., Xu, D., Muller, W., Moncada, S. & Liew, F. Y. (1995). Altered immune response in mice lacking inducible nitric oxide synthase. Nature 375, 408–411.

Wooldridge, M. (2002). Inroduction to Multiagent Systems. New York: John Wiley & Sons, Inc.

Ye, K. Q., Li, W. & Sudjianto, A. (2000). Algorithmic construction of optimal symmetric Latin hypercube designs. J. Stat. Plann. Inf. 90, 145–159. 33

CHAPTER 2. Parameter estimation and sensitivity analysis in an agent-based model of Leishmania major infection

To be submitted to the Journal of Theoretical Biology

Garrett M. Dancik4, Douglas E. Jones3,4, Karin S. Dorman1,2,4

1Departments of Statistics, 2Genetics, Development & Cell Biology, 3Veterinary Pathology,

and the 4Program in Bioinformatics and Computational Biology, Iowa State University

[email protected], [email protected], [email protected]

Abstract

Computer models of disease take a systems biology approach toward understanding host-pathogen interactions. In particular, data driven computer model calibration is the ba- sis for inference of immunological and pathogen parameters, assessment of model validity, and comparison between alternative models of immune or pathogen behavior. In this paper we describe the calibration and analysis of an agent-based model of Leishmania major infec- tion. A model of macrophage emigration following uptake of necrotic tissue is proposed to explain macrophage depletion following peak infection. Using Gaussian processes to approx- imate the computer code, we perform a sensitivity analysis to identify important parame- ters and to characterize their influence on the simulated infection. Subsequent calibration of 34 the model against previously published biological observations suggests that L. major has a relatively slow growth rate and can replicate for an extended period of time before damag- ing the host cell.

2.1 Introduction

Leishmania are protozoan parasites that are transmitted by bites of infected sand-

flies. The macrophage is the primary host cell. Over 20 species of Leishmania, endemic in

88 countries, are capable of causing human disease. Disease is either cutaneous, where skin ulcers occur on exposed surfaces of the body, or visceral, with near certain mortality if left untreated. In mice that are able to control L. major infection, the resistance mechanism is well understood: secretion of IL-12 by dendritic cells promotes a CD4+ Th1 response, Th1 cells activate macrophages through IFN-γ production, and activated macrophages clear the parasite. However, there are other species of the Leishmania parasite, such as L. amazo- nensis, which cause chronic disease in mice (McMahon-Pratt & Alexander, 2004), and the development of successful vaccines for any Leishmania species has remained elusive (Van- loubbeeck & Jones, 2004).

Computer models of disease take a systems biology approach toward understanding adverse or inefficient immune responses by integrating multiple sources of knowledge about host-pathogen interactions and immune cell function in order to study the collective, emer- gent behavior of a population of immune cells, i.e., the immune response. Such computer models have been used to gain insight into a variety of diseases. For example, in a model of Mycobacterium tuberculosis infection, Segovia-Juarez et al. (2004) identify chemokine diffusion rates and the arrival time, location, and macrophage activation efficiency of T cells as important factors in granuloma formation. In a model of systemic inflammatory response and multiple organ failure, An (2002) reproduces outcomes of unsuccessful clin- 35 ical trials involving blockage of proinflammatory mediators. In a model of influenza A in- fection, Beauchemin (2006) show how infection dynamics depend on the spatial structure of initially infected cells. Finally, in a model of antigen escape in HIV infection, Bernaschi

& Castiglione (2002) find that escape mutants with low transcription rate can explain the long-term asymptomatic phase of disease. For a recent review of computational immune system models, see Forrest & Beauchemin (2007).

A challenge faced when working with computer models is the need to choose values for model parameters. Typically, plausible choices for parameters are determined through literature searches and expert consultation, yet often the best available information are plausible ranges for model parameters rather than single values. In order to validate the model, one must ultimately fit the computer output to field data, choosing parameter val- ues that yield the best match between simulation output and biological observations. Model calibration has proven to be difficult in practice, especially since most computer models are high dimensional, non-linear, and resource-intensive. As a result, computer modelers tradi- tionally employ ad-hoc approaches to parameter estimation (Kennedy & O’Hagan, 2001), where model validation is based on the qualitative comparison of model predictions with

field data. In the field of population ecology, this approach is known as pattern-oriented modeling (reviewed in Grimm et al., 2005).

Nevertheless, data-driven model calibration is a fundamental and necessary step for making meaningful inference on model parameters, assessing the validity of a model, and performing model comparisons (Bayarri et al., 2002). When the computer model is expen- sive to run, this process generally entails 1) specification of the relationship between the computer model output and data that is observed in the field, 2) approximation of the com- puter code through a surrogate statistical model (e.g., a Gaussian process), and 3) estima- tion of computer model parameters via the surrogate model. The Gaussian process approx- 36 imation also facilitates sensitivity analysis which identifies important model parameters and the relationships between parameters and simulation output. Schonlau & Welch (2006), for example, use a Gaussian process emulator of a computer model of global human develop- ment to identify important parameters and characterize their effects. These sensitivity anal- ysis tools are particularly useful for types of computer models, such as agent-based and cel- lular automata models, where it is not clear how the computer code itself can be exploited, without running the model, to quantitatively understand the system.

Computer model calibration through a Gaussian process intermediate has been demon- strated successfully in several disciplines. For example, Kennedy & O’Hagan (2001) cali- brate a computer model of radionucleotide exposure following a chemical accident; Bayarri et al. calibrate a vehicular suspension system model (2006), and a vehicle collision model

(2002); Higdon et al. (2004) calibrate a spot welding model; and Heitmann et al. (2006) cal- ibrate a cosmological model. A recent formulation of the computer model calibration and validation approach is provided in Bayarri et al. (2007).

In this paper we describe the sensitivity analysis and calibration of an agent-based model of Leishmania major infection. A model of macrophage emigration triggered by necrotic tissue production is proposed for explaining macrophage depletion after peak infection. We

find that pathogen growth rate and host cell carrying capacity both affect macrophage lev- els at five weeks post infection (wpi), though not independently. Increasing parasite growth rate can both augment and paradoxically, suppress parasite loads, depending on the stage of infection and the ability of the pathogen to avoid detection. Furthermore, the timing of the adaptive immune response has a large effect on macrophage levels at 6.5 wpi. We verify that parameter estimation using the Gaussian process intermediate is accurate by calibrat- ing the computer model using simulated field data, and then calibrate our model using field data from Belkaid et al. (2000). Parameter estimates suggest that intracellular pathogen 37 replicates extensively before spreading to additional cells, a finding consistent with obser- vations of Leishmania growth in cultured macrophages (Chang et al., 2002). Finally, model comparison supports our proposal of explicit macrophage emigration against a model where emigration does not occur.

The rest of this paper is organized as follows: in section 2.2 we describe the agent- based model; in section 2.3 we describe the statistical methods used in Gaussian process computer model approximation (2.3.1), Gaussian process diagnostics (2.3.2), sensitivity analysis (2.3.3), calibration (2.3.4), implementation via Markov Chain Monte Carlo (2.3.5), and model comparison (2.3.6); in section 2.4 we report results obtained by applying these statistical methods to the agent-based model of Leishmania infection in mice; and in section

2.5 we discuss our findings.

2.2 An agent-based model of Leishmania major infection

Agent-based models (ABMs) capture the dynamics of complex systems whose prop- erties depend on the collective behavior of their interacting components. An ABM con- tains distinct entities, or agents, that inhabit a spatial environment. A simulation visualizes agents as they move and interact according to update rules executed at discrete time steps

(Wooldridge, 2002).

We describe an ABM of the immune response to L. major infection in the ear of a

C57BL/6 mouse, a system where experimental data are available (Belkaid et al., 2000). Our computer model is stochastic, so that multiple simulations will give different results even when the same model parameters and starting conditions are used. The current model is an extension of an earlier ABM of L. major infection (Dancik et al., 2006), and is largely based on the work of Segovia-Juarez et al. (2004), who explore granuloma formation during infection with another macrophage-tropic parasite, Mycobacterium tuberculosis. All model 38 parameters are listed in Table 2.6 of the appendix.

2.2.1 The model environment

We model a 2mm × 2mm cross section of the ear as a 100 × 100 lattice of square micro-compartments. This lattice represents an internal portion of a uniform infection area, and is therefore toroidal, so an object leaving the lattice will re-enter at the opposite end.

We select four evenly distributed micro-compartments to serve as source compartments where new cells enter the system.

We label micro-compartments (i, j), starting from (0, 0) at the bottom left. Define a

Moore neighborhood of length r at position (x0, y0) to be the local collection of points in the lattice given by

Mr(x0, y0)= {(x, y): |x − x0|≤ r and |y − y0|≤ r} .

Furthermore, define M1(x0, y0) to be the immediate Moore neighborhood of a micro-compartment

(x0, y0).

2.2.2 Stages of infection

2.2.2.1 Initial conditions

We take the starting point of our simulation to be 3.5 weeks post infection (wpi) of the experiment described in Belkaid et al. (2000), near the beginning of the dramatic phase of parasite growth. Data for the number of infected macrophages at this time point are not available. However, because pathogen load dynamics during the dramatic phase of para- site growth are similar to pathogen load dynamics following high dose infection (Lira et al.,

2000), we assume that conditions at the start of our simulation match conditions just two days post high dose infection. Based on parasite and expected cell counts observed two days post high-dose infection (Doug Jones, unpublished observations), we randomly place 105 39

macrophages on the lattice. A random number of parasites between 1 and Pm infect ran- domly selected resting macrophages until total pathogen load on the lattice exceeds 50−Pm.

A final, random resting macrophage receives 50 − Pm parasites, so the initial pathogen load is 50. All uninfected macrophages are given random life spans uniform between 0 and 100 days (Furth et al., 1973).

2.2.2.2 Infection of macrophages

Macrophages are the primary cells that L. major parasites infect (Belkaid et al., 2000) and the only host cells that support pathogen replication (Naderer et al., 2006). We use the term resting macrophage to refer to an uninfected, unactivated cell. We assume that intra- cellular parasites experience logistic growth at rate αI with carrying capacity kI + 30 . If the macrophage is not activated, intracellular parasites grow until their number exceeds the transfer threshold kI . Then, the macrophage enters a dying state, where it begins transfer- ring parasite to macrophages in its length two Moore neighborhood. Since extracellular par- asites are seldom observed in Leishmania infection (Chang et al., 2002), parasite transmis- sion is thought to occur when nearby macrophages phagocytize dying host cells. Segovia-

Juarez et al. (2004) use a similar model, but allow extracellular parasite and restrict par- asite uptake to macrophages in their length 1 Moore neighborhood. All macrophages that are not dying, including those already infected, can take up parasite. A dying macrophage is removed from the system once all of its intracellular parasites are consumed by surround- ing macrophages.

2.2.2.3 Chemokines and cell movement

Chemokines, chemical attractants that influence cell movement, play an important role in Leishmania infection (Roychoudhury & Roy, 2004). We include one generic chemokine 40 as an attractant for both macrophages and T cells. Its diffusion and decay properties are based on interleukin-8, an important chemokine involved in early infection. Infected and activated macrophages release cI units of chemokine per time step. In addition to direct- ing cell movement toward infected areas (see below), chemokine also increases the rate of macrophage recruitment (section 2.2.2.4).

Cell movement has been described as a biased-random walk in the presence of chemokine

(Tranquillo et al., 1988), and we model this random walk on the ABM lattice. Let

Ci,j = amount of chemokine in micro-compartment (i, j)

CM1 (i, j) = amount of chemokine in immediate Moore neighborhood M1(i, j)

A cell currently in micro-compartment (i, j) moves to micro-compartment (k,l) ∈ M1(i, j), (k,l) 6= (i, j) with probability Ck,l . All cells in our model move in this way, but at different CM1 (i,j)−Ci,j speeds which depend on cell type and state (section 2.2.2.7 and Table 2.6).

2.2.2.4 Recruitment and the adaptive immune response

T cells and macrophages are recruited to infected areas during infection. In the ABM, source compartments represent blood vessels where recruited cells arrive. At each time-step, macrophages enter with probability based on source compartment chemokine levels (see be- low), and T cells, after a delay, enter with constant probability pTrecr. Macrophage recruitment in response to chemokine: Macrophages are actively recruited around two days post infection following high-dose L. major inoculation (Sunderkotter et al.,

1993), and after a delay in low dose infection (Belkaid et al., 2000). This recruitment is modulated by chemokines such as macrophage chemotactic and activating factor (Badolato et al., 1996). Certain chemokines trigger the activation of endothelial cells and upregulation of monocyte receptors, which facilitate extravasation of cells from the circulatory system 41 into infected tissue. Other chemokines, such as granulocyte-macrophage colony-stimulating factor, upregulate macrophage production from the bone marrow (Janeway et al., 2005).

Therefore, chemokines collectively increase the amount of circulating macrophages as well as the rate of influx of these macrophages into the tissue.

We use a logistic function to model macrophage recruitment in response to the generic chemokine in our model. This function captures initial exponential increase in macrophage numbers due to macrophage production and influx, as well as the eventual loss of chemokine potency at high chemokine concentrations due to the saturation of chemokine receptors

(Rajotte et al., 1997). Let

mMrecr×exp(αMrecr+βMrecr×x) x > CM 1+exp(αMrecr+βMrecr×x) pMrecr,i(x)=  (2.1) −5  1.74 × 10 x ≤ CM

 where pMrecr,i(x) is the probability that a macrophage is recruited when there are x units of chemokine at source compartment i. The parameter mMrecr represents the maximum re- cruitment probability, αMrecr determines the minimum probability of recruitment as x → 0, and βMrecr determines how sharply recruitment increases in response to chemokine. This active macrophage recruitment mechanism is used as long as x is above a chemokine thresh- old, CM ; otherwise we use a base level of recruitment that will maintain macrophage equi- librium in the absence of infection. Because active macrophage recruitment does not appear to initiate until after 3.5 wpi in the experimental infection we are modeling (Belkaid et al.,

2000), which is the start of our simulation, we set CM dynamically to correspond to source compartment chemokine levels after 1 day of simulation.

The adaptive immune response: During infection, antigen-presenting cells take up pathogen from the site of infection, migrate to the draining lymph node, and present anti- gen (processed pathogen) to na¨ıve T cells. Na¨ıve T cells, in response, proliferate and ma- 42 ture into several classes of T cells, including Th1 CD4+ cells. In L. major infection, Th1

CD4+ cells migrate to the infected area and activate infected macrophages. The initiation of the adaptive immune response requires a threshold level of antigen (Janeway et al., 2005) and occurs only after a prolonged period of parasite growth in low dose L. major infection

(Belkaid et al., 2000). We assume that this threshold level of antigen is related to pathogen load at the infection site, and recruit T cells once a threshold pathogen level, Tthreshold, is achieved. The threshold is a one way trigger so T cells will continue to enter at each time step and at each source compartment with probability pTrecr as long as source compartment chemokine levels are greater than a threshold of CT chemokine units.

2.2.2.5 Macrophage emigration following uptake of necrotic tissue

In L. major infection, there is a sharp decrease in macrophage levels near the time that the T cell response reaches its peak (Belkaid et al., 2000). In our model, this decrease cannot be explained merely from the death of infected and activated macrophages (data not shown). Studies suggest that as the inflammatory response resolves, macrophages system- atically migrate to the draining lymph node (Bellingan et al., 1996), an event known to oc- cur following macrophage uptake of foreign antigen, apoptotic immune cells, and necrotic tissue (Winchester et al., 1984). In the ABM, we model the emigration of inflammatory macrophages by also tracking the production and uptake of necrotic tissue, a side effect of macrophage activation (Billack, 2006). Subsequent to macrophage activation, Anecr units of non-diffusive necrotic tissue are released. Uninfected macrophages in the length two Moore neighborhood consume one unit of this tissue and disappear from the lattice, representing rapid migration to the draining lymph node. 43

2.2.2.6 Macrophage activation

All T cells in our model are considered to be antigen specific Th1 CD4+ cells that are equally capable of activating infected macrophages. Macrophage activation is cell-mediated

(Sypek et al., 1984), and T cells within the immediate Moore neighborhood of an infected macrophage activate it with probability Tactm. In L. major infection, macrophage activation is sufficient for elimination of intracellular parasite (Wei et al., 1995). Mals days after T cell activation, the macrophage finishes destroying all intracellular parasite, undergoes apopto- sis, produces Anecr units of necrotic tissue, and is removed from the lattice.

2.2.2.7 Time scales

Each time step in the ABM is equal to approximately six seconds of real time. Chemokine diffusion and decay as well as parasite growth occur each time step. Following Segovia-

Juarez et al. (2004), T cells move every 200 time steps (20 minutes) and macrophages move on slower time scales that vary according to macrophage state (see Table 2.6). Additional update rules, such as the take up of parasite and activation of macrophages, occur every 10 minutes.

2.3 Statistical methods

Our primary objective is to estimate the computer model parameter vector θ = (θ1, . . . , θp)

(p parameters from Table 2.6) given a vector yF of biological field observations. The cali- bration process (see section 2.3.4) requires a large number of computer model evaluations to explore the parameter space. For computer models that are expensive (i.e., time con- suming) the number of evaluations is limited, and calibration demands a fast and accurate emulator of the computer code. The type of emulator that we use is the Gaussian process.

In addition to its role in parameter calibration, the Gaussian process emulator is utilized in 44 sensitivity analysis to explore the influence of parameters on the simulated response. In this section we give a brief overview of the statistical methods we use, including Gaussian pro- cess prediction of scalar and functional computer model output, sensitivity analysis, com- puter model calibration, and model comparison, with particular attention paid to notions relevant to our ABM. The reader is encouraged to consult the cited references for more de- tails and theory.

2.3.1 The Gaussian process as a computer model surrogate

Gaussian processes (GPs) are flexible and useful tools for approximating computer models when the computer model response surface of interest is a smooth function of the parameter space (Sacks et al., 1989; Kennedy & O’Hagan, 2001; Bayarri et al., 2002). In this framework, a collection of responses (i.e., scalar computer model outputs or principle component weights) is assumed to be multivariate normal with a correlation structure that depends on the parameter vector θ. In our computer model, output consists of pathogen load and macrophage counts during the course of the infection. GP prediction requires a collection of m known computer model outputs, zknown. The form of zknown depends on whether computer model output is scalar (section 2.3.1.1) or functional (section 2.3.1.2).

(new) Computer model output at an untried parameter θ , conditioned on zknown (for scalar output) or zknown’s most important principle component weights (for functional output), is predicted using standard multivariate normal distribution theory (Rencher, 2002).

2.3.1.1 Gaussian process prediction of scalar computer model output

The Gaussian process uses m scalar computer model observations, zknown = z θ(1) ,...,z θ(m) at m parameter vector inputs, and interpolates all other possible      model output while smoothing the dataset. We adopt a GP with constant mean, µ, a Gaus- 45 sian or product correlation structure of the form

p 2 (i) (j) (i) (j) Cij ≡ cor(z(θ ),z(θ )) = exp − βk θk − θk (2.2) ! kX=1  

2 and a constant nugget term σe to account for stochasticity in the computer model. The variance-covariance matrix has the form

2 2 Vm×m = σGPCm×m + σe Im×m,

2 where Cm×m is the matrix formed from (2.2) and σGP is the unconditional variance of mean computer model output at any parameter vector θ. The GP uses the observed computer

(new) (new) model outputs zknown to predict output,z ˆ θ = E z θ |zknown at a previously   h   i unobserved parameter setting θ(new). Specifically, predicted computer model output is nor- mally distributed with mean

2 ′ −1 µ + σGPr1×mVm×m(zknown − µ1m×1) (2.3) and variance

2 2 4 ′ −1 σGP + σe − σGPr1×mVm×mrm×1, (2.4)

′ (new) (i) where rm×1 = (r1,...,rm) , with the correlation ri = cor z θ ,z θ defined in      (2.2).

2 If the computer model output is deterministic, the nugget term σe is not needed, and the predictor will simply interpolate the known output. In the case of our ABM, the com- puter model output is stochastic, and the GP, with positive nugget term, interpolates and smooths the known output. We fit a GP to the m computer model outputs by obtaining

2 2 maximum likelihood estimates of the hyperparameters α,β,µ,σGP,σe using the software  46

PeRK (Santner et al., 2003).

2.3.1.2 Gaussian process prediction of functional computer model output

We predict complete time series output (pathogen load and macrophages observed every day from 3.5 to 8.5 wpi) from the computer model by using singular value decom- position (SVD) to reduce the dimensionality of the output and fitting independent GPs to the most important principle component weights (Heitmann et al., 2006). The collection of observed outputs [zknown]i,j , i = 1, ..., k, j = 1, ..., m is a matrix of m functional re- sponses, where column j of the matrix contains the k-dimensional functional response corre- sponding to the input parameter θ(j). Also let r = min(k, m). Using singular value decom- position, r ′ [zknown]i,j = UkxrDrxrVrxm i,j = λp [αp]i [wp(θ)]j, (2.5) p=1   X th th th ′ where λp is the p singular value, αp is the p column of U, and wp(θ) is the p row of V .

th ′ We will refer to the j column of V , which contains the elements [wp(θ)]j, p = 1,...,r, as a vector of principle component (PC) weights corresponding to the jth observation. The output zknown is approximated by keeping the l

2 λp variance explained(p)= r 2 i=1 λi P In our analysis, we keep the largest PC weights whose corresponding singular values ac- count for > 99% of the variation in the output, and fit GPs using the R package mlegp

(Dancik & Dorman, 2008). 47

2.3.2 Gaussian process diagnostics

Cross-validation techniques are used to assess the accuracy of the GP approxima-

(i) (i) tion. Letz ˆ−i θ be the predicted computer model output at θ after removing known   computer model runs at θ(i) from the simulation data set. Then the standardized cross- validated residual for known observation z θ(i) is  

(i) (i) z θ − zˆ−i θ (i) ,  se zˆ−i θ     (i) (i) where se zˆ−i θ is the standard error of the cross-validated predictionz ˆ−i θ . h  i   2.3.3 Sensivity analysis using the Gaussian process

Sensitivity analysis includes a broad range of statistical techniques for measuring a parameter’s ability to influence the output of a model. Typically, sensitivity analysis is used in factor screening methods for identifying the most important parameters of a system, as well as for assessing how the response of a model depends on its input parameters (Saltelli et al., 2000). The approach that we use follows that of Schonlau & Welch (2006). Of inter- est is how the computer model parameter vector θ = (θ1, . . . , θp) influences computer model output. Inferences are based on a prior π(θ) on model parameters, and the observed com- puter model output vector zknown (for scalar output) or the zknown’s most importance PC weights (for functional output). In our sensitivity analysis we use independent Unif(0,1) pri- ors on all (scaled) components of θ.

2.3.3.1 Sensitivity analyis for scalar computer model output

th For θk, the k component of the parameter vector, the main effect g(θk) is the ex- pected computer model output, E[z(θ)|zknown, θk], averaged over the joint prior distribution 48

th π(θ−k) for all components of θ except the k . Two-way interaction effects g(θk, θl) are de- fined similarly, except integration is with respect to a joint prior on all parameter vector

th th components but the k and the l , k 6= l. A main effects graph plots g(θk) against θk, thereby illustrating the effect of varying the single parameter θk on the simulation output. Contour plots conveniently illustrate two-way interaction effects.

The GP is a functional approximation of the computer model, and therefore lends it- self to functional analysis of variance (FANOVA) decomposition when marginal priors on the individual components of θ are independent (Schonlau & Welch, 2006). In FANOVA de- composition, the total functional variance of the GP emulator is decomposed into variance due to the main and interaction effects of the parameters. The percentage of the total func- tional variance accounted for by a particular effect provides a measure of the importance of that effect. We report the percent of total functional variance that is accounted for by main effects and two-way interactions, using independent uniform priors for all components of θ, over the parameter ranges given in Table 2.6. Details can be found in Schonlau & Welch

(2006).

2.3.3.2 Sensitivity analysis for functional computer model output

When zknown is a matrix of functional observations, as described in section 2.3.1.2, then the main effect of the parameter θk on each dimension of output is

r

gi(θk)= λp [αp]i wp(θ), pX=1 b th where λp is the p singular value, αp is a known vector, (see section 2.3.1.2), and

wp(θ)= E [wp(θ)|wknown,p, θk] ,

b 49

th where wp is the GP predictor of the p PC weight, wknown,p are the observed PC weights corresponding to the pth singular value, and expectation is with respect to the joint prior

th distribution π(θ−k) for all components except for the k . For time course data, a main ef-

th fects graph plots the i time point against gi(θk) for various values of θk.

2.3.4 Computer model calibration using field observations

Typically, the relationship between a computer model and reality is defined according to

yF (xi)= yM (xi, θ)+ b(xi)+ ǫi, i = 1,...,n, (2.6)

th where xi is a vector of known, user-controlled input variables corresponding to the i ob- servation, θ is the true value of computer model calibration parameters, yF (xi) denotes a scalar field observation, yM denotes computer model output, b is a bias term, and ǫi are ob- servation errors (Higdon et al., 2004). For simplicity, we assume that the computer model is an accurate, unbiased representation of reality (i.e., b(·) = 0). As is typical, we assume

2 measurement errors are independent and identically distributed as N(0,σF), so that all field observations follow a normal distribution, namely

2 yF (xi) ∼ N yM (xi, θ) ,σF (2.7)  Parameter calibration is the process of using field observations to estimate or cali- brate the unknown computer model parameters so that computer model output is consis- tent with field data. Calibration can be achieved by maximizing the likelihood of the field observations assuming (2.7), or in a Bayesian context, by specifying priors on unknown cal- ibration parameters in order to obtain a posterior distribution. For the Bayesian approach,

Markov Chain Monte Carlo can be used to simulate from the posterior distribution of θ 50 using the computer model directly if the simulation is cheap (Gilks et al., 1998, Higdon et al., 2004, for an example). Instead, we use the GP emulator yGP in place of the com- puter model yM in equation (2.6), and after removing the bias term, obtain

yF (xi)= yGP (xi, θ)+ ǫi, i = 1,...,n,

2 2 where the GP emulator depends on the hyperparameters (α,β,µ,σGP,σe ). The above assumption on the error term leads to a joint distribution of the field data vector yF ≡ (yF (x1),...,yF (xn)) and known computer model outputs zknown that is multi- variate normal. The posterior distribution of computer model parameters is given by

π(θ|yF ,zknown) ∝ f(yF |θ,zknown)π(θ), (2.8)

where π(θ) is a prior on computer model parameters, and f(yF |θ,zknown) is the multivariate normal likelihood function with a mean matrix given in equation (2.3), with µ replaced by

(j) µ1n×1 and r replaced by the matrix [R]ij ≡ cor z(θ),z(θ ) ,i = 1,...,n,j = 1,...,m, for   correlation function defined in equation (2.2), where z(θ) is (unobserved) mean computer

(j) th model output at the true parameter vector, and z θ is the j element of zknown. The

  2 2 variance-covariance matrix follows the form of equation (2.4), with σGP + σe replaced by 2 2 σGPJn×n + σFIn×n, where Jn×n is a unit matrix, and r replaced by R. The above equations assume scalar field data, i.e., observations from a single time point. When field data from multiple time points are independent, it is straightforward to calibrate against data from all time points using independent GPs to predict computer

(1) (mj ) model output at each time point. Let zknown,j = zj(θ ),...,zj(θ ) be a set of ob-   served computer model outputs at time point tj. For each zknown,j, j = 1,...,T , fit inde-

2 2 pendent GPs with hyperparameters αj,βj, µj,σGP,j,σe,j . Define yF,j =   51

yF,j(x1),...,yF,j(xnj ) to be a vector of independent field observations at time point tj,   where yF,j is independent of yF,k ∀j 6= k. Then given a collection of field data yF,1,...,yF,T , the posterior distribution of the true calibration parameter vector is

T π (θ|yF,1,...,yF,T ,zknown,1,...,zknown,T ) ∝ (f (yF,j|θ,zknown,j)) π(θ) (2.9) jY=1

2.3.5 Markov Chain Monte Carlo

Markov Chain Monte Carlo (MCMC) is a sampling method used to simulate from a probability distribution that is known up to a multiplicative constant (Gilks et al., 1998).

In computer model calibration, the goal is to obtain the posterior distribution (equation 2.8 or 2.9) for the computer model parameter vector θ, given a prior π(θ) on computer model parameters. In our computer model, θ = (θ1, . . . , θp). For convenience in the statistical analysis, we scale each component of θ from the biological ranges given in Table 2.6 to be between 0 and 1. We assume independent (discrete) uniform priors on components of θ p such that π(θ) = πi(θi), where πi,i = 1,...,p, is the Unif(0,1) probability density func- iY=1 tion for continuous θi, and the discrete uniform probability mass function for discrete θi. Let θ(d) be the dth draw from the posterior distribution of equation (2.8). Then the

Metropolis-Hastings algorithm (Chib & Greenberg, 1995) is used to simulate from

π(θ|yF ,zknown) as follows:

1. Select a starting value for θ(0)

2. Given the current realization θ(d−1), generate θ∗ according to a jumping or proposal

distribution, θ∗ ∼ J ·|θ(d−1)   3. Calculate the acceptance probability

∗ ∗ ∗ (d−1) f(yF |θ ,zknown)π(θ )/J(θ |θ ) (a) Let r = min (d−1) (d−1) (d−1) ∗ , 1 f(yF |θ ,zknown)π(θ )/J(θ |θ )   52

θ∗ with probability r (d) (b) Set θ =  (d−1)  θ with probability 1 − r

4. Repeat steps 2-3 until convergence and the desired number of posterior samples is ob- tained.

To simulate from the posterior distribution of equation (2.9), replace the likelihood function T f(yF |θ,zknown) with the joint likelihood (f (yF,j|θ,zknown,j)) in step 3(a). jY=1 As written above, the jumping distribution J(·|θ) is multidimensional. However, com- ponents of θ can be updated individually through the use of a Gibbs sampling algorithm

(Casella & George, 1992; Gelman et al., 2004). We use a Gibbs sampling algorithm to pro-

∗ (d−1) pose θi ,i = 1,...,p according to independent jumping distribution Ji(·|θi ) that is specifically Unif(0,1) for continuous θi and discrete uniform for discrete θi. We verify that the MCMC sampler has converged by running three independent chains of the sampler, us- ing overdispersed initial values of θ(0) for each chain, and by comparing within chain and between chain variances using the method of Gelman & Rubin (1992), as implemented in the R package coda (Plummer et al., 2006). In our analysis, we use the modular (i.e., em- pirical Bayesian) approach advocated by Bayarri et al. (2007), and fix the GP hyperparam-

2 2 eters (α,β,µ,σGP,σe ) at their maximum likelihood estimates so that the (joint) likelihood T function f(yF |θ,zknown) (or (f (yF,j|θ,zknown,j))) is fully specified. However, a Bayesian jY=1 treatment of these hyperparameters is also possible (for example, see Higdon et al., 2004).

2.3.6 Bayesian model comparison

Bayes factors (BFs) are commonly used for assessing the evidence in favor of one model against an alternative (Kass & Raftery, 1995). Specifically, we are interested in as- sessing the evidence in favor of model M1 against the nested null model M0. In both mod- els, the full computer model parameter vector θ = (Anecr, λ). The continuous parameter 53

vector λ ∈ Λ is common to both models. The single parameter Anecr ∈ {1,..., 6} in M1 and

Anecr =0 in M0.

For observed data y, the Bayes factor BF10 is the ratio of the marginal likelihood of the data under M1 to the marginal likelihood of the data under M0. Specifically,

f(y|M1) Λ f(y|λ, M1)π(λ|M1) dλ BF10 = = , f(y|M0) RΛ f(y|λ, M0)pi(λ|M0) dλ R where π(λ|Mi) is a prior on λ under model Mi, i = 1,2.

The relationship between the posterior odds in favor of M1 and the Bayes factor is usually written as

π(M1|y) π(M1) = x BF10 , (2.10) π(M0|y) π(M0) where π(Mi) is the apriori probability assigned to Mi, i = 0, 1, with p(M0) = 1 − p(M1). Rearrangemet of (2.10) yields π(M1|y) π(M0|y) BF10 = , π(M1) π(M0) and the Bayes factor is the ratio of the posterior odds of M1 to its prior odds. If the two models are apriori equally likely, which is what we assume, then the Bayes factor is the pos- terior odds in favor of M1. In our analysis, we estimate the probabilities π(Mi|y),i = 1, 2 from the posterior distribution of the computer model parameter vector θ that we obtain using MCMC. 54

2.4 Results

2.4.1 Experimental design and computer model output

Ideally we would like to calibrate all 25 parameters in our computer model, but be- cause field data is limited we will focus our attention on the five parameters we believe to be the most interesting and the most relevant given the field data available. Using data from Belkaid et al. (2000), our goal is to estimate the parameter vector θ =

(αI , kI ,βMrecr,pTrecr, Tthreshold) whose true value we assume exists somewhere in the hy- percube formed by allowing each parameter to vary in the ranges specified in Table 2.6 of the appendix. The remaining parameters are fixed at values given in the table. Note that the parameter Anecr is currently fixed at 2; we vary this parameter in section 2.4.6.

The parameters αI , kI , and Tthreshold are pathogen-dependent parameters which determine growth rate, the lethal parasite density of infected macrophages, and the threshold level of pathogen that triggers the adaptive immune response. βMrecr is a macrophage recruitment parameter and pTrecr determines the rate of T cell recruitment. We choose all parameter values and ranges to be consistent with L. major infection, and display references for these choices in Table 2.6.

In order to estimate θ, we must obtain computer model output for m representative choices of the parameter vector θ. In the absence of informative prior information about the true value of θ, the desired approach is to choose a space-filling design that adequately covers the parameter space. We choose 50 design points using a maximin Latin hypercube design (Morris & Mitchell, 1995) and run two replicates at each design point.

Output from the simulation consists of total macrophage and parasite counts on the lattice over time (Fig. 2.1, black dashed lines), and our primary objective is to estimate

θ so that our simulation output is consistent with the data of Belkaid et al. (2000) (Fig. 55

2.1, solid black circles). Variation in macrophage counts and pathogen load is evident in the values of and locations of the peaks. Peak macrophage levels range from 112 to 1903 macrophages, with peak locations ranging from 4.2 to 6.9 wpi. Peak pathogen load ranges from 3263 to 15610 (57.1 and 125 on the scale of Fig. 2.1B, with peak locations ranging from 4.5 to 7.0 wpi. In all of our simulations, infection is cleared by approximately 8 wpi.

In order to get a broad sense of the effect of computer model parameters on simula- tion output, we first calculate the main effects of the five model parameters on pathogen load over time, using the methods explained in sections 2.3.1.2 and 2.3.3.2, by fitting GPs to the six most important PC weights which explain 99.9% of the variation in the output.

Figure 2.2 illustrates the main effects of each parameter. Qualitatively, the most impor- tant parameters are growth rate (αI ), transfer threshold (kI ), and the threshold level of pathogen that triggers the adaptive immune response (Tthreshold). Interestingly, high growth rate favors pathogen load early in infection but not at later time points. When growth rate is high, there is a sharp increase in pathogen load and a higher peak, but a faster resolution of infection. Infection is prolonged when growth rate is low, although the peak is lower and the infection is eventually cleared. A similar dynamic is seen with transfer threshold, where high values favor the pathogen early in infection but not at later time points. Lastly, when

Tthreshold is high, peak pathogen levels are higher and the pathogen survives longer than when Tthreshold is low. However, the timing of the peak (∼5.5 wpi) is similar regardless of the value of Tthreshold.

2.4.2 Fitting the Gaussian processes

In the calibration process, we consider the following field data: macrophage counts at t1 = 5, t2 = 5.5, t3 = 6.5 and t4 = 8 wpi; and parasite load at t5 = 4.5 and t6 = 6.5 wpi(Fig. 2.1, solid black circles). Therefore, our GP emulators must be able to predict com- 56

∗ (i) puter model output at these six time points. Let zj θ be the untransformed computer   model output for simulation i at time point tj, and define

log z∗ θ(i) , j = 1,..., 5 (i) j zj θ =  (2.11) ∗ (i)    log zj θ + 1 , j = 6.      The simulation output is adjusted at t6 to prevent numerical issues for simulations where the infection is cleared by 6.5 wpi. We fit independent GPs to each vector zknown,j as de- scribed in section 2.3.1.1, for the six time points using the software PErK (Santner et al.,

2003). To check that the GP emulator accurately predicts computer model output, we cal- culate cross-validated predictions and cross-validated standardized residuals according to section 2.3.2 for the 100 computer model runs. Diagnostic plots that compare cross-validated predictions with observed values for all time points are provided in Fig. 2.3, which indicate that cross-validated predictions from the GP emulators are close to the observed values.

One exception is the GP emulator for time point t6, and this is discussed further in section 2.5.1. Nevertheless, at least 94% of the standardized cross-validated residuals fall between

±2 for all time points. For time points t1 and t2, at least 96% of the standardized residuals fall between ±2. These results are summarized in Table 2.1.

2.4.3 Sensitivity analysis

In order to quantify the importance of computer model parameters on simulation output and to characterize their effects, we perform sensitivity analysis as described in sec- tion 2.3.3.1. Because we are interested in calibrating the computer model to time points t1,

... t6, we perform sensitivity analysis on the GP predictors for those time points. The per- centage contributions of main and important two-factor interaction effects are provided in

Table 2.2. Together, main effects and two-factor interactions account for more than 97% 57 of the total functional variance of the GP emulators for all time points. For time points where main effects account for more than 80% of the total functional variance of the GP predictor, we also plot the most important main effects. Main effects for αI and kI at t1, and for kI , βMrecr, and Tthreshold at t3 are plotted in Fig. 2.4A. Main effects for αI and kI at t5 and t6, and for Tthreshold at t6 are plotted in Fig. 2.4B. Interaction effects for αI × kI and αI × Tthreshold are plotted in Fig. 2.5.

The most influential parameter for any time point is pathogen growth rate (αI ), which accounts for 93.5% of the total functional variance for the GP predictor at t5. Growth rate is also the most important parameter at t1 and t6, accounting for 54.1% and 54.6% of the total functional variance of the GP predictors, respectively (Table 2.2). For (scaled) val- ues of αI > 0.2, the relationship between pathogen growth rate and macrophage levels at

5.5 wpi (t1) is positive, whereas transfer threshold (kI ) is negatively related to the response

(Fig. 2.4A). However, these effects are not independent; the αI × kI interaction accounts for

11.3% of the total variance of the GP predictor (Table 2.2), and kI has little effect when αI is below 4 × 10−5 / minute (Fig. 2.5A).

Interestingly, the main effect of αI on pathogen load changes over time (Fig. 2.4B).

The parameter αI has a positive effect on pathogen load at 4.5 wpi but a negative effect on pathogen load at 6.5 wpi (t5 and t6, respectively). At t6, there is an interaction between αI and Tthreshold which accounts for 4.5% of the total functional variance of the GP predictor

−5 (Table 2.2), and Tthreshold has only a minor effect when αI is above 6.0 × 10 (Fig. 2.5B).

The parameter Tthreshold is the most important parameter at t3, accounting for 39.3% of the total functional variance of the GP predictor (Table 2.2), and has a strong positive relationship with the response (Fig. 2.4). Transfer threshold (kI ) is the most important pa- rameter at t2 and t4, accounting for 29.7% and 33.5% of the total functional variance, re- spectively (Table 2.2), and is negatively related with all macrophage time points (t1 through 58

t4; Fig. 2.4A and data not shown).

2.4.4 Calibration using simulated data

The amount of field data available for calibration is limited (e.g. macrophage data are representative of 10 experiments; Belkaid et al., 2000), and we do not have accurate information about field error. In order to verify that accurate calibration is possible us- ing the type of data available, we first apply the calibration approach to simulated field data. We use the ABM to independently simulate a single observation at each of the six

∗ time points using a randomly chosen but known parameter vector θ = (θ1, . . . , θ5), where

θi ∼ Unif(0.1, 0.9), for i = 1,..., 5. This truncated range is chosen to ensure that the prior

∗ distribution of θ (θi ∼ independent U(0,1) ∀i) adequately covers the true value θ . The time points and number of observations are the same as those that will be used in the cal-

2 ibration with real field data in section 2.4.5. In order to estimate field variance (σF) we as- sume that error from the measurement process is negligible in comparison to the inherent

2 variability of the biological system, and estimate σF for each time point by using the ABM to generate three independent observations for each time point and calculating the sample variance.

Calibration is carried out using the methods described in sections 2.3.4 and 2.3.5. For each of three independent MCMC chains, we obtain 200,000 samples with 10,000 sample burn-in to obtain the posterior distribution of the computer model parameter vector. A summary of the marginal posterior distributions for all parameters from the three combined chains is given in Table 2.3. The posterior distributions for αI , kI , βMrecr, and Tthreshold are shown in Fig. 2.6. The posterior distribution of pTrecr is relatively non-informative (Table 2.3 and data not shown).

The 90% posterior credible intervals for all five parameters capture the true parame- 59

ter vector (Table 2.3), and the marginal posterior distributions for αI , kI , and Tthreshold are centered around their true values (Fig. 2.6). Importantly, more precise information about model parameters is obtained. For example, the posterior 90% credible intervals for αI and kI are both more than 72% smaller than their prior 90% credible interval ranges (Table 2.3).

2.4.5 Calibration using field data

Belkaid et al. (2000) provide data plots for parasite and macrophage counts over time following low dose L. major infection in the ear. Because we are modeling the center of the infection, and not the entire ear, we scale these observations by a factor of 71−1, based on expected macrophage counts at two days post high dose infection (i.e., the start of our sim- ulation), and transform these observations as we did for simulation data following equation

(2.11). In the absence of field-estimated error, we assume, as we did in the calibration of

2 section 2.4.4, that σF is equal to the sample variance of observed computer model output. Posterior distributions were obtained using the same procedures used for simulated data in section 2.4.4. A summary of the resulting marginal posterior distributions is given in Table 2.4, and the distributions for αI , kI , and Tthreshold are plotted in Fig. 2.7. Using posterior medians as point estimates, we estimate a pathogen growth rate (αI ) of 7.73 ×

−4 10 / minute (a doubling time of 15 hours), and a transfer threshold (kI ) of 332 parasites.

The posterior distributions of βMrecr and pTrecr are relatively non-informative (Table 2.4 and data not shown). To determine whether our calibrated model could accurately repro- duce field observations, we ran 100 simulations using different parameter vectors randomly selected from the posterior distribution of θ. Graphs of mean computer model output, ± 1 standard deviation, along with the observed field data, are given in Fig. 2.8. All field ob- servations are within one standard deviation of mean computer model output from the cali- 60 brated computer model.

2.4.6 Comparison of alternative models of macrophage behavior

In the previous model, we arbitrarily fixed the parameter Anecr at a value of two in an attempt to capture the decrease in macrophage counts between 5.5 and 8 wpi that

Belkaid et al. (2000) observes. We next wanted to estimate the value of Anecr and to as- sess whether or not a model with necrotic tissue fits the data better than a model with- out necrotic tissue, and if so, whether or not the difference is statistically significant. For- mally, we consider the alternative model M1 : Anecr ∈ {1,..., 6} and the null model

M0 : Anecr = 0. Because the posterior distributions for αI , kI , and Tthreshold had poste- rior distributions concentrated at the high end of their ranges (Table 2.4 and Fig. 2.7), we updated our prior ranges for the parameter vector θ = (αI , kI ,βMrecr,pTrecr, Tthreshold) by shifting the range of each parameter such that its 99.99% highest posterior density credible interval is included in the new range. This shift did not effect the parameters βMrecr and pTrecr. The width of the ranges were kept the same in order to maintain the same level of conservativeness in our priors. Updated ranges for αI , kI , and Tthreshold and the range for

Anecr, are given in Table 2.7 in the appendix. We use a 70 point maximin Latin hypercube design to vary the updated computer model parameter vector θ = (Anecr, αI , kI ,βMrecr,pTrecr, Tthreshold), again with two replicate runs per design point. Because Anecr is discrete, we map its scaled values from the Latin

1 hypercube design { 70 ,..., 1} to the discrete values {0,..., 6} so that each possible value of

Anecr is observed 10 times in the design. For time points tj, j = 1,... 6, we fit independent

GPs to zknown,j and calibrate our computer model using the methods described in sections

2.3.4 and 2.3.5. We assume that the two models are apriori equally likely (i.e., π(M0) =

1 π(M1)= 2 ), and that 61

1 π(Anecr) = , Anecr = 0 2 (2.12) 1 12 , Anecr = 1,..., 6 and Anecr is independent of the other parameters. The remaining (scaled) parameters are again given independent Unif(0,1) priors.

Posterior distributions were obtained using the same procedures described in section

2.4.4. A summary of the resulting marginal posterior distributions is given in Table 2.5, and the posterior distribution of Anecr is given in Fig. 2.9. Using posterior medians as point

−4 estimates, we estimate a pathogen growth rate (αI ) of 7.38×10 / minute (a doubling time of 15.7 hours), a transfer threshold (kI ) of 317 parasites, and a production (Anecr) of three units of necrotic tissue following macrophage activation. The Bayes factor for M1 against

M0 is BF10 = 7.3 which indicates ‘substantial’ evidence in favor of the model with necrotic tissue against the model without necrotic tissue, using the scale of Kass & Raftery (1995).

The main effect of all computer model parameters (Anecr, θ) on macrophage counts over time is seen in Fig. 2.10. Notably, the sharpest decrease in macrophage levels after 5.5 wpi occurs when Anecr is high, while this decrease (i.e, the shape of the curve) is similar between low, medium, and high values of the other computer model parameters.

2.5 Discussion

In this paper, we have described an agent-based model of L. major infection, fit mul- tiple GPs to computer model outputs of pathogen load and macrophage counts at various time points, performed sensitivity analysis on the GP emulators to identify and character- ize the effects of important model parameters, and calibrated the computer model using the data of Belkaid et al. (2000). Sensitivity analysis indicates that pathogen growth rate and macrophage carrying capacity are important parameters that affect macrophage levels at 62

five wpi, though not independently; pathogen growth rate is positively related to pathogen load at 4.5 wpi but negatively related to pathogen load at 6.5 wpi, and the extent of the latter effect depends on the ability of the pathogen to avoid detection. Parameter estimates of computer model parameters indicate that L. major can replicate extensively before de- stroying the host cell. Model comparison indicates that a model with macrophage emigra- tion following uptake of necrotic tissue provides a substantially better fit to the data than a model without macrophage emigration. GP diagnostics, sensitivity analysis, calibration with simulated and field data, and model comparison are discussed in detail in the following section.

2.5.1 Gaussian process computer model approximation

The GP predictors of computer model output at time points t1,...,t6 accurately cap- ture the computer model output over the plausible parameter space (Fig. 2.3 and Table

2.1). Specifically, when uncertainty in the predictor is considered, as determined through analysis of cross-validated standardized residuals, the GP predictors accurately capture all known computer model outputs for all time points, except for time point t5, where 98% of the standardized residuals fall between ±3 (Table 2.1). The predictor for the first pathogen time point t5 is extremely precise, with (near) perfect correlation between observed com- puter model output and cross-validated predictions (Fig. 2.3, r = 1.00), likely because 93.5% of the total variation in output is explained by the single parameter αI (Table 2.2). Predic- tions from the GP emulator at the second pathogen time point t6 are less precise (Fig. 2.3, r = 0.84), but still reasonable. Despite the tendency for predicted values to be biased high when the observed computer model output is zero, all predictions are within an acceptable margin of error (Table 2.1), and predictions are generally accurate when the observed out- put is high. Parasite load at this late time point is inherently difficult to predict because 63 parasite clearance becomes possible. Since GP models are not appropriate for modeling such outcomes, it may be more reasonable to model the logit probability of clearance and, separately, parasite load conditional on non-clearance.

2.5.2 Sensitivity analysis

The FANOVA decomposition (Table 2.2) illustrates how uncertainty in computer model output can be attributed to uncertainty in computer model input. The uncertainty in computer model input is specified through the prior distribution of computer model pa- rameters. Using the language of Kennedy & O’Hagan (2001), the prior distribution may represent 1) current belief about the true value of a parameter (parameter uncertainty), or 2) environmental variability in a parameter that cannot be controlled (parametric vari- ability). Our choices of priors are primarily based on the former, although we also expect environmental variability to be present. Interestingly, for prior distributions that represent only parametric variability, e.g., the within-species variability of a particular parameter, sensitivity analysis will reflect how within-species differences can contribute to the variation often observed in biological studies. We expect the calibration process to provide the most information about parameters with large FANOVA contributions, whereas we expect to ob- tain little information about parameters with minor contributions to the overall functional variance of the GP emulators. For example, we expect a narrower 90% posterior credible in- terval for αI which accounts for 54.13% of the variation at the first macrophage time point t1 and 93.47% of the variation at the first pathogen time point t5, than for the parameter pTrecr, whose only notable contribution to functional variance is 2.46% at time point t6 (Ta- ble 2.2). In section 2.5.3 we discuss the idea of using sensitivity analysis to guide data col- lection so that model calibration is more efficient.

The main and interaction effects plots (Figs. 2.4-2.5) help illustrate the complex re- 64 lationship between pathogen behavior and host response. For example, the main effect of growth rate (αI ) is positive for pathogen load at 4.5 wpi (time point t5) but negative at

6.5 wpi (time point t6; Fig. 2.4A), indicating that although high growth rate may favor pathogen survival early in infection, it can be detrimental to long term survival. This para- dox results because a pathogen that replicates rapidly risks early detection and may quickly trigger the adaptive immune response. When growth rate is high, pathogen load increases and peaks quickly, and is quickly controlled, while slow growth rate leads to gradual in- crease in pathogen load and a longer infection (Fig. 2.2). In their model of Mycobacterium tuberculosis, Segovia-Juarez et al. (2004) report a similar finding (though for different rea- sons), observing that the sign of the partial rank correlation coefficient between growth rate and extracellular bacteria changes from positive to negative over time.

There are many parameters, both pathogen- and host-related, that influence the amount of macrophages at 5.5 and 6.5 wpi (time points t1 and t3, respectively). One surprising result is that for both time points, the macrophage recruitment parameter βMrecr is not the most influential (Table 2.2), despite being the only parameter that directly influences macrophage levels. This result is partly explained by the relatively large uncertainty ranges for some input parameters, such as αI , though the sensitivity analysis also identifies im- portant system dynamics that we had not necessarily anticipated. For example, Tthreshold, which determines the threshold pathogen level that triggers the T cell response, is the most important parameter at macrophage time point t3, accounting for 39.3% of the overall func- tional variance of the GP predictor (Table 2.2). Although the primary role of T cells is in macrophage activation, activated macrophages do not live very long, and they also leave behind necrotic tissue that triggers nearby resting macrophages to depart for the draining lymph node (see section 2.2.2.5). Also, as the infection resolves, cytokine levels decrease, re- ducing macrophage recruitment. In combination, T cell actions and therefore the parameter 65

Tthreshold have a significant effect on macrophage numbers (Fig. 2.4A).

At macrophage time point t1, growth rate (αI ) is the most important parameter, ac- counting for 54.1% of the variation of the GP predictor. When αI is greater than 0.2 (Fig.

−4 2.4A, αI = 3.9 × 10 / minute on its biological scale, or a doubling time faster than 30 hours), the number of macrophages is expected to increase sharply as growth rate increases, probably because rapid growth allows infected macrophages to reach transfer threshold (kI ) quickly, causing the infection to spread to more macrophages that each release chemokine to boost macrophage recruitment (section 2.2.2.4). Because an increase in macrophage re- cruitment is dependent on the infection spreading to other cells, low growth rates below a certain threshold (i.e., below 0.2, or 3.9 × 10−4 / minute on its biological scale) lead to little spread by time point t1, and little impact on macrophage counts (Fig. 2.4A). The small decrease in the main effect of αI towards its lower range is likely an artifact due to the stochasticity of the ABM, and the fact that we only consider two replicates for each de- sign point when fitting the GP.

We also expect transfer threshold kI to effect infection spread and consequently, macrophage recruitment. When kI is low, macrophage death and pathogen transmission is more frequent, leading to increased macrophage recruitment. Such a relationship ap- pears in the main effect plot (Fig. 2.4B). However, the role of parameters αI and kI on macrophage recruitment is complex. Their two-way interaction accounts for 11.3% of the total functional variance of the GP predictor (Table 2.2), with kI having only a minor ef-

−5 fect on macrophage counts when growth rate (αI ) is slow (< 3.9 × 10 / .1 min) or fast (> 8.0 × 10−5 / .1 min), corresponding to a doubling time slower than 30 hours or faster than 15 hours (Fig. 2.5A). For moderate values of growth rate, the value of kI is also im- portant for predicting macrophage counts, and although macrophage numbers increase with αI , they decrease as kI increases. Therefore, the effect of moderate growth rate on 66

macrophage levels can be offset by a change in kI .

At the last pathogen time point t6, a strong interaction exists between growth rate

(αI ) and the threshold level of pathogen that triggers the adaptive immune response (Tthreshold; Fig. 2.5B). In the computer model, we assume that antigen presentation in the lymph node is a function of pathogen load, and the adaptive immune response is triggered once a thresh- old pathogen level is reached. At this time point, expected pathogen load decreases as αI increases. However as Tthreshold increases (i.e., the pathogen becomes more efficient at avoid- ing detection), the infection is prolonged and expected pathogen load is higher at the later time point t6. We now consider these dynamics in the context of the evolution or co-evolution of pathogen genes and pathogen-host relationships. There are many mechanisms utilized by parasites to escape immune detection in their hosts (reviewed in Damian, 1997). For exam- ple, L. major is known to inhibit antigen presentation in infected macrophages (Fruth et al.,

1993), and Toxoplasma gondii, another parasite that infects macrophages, downregulates

MHC class II expression, which is required for antigen presentation (Luder et al., 1998).

Given the above relationship between growth rate (αI ) and sensitivity to immune detection

(Tthreshold), we expect that a pathogen efficient at escaping immune detection will be less evolutionarily constrained at genes related to growth. In this case, the αI ‘gene’ would be free to evolve. In general, we expect a positive correlation between growth rate and immune evasion phenotypes.

2.5.3 Calibration using simulated data

By testing parameter calibration with simulated data, it is clear that we can accu- rately estimate computer model parameters using the type of field data available. The pos- terior credible intervals for all five parameters capture the true parameter vector (Table 67

2.3), and posterior distributions for αI , kI , and Tthreshold are centered around their true val- ues (Fig. 2.6). Importantly, more precise information about model parameters is obtained.

Whereas we start with uninformative prior information (∼Unif(0,1)) for all parameters, the posterior 90% credible interval for αI , for example, is between 0.491 and 0.701, and has a width that is 76.6% narrower than the prior 90% credible interval.

The posterior distribution for pTrecr is the least informative (Table 2.3), but not sur- prising given the weak contribution of pTrecr to the functional variance of the GP emulators

(2.5% at time point t6; < 1% for all other time points; see Table 2.2). Ideally, both the type of data and time points for data collection should be chosen so that the information ob- tained about model parameters during calibration is maximized. Using the computer model and applying sensitivity analysis techniques, one can identify time points and responses that depend significantly on the choice of model parameter values (i.e., where FANOVA contributions are large). For example, pTrecr accounts for 18.4% of the overall functional variance of a GP predictor for T cell counts at approximately 6 wpi. When a single obser- vation at this time point is included in the calibration process, the posterior distribution of pTrecr has a sharper peak, and the width of the 90% credible interval is 6.2% narrower than the interval obtained using only macrophage counts and pathogen load (data not shown).

This observation suggests a practical use for GP-based sensitivity analysis as a guide in data collection, particularly when data generation is expensive, for efficient calibration of complex computer models.

2.5.4 Calibration using field data

Computer modelers should be cautious in the interpretation of calibrated model pa- rameters (for an important discussion of this issue, see Kennedy & O’Hagan, 2001). The es- sential point is that since no model is correct, the ‘true’ value of a parameter is likely to dif- 68 fer from its calibrated (i.e., best fitting) value. However, this is true of all models, whether mathematical, computational, or statistical, and our intent is to use the calibrated com- puter model as a measure of our systems-level understanding of parasitic infection and the host immune response. If our prior parameter ranges are reasonable and the model is cor- rect, then the posterior distributions reflect, to the best of our knowledge, the ‘true’ values of model parameters. Posterior estimates that are surprising or interesting may warrant special attention, and should be further analyzed through biological experiments. Transfer threshold is one such parameter in our model. Relatively little is known about its in vivo value. What follows is a discussion of the calibration results of section 2.4.5, where we let θ vary according to Table 2.6. However, the discussion is applicable to the calibration results of section 2.4.6, since posterior estimates are similar in both calibrations (Tables 2.4 and

2.5). We discuss model comparison specifically in section 2.5.5.

Despite the limited amount of field data used, we were able to obtain information about most of the computer model parameters we estimated. For example, the width of the 90% posterior credible intervals for αI and kI are approximately 74 and 64% of their prior 90% credible intervals, respectively (Table 2.4). The high posterior median for transfer threshold (kI = 332) suggests that L. major is capable of extensive intracellular replication

−4 without destroying the host cell, while the median growth rate (αI = 7.73 × 10 / minute, or a doubling time of 15 hours), is relatively slow. These properties, which enable a par- asite to replicate for extended periods in the absence of immune-mediated pathology, are believed fundamental to the microbial virulence of L. amazonensis (Chang et al., 2002), and may partially explain the silent phase of L. major growth observed by Belkaid et al. (2000), although our calibration results did not consider this slow phase.

The calibrated computer model predicts pathogen and macrophage observations quite well, although some aspects of macrohage emigration are likely incorrect. When mean com- 69 puter model behavior is considered (and under the assumption that field data is unbiased), predictions from the calibrated computer model are accurate for macrophages at 5.5 wpi, and for pathogen load at 4.5 and 6.5 wpi (time points t1, t5, and t6), but appear biased high for macrophages at 6.5 and 8 wpi (time points t3 and t4) (Fig. 2.8). These latter two time points correspond to macrophage time points after peak infection, suggesting that at least one component of inflammatory macrophage behavior is missing or incomplete. Specif- ically, although our model accounts for a decrease in macrophage numbers as the infection is resolved, this decrease is not as extreme as that observed by Belkaid et al. (2000). In this calibration we assume that following macrophage activation, two units of necrotic tissue are released, each is taken up by a nearby macrophage, and the two macrophages are removed.

These results (also see section 2.5.5) suggest that this model of macrophage emigration is too mild. Indeed, in vivo experiments involving peritonitis and the adoptive transfer of in-

flammatory macrophages show that inflammatory macrophages are no longer detectable

96 hours following the resolution of inflammation (Bellingan et al., 1996). The exact sig- nals and mechanisms for macrophage emigration are unclear, but macrophage emigration could be triggered by long-range signals, such as decreasing cytokine levels, in addition to local phagocytotic events. In an earlier version of the model that assumed a constant macrophage recruitment rate, we modeled this behavior by arbitrarily assigning a short macrophage life span to inflammatory macrophages (Dancik et al., 2006).

In the calibration process, we have made the standard assumption on measurement

2 error i.e., ǫi ∼ N(0,σF) , which implies that field measurements are not biased. This as- sumption is generally reasonable in practice, though importantly it may not hold for certain types of biological data, such as parasite counts, that are estimated through serial dilution assays. Typically, biologists use ad hoc methods to estimate concentrations in this manner, but more precise estimates with corresponding standard errors can be obtained through for- 70 mal statistical treatment of the data (Gelman et al., 2004; Mehrabi & Matthews, 1995).

Accurate modeling of field data and measurement error is essential for accurate calibration, especially when measurement bias and/or deviation from normality are expected.

In our calibrations, because the raw replicated field data were not available, we esti- mated field variance based on the stochasticity of the computer model. Although Belkaid et al. (2000) do not provide the raw data, they do provide error bars for the two pathogen time points in their Fig. 1. Although it is very difficult to get accurate estimates of stan- dard deviation from these graphs, it appears that the standard deviations from the paper are larger than those estimated from the computer model. The difference in variance esti- mates is not surprising, since variability observed by Belkaid et al. (2000) is a combination of both measurement error and inherent variability of the biological system. Furthermore, we fix all parameter values when estimating variability, whereas immunologists work with multiple pathogen and host organisms that may behave differently (i.e., have different pa- rameter values) even if the organisms are genetically identical and their environments are the same (Wong et al., 2005). If our estimates of field variance are lower than the truth, our posterior distributions will be misleadingly narrow. For these reasons, it is important that replicate field observations be available, so that field variance can be estimated directly, or handled in a Bayesian context as part of the calibration process. Including replicate obser- vations in calibration will also decrease the amount of uncertainty in our posterior estimates of model parameters. In our calibration with field data, using three replicate observations for each time point instead of a single observation results in an additional 8 and 9% de- crease in the length of the 90% posterior credible intervals for αI and kI , respectively. Lastly, we note that although Belkaid et al. (2000) provide data for up to 10.5 wpi, we have chosen not to use any data after 8 wpi in our analysis. After 8 wpi, despite the fact that the infection is controlled, low levels of parasite persist (Belkaid et al., 2000), an ob- 71 servation that the current model cannot reproduce. It is now known that regulatory T cells

(which we do not model) are required to maintain this low-level parasite persistence, which also provides host immunity to re-infection (Belkaid et al., 2003). Without modeling regu- latory T cells, sterile cure (i.e., complete parasite elimination) is always achieved (data not shown), an outcome consistent with depletion studies following L. major infection (Belkaid et al., 2003). We expect that inclusion of regulatory T cell function into our model will decrease the rate of parasite clearance, achieve a better fit at t6, and capture the low level parasite persistence (after 8 wpi) that we currently do not achieve in simula- tion.

However, the nature of pathogen persistence is curious in light of our posterior esti- mate of kI . Specifically, Belkaid et al. (2000) observe the persistence of 100-10,000 parasites in the skin. Our results suggest that at the site of infection, pathogens survive in a very small population of infected macrophages (between 1 and 30), independent of pathogen that persists in fibroblasts in the lymph node (Bodgan et al., 2000). The in vivo value of kI is not known, but Belkaid et al. 2000 observe ”heavily infected macrophages” during the per- sistent stage of infection.

2.5.5 Model comparison and macrophage emigration

In our computer model, we have proposed that macrophage activation results in the production of Anecr units of necrotic tissue, and that macrophages emigrate to the drain- ing lymph node following uptake of necrotic tissue. This mechanism of macrophage emi- gration is supported by the high Bayes factor (BF10 = 7.3) that indicates ‘substantial’ ev- idence in favor of a model where necrotic tissue is produced (i.e., macrophage emigration occurs) against a model where necrotic tissue is not produced (i.e., where macrophage emi- gration does not occur). Although it is possible to use other macrophage recruitment func- 72 tions than the one we have used (Equation 2.1), the decrease in macrophage numbers ob- served by Belkaid et al. (2000) cannot be explained by ceasing macrophage recruitment af- ter 5.5 wpi unless macrophage lifespan is changed (data not shown). Our posterior estimate of Anecr is three, which is one unit higher than the value of Anecr used in the calibration of section 2.4.5. We expect that fixing Anecr at three, rather than at two, would have little ef- fect on macrophage levels before 5.5 wpi but would lower the peak number of macrophages and increase the macrophage depletion rate after 5.5 wpi (Fig. 2.10).

The large amount of uncertainty in our posterior estimate of Anecr (Table 2.5) is the result of assuming that both models are apriori equally likely, i.e, π(Anecr =0)= π(Anecr ∈

1 1 {1,..., 6}) = 2 . Using a uniform prior π(Anecr) = 7 , Anecr ∈ {0,..., 6} gives a 90% poste- rior credible interval of (1,4) and has little effect on the Bayes factor (data not shown).

It is also possible that inclusion of regulatory T cells in the model would change the posterior estimate of Anecr. However, the inclusion of regulatory T cells would facilitate pathogen survival, increase macrophage recruitment, and therefore require a stronger macrophage emigration mechanism (e.g. Anecr > 3) in order to capture the observed decrease in macrophage counts. Interestingly, macrophage take-up and presentation of necrotic tissue (i.e., self- antigen) may partially explain the increase in natural regulatory T cells observed at the infection site during resolution of the infection (Cozzo et al., 2003; Mendez et al., 2004). Al- though L. major-specific natural regulatory T cells have been observed (Suffia et al., 2006), it is unclear why regulatory T cell levels do not increase until the infection is nearly re- solved (Mendez et al., 2004).

Mathematical models of disease generally assume that the average lifespan (or nat- ural death rate) of host cells are constant (a notable exception are models attempting to explain CD4+ T cell depletion in HIV (Perelson et al., 1992; Bernaschi & Castiglione, 2002;

Yates et al., 2007)). Our results indicate that in L. major infection, a change in macrophage 73 behavior (e.g., emigration following uptake of necrotic tissue) is necessary to account for the decrease in macrophages observed after 5.5 wpi. This emigration of inflammatory macrophages is not specific to L. major infection (Bellingan et al., 1996), and we would expect similar macrophage behavior in other infections where necrotic tissue (or more generally a high level of self antigen) is present. In mathematical models, because the reproductive num- ber R0 is inversely related to the natural death rate of the host cell, and when R0 > 1 peak pathogen levels depend on the number of host cells available (Wodarz & Nowak, 2002), it may be important to take macrophage emigration into account when modeling infection with other macrophage-tropic organisms, such as M. tuberculosis and T. gondii.

2.6 Conclusions and future work

In this work we described an agent-based model of L. major infection, used a GP to approximate and analyze the computer model, and estimated six immunological and pathogen-related parameters using field data. Simulations from the calibrated model are generally consistent with biological data, although more field data is needed for further val- idation. In the future we will incorporate regulatory T cells into our model in order to cap- ture the low-level parasite persistence that has been observed (Belkaid et al., 2003). We will continue to use model comparison methods to compare alternate models of regulatory T cell recruitment, as well as inflammatory macrophage behavior.

Acknowledgements

GMD and KSD were partially supported by PHS grant GM068955, USDA IFAFS

Multidisciplinary Graduate Education Training Grant (2001-52100-11506), and an ISU

CIAG Research Support Grant. 74

2.7 Appendix: Agent-based model parameters

Range / Class Param Description Value Unit Ref λ Chemokine diffusion coef- 0.64 / .1 min 1 ficient Chemokine γ Chemokine degradation 0.001 / .1 min 1 coefficient CR Minimum amount of 1 scalar 1 chemokine necessary to influence cell movement cI Amount of chemokine 5000 units / 1 released by infected and time step activated macrophages P0 Initial number of parasite 50 parasites e −5 α Intracellular parasite (2.41, 9.63) ×10 2,e Leishmania I /.1min growth rate kI Transfer threshold of (50,400) parasites e infected macrophages Tthreshold Amount of total pathogen (1500,7000) scalar e that triggers the T cell response Tactm Probabilty a T cell will 100 % e activate a macrophage Mrls Resting, non- 100 days 1 inflammatory macrophage lifespan M Activated macrophage 2 days e Macrophages als lifespan Mrsp Resting macrophage speed 1.0 µm/ min 1 Masp Activated macrophage 0.25625 µm/ min 1 speed Misp Infected macrophage 0.0007 µm/ min 1 speed Minit Initial number of 105 scalar 1,e macrophages 75

Range / Class Param Description Value Unit Ref

Pm Maximum number of par- 4 scalar e asites per initially infected macrophage Anecr Amount of necrotic tis- 2 scalar e sue released following macrophage activation T T cell lifespan 3 days 1 T cells ls Tsp Tcellspeed 2 µm/ min 1 βMrecr Macrophage recruitment (0.0002, see sec 2,e hyperparameter 0.0003) 2.2.2.4 α Macrophage recruitment -5.66 see sec 2,e Recruitment Mrecr hyperparameter 2.2.2.4 mMrecr Macrophage recruitment 0.005 see sec 2,e hyperparameter 2.2.2.4 pTrecr Probability of T cell re- (0.0150,0.025) % /time 2,e cruitment step CM Chemokine threshold Chemokine scalar 2,e for logistic macrophage levels at recruitment source 1 day post simulated infection CT Chemokine threshold for 1 scalar 1 T cell recruitment Miscellaneous Update Time scale for additional 10 minutes e interactions

Table 2.6: ABM parameters and values used in initial calibration. Reference codes: 1. Segovia-Juarez et al. (2004) and references within; 2. Belkaid et al. (2000); e. estimate. Parameters αI , kI , βMrecr, pTrecr, and Tthreshold are calibrated using field data. 76 Sqrt(Pathogen Load) Simulated Macrophages 0 20 40 60 80 100 120 0 500 1000 1500

4 5 6 7 8 4 5 6 7 8

A.Weeks Post Infection B. Weeks Post Infection

Figure 2.1: Simulated output over time (black dashed lines), and the field data (solid black circles) used in calibration. A. Simulated macrophages; field data for time points t1 through t4. B. Square root of simulated pathogen load; field data for time points t5 and t6. Simulated data was obtained from a 50 point Latin hypercube design with two replicates per design point, over the parameters αI , kI , βMrecr, pTrecr, Tthreshold. 77

h h h α I h h h h k I h h T threshold h p Trecr m h h β Mrecr h mh m mmh hmhl mhl h h l hl h l m mm l mhl mhl l h l ml l l = low h l l hmhl l l m = medium h m l m l h l mhl hm h = high l l mhl m l l l h l ml h l l mhl l

Pathogen load m l l ml l hmmh l h h mhl l l h ml l h m h l mhl l m l hmhm h l mhl l l h l h hmhl l h mh l mm l hmhl l l h hmhl l l l h m l h l l l mh hmhl l h mhl l hmh hmhl l l l h l l l mhmmhl l hl mhl l l l h hmhmhmhl mhl hl l mhl l l l l l l l hl hl mhl mhl mhl mhl h l 0mhl l l 2000 4000 6000 8000 hl mhl mhl mhl mhl mmmmmmhlh l h l h l h l h l

4 5 6 7 8

Weeks post infection

Figure 2.2: Main effects of computer model parameters on simulated pathogen load over time. Main effects are calculated using low (l), medium (m), and high (h) val- 10−4 ues of each parameter, with (l,m,h) values of αI : (2.41, 6.02, 9.63) × min ), kI : (50, 175, 400) parasites, Tthreshold: (1500, 4250, 7000) parasites, pTrecr: (0.0150, 0.0200, 0.0250) % / time step, βMrecr: (0.0002, 0.00025, 0.0003) / unit chemokine. 78

t1: Macrophages (5 wpi) t4: Macrophages (8 wpi) observed observed

5.0 6.0 r = 0.87 r = 0.89 3 4 5 6 7

5.0 5.5 6.0 6.5 4.0 4.5 5.0 5.5 6.0 6.5 7.0

predicted predicted

t2: Macrophages (5.5 wpi) t5: Pathogen (4.5 wpi)

observed r = 0.91 observed r = 1.00 4.5 6.0 5.5 7.0 8.5 4.5 5.0 5.5 6.0 6.5 7.0 5.5 6.0 6.5 7.0 7.5 8.0 8.5

predicted predicted

t3: Macrophages (6.5 wpi) t6: Macrophages (6.5 wpi) observed observed

4 5 6 7 r = 0.90 r = 0.84 0 2 4 6 8

4.0 4.5 5.0 5.5 6.0 6.5 7.0 0 2 4 6 8

predicted predicted

Figure 2.3: Gaussian process diagnostics. Observed computer model output vs. cross- validated predictions using the Gaussian process emulators at time points t1 through t6. Cross-validated predictions are plotted against the corresponding observed values (open circles) and each observed value is plotted against itself (black lines). The value r measures the correlation between observed values and corresponding cross-validated predictions. 79

Time point Macrophage (wpi) Pathogen (wpi) Time point t1(5) t2(5.5) t3(6.5) t4(8) t5(4.5) t6(6.5) ±2 0.96 0.98 0.95 0.94 0.95 0.94 ±3 1.00 1.00 1.00 1.00 0.98 1.00

Table 2.1: Gaussian process diagnostics. For each time point, proportion of cross- validated standardized residuals that fall between ±2 and ±3.

Time point Macrophage (wpi) Pathogen (wpi) Effect t1(5) t2(5.5) t3(6.5) t4(8) t5(4.5) t6(6.5) αI 54.13 24.74 1.32 3.08 93.47 54.61 kI 18.68 29.73 36.65 33.53 5.09 12.44 βMrecr 9.76 6.48 6.09 8.44 - 1.68 pTrecr ---- - 2.46 Tthreshold 4.15 17.48 39.25 32.08 - 23.32 αI × kI 11.27 11.30 5.42 5.04 1.31 - αI × Tthreshold 1.43 5.48 - - - 4.47 kI × Tthreshold - 2.50 - 15.12 -- Total 99.42 97.72 97.30 97.40 99.88 98.98

Table 2.2: FANOVA decomposition. For each time point, the percentage contribution of main effects and two-way factor interaction effects to the total functional variance of the Gaussian process predictor. Only effects accounting for > 1% of the total variance for at least one time point are shown, with ‘-’ indicating < 1%.

% decrease in width from Posterior estimates prior 90% credible interval Parameter True value Median 90% credible interval to posterior 90% credible interval αI 0.6267 0.5847 (0.4906,0.7009) 76.63 kI 0.1037 0.1421 (0.0306, 0.2785) 72.46 βMrecr 0.6940 0.5374 (0.1575, 0.9348) 13.63 pTrecr 0.2782 0.5179 (0.0569, 0.9508) 0.68 Tthreshold 0.2957 0.4668 (0.1951, 0.8940) 22.34

Table 2.3: Posterior summary of computer model parameters following cali- bration using simulated field data. Posterior summary of θ, averaged over 3 chains of 200,000 samples each, including a 10,000 sample burn-in, compared to the true value. All parameters are scaled to be between 0 and 1 and have independent Unif(0,1) priors. 80

kI (t3) Tthreshold(t3) αI (t5)

αI (t6)

kI (t5) βMrecr(t3)

Tthreshold(t6) kI (t1) predicted output predicted output

kI (t6) 2 4 6 8

αI (t1) 5.0 5.5 6.0 6.5

0.0 0.4 0.8 0.0 0.4 0.8

A.parameter value B. parameter value

Figure 2.4: Main effects of the most influential parameters. A. Log macrophages at 5.0 (t1) and 6.5 (t3) wpi. B. Log pathogen load at 4.5 (t5) and 6.5 (t6) wpi. The biological ranges of model parameters, corresponding to (0,1), are as follows: −5 −5 αI , (2.40705 , 9.6317 ) / .1 min.; kI , (50, 400) parasites; βMrecr, (0.0002, 0.0003) / unit chemokine; Tthreshold, (1500, 7000) parasites. 81 I k threshold T 2000 3000 4000 5000 6000 7000 50 100 150 200 250 300 350 400

4e−05 8e−05 4e−05 8e−05

−5 −5 A.αI (×10 ) B. αI (×10 )

Figure 2.5: Two-way interaction plots. A. Log macrophages at 5 wpi (t1) for parame- ters αI and kI ; B. Log pathogen load at 6.5 wpi (t6) for parameters αI and Tthreshold. Units are rate / .1 minutes for αI and parasites for kI and Tthreshold. 82

αI kI Frequency Frequency 0 20000 50000 0 50000 150000

0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0

βMrecr Tthreshold Frequency Frequency 0 20000 40000 0 20000 50000

0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0

Figure 2.6: Posterior distributions of αI , kI , βMrecr, and Tthreshold following calibration using simulated field data. The dashed black lines correspond to the 90% posterior credible intervals; the true value of each parameter is represented by a black dot- ted line. The solid black lines that surround the histograms indicate the a priori parameter ranges, from prior distributions that are independent Unif(0,1) for all parameters. 83

% decrease in width from prior 90% credible interval to Posterior estimates posterior 90% Parameter Median 90% credible interval Units credible interval 7.73 (7.24, 8.88) / .1 min. (×10−5) αI 1.50 (1.30, 1.60) doubling time (hours ×101) 74.73 2 kI 3.32 (2.70, 3.84) parasites (×10 ) 63.79 βMrecr 2.31 (2.03, 2.81) / unit chemokine (×10−4) 13.58 −2 pTrecr 1.93 (1.54, 2.43) % / time step (×10 ) 0.79 −4 Tthreshold 5.57 (3.92, 6.82) parasites(×10 ) 41.46

Table 2.4: Posterior summary of computer model parameters following calibra- tion using field data. Posterior summary of θ, averaged over 3 chains of 200,000 samples each, including a 10,000 sample burn-in. All parameters are scaled back to their biological values and are reported in the units indicated.

% decrease in width from prior 90% credible interval to Posterior estimates posterior 90% Parameter Median 90% credible interval Units credible interval 7.38 (6.97, 8.07) / .1 min. (×10−5) αI 1.57 (1.43, 1.66) doubling time (hours ×101) 83.05 2 kI 3.17 (2.26, 3.89) parasites (×10 ) 48.38 βMrecr 2.58 (2.06, 2.94) / unit chemokine (×10−4) 2.38 −2 pTrecr 1.86 (1.53, 2.41) % / time step (×10 ) 2.81 −4 Tthreshold 5.27 (2.53, 6.80) parasites(×10 ) 13.61 Anecr 3.00 (0.00, 6.00) units of necrotic tissue 0

Table 2.5: Posterior summary of computer model parameters following calibra- tion using field data. Posterior summary of (Anecr, θ), averaged over 3 chains of 200,000 samples each, including a 10,000 sample burn-in. All parameters are scaled back to their biological values and are reported in the units indicated. 84

αI Frequency 0 30000

4e−05 6e−05 8e−05

kI Frequency 0 60000

50 100 150 200 250 300 350 400

Tthreshold Frequency 0 20000

2000 3000 4000 5000 6000 7000

Figure 2.7: Posterior distributions of αI , kI , and Tthreshold following calibration using field data. The dashed black line corresponds to the 90% posterior credible interval; the median value of the parameter is represented by a thick black dotted line. The solid black lines that surround the histograms indicate the a priori parameter ranges, according −5 −5 to independent prior distributions: αI ∼ Unif(2.407 × 10 , 9.632 × 10 ) /.1 minutes), kI ∼ Unif(50,400) parasites, and Tthreshold ∼ Unif(1500,7000) parasites. 85 Sqrt (Parasite Load) Number of Macrophages 200 400 600 800 0 20 40 60 80 100

4 5 6 7 8 4 5 6 7 8

A.Weeks Post Infection B. Weeks Post Infection

Figure 2.8: Posterior predictive distribution of macrophage counts (A) and par- asite load (B). One hundred simulations were run using randomly selected parameter vectors from the posterior distribution of θ. Solid black lines, mean computer model output ± 1 standard deviation; solid black circles, field data (A, time points t1 through t4; B, time points t5 and t6) used during calibration; vertical dashed black lines indicate ± 1 standard deviation corresponding to estimated field variance. 86

BF10 = 7.3 0 50000 100000 150000 200000

0 1 2 3 4 5 6 Anecr

Figure 2.9: Posterior distribution of Anecr following calibration using field data. Posterior distribution is averaged over 3 chains of 200,000 samples each, including a 10,000 sample burn-in. BF10 , Bayes factor of M1 : Anecr ∈ {1,..., 6} against M0: Anecr= 0, αI n = 2. 87

l l l l l l l l l hl hh l l l h l l l h h l l hl l h l l l hl l h h l l l l l l l l l l hl h l l hl l l l l l l hhmhmm h l h l l hlmlmmhlml α I l l l m mhml l h l h l l hm l ml k I ml hh mhm l h l h hmhhh mml h l l h l h hml h T threshold lmh l l l h mmml h l h hml l ml h l h p Trecr hmh l l h mhmml hh l hh hl mml h l h l h l hhh mh ml h l hh βMrecr ml l l hmml hhl h hmhl l h l hhl m mmml hl h A necr l h h h mmlml hl hhh h l h l hhl hmh mmlmlm l hl hhh h l l hl hmh mmlmm l hhhhh mh l hl h h mlmlmm l l hl hhhhh Macrophage count l h hmh h ml m h hmhl l hmh hh l lmlmlmmml l l hhhhhh l l h l l mhh hhh l l lmlmmmmml l l hhhhhh h l hl l l lmhhh hhhhl l lmlmlmmmmlml l hhhhhh h l l lmlmhhhhh hhl hl hl l lmlmmmmlmlmlml hh mhl h h l l l lmlml hhhhhh hhl hl hl hl l lmlmmmlml hml h l l l l mlmlml l hhhhhhhhl hl hl hl hl hl lml hl l l h l l l lmlmlmlml l l hhhhhhhhl hl hl mhl h h l l lmlmlmlml l l l l hhhhh hmhl l h lmlmlmlmlmlml l l l 200 400hmhl hl 600 800 1000 h mmlml hmhlmhl hh mhlmhlmhl l hh mhlmhl hl hhh mhl l hhhhhh hhhhhhhhhhh

4 5 6 7 8 9 10 11

Weeks post infection

Figure 2.10: Main effects of computer model parameters on simulated macrophage counts over time. Main effects are calculated using low (l), medium (m), and high (h) values of each parameter, with (l,m,h) values of αI : (6.00, 9.66, 13.32) 10−4 × min ), kI : (100, 275, 450) parasites, Tthreshold: (2500, 5250, 8000) parasites, pTrecr: (0.0150, 0.0200, 0.0250) % / time step, βMrecr: (0.0002, 0.00025, 0.0003) / unit chemokine; and Anecr: (0, 3.5, 6) units of necrotic tissue. 88

Class Param Description Range Unit −5 α Intracellular parasite (6.00, ×10 I /.1min growth rate 13.32) Leishmania kI Transfer threshold of (100,450) parasites infected macrophages Tthreshold Amount of total pathogen (2500,8000) scalar that triggers the T cell response Macrophages Anecr Amount of necrotic tis- (0,6) scalar sue released following macrophage activation

Table 2.7: Updated ranges for ABM parameters following initial calibration. All other parameters are fixed or varied as specified in table 2.6. 89

Bibliography

An, G. (2002). Agent-based computer simulation and SIRS: building a bridge between basic science and clinical trials. Shock 16, 266–273.

Badolato, R., Sacks, D. L., Savoia, D. & Musso, T. (1996). Leishmania major: in- fection of human monocytes induces expression of IL-8 and MCAF. Exp. Parasitol. 82, 21–26.

Bayarri, M., Berger, J. O., Cafeo, J., Garcia-Donato, G., Liu, F., Palomo, J., Parthasarathy, R. J., Paulo, R., Sacks, J. & Walsh, D. (2006). Computer model validation with functional output. Technical Report Number 165, National Institute of Statistical Sciences 70, 102.

Bayarri, M., Berger, J. O., Higdon, D., Kennedy, M., Kottas, A., Paulo, R., Sacks, J., Cafeo, J., Cavendish, J. & Tu, J. (2002). A framework for the validation of computer models. In: Proceedings of the Workshop on Foundations for V&V in the 21st Century (Pace, D. & Stevenson, S., eds.). San Diego, CA: Society for Modeling and Simulation International.

Bayarri, M., Berger, J. O., Paulo, R., Sacks, J., Cafeo, J. A., Cavendish, J., Lin, C.-H. & Tu, J. (2007). A framework for validation of computer models. Techno- metrics 49, 138–154.

Beauchemin, C. (2006). Probing the effects of the well-mixed assumption on viral infec- tion dynamics. J. Theor. Bio. 242, 464–477.

Belkaid, Y., Mendez, S., Lira, R., Kadambi, N., Milon, G. & Sacks, D. (2000). A natural model of Leishmania major infection reveals a prolonged “silent” phase of para- site amplification in the skin before onset of lesion formation and immunity. J. Immunol. 165, 969–977.

Belkaid, Y., Piccirillo, C. A., Me´ndez, S., Shevach, E. M. & Sacks, D. L. (2003). CD4+CD25+ immunoregulatory T cells control Leishmania major persistence and the development of concomitant immunity. Nature 420, 502–507. 90

Bellingan, G. J., Caldwell, H., Howie, S. E. M., Dransfield, I. & Haslett, C. (1996). In vivo fate of the inflammatory macrophage during resolution of inflammation. J. Immunol. 157, 2577–2585.

Bernaschi, M. & Castiglione, F. (2002). Selection of escape mutants from immune recognition during HIV infection. Immunol. Cell. Bio. 80, 307–313.

Billack, B. (2006). Macrophage activation: role of Toll-like receptors, nitric oxide, and nuclear factor kappa B. Am. J. Pharm. Educ. 70, 102. .

Bodgan, C., Donhauser, N., Doring,¨ R., Rollinghoff,¨ M., Diefenbach, A. & Rittig, M. G. (2000). Fibroblasts as host cells in latent leishmaniasis. J. Exp. Med. 191(12), 2121–2129.

Casella, G. & George, E. I. (1992). Explaining the Gibbs sampler. Amer. Statistician 46, 167–174.

Chang, K., Reed, S. G., McGwire, B. S. & Soong, L. (2002). Leishmania model for microbial virulence: the relevance of parasite multiplication and pathoantigenicity. Acta Trop. 85, 375–390.

Chib, S. & Greenberg, E. (1995). Understanding the Metropolis-Hastings algorithm. Amer. Statistician 49, 327–335.

Cozzo, C., Larkin, J. I. & Caton, A. J. (2003). Cutting edge: self-peptides drive the peripheral expansion of CD4+CD25+ regulatory T cells. J. Immunol. 171, 5678–5682.

Damian, R. T. (1997). Parasite immune evasion and exploitation: reflections and projec- tions. Parasitology 115, 169–175. .

Dancik, G. M. & Dorman, K. S. (2008). mlegp: statistical analysis for computer models of biological systems using R. Bioinformatics Submitted.

Dancik, G. M., Jones, D. E. & Dorman, K. S. (2006). An agent-based model for Leishmania infection. Interjournal of Complex Systems , 1853.

Forrest, S. & Beauchemin, C. (2007). Computer immunology. Immunol. Rev. 216, 176–197.

Fruth, U., Solioz, N. & Louis, J. A. (1993). Leishmania major interferes with antigen presentation by infected macrophages. J. Immunol. 150, 1857–1864.

Furth, R. V., Dulk, M. D. D. & Mattie, H. (1973). Quantitative study on the pro- duction and kinetics of mononuclear phagocytes during an acute inflammatory reaction. J. Exp. Med. 138, 1314–1330. 91

Gelman, A., Carlin, J. B., Stern, H. S. & Rubin, D. B. (2004). Bayesian Data Anal- ysis. London: Chapman & Hall.

Gelman, A. & Rubin, D. B. (1992). Inference from iterative simulation using multiple sequences. Stat. Sci. 7, 457–511.

Gilks, W. R., Richards, S. & Spiegelhalter, D. J. (1998). Markov Chain Monte Carlo in practice. London: Chapman & Hall.

Grimm, V., Revilla, E., Berger, U., Jeltsch, F., Mooij, W. M., Railsback, S. F., Thulke, H.-H., Weiner, J., Wiegand, T. & DeAngelis, D. L. (2005). Pattern-oriented modeling of agent-based complex systems: lessons from ecology. Science 310, 987–991.

Heitmann, K., Higdon, D., Nakhleh, C. & Habib, S. (2006). Cosmic calibration. ApJ. 646, L1–L4.

Higdon, D., Kennedy, M., C, C. J., Cafeo, J. A. & Ryne, R. D. (2004). Combining field data and computer simulations for calibration and prediction. Siam J. Sci Comput. 26, 448–466.

Janeway, C. A., Travers, P., Walport, M. & Shlomchik, M. J. (2005). Immuno- biology: The Immune System in Health and Disease, 6th ed. New York: Garland Science Publishing.

Kass, R. E. & Raftery, A. E. (1995). Bayes Factors. Journal of the American Statisti- cal Association 90(430), 773–795. URL citeseer.ist.psu.edu/539880.html.

Kennedy, M. C. & O’Hagan, A. (2001). Bayesian calibration of computer models. J.R. Statist. Soc. B 63, 425–464.

Lira, R., Doherty, M., Modi, G. & Sacks, D. (2000). Evolution of lesion formation, parasitic load, immune response, and reservoir potential in C57BL/6 mice following high- and low-dose challenge with Leishmania major. Infect. Immun. 68, 5176–5182.

Luder, C. K., Lang, T., Beuerle, B. & Gross, U. (1998). Down-regulation of MHC class II molecules and inability to up-regulate class I molecules in murine macrophages after infection with Toxoplasma gondii. Clin. Exp. Immunol. 112, 308–316.

McMahon-Pratt, D. & Alexander, J. (2004). Does the Leishmania major paradigm of pathogenesis and protection hold for New World cutaneous leishmaniases or the visceral disease? Immunol. 201, 206–224.

Mehrabi, Y. & Matthews, J. N. S. (1995). Likelihood-based methods for bias reduction in limiting dilution assays. Biometrics 51, 1543–1549. 92

Mendez, S., Reckling, S. K., Picciriloo, C. A., Sacks, D. & Belkaid, Y. (2004). Role for CD4+CD25+ regulatory T cells in reactivation of persistent leishmaniasis and control of concomitant immunity. J. Exp. Med. 200, 201–210.

Morris, M. D. & Mitchell, T. J. (1995). Exploratory designs for computational experi- ments. J. Stat. Plan. Infer. 43, 381–402.

Naderer, T., Ellis, M. A., Sernee, M. F., De Souza, D. P., Curtis, J., Handman, E. & McConville, M. J. (2006). Virulence of Leishmania major in macrophages and mice requires the gluconeogenic fructose-1,6-bisphosphatase. PNAS 103, 5502– 5507. .

Perelson, A. S., Kirschner, D. E. & Boer, R. D. (1992). Dyanmics of HIV infection of CD4+ T cells. Math. Biosci. 114, 81–125.

Plummer, M., Best, N., Cowles, K. & Vines, K. (2006). Coda: Output analysis and diagnostics for MCMC URL http://cran.r- project.org/src/contrib/Descriptions/coda.html.

Rajotte, D., Cadieux, C., Haman, A., Wilkes, B. C., Clark, S. C., Hercus, T., Woodcock, J. A., Lopez, A. & T, H. (1997). Crucial role of the residue R280 at the F-G loop of the human granulocyte/macrophage colony-stimulating factor receptor chain for ligand recognition. J. Exp. Med. 185, 1939–1950.

Rencher, A. C. (2002). Methods of Multivariate Analysis, 2nd Edition. New York: Wiley.

Roychoudhury, K. & Roy, S. (2004). Role of chemokines in Leishmania infection. Curr. Mol. Med. 4, 691–696.

Sacks, J., Welch, W. J., Mitchell, T. J. & Wynn, H. P. (1989). Design and analy- sis of computer experiments. Stat. Sci. 4(4), 409–435.

Saltelli, A., Chan, K. & Scott, E. M. (2000). Sensitivity Analysis. New York: Wiley.

Santner, T. J., Williams, B. J. & Notz, W. (2003). The Design and Analysis of Com- puter Experiments. New York: Springer.

Schonlau, M. & Welch, W. (2006). Screening the input variables to a computer model via analysis of variance and visualization. In: Screening: Methods for Experimentation in Industry, Drug Discovery, and Genetics (Dean, A. & Lewis, S., eds.). New York: Springer, pp. 308–327.

Segovia-Juarez, J. L., Ganguli, S. & Kirschner, D. (2004). Identifying control mechanisms of granuloma formation in M. tuberculosis infection using an agent-based model. J. Theor. Biol. 231, 357–376. 93

Suffia, I. J., Reckling, S. K., Piccirillo, C. A., Goldszmid, R. S. & Belkaid, Y. (2006). Infected site-restricted Foxp3+ natural regulatory T cells are specific for micro- bial . J. Exp. Med. 203(3), 777–788.

Sunderkotter, C., Kunz, M., Steinbrink, G., Meinardus-Hager, M., Goebeler, H. B. & Song, C. (1993). Resistance of mice to experimental leishmaniasis is associated with more rapid appearance of mature macrophages in vitro and in vivo. J. Immunol. 151, 4891–4901.

Sypek, J. P., Panosian, C. B. & Wyler, D. J. (1984). Cell contact-mediated macrophage activation for antileishmanial defense. J. Immunol. 133, 3358–3365.

Tranquillo, R. T., Lauffenburger, D. A. & Zigmond, S. H. (1988). A stochastic model for leukocyte random motility and chemotaxis based on receptor binding fluctua- tion. J. Cell. Bio. 106, 303–309.

Vanloubbeeck, Y. & Jones, D. E. (2004). The immunology of Leishmania infection and the implications for vaccine development. Ann. N.Y. Acad. Sci. 1026, 267–272.

Wei, X. Q., Charles, I. G., Smith, A., Feng, G. H., Huang, F. P., Xu, D., Muller, W., Moncada, S. & Liew, F. Y. (1995). Altered immune response in mice lacking inducible nitric oxide synthase. Nature 375, 408–411.

Winchester, G., Sunshine, G., Nardit, N. & Mitchison, A. (1984). Antigen- presenting cells do not discriminate between self and nonself. Immunogenetics 19, 487– 491. .

Wodarz, D. & Nowak, M. A. (2002). Mathematical models of HIV pathogenesis and treatment. BioEssays 24, 1178–1187.

Wong, A., Gottesman, I. & Petronis, A. (2005). Phenotypic differences in genetically identical organisms: the epigenetic perspective. Hum. Mol. Genet. 14(Review Issue 1), R11–R18.

Wooldridge, M. (2002). Inroduction to Multiagent Systems. New York: John Wiley & Sons, Inc.

Yates, A., Stark, J., Klein, N., Antia, R. & Callard, R. (2007). Understanding the slow depletion of memory CD4+ T cells in HIV infection. Plos Med 4(5), e177. 94

CHAPTER 3. Identifying correlates of optimal immune escape strategies and immune escape success in an in silico model of viral infection

To be submitted to the Journal of Theoretical Biology

Garrett M. Dancik3, Susan Carpenter4, Karin S. Dorman1,2,3

1Departments of Statistics, 2Genetics, Development & Cell Biology, and the 3Program in

Bioinformatics and Computational Biology, Iowa State University; and 4Department of

Veterinary Microbiology & Pathology, Washington State University

Abstract

The adaptive immune response is made up of the cellular arm which eliminates in- fected cells and the humoral arm which eliminates free virus. Escape mutants arise when mutations in cellular or humoral epitopes create viral quasispecies that are not recognized by the current response, and are commonly observed in persistent infections such as HIV and EIAV. In the current study, we perform in silico experiments in a computer model of viral infection and identify correlates of optimal immune escape strategy and immune escape success. Cellular or humoral escape mutants are injected at various time points, and peak plasma viral load is used as a measure of escape success. Using the amount of killing by each arm in a controlled infection as a measure of the relative importance of each arm, we find that the overall relative importance of the cellular (humoral) response is more 95 strongly correlated with both optimal escape strategy and immune escape success than the current relative importance of the cellular (humoral) response at the time the escape mu- tant is introduced. Furthermore, the optimal escape strategy generally does not depend on the time the escape mutant arises. Mutants that simultaneously escape both cellular and humoral responses perform substantially better than cellular or humoral escape mutants alone, highlighting the importance of both arms in controlling viral infections. Finally, loss of infectivity in humoral escape mutants favors the virus, likely because the strength of the cellular response weakens as infectivity decreases.

3.1 Introduction

The adaptive immune response is made up of the cellular arm which eliminates in- fected cells and the humoral arm which eliminates free virus. Within each arm, lympho- cytes proliferate in response to the specific antigens they recognize. Cytotoxic T lympho- cytes (CTLs) recognize viral peptides presented on surface MHC class I (MHCI) molecules of infected cells and lyse the host cell. B cells recognize viral surface proteins, phagocytize the virus, and present antigen on MHC class II (MHCII) molecules to helper T (TH) cells.

TH cells that recognize the MHCII:peptide complex activate B cells, which leads to the production of antibodies that bind free viral particles, inducing uptake and disposal from phagocytotic cells or blocking viral entry into susceptible host cells.

Not surprisingly, viruses can ‘escape’ recognition from the cellular and humoral re- sponses if key proteins are mutated. Mutations in epitopes recognized by lymphocytes al- low the virus to remain hidden from the adaptive immune response until another antigen- specific response develops. Antigen escape has been observed in several viral infections.

For example, in HIV-1, equine infectious anemia virus (EIAV), SIV, and Hepatitis B virus, both CTL escape (Mealy et al., 2003; Chen et al., 2000; Kelleher et al., 2001; Cuestes et al., 96

2006) and humoral escape (Wei et al., 2003; Richman et al., 2003; Rwambo et al., 1989;

Cuestes et al., 2006) have been observed. For the virus, immune escape often incurs a fit- ness cost. HIV mutants have decreased replicative capacity in vitro compared to wild type virus (Martinez-Picado et al., 2007; Zennou et al., 1998), and reversion of mutants to wild type is commonly observed following transmission of the mutant into a new host (Friedrich et al., 2004; Leslie et al., 2004).

Along with mutation, viruses can prolong escape by interfering with a host’s antigen presentation machinery or blocking physical access to an . Both herpes simplex virus type I and HIV-1 decrease MHCI:peptide expression on the surface of infected cells (Ploegh,

1998; Schwartz et al., 1996; Howcroft et al., 1993), which weakens the efficacy of the cellular response. In HIV, an evolving glycan shield on the surface of the virus particle partially blocks antibody access to epitope (Wei et al., 2003), weakening the efficacy of the humoral response.

Mathematical and computer (e.g., cellular automata, agent-based) models of the im- mune response to viral infections have been used to gain insight into immune behavior and virus evolution. For example, in a mathematical model of the adaptive immune response,

Arnaout & Nowak (2000) explored how competition between the humoral and cellular re- sponses for proliferative stimuli leads to a cooperative humoral and cellular response that benefits the host. Iwasa et al. (2005) showed that for a wide class of mathematical models, within-host evolution results in a greater viral pathogenicity. Using a mathematical model of HIV infection, Nowak et al. (1991) proposed a viral diversity threshold as a trigger for progression to AIDS. In a computational model of HIV infection, Bernaschi & Castiglione

(2002) attributed the asymptomatic phase of disease to escape mutants with low transcrip- tion rates, and Castiglione et al. (2004) found that viral diversity is primarily the result of cellular escape. Mathematical models of HIV are reviewed in Callaway & Perelson (2002) 97 and Wodarz & Nowak (2002); computer models of the immune response are reviewed in

Forrest & Beauchemin (2007).

The objective of the current study is to identify correlates of optimal immune escape strategy and immune escape success by performing in silico experiments using a modified version of the Celada-Seiden stochastic cellular automata model of the immune response

(Seiden & Celada, 1992). We find that the overall relative importance of the cellular (hu- moral) response in a controlled infection is more strongly correlated with both optimal es- cape strategy and immune escape success than the current relative importance of the cellu- lar (humoral) response at the time the escape mutant is introduced. We also find that the optimal escape strategy generally does not depend on the time the escape mutant arises, and that simultaneous cellular and humoral escape is substantially more successful than escape from either the cellular or humoral arm alone. Finally, decreasing infectivity in hu- moral escape mutants favors the virus, likely because the strength of the cellular response weakens as infectivity decreases.

3.2 Materials and Methods

3.2.1 A computer model of viral infection

The computer model that we use is a modified version of the C-IMMSIM model of

HIV infection (Bernaschi & Castiglione, 2002; Castiglione et al., 2004). Specifically, our computer code is developed from the source code C-IMMSIM v.6.3 provided by Filippo

Castiglione. The immune simulator (IMMSIM), originally described in Seiden & Celada

(1992), is a stochastic cellular automata model of the humoral immune response that in- cludes B cells, TH cells, antibodies, and generic antigen presenting cells (APCs). Bit strings are used to model receptors and their ligands. Interactions are then modeled probabilisti- cally through an affinity function pa(h), where the input h is the number of bits that differ 98 between the receptor and ligand bit strings (i.e., the Hamming distance). The use of bit strings allows the model to explicitly incorporate the clonal selection of lymphoctyes and provides a framework for antigen escape through the mutation (flipping) of viral peptides

(bit strings).

C-IMMSIM is an enhanced C language version of IMMSIM that incorporates both the humoral and cellular arms of the adaptive immune response. Macrophages, dendritic cells, B cells, TH cells, CTLs, plasma secreting B cells, and antibodies are included. Bit strings are used to encode lymphocyte (i.e., and T cell) receptors, antibody antigen- binding sites, and MHC molecules. Antigens that are recognized by B cells and antibodies are referred to as epitopes, and antigens that are recognized by CTLs are referred to as a peptides (Bernaschi & Castiglione, 2002). Each virus particle is specified by a single epitope and a single peptide, which determines the virus strain. Humoral and cellular escape mu- tants are created by flipping bits in (i.e., mutating) the appropriate bit string. We use a bit string length of 12 in all of our simulations.

Simulated infection proceeds according to the following set of rules, starting with an injection of 100 virus particles into the plasma. A virus infects a macrophage with proba- bility pMA Virus. Within an infected cell, viral transcription begins with probability pw. New virus particles are produced at rate αI . A single virus particle buds off of the host cell with probability BuddingProb. Once the amount of intracellular virus reaches the carrying capac- ity NmaxVirus, lysis occurs, and viral particles are released into the plasma. Plasma virus is also phagocytized by dendritic cells with probability pDC Ag, by macrophages with prob- ability pMA Ag, and by B cells with a probability that depends on the affinity of the B cell receptor for the viral epitope (see below). Virus that is phagocytized is no longer infectious, and dendritic cells, macrophages, and B cells can all function as APCs. Clonal expansion occurs when a lymphocyte recognizes a MHCI:peptide or MHCII:peptide complex. Each 99 individual CTL divides a maximum of DupStepTC times, and each B cell divides a maxi- mum of DupStepB times. B cells differentiate into plasma secreting B cells, which produce

AbMultFact antibodies. A summary of the relevent C-IMMSIM rules is given in Fig. 3.1.

A summary of the computer model parameters relevant to the current work along with the values we use is given in Table 3.1.

We use three separate affinity functions in our model: the cellular affinity function, ptc(·), for interactions between CTL receptors and MHCI:peptide complexes; a second cellu- lar affinity function, pth(·), for interactions between TH receptors and MHCII:peptide com- plexes on APCs; and the humoral affinity function, ph(·), for interactions between both B cell receptors or antibody and surface viral epitopes. Affinity between two molecules can be

‘strong’ or ‘weak’, which we model using the following low and high affinity functions:

0.000 h ≤ 8 0 h ≤ 9   0.050 h = 9    0.005 h = 10    pa,low(h)=  pa,high(h)=  0.417 h = 10    0.010 h = 11  1.000 h = 11    0.250 h = 12    1.000 h = 12       The affinity function returns a binding probability for a given Hamming distance h. We use the high affinity function pa,high(·) for pth(·) and let ptc(·) and ph(·) vary among pa,low(·) and pa,high(·). Because we do not model multiple binding sites (e.g., multiple MHCI:peptide complexes on a single cell), the affinity function specifies an overall binding probability and also determines the avidity of an interaction (Gallimore et al., 1998).

In viral infections, the eclipse phase refers to the time period after a virus enters a cell and before progeny virus is made. We incorporate an eclipse phase by having viral tran- scription begin with probability pw at each time step. Our interpretation of pw is differ- 100 ent from that of Castiglione et al. (2004) and Bernaschi & Castiglione (2002), who use it to model latent infection of CD4+ T cells. Finally, in the experiments we perform, our objec- tive is to systematically inject viral escape mutants and assess their success. We therefore do not allow random virus mutation in our model in order to control when escape happens and to facilitate the identification of measures of success.

3.2.2 Computer model analysis

3.2.2.1 Quantitative characterization of the immune response

We define several immune response variables for the purpose of identifying relation- ships between the immune response and escape mutant success. Let h(t) be the cumulative number of viral particles killed by the humoral response (i.e., by B cells and antibodies) at time t, and let c(t) be the cumulative number of viral particles killed by the cellular re- sponse (i.e., by CTLs), which includes all virus particles inside a cell when the cell is killed.

Then the relative cumulative proportion of humoral (H(t)) and cellular (C(t)) killing at time t are as follows:

h(t) h(t)+c(t) h(t) > 0,c(t) > 0 H(t)=  , C(t) = 1 − H(t) (3.1)  0.5 h(t)= c(t) = 0

In this way, H(t) and C(t) quantify the relative importance of each arm of the immune sys- tem at time t. If the infection ends (is cleared) by time t, then H(tend) and C(tend) quantify the relative importance of each arm of the immune system in resolving the infection.

We define h′(t) to be the amount of viral particles killed by the immune response dur- ing time step t, and c′(t) to be the amount of viral particles killed by the cellular immune response during time step t. Using these measures, we then define the humoral and cellular 101 killing efficiencies as

h′(t) h′(t) heff(t)= total amount of plasma virus at t and ceff(t)= total amount of intracellular virus at t .

Other measures we consider are nc(t), the number of CTLs present at time t; nh(t), the combined number of B cells and antibodies present at time t, and nth(t), the number of TH cells present at time t.

3.2.2.2 K-means clustering

The objective of k-means clustering is to partition a collection of n (multidimensional) observations into k mutually exclusive subsets (i.e., clusters), C1,...,Ck, by minimizing the within cluster variance (Hartigan & Wong, 1979). Specifically, for a collection of observa- tions, x1,...,xn, where each observation is m-dimensional, the goal is to minimize

k 2 V = |xj − µi| , Xi=1 xXj ∈Ci where µi is the centroid of all observations in cluster Ci, and |a − b| is the Euclidian dis- tance between a and b. The algorithm proceeds by randomly assigning observations to clus- ters C1,...Ck, calculating the cluster centers µ1,...,µk, and then reassigning each obser- vation to the cluster whose center is the closest, in terms of Euclidian distance. Calculation of cluster centers and reassignment of observations to clusters is repeated until the clus- ter assignments no longer change. In our analysis, we cluster computer model output using the function ‘kmeans’ in the R package stats (R Development Core Team and contributors worldwide, 2007). 102

3.2.2.3 Classification and regression trees

Both classification and regression trees are used to model the relationship between a response variable and a collection of inputs. Classification trees are used for categorical ob- servations where, traditionally, observed outcomes are referred to as classes; regression trees are used for observations that are continuous. Tree construction is similar in both cases.

Starting with all observations at the root node, a tree is grown using binary recursive parti- tioning. At each node, an input p and a splitting criteria is chosen that splits the remaining observations into two groups and maximizes the reduction in impurity. If p is binary, then observations are split based on ‘low’ and ‘high’ values of p. If p is continuous, then obser- vations are split based on px for some optimized value x. For classification trees, impurity is calculated using the Gini index (Breiman et al., 1984), which measures the amount of disagreement among classes in the two groups resulting from the split. For regression trees, impurity is based on a sum of squared error calculation. The process re- peats until convergence conditions are satisfied. For classification trees, leaves are labeled based on the dominating class. For regression trees, each leaf is labeled with a fitted value that is the arithmetic mean of all observations at that leaf. In our analysis, we fit classifica- tion and regression trees using the R package tree.

We also report several statistics with each tree. For classification trees, we calculate the percent accuracy of the classification for each leaf l as

number of observations in l of the dominating class accuracy of l = × 100% total number of observations in l and a percent coverage probability of the dominating class as

number of observations in l of the dominating class c coverage by l = × 100% total number of observations of class c 103

If a leaf has 100% accuracy and coverage then the decisions along the branches leading to it completely and exactly define a class. The percent coverage for a class c is the sum of the percent coverages over all leaves where c is the dominating class. The overall accuracy of the tree T is defined as

number of correctly classified observations in T accuracy of T = × 100% total number of observations in T

.

For regression trees, we report the root mean square error (RMSE) as

n RMSE = (y − yˆ )2, v i i ui=1 uX t where y1,...,yn are the n observations, andy ˆ1,... yˆn are their corresponding fitted values.

3.2.2.4 Statistical significance

Bootstrap estimates of standard deviations for rank correlations between x and y are found by resampling from pairs (x, y) with replacement 10,000 times, calculating the rank correlation for each sample, and taking the standard deviation of the sample rank corre- lations (Efron, 1981). We use a permutation test for assessing whether or not the differ- ence between two rank correlations is statistically significant (Good, 2005). Formally, in- terest is in the null hypothesis cor(x1, y) = cor(x2, y). First, calculate the test statistic

T = cor(x1, y) − cor(x2, y). Permute the vector y 10,000 times, and for each permutation

(i) (i) (i) (i) th i, calculate the sample statistic T = cor(x1, y ) − cor(x2, y ), where y is the i per- mutation of y. The bootstrap p-value is the proportion of times T (i) is more extreme than

T,i = 1,... 10, 000. We consider the difference between cor(x1, y) and cor(x2, y) to be sta- tistically significant when p < 0.001. P -values for computer model parameters in a main 104 effects model are calculated using the function ‘lm’ in the R package stats (R Development

Core Team and contributors worldwide, 2007).

3.3 Results

3.3.1 Effect of computer model parameters on simulation output

We vary the computer model parameters NmaxVirus, BuddingProb, pMA Virus,

8 DupStepB, DupStepTC, ptc(·), and ph(·) according to a 2 (n = 256) factorial design, with 10 replicates per design point, using the high and low parameter values given in Table 3.1.

The remaining parameters are fixed at the values specified in the table. All additional C-

IMMSIM parameters that we do not mention are fixed at their standard values (Castiglione,

2008). Each simulation is run for a total of tend= 365 days. Output from the simulation consists of plasma viral load over time and the immune response measures explained in section 3.2.2.1. The choice of computer model parameters leads to distinct plasma viral load profiles and immune responses (see below). In 83% of the simulations, the infection is cleared by tend; infections that are not cleared have a mean plasma viral load of 421 at tend and are clearly trending toward ultimate clearance.

Plasma viral load

Fig. 3.2 shows graphs of computer model output clustered according to plasma viral load. A classification tree describing the relationship between the computer model param- eters and cluster outcomes is given in Fig. 3.3. The tree has an overall accuracy of 96%, indicating that changes in immune and viral parameters largely explain the distinct viral profiles. Virus infectivity (pMA Virus) is the most important parameter, with low pMA Virus generally leading to sharply peaking infections that are quickly resolved (clusters 1,3,5) and high pMA Virus leading to recurrent cycles of viremia (clusters 2,6,7,8). Cluster 4, which 105 is accurately classified (accuracy = 100%) but poorly covered (coverage = 40%) includes simulations with both high and low levels of pMA Virus. High humoral (ph(·)) and cellular

(ptc(·)) affinities leads to simulations with lower peak plasma virus levels. For example, the lowest overall plasma viral loads are observed in cluster 4, where ph(·) or ptc(·) is strong.

Clusters 8 and 6 are distinguished based on ph(·), and peak plasma viral load is higher when ph(·) is weaker. The carrying capacity of infected host cells, NmaxVirus, effects the timing of the peak plasma levels, with the first peak occurring later when NmaxVirus is higher (e.g., compare cluster 1 and 5). Finally, peak plasma viral load is higher when the B cell proliferation rate (DupStepB) is lower (cluster 3 vs. cluster 5).

Relative importance of cellular and humoral responses

Both viral and immune parameters effect the overall relative importance of the hu- moral (H(tend)) and cellular (C(tend)) responses. A regression tree for H(tend) against com- puter model parameters is shown in Fig. 3.4. Only the parameters pMA Virus, ptc(·), and ph(·) appear in the tree. High values of ph(·) favor the humoral response, while high values of pMA Virus and ptc(·) favor the cellular response (i.e., result in lower values of H(tend)). All parameters in the regression tree are statistically significant (p < 0.001) based on a main effects model. The tree has a RMSE of 0.14.

3.3.2 Correlates of successful immune escape

In order to identify correlates of successful immune escape strategies, we perform sev- eral in silico experiments involving the injection of cellular escape mutants (CEM) and hu- moral escape mutants (HEM) at various time points. The immune response measures that

′ ′ we consider as correlates include H(tend), C(tend),H(tinj), C(tinj), h (tinj),c (tinj), heff(t),ceff(t),

nh(tinj) nc(tinj),nh(tinj),nth(tinj) and (see section 3.2.2.1 for details). Escape mutants are nc(tinj) 106 created by randomly flipping 11 bits in the appropriate bit string of the virus (i.e., the pep- tide or epitope). We choose to use a large number of bit mutations so that cross-reactivity between the mutant and wild type strain is minimal, which ensures that the mutant strain definitely escapes the desired immune response. For each point in the factorial design, we

8 find the average time of peak infection, tpeak. Using the same 2 factorial design with 10 replicates and the same random number seeds for each replicate, we inject 50 HEM at the

1 1 appropriate tpeak. The experiment is repeated for injection times 4 × tpeak, 2 × tpeak, and 3 4 × tpeak, and the entire process is repeated to inject CEM at the four injection times. Again, the same random number seeds are used, which allows us to perform pairwise com- parisons between corresponding HEM and CEM simulations.

We measure escape mutant success using peak plasma viral load. HEM had a higher peak than CEM in 54% of the simulations, while CEM had a higher peak than HEM in

41% of the simulations. In the remaining 5% of the simulations there was no difference in peak plasma viral load between HEM and CEM; in 93% of these simulations, peak plasma levels of the mutant strain did not exceed the initial inoculation of 50 viruses. When plasma viral load of the escape mutant fails to exceed the initial inoculation size, we say that the escape mutant fails to establish a secondary infection. Table 3.4 shows the proportion of

CEM and HEM that fail to establish a secondary infection in each experiment. In general,

HEM are always less likely than CEM to establish a secondary infection, and the probabil- ity of either failing to establish a secondary infection increases as the injection time moves closer to tpeak. 107

Overall relative humoral and cellular importance as correlates of escape suc- cess

In order to characterize the relationship between immune response measures and

CEM and HEM success, we calculate rank correlations of various immune response mea- sures with peak plasma viral load of CEM and HEM over time (Figure 3.5). Because in some cases a large proportion of runs did not establish a secondary infection (Table 3.4), the correlations are calculated after removing these simulations from the data set.

H(tend) and equivalently, C(tend), are consistently more strongly correlated with peak plasma HEM than are the other measures we consider. Notably, H(tend) (cor = 0.81 at tpeak) is a better correlate than H(tinj) (cor = 0.68 at tpeak) at all time points, and these 1 3 differences are statistically significant (p < 0.0001) at injection times 4 × tpeak, 4 × tpeak, and tpeak. Similarly, C(tend) (cor = 0.68 at tpeak) is generally more strongly correlated with peak plasma CEM than C(tinj) (cor = 0.62 at tpeak) is, and these differences are statisti- 1 3 cally significant at injection times 4 × tpeak, 4 × tpeak, and tpeak. Together, these results suggest that the success of an escape mutant depends more on the relative overall impor- tance of each arm of the immune response than on the relative importance of each arm at the time the escape mutant arises (i.e., the time of the injection).

Overall relative humoral and cellular importance as indicators of optimal es- cape strategy

In order to compare the relative success of HEM to CEM we calculate

dpeak = peak plasma viral load for CEM - peak plasma viral load for HEM 108 for each pair of simulations for all injection times. Simulations where CEM and DEM were equally successful (i.e., dpeak = 0) did not appear to correlate with any immune response measures or parameters (data not shown) and were removed from the data set. In 94% of the simulations that were removed, secondary infection was not established.

Fig. 3.6 shows the rank correlation of dpeak with various immune response measures 1 1 3 when the mutant strain is injected at 4 × tpeak, 2 × tpeak, 4 × tpeak and tpeak. C(tend) and equivalently, H(tend), are consistently more strongly correlated with dpeak than the other measures we consider. Notably, C(tend) (cor = 0.86 at tpeak) has a stronger correlation than

C(tinj) (cor = 0.78 at tpeak) at each injection time, and these differences are statistically significant (p < 0.0001). The correlation indicates that stronger relative cellular immune responses (i.e., high values of C(tend)) are associated with increased CEM success relative to

HEM (i.e., high values of dpeak), and that stronger relative humoral immune responses are associated with increased HEM success relative to CEM.

We say that there is an optimal escape strategy when, in a pair of cellular and hu- moral escape mutants, one escape mutant achieves greater peak plasma viral load than the other. Table 3.4 shows the percentage of times optimal escape strategies are (indepen- dently) in agreement for all pairs of injection times. The optimal strategy at tpeak is also 3 1 the optimal escape strategy at 4 × tpeak and 2 × tpeak 90% of the time, and is the optimal 1 escape strategy at 4 × tpeak 81% of the time. A classification tree for successful escape strat- egy, based on H(tend), and which combines simulations from all injection times, is given in Fig. 3.7. The tree, which has an overall accuracy of 88%, indicates that when the cellular response is more important overall (i.e., H(tend) < 0.48), cellular escape is a better strategy, and when the humoral response is more important overall (i.e., H(tend) ≥ 0.48), humoral escape is a better strategy. Together, these results suggest that the optimal escape strategy depends on the relative overall importance of each arm of the immune response, and not 109 on the relative importance of each arm of the immune response at the time the escape mu- tant arises. In general, a single strategy is consistently optimal throughout the course of the infection.

3.3.3 Simultaneous cellular and humoral escape

Dual escape mutants (DEM) are created by randomly flipping 11 bits in both the pep- tide and epitope bit strings. Using the same experimental design as in section 3.3.2, we in-

1 1 3 ject 50 DEM at 4 × tpeak 2 × tpeak, 4 × tpeak, and tpeak. Table 3.4 shows the proportion of times DEM fails to establish secondary infection for all time points. At tpeak, DEM fails to establish secondary infection only 12% of the time, whereas CEM and HEM fail to establish secondary infection 34% and 40% of the time, respectively.

In a majority of the trials, the peak plasma viral load of DEM was higher than the combined peak plasma viral loads of HEM and CEM, indicating that cellular and humoral escape have a synergistic effect. Fig. 3.8 compares mean peak plasma viral load for DEM,

CEM, and HEM. The figure shows that the synergy does not depend on which escape mu- tant (CEM or HEM) is more successful alone, and that synergy occurs when both CEM and

HEM are not successful in isolation. Averaged over all simulations, mean peak plasma viral load for DEM is 3.12 times greater than peak plasma viral load for HEM, the more success- ful escape mutant. When peak plasma viral load is the same for CEM and HEM, the mean peak plasma viral load is 52 and the mean peak plasma viral load for DEM is more than

1100 times greater. Two representative simulations, with corresponding HEM, CEM, and

DEM peak plasma viral load outputs are shown in Fig. 3.9. 110

3.3.4 Viral fitness and immune escape success

Because escape mutants could be inherently less recognizable to the immune system, and because escape mutants often have reduced viral fitness, we analyze immune escape in the context of changing viral phenotypes. Specifically, we are interested in cellular escape mutants that become less recognizable to the cellular response (i.e., where ptc(·) is high for the wild type virus but low for the mutant); in humoral escape mutants that become less recognizable to the humoral response (i.e., where ph(·) is high for the wild type virus but low for the mutant); and in escape mutants that are both less recognizable and less infec- tious than the wild type virus (i.e., where the appropriate affinity function is lower in the mutant and pMA Virus is high for the wild type but low for the mutant). We will refer to an escape mutant whose phenotype changes as a super escape mutant, and an escape mu- tant whose phenotype does not change as a standard escape mutant. Super escape mutants whose affinity functions differ from the wild type will be denoted as CEM+ and HEM+; su- per escape mutants that are also less infectious will be denoted as CEM+− and HEM+−.

Two experiments are performed. In the first experiment, for previous simulations (section

+ 3.3.1) where ptc(·) is high, we inject 50 CEM at tpeak, and for all simulations where ph(·)

+ is high, we inject 50 HEM at tpeak. In the second experiment, we take simulations from

+− the previous experiment which have high values of pMA Virus, and inject 50 CEM or 50

+− HEM at tpeak. Table 3.4 shows the proportion of times that peak plasma viral load for super escape mutants is greater than, less than, or equal to peak plasma viral load of standard escape mutants and super escape mutants that are less infectious. Not surprisingly, super escape mutants (CEM+ and HEM+) generally have higher peak plasma viral loads than standard escape mutants (CEM and HEM), and this is true more often for cellular escape mutants.

Peak plasma viral load of CEM+ is greater than peak plasma viral load of CEM 70% of the 111 time, whereas peak plasma viral load of HEM+ is greater than peak plasma viral load of

HEM 56% of the time. Surprisingly, humoral super escape mutants that are less infectious

(HEM+−) have higher peak plasma viral loads than super escape mutants that are equally infectious (HEM+) 63% of the time. A representative example comparing the simulation for

HEM+− with its corresponding HEM+ simulation is given in Fig. 3.10.

3.4 Discussion

In this paper, we performed in silico experiments involving the injection of cellular, humoral, and dual escape mutants under a variety of conditions (i.e., for various virus and immune parameters) and at various stages of infection. We identify the overall relative im- portance of the cellular (or humoral) response as a strong correlate of optimal immune es- cape strategy and individual escape success that is relatively robust to the time in which the escape mutant arises. Dual escape mutants are substantially better than either cellular escape or humoral escape alone, and loss of infectivity in humoral escape mutants strength- ens, not weakens, the virus.

Whether the cellular or humoral response is relatively more important (i.e., kills a larger number of viruses) depends on both virus and immune parameters. More infectious viruses and immune systems with stronger cellular affinities will result in more important cellular responses, while stronger humoral affinities result in more important humoral re- sponses (Fig. 3.4), an observation in agreement with Arnaout & Nowak (2000). The more important a particular arm of the adaptive immune response is, the more successful an es- cape mutant from that arm will be (Fig. 3.5). Similarly, the optimal escape strategy is to escape from the arm that is the most important (Fig. 3.7), and the relative success of either escape strategy increases with the relative importance of the arm being escaped from (Fig.

3.6). The lower correlations between H(tend) and both the relative and absolute success of 112 immune escape from either arm (Figs. 3.6 and 3.5) can be attributed to the fact that there is a delay before the adaptive immune response develops (Dustin & Rice, 2007), particularly when initial antigen load is low (Russell et al., 2007). Prior to the development of the adap- tive immune response, we would not expect CEM with the same phenotype as wild type virus to have any systematic advantage over HEM, and vice-versa.

Although cellular and humoral responses are often considered separately in viral in- fections, Arnaout & Nowak (2000) use a mathematical model to demonstrate that both responses are important for controlling infection and immunopathology, and that both re- sponses can coexist despite the fact that they compete for proliferative stimuli. The sub- stantial success of dual escape mutants in our simulations highlights the fact that both re- sponses are important for protecting the host and supports vaccination strategies that in- duce both responses (Wang et al., 2008). During the course of an infection, a virus is ex- posed to both the cellular and humoral arms of the immune response. A replicating virus inside an infected cell can be recognized and eliminated by the cellular arm. A virus parti- cle that buds off or bursts out of an infected cell can be recognized and eliminated by the humoral arm.

Therefore, a virus that escapes only one response is still vulnerable to the other. Specif- ically, when in the plasma an individual HEM is more likely to survive and reinfect a cell than the wild type virus, but is just as likely to be eliminated when inside an infected cell.

Analogously, a CEM is more likely to survive in, proliferate, and therefore be released from an infected cell, but is just as likely to be eliminated as the wild type virus when in the plasma. A DEM is substantially more successful because of the multiplicative effect of be- ing able to infect more cells (i.e., escaping the humoral response) and producing more virus

(i.e., escaping the cellular response). This synergistic effect is observed regardless of whether

1 HEM or CEM is more successful alone (Fig. 3.8), and also occurs for injection times 4 × 113

1 3 tpeak, 2 × tpeak, and 4 × tpeak (data not shown). In our simulations, we assume that the cellular and humoral responses recognize dis- tinct epitopes so that there is no cross-reactivity between cellular and humoral escape. How- ever, cellular and and humoral epitopes may overlap in the viral genome (Heber-Katz et al.,

1988; Hioe et al., 1990). In this case, a mutation in the overlapping region will give rise to a DEM, which can be highly successful even when escape from the cellular or humoral re- sponse alone is not (Fig. 3.8). Therefore, in addition to promoting both responses, success- ful vaccination of rapidly evolving viruses will likely require that the epitopes recognized by each response be distinct.

Replicative capacity is commonly used to measure the fitness of a virus (Qui˜nones

Mateu & Arts, 2002; Sheldon et al., 2006). In growth competition and infection assays, it follows that fitness will be positively related to infectivity. However, our results suggest that the fitness of a virus in vivo depends on both the phenotype of the virus and the (current) ability of each arm of the adaptive immune response to recognize it. Specifically, we find that the success of a humoral escape mutant increases when the mutant is less infectious.

These results, although counterintuitive, can be explained by considering the competition that exists between the cellular and humoral responses and how viral parameters effect the importance of each arm (Arnaout & Nowak, 2000).

Both the less infectious super escape mutant (HEM+−) and the super escape mutant with the same infectiousness as the wild type virus (HEM+) are equally vulnerable to elim- ination by the cellular response, which is relatively important because infection rate is high

(Fig. 3.4). Highly infectious HEM+, therefore, are efficient at infecting cells but risk be- ing rapidly eliminated by the cellular response. The loss of infectivity in HEM+−, however, weakens the cellular response, which increases the likelihood of successful virus production within an infected cell. Because the less infectious escape mutant is a humoral escape mu- 114 tant, HEM+− is not immediately at risk in the plasma, even though it is potentially ex- posed to the humoral response longer than HEM+ is. In this case, HEM+− simultaneously escapes one arm of the immune response while weakening the other through a change in phenotype. Similar arguments support the observation that a super humoral escape mutant is less successful when infectivity increases, while super cellular escape mutants are more successful when infectivity increases (data not shown). Therefore, although it is expected that in a na¨ıve host high infectivity favors the virus (Fig. 3.2), viruses that appear less fit in vitro may actually be more fit in vivo.

Finally, we note that we have chosen to use peak plasma viral load as a measure of escape mutant success. Plasma viral load, in addition to being relatively easy to measure, is a common predictor for rate of CD4+ T cell decrease and progression to AIDS in HIV infection (Mellors et al., 1997) and has been used to monitor antiviral therapy in Hepatitis infection (Lewin et al., 2003). Other measure of viral success are possible, for example peak total viral load and time until clearance. For our wild type simulations, the rank correla- tion between peak plasma viral load and peak viral load is 0.83, but the rank correlation between peak plasma viral load and time until clearance is 0.40. Therefore, although peak plasma virus is a good indicator of total virus it is not necessarily a good indicator of the length of the infection. Fig. 3.8A is an example where peak plasma viral load is higher for the DEM than it is for the CEM, but the the CEM is able to survive longer.

In conclusion, in silico experiments involving escape mutants can help to gain insight into the complexity of immune escape at the systems level. For a specific type of infection, a calibrated model could be used to make predictions about successful escape strategies. A general prediction is that escape from the arm which does more killing (i.e., the more im- portant arm) in situations where infection is controlled is the optimal strategy, independent of when escape occurs, provided that ‘escape’ does not occur before the adaptive immune 115 response is sufficiently developed. This prediction supports the notion that if cellular (hu- moral) escape mutants are commonly observed in a particular infection, then the cellular

(humoral) arm is more important for controlling the infection. However, functional con- straints can also favor certain types of mutations (Peyerl et al., 2004). Ultimately, informa- tion about functional constraints and escape mutant phenotypes could be combined to test vaccine and treatment strategies in silico, with goals of optimizing the balance between in- duced cellular and humoral responses in order to efficiently control infection and limit the possibility of immune escape.

Acknowledgements

GMD and KSD were partially supported by PHS grant GM068955 and an ISU CIAG

Research Support Grant. 116

Infection Virus Recognition and proliferation Binding probability

TCR CTL NmaxVirus DupStepTC ptc(·) peptide p MA MA Virus MHCI:peptide MA

BuddingProb virus TCR TH DupStepTH pth(·) Antigen take up DupStepB B cell (APC) MHCII:peptide

MA pMA Ag epitope B cell (antibody)

BCR ph(·) DC pDC Ag virus epitope

Figure 3.1: Summary of relevant C-IMMSIM rules and parameters. 117

Range / Class Param Description Value Units αI Intracellular virus growth (1,2) viruses / 8 rate hours NmaxVirus Carrying capacity of (50,100) viruses infected macrophages Virus BuddingProb Probability that a virus (0, 10) % / 8 buds from an infected cell hours pw Probability that viral 33 % / 8 transcription begins hours pMA Virus Probability that a virus (1, 90) % / 8 infects a macrophage hours 1 pMA Ag Probability that a 2 × pMA Virus % macrophage phagocy- tizes a virus pDC Ag Probability that a den- pMA Virus % / 8 dritic cell phagocytizes a hours virus pm Per bit mutation rate 0 % / in- fection event ptc(·) Probability a cytotoxic T (low, high) Affinity lymphoctye recognizes an see main MHCI:peptide complex text for ph(·) ProbabilityaBcellor (low, high) details antibody recognizes a surface viral epitope pth(·) ProbabilityaThelper high cell recognizes an MHCII:peptide complex DupStepB Maximum number of (3,4) duplications Immune response duplications allowed per / cell activated B cell DupStepTC Maximum number of (3,4) duplications duplications allowed per / cell activated cytotoxic T Cell DupStepTH Maximum number of 4 duplications duplications allowed per / cell activated helper T cell AbMultFact Number of antibodies 3 antibodies produced by a plasma B / cell cell

Table 3.1: Subset of computer model parameters. Only parameters that we intro- duce or vary are described. Remaining parameters are set at their standard values (Cas- tiglione, 2008). 118

1 2 3 4 0 50000 100000 150000 5 6 7 8 plasma viral load 0 50000 100000 150000

0 100 200 300 0 100 200 300 0 100 200 300 0 100 200 300 days post infection Figure 3.2: Simulation output clustered according to plasma viral load. Graphs show mean plasma viral load over time at each design point (black lines) and cluster centers (red lines).

proportion tinj CEM HEM CEM + HEM DEM 1 4 × tpeak 0.00 0.00 0.00 0.00 1 2 × tpeak 0.00 0.00 0.00 0.00 3 4 × tpeak 0.11 0.20 0.07 0.10 1 × tpeak 0.34 0.40 0.09 0.12

Table 3.2: Proportion of escape mutants that fail to establish secondary infec- tion. Failure is defined as having a peak plasma viral load equal to the initial inoculation of 50 viruses. 119

p MA Virus|

split term

low high a a (b) (b) (c) (c) ph (·) NmaxVirus

NmaxVirus 4

(100) ptc(·) ptc(·) (53.78)

p (·) ptc(·) ptc(·) h 7 DupStepB (78.12) 5 (60.98) ph(·) 3 5 (100) 2 7 4 (100) (87.5) (69.57) 1 4 (100) (100) (100) (88.89) (30.43) (100) (100) (100) (39.02) (100) (13.45) (26.89) 8 6 (100) (100) (100) (100)

Figure 3.3: Classification tree for simulated plasma virus load profiles. Classifi- cation of cluster outcomes (Fig. 3.2) using computer model parameters, with leaves labeled with the dominant cluster number (a); for each leaf, the (b) percent accuracy and (c) per- cent coverage is reported. 120

ptc (·) |

split term

low high

fitted fitted value value

pMA Virus ph (·) p (·) h 0.968 pMA Virus pMA Virus 0.513 0.871 0.257 0.010 0.695 0.262

Figure 3.4: Regression tree for overall relative humoral importance (H(tend)) against computer model parameters. All parameters in the tree (pMA Virus, ptc(·), and ph(·)) are statistically significant (p < 0.001), according to an analysis of variance of a main effects model for all computer model parameters.

1 1 3 tpeak 4 2 4 1 1 4 1 83 80 81 1 2 1 90 90 3 4 1 90 1 1

Table 3.3: Consistency of optimal escape strategies for each pair of injection times. Data correspond to the percentage of times optimal escape strategies (cellular or humoral escape) are in agreement for each pair of injection times. 121

HEM CEM

H(tend) * *

H(t ) * inj * C(tend)

C(tinj) nc(tinj) * ′ c (tinj)

ceff(t) 0.4 0.5 0.6 0.7 0.8 0.9

nh(tinj)/ct(tinj) ′ h (t) c (tinj) eff

rank correlation ′ * h (tinj) c (t) nc(tinj) eff H(tinj)

* H(tend) C(tinj) *

* C(tend) * −0.9 −0.8 −0.7 −0.6 −0.5 −0.4 0.25 0.50 0.75 1.00 0.25 0.50 0.75 1.00

time of challenege (×tpeak) Figure 3.5: Correlation of various immune response characteristics with peak plasma HEM and CEM over time.. Data shown are rank correlations ± 2 bootstrap standard deviations. Immune response measures are defined in section 3.2.2.1. *, statisti- cally significant difference (p < 0.0001) between correlates H(tend) and H(tinj) or C(tend) and C(tinj).

CEM+ HEM+ greater less equal greater less equal CEM 70 4 26 56 22 22 HEM CEM+− 65 25 9 9 63 28 HEM+−

Table 3.4: Effect of viral fitness on immune escape success. Percentage of super escape mutants (CEM+, HEM+) with peak plasma viral load greater than, less than, or equal to peak plasma viral load of standard escape mutants (CEM,HEM) or super escape mutants with decreased infectivity (CEM+−, HEM+−). 122

* * * C(tend) * C(tinj) ceff(t) ′ c (tinj) nc(tinj) rank correlation 0.4 0.5 0.6 0.7 0.8 0.9

nh(tinj) heff(t) rank correlation nh(tinj)/nc(tinj) ′ h (tinj)

H(tinj)

rank correlation * H(tend) * * * −0.9 −0.7 −0.5 0.25 0.50 0.75 1.00

time of challenege (×tpeak)

Figure 3.6: Correlation of various immune response characteristics with dpeak = peak plasma viral load (CEM) - peak plasma viral load (HEM). Data shown are rank correlations ± 2 bootstrap standard deviations. Immune response measures are de- fined in section 3.2.2.1. *, statistically significant difference (p < 0.0001) between correlates at H(tend) and H(tinj) or C(tend) and C(tinj). 123

H(tend) < 0.48 |

true false

CEMa HEM (87.19)b (88.55) (84.36)c (90.71)

Figure 3.7: Classification tree for optimal immune escape strategy using overall 1 1 3 relative importance of the humoral response. CEM and HEM are injected at 4 , 2 , 4 , and tpeak. The optimal escape strategy is the strategy that produces a higher peak plasma virus load. Leaves are labeled with the dominant escape strategy; for each leaf, the (b) percent accuracy and (c) percent coverage is reported. 124

R=2.48

R=3.12 CEM HEM DEM

R=1171 R=3.12 0 50000 100000 150000 200000

All runs peak(CEM) > peak(HEM) > peak(CEM) = (P = 100%) peak(HEM) peak(CEM) peak(HEM) (P = 35%) (P = 51%) (P = 14%)

Figure 3.8: Peak plasma viral load for cellular (CEM), humoral (HEM) and dual cellular and humoral (DEM) escape mutants when injected at tpeak. For each category, we report mean plasma viral load ± 1 standard error of the mean, where standard error is with respect to CEM + HEM combined or DEM; R = ratio of DEM mean peak plasma viral load to the larger of CEM/HEM mean peak plasma viral load; and P = percentage of total runs. Mutant strains are injected at tpeak. 125

A m[tps, ] 0 20000 40000 60000

B WT DEM CEM HEM t inj m[tps, ] 0 5000 15000 25000

0 100 200 300

Figure 3.9: Cellular and humoral escape mutants have synergistic effects. Rep- resentative examples of escape mutant plasma viral load when (A) peak plasma viral load for the cellular escape mutant (CEM) is greater than the peak plasma viral load for the hu- moral escape mutant (HEM) and (B) peak plasma viral load for HEM is greater than peak plasma viral load for CEM. In both cases, peak plasma viral load for a dual escape mutant (DEM) is higher than peak plasma viral load for CEM and HEM combined. Escape mu- tants are injected at tinj = tpeak. WT, plasma viral load of wild type virus when no escape mutants are injected. 126

WT HEM + HEM+ − t inj plasma viral load 0 10000 20000 30000 40000 50000

0 100 200 300

days post infection

Figure 3.10: Humoral escape mutants that are less infectious (HEM+−) are more successful than escape mutants (HEM+) with the same infectivity as wild type (WT) virus.. Escape mutants are injected at tinj = tpeak. 127

Bibliography

Arnaout, R. A. & Nowak, M. A. (2000). Competitive Coexistence in Antiviral Immu- nity 204, 431–441.

Bernaschi, M. & Castiglione, F. (2002). Selection of escape mutants from immune recognition during HIV infection. Immunol. Cell. Bio. 80, 307–313.

Breiman, L., Friedman, J. H., Olshen, R. A. & Stone, C. J. (1984). Classification and Regression Trees. New York: Chapman & Hall.

Callaway, D. S. & Perelson, A. S. (2002). HIV-1 infection and low steady state viral loads. Bull. Math. Biol. 64, 29–64.

Castiglione, F. (2008). CIMMSIM Standard Parameter Values. URL http://www.iac.rm.cnr.it/ filippo/paramter-page.html".

Castiglione, F., Poccia, F., D’Offizi, G. & Bernaschi, M. (2004). Mutation, fit- ness, viral diversity, and predictive markers of disease progression in a computational model of HIV type 1 infection. AIDS Res. Hum. Retroviruses 20(12), 1314–1323.

Chen, Z. W., Craiu, A., Shen, L., Kuroda, M. J., Iroku, U. C., Watkins, D. I., Voss, G. & Letvin, N. L. (2000). Simian immunodeficiency virus evades a dominant epitope-specific cytotoxic T lymphocyte response through a mutation resulting in the ac- celerated dissociation of viral peptide and MHC class I. J. Immunol. 164, 6474–6479.

Cuestes, M. L., Mathet, V. L., Ruiz, V., Minassian, M. L., Rivero, C., Sala, A., Corach, D., Alessio, A., Pozzati, M., Frider, B. & Oubina,˜ J. R. (2006). Un- usual naturally occurring humoral and cellular mutated epitopes of hepatitis B virus in a chronically infected argentine patient with anti-HBs antibodies. J. Clin. Microbiol. 44, 2191–2198.

Dustin, L. B. & Rice, C. M. (2007). Flying under the radar: the immunobiology of hep- atitis C. Annu. Rev. Immunol. 25, 71–99.

Efron, B. (1981). Nonparametric estimates of standard error: the jackknife, the bootstrap, and other methods. Biometrika 68(3), 589–599. 128

Forrest, S. & Beauchemin, C. (2007). Computer immunology. Immunol. Rev. 216, 176–197.

Friedrich, T. C., Dodds, E. J., Yant, L. J., Vojnov, L., Rudersdorf, R., Cullen, C., Evans, D. T., Desrosiers, R. C., Mothe,´ B. R., Sidney, J., Sette, A., Kunstman, K., Wolinsky, W., Piatak, M., Lifson, J., Hughes, A. L., Wil- son, N., O’Connor, D. H. & Watkins, D. I. (2004). Reversion of CTL escape - vari- ant immunodeficiency viruses in vivo. Nat. Mde. 10, 275–281.

Gallimore, A., Dumrese, T., Hengartner, H., Zinkernagel, R. M. & Ramensee, H. G. (1998). Protective immunity does not correlate with the hierarchy of virus-specific cytotoxic T cell responses to naturally processed peptides. J. Exp. Med. 187, 3301–3311.

Good, P. I. (2005). Permutation, Parametric, and Bootstrap Tests of Hypotheses. New York: Springer-Verlag.

Hartigan, J. A. & Wong, M. A. (1979). A K-means clustering algorithm. Appl. Statist. 28, 100–108.

Heber-Katz, E., Dietzschold, B. & Burns-Purzycki, C. (1988). Overlapping T cell antigenic sites on a synthetic peptide fragment from herpes simplex virus glycoprotein D, the degenerate MHC resctriction elicitied, and functional evidence for antigen-Ia interac- tion. J. Exp. Med. 167, 275–287.

Hioe, C. E., Dybdahl-Sissoko, N., Philpott, M. & Hinshaw, V. S. (1990). Overlap- ping cytotoxic T-lymphocyte and B-cell antigenic sites on the influenza virus H5 hemag- glutinin. J. Virol. 64, 6246–6251.

Howcroft, T. K., Strebel, K., Martin, M. A. & Singer, D. S. (1993). Repression of MHC class I gene promoter activity by two-exon Tat of HIV. Science 260, 1320–1322.

Iwasa, Y., Michor, F. & Nowak, M. A. (2005). Virus evolution within patients in- creases pathogenicity. J. Theor. Biol. 232, 17–26.

Kelleher, A. D., Long, C., Holmes, E. C., Allen, R. L., Wilson, J., Conlon, C., Workman, C., Shaunak, S., Olson, K., Goulder, P., Brander, C., Ogg, G., Sullivan, J. S., Dyer, W., Jones, I., McMichael, A. J., Rowland-Jones, S. & Phillips, R. E. (2001). Clustered mutations in HIV-1 gag are consistently required for escape from HLA-B27-restricted cytotoxic T lymphocyte responses. J. Exp. Med. 193, 375–386.

Leslie, A. J., Pfafferott, K. J., Chetty, P., Draenert, R., Addo, M. M., Feeney, M., Tang, Y., Holmes, E. C., Allen, T., Prado, J. G., Altfeld, M., Brander, C., Dixon, C., Ramduth, D., Jeena, P., Thomas, S. A., St. John, 129

A., Roach, T. A., Kupfer, B., Luzzi, G., Edwards, A., Taylor, G., Lyall, H., Tudor-Williams, G., Novelli, V., Martinez-Picado, J., Kiepiela, P., Walker, B. D. & Goulder, P. J. (2004). HIV evolution: CTL escape mutation and reversion after transmission. Nat. Med. 10, 282–289. Lewin, S. R., Ribeiro, R. M., Walters, T., Lau, G. K., Bowden, S., Locarnini, S. & Perelson, A. S. (2003). Analysis of hepatitis B viral load decline under potent therapy: complex decay profiles observed. Hepatology 34, 1012–1020. Martinez-Picado, J., Prado, J. G., Fry, E. E., Pfafferott, K., Leslie, A., Chetty, S., Thobakgale, C., Honeyborne, I., Crawford, H., Matthews, P., Pillay, T., Rousseau, C., Mullins, J. I., Brander, C., Walker, B. D., Stuart, D. I., Kiepiela, P. & Goulder, P. (2007). Fitness Cost of Escape Mutations in p24 Gag in Association with Control of Human Immunodeficiency Virus Type 1. J. Virol. 80, 3617–3623.

Mealy, R. H., Zhang, B., R, L. S., Littke, M. H. & McGuire, T. C. (2003). Epi- tope specificity is critical for high and moderate avidity cytotixic T lymphocytes associ- ated with control of viral load and clinical disease in horses with equine infectious anemia virus. Virology 313, 537–552. Mellors, J., Munoz, A., Giorgi, J. V., Margolick, J. B., Tassoni, C. J., Gupta, P., Kingsley, L. A., Todd, J. A., Saah, A. J., Detels, R., Phair, J. P. & Ri- naldo, J. C. R. (1997). Plasma Viral Load and CD4+ Lymphocytes as Prognostic Markers of HIV-1 Infection. Ann. Intern. Med. 12, 946–954. Nowak, M. A., Anderson, M. R., McLean, A. R., Wolfs, T. F., Goudsmit, J. & May, R. M. (1991). Antigenic diversity thresholds and the development of AIDS. Sci- ence 254, 963–969. Peyerl, F. W., Bazick, H. S., Newberg, M. H., Barouch, D. H., Sodroski, J. & Letvin, N. L. (2004). Fitness costs limit viral escape from cytotoxic t lymphocytes at a structurally constrained epitope. J. Virol. 78, 13901–13910.

Ploegh, H. L. (1998). Viral strategies of immune evasion. Science 280, 248–253.

Quinones˜ Mateu, M. E. & Arts, E. J. (2002). Fitness of drug resistant HIV-1: methodology and clinical implications. Drug Resist. Updat. 5, 224–233.

R Development Core Team and contributors worldwide (2007). The R stats package URL http://cran.r.project.org/.

Richman, D. D., Wrin, T., Little, S. J. & Petropoulos, C. J. (2003). Rapid evolu- tion of the neutralizing antibody response to HIV type 1 infection. Proc. Natl. Acad. Sci. 100, 4144–4149. 130

Russell, M. S., Iskandar, M., Mykytczuk, O. L., Nash, J. H., Krishnan, L. & Sad, S. (2007). A reduced antigen load in vivo, rather than weak inflammation, causes a substantial delay in CD8+ T cell priming against Mycobacterium bovis (bacillus Calmette-Guerin. J. Immunol. 179, 211–220.

Rwambo, P. M., Issel, C. J., Hussain, K. A. & Montelaro, R. C. (1989). In vitro isolation of a neutralization escape mutant of equine infectious anemia virus (EIAV). Arch of Virol 111(4), 275–280.

Schwartz, O., Marechal, V., Le Gall, S., Lemonnier, F. & Heard, J. M. (1996). Endocytosis of major histocompatibility complex class I molecules is induced by the HIV- 1 Nef protein. Nature Med. 2, 338–342.

Seiden, P. & Celada, R. (1992). A model for simulating cognate recognition and re- sponse in the immune system. J. Theor. Biol. 158, 329–357.

Sheldon, J., Rodes,´ B., Zoulim, F., Bartholomeusz, F. & Soriano, A. (2006). Mu- tations affecting the replication capacity of the hepatitis B virus. J. Viral Hepatitis 13, 427–434.

Wang, S., Kennedy, J. S., West, K., Montefiori, D. C., Coley, S., Lawrence, J., Shen, S., Green, S., Rothman, J. A., A L andEnnis, Pal, R., Markham, P. & Lu, S. (2008). Cross-subtype antibody and cellular immune responses induced by a polyvalent DNA prime-protein boost HIV-1 vaccine in healthy human volunteers. Vaccine 26, 1098–1110.

Wei, X., Decker, J. M., Wang, S., Hui, H., JC, K., Wu, X., Salazar-Gonzalez, J. F., Salazar, M. G., Kilby, J. M. & Saag, M. S. (2003). Antibody neutralization and escape by HIV-1. Nature 422, 307–312.

Wodarz, D. & Nowak, M. A. (2002). Mathematical models of HIV pathogenesis and treatment. BioEssays 24, 1178–1187.

Zennou, V., Mammano, F., Paulous, S., Mathez, D. & Clavel, F. (1998). Loss of viral fitness associated with multiple Gag and Gag-Pol processing defects in human immunodeficiency virus type 1 variants selected for resistance to protease inhibitors in vivo. J. Virol. 72, 3300–3306. 131

CHAPTER 4. mlegp: statistical analysis for computer models of biological systems using R

Accepted for publication in Bioinformatics subject to minor software revisions

Garrett M. Dancik,1,2, Karin S. Dorman,1,2,3

1Program in Bioinformatics & Computational Biology, 2Department of Statistics, and

3Department of Genetics, Development & Cell Biology, Iowa State University, Ames, IA

50010.

Abstract

Summary: Gaussian processes (GPs) are flexible statistical models commonly used for predicting output from complex computer codes. As such, GPs are well-suited for the anal- ysis of computer models of biological systems, which have been traditionally difficult to an- alyze due to their high-dimensional, non-linear, and resource-intensive nature. We describe an R package, mlegp, that fits GPs to computer model outputs and performs sensitivity analysis to identify and characterize the effects of important model inputs.

Availability: http://www.biomath.org/mlegp

Contact: [email protected]

Supplementary information: See http://www.biomath.org/mlegp for a user manual and examples. 132

4.1 Introduction

Gaussian processes (GPs) commonly facilitate analysis of computer models that are high-dimensional, non-linear, and resource intensive (Santner et al., 2003) by serving as fast and accurate emulators of these models. GPs play a prominent role in computer model calibration and validation (Kennedy and O’Hagan, 2001; Bayarri et al., 2007), as well as sensitivity analysis to rank inputs in order of importance (based on Functional Analysis of

Variance decomposition) and to characterize their effects (through visual plots of main and two-way factor interactions) (Schonlau and Welch, 2006).

We describe an R package, mlegp, that implements GP modeling with power exponen- tial correlation structure (Santner et al., 2003); the sensitivity analysis methods described in Schonlau and Welch (2006); and the modeling of functional computer model output de- scribed in Heitmann et al. (2006). In addition, mlegp extends previous GP models to handle stochastic computer output with non-constant (heteroscedastic) variance by no longer re- quiring a constant nugget term across observations.

The package is appropriate for what Kitano (2002) describes as simulation-based re- search in systems biology. In this context, computer models have been used to simulate gene expression and signal transduction pathways, e.g., in Escherichia coli (Dobrzy´nski et al., 2007); and infectious disease at the cellular level, e.g., Mycobacterium tuberculosis in- fection (Segovia-Juarez et al., 2004). We demonstrate the capabilities of mlegp by analyzing a computer model of parasitic infection. 133

4.2 Statistical methods

4.2.1 The Gaussian process

(1) (m) Let zobs = [z(x ),...,z(x )] be a vector of observed computer model outputs for

(i) (i) (i) 1 m choices of the input vector x = [x1 ,...,xp ] . We are interested in predicting output z(t) at untried input t.

The correlation between any two computer model outputs is assumed to have the form

p (i) (j) 2 (i) (j) − βk(x −x ) Cij ≡ cor(z(x ),z(x )) = e k=1 k k (4.1) P Let µ(x) be the unconditional mean E[z(x)]. Define the mean matrix

(1) (m) M ≡ E[zobs]=[µ(x ),...,µ(x )].

Under the GP model, computer output follows a multivariate normal distribution

2 2 zobs ∼ MVNm (M,σGP C + σe I), (4.2)

2 where I is the mxm identity matrix, C ≡ {Cij} from equation (4.1), σGP is the uncondi- 2 tional variance of mean computer model output, and the nugget σe accounts for computer model stochasticity. For convenience, denote the variance-covariance matrix as V . Then, the GP predictive distribution of z(t) is normal with mean and variance

2 ′ −1 E[z(t)|zobs] = µ(t)+ σGP r V (zobs − M)

2 2 4 ′ −1 Var[z(t)|zobs] = σGP + σe − σGP r V r, 1In this chapter, we denote the computer model parameter as x, not θ as was done previously 134

′ (i) where r = [r1,...,rm] , with ri = cor(z(t),z(x )) following equation (4.1). For more de- tails, see Santner et al. (2003).

Legend Legend (p1) x1 (85.643) x2 ( 6.885) 0 x3 ( 0.004) 0.5 x4 ( 0.014) 1 x5 ( 0.000) pathogen load pathogen load pathogen load 500 1500 2500 0 5000 10000 0 2000 6000

0 5 10 15 20 25 30 35 0.0 0.2 0.4 0.6 0.8 1.0 0 5 10 15 20 25 30 35

(a) days post infection (b) param value (c) days post infection

Figure 4.1: Sensitivity analysis of a computer model of parasitic infection using the R package mlegp. (a) Computer model output, consisting of pathogen load over time, for 100 simulations obtained by varying inputs x1 through x5, (b) main effects for all inputs on pathogen load at five dpi, along with the percentage contribution of each effect to the total functional variance of the GP predictor, and (c) main effect of x1 on pathogen load over time.

4.2.2 Sensitivity analysis

Schonlau and Welch (2006) describe sensitivity analysis (SA) of computer models us- ing GPs. For independent marginal priors on the components of x, the total variance of the

GP predictor can be decomposed into contributions from single and interacting inputs, a technique called Functional Analysis of Variance (FANOVA) decomposition. The percent of total functional variance attributed to the main effect of an input or the interaction ef- fect among inputs provides a measure of the importance of that effect. The main effect

th E[z(x)|zobs,xk] of the k input variable predicts output, given xk and known results zobs, by integrating against a prior π(x−k) on all remaining variables in x. The two-way interac- tion effect E[z(x)|zobs,xk,xl] is similarly defined. Main effects plots and contour plots con- veniently illustrate main effects and two-factor interactions as functions of the model inputs xk and (xk,xl). 135

4.3 Software

The package mlegp, available in R (R Development Core Team, 2007), finds maximum likelihood estimates of Gaussian process parameters using the likelihood that follows from eq. (4.2). The package extends previous GP models by allowing the user to replace the identity matrix I in (4.2) with a user-defined diagonal matrix N. This extension leads to more accurate GP emulators of heteroscedastic computer models when the variance is known or well-estimated. Another approach to this problem is implemented in the R package tgp, which fits separate GPs to a partitioned input space using a fully Bayesian approach (Gra- macy, 2007). Not all nonconstant variance can be partitioned in this way. On the other hand, our model requires knowledge of the nugget matrix up to a multiplicative constant.

GPs with constant mean functions (i.e., µ(x) ≡ µ0) or linear functions in x are sup- ported. For high dimensional or functional output such as time-series data, the user can opt to fit independent GPs to individual outputs or, instead, to the most important principle component weights following singular value decomposition of the output (Heitmann et al.,

2006). For each GP, the R package provides cross-validated diagnostics, performs FANOVA decomposition, and produces plots for all main and two-way factor interaction effects. Main effects for functional output can also be produced.

4.4 Example application: analysis of a computer model of disease

The SA methods we describe have been used to analyze computer models with up to

40 input variables (Schonlau and Welch, 2006). For demonstration purposes, we use mlegp to analyze the effects of five input variables (x =[x1,...,x5]) in a computer model of Leish- mania major infection (Dancik et al., 2006). Inputs are scaled between 0 and 1 and are described in Supplementary Materials. Computer model output consists of pathogen load over time (Fig. 4.1a). Using mlegp, we fit a GP to 100 observations of pathogen load at 136

five days post infection (dpi). Main effects for all inputs and their FANOVA contributions are provided in Fig. 4.1b. Lastly, we use mlegp to calculate the main effect of x1 (pathogen growth rate) on the temporal evolution of pathogen load by fitting independent GPs to the six most important principle component weights (Fig. 4.1c). The SA shows that the input variable x1 is the most important input for determining pathogen load at five dpi and has a positive relationship with this response (Fig. 4.1b). Low values of x1 result in a gradual increase in pathogen load and a relatively longer infection, whereas high values of x1 result in a sharp increase in pathogen load and a higher peak, but a fast resolution of the infection

(Fig. 4.1c).

In Supplementary Materials, we report additional output from mlegp, including GP diagnostic plots, the FANOVA decomposition, and two-way interaction contour plots, for pathogen load at both five and 18 dpi. We also illustrate the advantage of using a non- constant nugget term for heteroscedastic computer model output.

Funding

National Institutes of Health (GM068955); United States Department of Agriculture

(2001-52100-11506).

Acknowledgement

Partially supported by PHS GM068955 (both) and USDA IFAFS Project 2001-52100-

11506 (GMD). 137

Bibliography

Bayarri, M., Berger, J.O., Paulo, R., Sacks, J., Cafeo, J., Cavendish, J., Lin, C-H., Tu, J. (2007) A framework for validation of computer models. Technometrics, 49, 138-154.

Dancik, G.M., Jones, D.E., Dorman, K.S. (2006) An agent-based model for Leishmania in- fection. Interjournal of Complex Systems, 1853.

Dobrzy´nski, M., Rodr´ıguez, J.V., Kaandorp, A., Blom, J.G. (2007). Computational meth- ods for diffusion-influenced biochemical reactions. Bioinformatics, 23, 1969-1977.

Gramacy, R.B. (2007) tgp: An R package for Bayesian nonstationary, semiparametric non- linear regression and design by treed Gaussian process models. J. Stat. Soft., 19(9).

Heitmann, K., Higdon, D., Nakhleh, C., Habib, S. (2006) Cosmic Calibration, Astrophys. J., 646, L1-L4.

Kennedy, M.C., O.Hagan, A. (2001) Bayesian calibration of computer models. J.R. Statist. Soc. B, 63, 425-464.

Kitano, H. (2002) Computational systems biology. Nature, 420, 206-210.

R Development Core Team. (2007) R: A Language and Environment for Statistical Com- puting. R Foundation for Statistical Computing, Vienna, Austria. ISBN 3-900051-07-0, URL http://www.R-project.org/.

Santner, T.J., Williams, B.J., Notz, W. (2003) The Design and Analysis of Computer Ex- periments. Springer, New York.

Schonlau, M. and Welch, W. (2006). Screening the input variables to a computer model via analysis of variance and visualization. In Dean, A., and Lewis., S. (eds), Screening: Meth- ods for Experimentation in Industry, Drug Discovery, and Genetics. Springer, New York, pp. 308-327.

Segovia-Juarez, J.L., Ganguli, S., Kirschner, D. (2004) Identifying control mechanisms of granuloma formation during M. tuberculosis infection using an agent-based model. J. Theor. Biol., 231, 357-376. 138

4.5 Supplementary Materials

This section contains supplementary material for the paper ‘mlegp: statistical analy- sis for computer models of biological systems using R’, by Garrett M. Dancik and Karin S.

Dorman.

4.5.1 Methods

Only methods not discussed in the main paper are described here. A cross-validated prediction of the observation z(x(i)) is the prediction obtained following the removal of all

(i) observations at x from the training set zobs. Denoting zobs,-i as zobs without observations

(i) (i) at x , the cross-validated prediction is E z(x ) | zobs,-i . Plotting cross-validated predic-

(i) h (i) i tions E z(x ) | zobs,-i against observation z(x ) produces diagnostic plots, with predicted

h (i) i variances Var z(x ) | zobs,-i used to add confidence bounds. h i To evaluate the benefit of accounting for heterscedasticity in the data, we analyze pathogen load at five days post infection. We fit one Gaussian process (GP) using a con- stant nugget term and another GP with a diagonal nugget matrix whose ith element is pro- portional to the sample variance estimated from two independent computer runs at param- eter setting x(i). Diagnostic plots are used to compare the performance of these two meth- ods.

4.5.2 Results and Discussion

In Table 4.1 we describe the five parameters x1,...,x5 varied during simulation of the computer model. After fitting a GP to the model, we examine diagnostic plots comparing cross-validated predictions from the GP predictors with observed values for these responses

(Fig. 4.5.2). The early time point is better predicted by the GP than the late time point; 139 nevertheless, 93% of true observations at the late time point fall in the 95% confidence in- tervals (data not shown).

We report the Functional Analysis of Variance (FANOVA) decomposition contribu- tions for all main and two-way factor interaction effects for pathogen load at five and 18 days post infection (dpi) in Table 4.1. Main effects of all model parameters on pathogen load at five and 18 dpi are given in Fig. 4.3, and interaction effects for the most important parameter interactions at each time point are provided in Fig. 4.4. The results from mlegp show that the growth rate parameter x1 is the most influential parameter for pathogen load at both five and 18 dpi (Table 4.1), but the effect of this parameter on pathogen load changes over time (Fig. 4.3). Two other important parameters are the carrying capacity of infected cells (x2), which is positively related to pathogen load at five dpi, and the thresh- old level of pathogen that triggers the adaptive immune response (x5), which is positively related to pathogen load at 18 dpi (Table 4.1, Fig. 4.3). However, both of these parameters have large interaction effects with x1 at their respective time points (Table 4.1), indicating that the nature of their effects depends on the value of x1. Based on the interaction effect plots (Fig. 4.4), x2 has little effect on pathogen load at five dpi when x1 is low, while x5 has little effect on pathogen load at 18 dpi when x1 is high. Lastly, we demonstrate the advantage of using an arbitrary diagonal nugget matrix, rather than a constant nugget term, in a GP model of pathogen load at five dpi, where out- put is heteroscedastic. Cross-validated diagnostic plots for both GPs are provided in Fig.

4.5, where we see the GP with diagonal nugget matrix is a more precise model. In partic- ular, confidence bands of the constant nugget GP are overly conservative for low values of the response, with 100% of the data points falling in the 95% confidence intervals when pathogen load is less than 1900, compared to 94.7% of the data points for the GP with a diagonal nugget matrix. 140 predicted predicted 0 1000 3000 −2000 2000 6000 1000 2000 3000 4000 0 2000 4000 6000 8000

observed observed

(a) GP emulator for pathogen load at 5 dpi (b) GP emulator for pathogen load at 18 dpi

Figure 4.2: Gaussian process diagnostic plots: solid black line, observed output vs. ob- served output; solid red lines, cross-validated prediction ± 1 standard deviation.

Legend Legend x1 (85.643) x1 (56.542) x2 ( 6.885) x2 ( 2.794) x3 ( 0.004) x3 ( 0.000) x4 ( 0.014) x4 ( 2.199) x5 ( 0.000) x5 (26.467) pathogen load pathogen load 1000 3000 500 1500 2500

0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0

(b) param value (b) param value

(a) 5 dpi (b) 18 dpi

Figure 4.3: Main effects of all parameters using the Guassian process predictors for pathogen load at five and 18 dpi. 141

description parameter 5 dpi 18 dpi pathogen growth rate x1 85.64 56.54 carrying capacity of infected host cell x2 6.89 2.79 (macrophage) macrophage recruitment parameter x3 0.00 0.00 T cell recruitment rate x4 0.01 2.20 threshold level of pathogen that triggers adap- x5 0.00 26.47 tive immune response x1:x2 7.40 0.39 x1:x3 0.00 0.00 x1:x4 0.00 0.08 x1:x5 0.00 11.38 x2:x3 0.00 0.00 x2:x4 0.00 0.03 x2:x5 0.00 0.06 x3:x4 0.00 0.00 x3:x5 0.00 0.00 x4:x5 0.00 0.06 All main and 2-way interactions 99.94 100.00

Table 4.1: FANOVA Decomposition of GP predictors for pathogen load at five and 18 dpi. The contribution of each effect is reported as a percentage of the total functional variance of each GP predictor. 142 x2 0.0 0.4 0.8

0.0 0.2 0.4 0.6 0.8 1.0

x1

(a) x1 and x2 at 5 dpi x5 0.0 0.4 0.8

0.0 0.2 0.4 0.6 0.8 1.0

x1

(b) x1 and x5 at 18 dpi

Figure 4.4: Interaction effects of the most important interactions for pathogen load at five and 18 dpi. 143 predicted 0 1000 3000

1000 2000 3000 4000

observed

(a) GP with constant nugget term predicted 1000 3000

1000 2000 3000 4000

observed

(b) GP with user-defined diagonal nugget matrix

Figure 4.5: GP diagnostic plots for pathogen load at 5 dpi, for (a) GP with constant nugget term and (b) GP with a diagonal nugget matrix whose ith element is proportional to the sample variance of two indepent computer model runs evaluated at x(i); solid black lines, observed output vs. observed output; solid red lines, predicted mean output ± 2 standard deviations. 144

CHAPTER 5. General conclusions and future work

The immune response is incredibly complex. At the cellular level, individual cells move, react, and interact in response to chemokines, cytokines, and other environmental signals, creating a chain of events that ultimately results in cellular and humoral immunity.

Despite this complexity, there has been a great deal of success in using computer models to study immune responses to infectious diseases (Bernaschi & Castiglione, 2002; Castiglione et al., 2004; Segovia-Juarez et al., 2004; Beauchemin, 2006). In this thesis, I have described the development and calibration of a computer model of Leishmania major infection, the identification of correlates of escape mutant success in a computer model of a viral infec- tion, and statistical software to aid in the analysis and calibration of computer models.

5.1 Host-pathogen interactions and emergent behavior in intracellular infections

The emergent behavior of a system, by definition, can only be understood through consideration of how system components interact. In this thesis, analysis of two computer models reveal insight into immune and pathogen behavior not obvious unless interactions between individual components or pathogen and immune parameters are considered.

In the agent-based model of L. major infection, growth rate has the largest effect on the first macrophage time point, despite the fact that growth rate is a pathogen parameter.

This happens because increasing growth rate causes infected macrophages to reach carry- 145 ing capacity sooner, which spreads the infection to additional cells. Infected macrophages release chemokine, which boosts macrophage recruitment. Increasing growth rate can also favor or suppress pathogen loads, depending on the stage of the infection and the ability of the pathogen to avoid detection. This is because a pathogen that proliferates rapidly risks being quickly detected and eliminated by the immune response. As a result, higher growth rates lead to higher peak pathogen loads but shorter infection lengths.

In in silico experiments involving the injection of escape mutants, a key finding is that loss of infectivity in humoral escape mutants benefits the virus, despite the fact that less infectious escape mutants would generally be considered less fit in vitro. This is because the strength of the cellular response depends on the infectiousness of the virus (Arnaout &

Nowak, 2000), and decreasing infectivity weakens the cellular response. We also find that simultaneous escape from both cellular and humoral responses is substantially more success- ful than escape from either response alone. This is because of the multiplicative effect of be- ing able to infect more cells (humoral escape) and producing more virus (cellular escape).

In general, this finding highlights the importance of both immune responses in control- ling infection, and supports the need for vaccination strategies that induce both responses

(Wang et al., 2008).

5.2 Future work

In the future, I plan on extending the computer models and the statistical software presented in this thesis. In L. major infection, low levels of pathogen persist following reso- lution of the infection, a stage of the infection that the current model cannot capture. It is known that natural regulatory T cells (Tregs) are required to maintain pathogen persistence

(Belkaid et al., 2003). However, the antigen-specificity of Tregs is not entirely clear. Sen- sitivity analysis could reveal implications of alternative models of Treg antigen-specificity. 146

The computer model could also be used to quantitatively compare alternative models.

Equine infectious anemia virus (EIAV) is a close relative of HIV that causes persis- tent infection in horses with disease characterized by acute, chronic, and asymptomatic stages (Leroux et al., 2004). Interestingly, evolution of the protein Rev, which facilitates nu- clear export of incompletely spliced viral RNA, correlates with the clinical stage of the dis- ease and Rev activity is highest during the chronic stage (Belshan et al., 2001). Our modi-

fied version of C-IMMSIM will be extended and calibrated to simulate EIAV infection, and will then be used to characterize the fitness advantage achieved through increased Rev ac- tivity.

Finally, a limitation of current Gaussian process (GP) predictors is the assumption that the computer model is either deterministic or has constant variance (i.e., is home- oscedastic). There is therefore a need for statistical methods that accurately predict het- eroscedastic computer model output. The R package mlegp, which allows the user to specify a diagonal nugget matrix up to a multiplicative constant, is a first step in this direction. In future versions of the package, I will allow for hierarchical GP models that use one GP to predict computer model variance and a second GP that predicts computer model output, conditional on the predicted computer model variance of known runs. For heterescedastic computer models, a hierarchical GP model will minimize bias in GP predictions, improve posterior predictions in calibration, and could also be adapted for use in computer model optimization (Jones et al., 2004) and adaptive sampling (Gramacy, 2005). 147

Bibliography

Arnaout, R. A. & Nowak, M. A. (2000). Competitive Coexistence in Antiviral Immu- nity 204, 431–441.

Beauchemin, C. (2006). Probing the effects of the well-mixed assumption on viral infec- tion dynamics. J. Theor. Bio. 242, 464–477.

Belkaid, Y., Piccirillo, C. A., Me´ndez, S., Shevach, E. M. & Sacks, D. L. (2003). CD4+CD25+ immunoregulatory T cells control Leishmania major persistence and the development of concomitant immunity. Nature 420, 502–507.

Belshan, M., Baccam, P., Oaks, J. L., Sponseller, B. A., Murphy, S. C., Cor- nette, J. & Carpenter, S. (2001). Genetic and biological variation in equine infec- tious anemia virus rev correlates with variable stages of disease in an experimentally in- fected pony. Virology 279, 185–200.

Bernaschi, M. & Castiglione, F. (2002). Selection of escape mutants from immune recognition during HIV infection. Immunol. Cell. Bio. 80, 307–313.

Castiglione, F., Poccia, F., D’Offizi, G. & Bernaschi, M. (2004). Mutation, fit- ness, viral diversity, and predictive markers of disease progression in a computational model of HIV type 1 infection. AIDS Res. Hum. Retroviruses 20(12), 1314–1323.

Gramacy, R. B. (2005). Bayesian treed Gaussian process models. Ph.D. thesis, University of California.

Jones, D. R., Schonlau, M. & Welch, W. J. (2004). Efficient global optimization of expensive black-box functions. J. Glob. Opt. 13, 455–492.

Leroux, C., Cadore, J.-L. & Montelaro, R. C. (2004). Equine anermica virus (EIAV): what has HIV’s country cousin got to tell us? Vet. Res. 35, 485–512.

Segovia-Juarez, J. L., Ganguli, S. & Kirschner, D. (2004). Identifying control mechanisms of granuloma formation in M. tuberculosis infection using an agent-based model. J. Theor. Biol. 231, 357–376. 148

Wang, S., Kennedy, J. S., West, K., Montefiori, D. C., Coley, S., Lawrence, J., Shen, S., Green, S., Rothman, J. A., A L andEnnis, Pal, R., Markham, P. & Lu, S. (2008). Cross-subtype antibody and cellular immune responses induced by a polyvalent DNA prime-protein boost HIV-1 vaccine in healthy human volunteers. Vaccine 26, 1098–1110.