A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood...

182
A Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical Engineering and Computer Sciences University of California at Berkeley Technical Report No. UCB/EECS-2008-165 http://www.eecs.berkeley.edu/Pubs/TechRpts/2008/EECS-2008-165.html December 17, 2008

Transcript of A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood...

Page 1: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

A Computational Model of Human Blood Clotting:Simulation, Analysis, Control, and Validation

Joseph Gerard Makin

Electrical Engineering and Computer SciencesUniversity of California at Berkeley

Technical Report No. UCB/EECS-2008-165

http://www.eecs.berkeley.edu/Pubs/TechRpts/2008/EECS-2008-165.html

December 17, 2008

Page 2: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Copyright 2008, by the author(s).All rights reserved.

Permission to make digital or hard copies of all or part of this work forpersonal or classroom use is granted without fee provided that copies arenot made or distributed for profit or commercial advantage and that copiesbear this notice and the full citation on the first page. To copy otherwise, torepublish, to post on servers or to redistribute to lists, requires prior specificpermission.

Acknowledgement

Thanks to Profs. Srini Narayanan and Jerome Feldman for theirsupervision of this thesis; to Profs. Bruno Olshausen and Jose Carmena forserving as readers; and to Profs. Shankar Sastry and Tom Budinger forserving as qualification examiners.

Page 3: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

A Computational Model of Human Blood Clotting: Simulation, Analysis,Control, and Validation

by

Joseph Gerard Makin

B.S., B.A. Swarthmore College 2003

A dissertation submitted in partial satisfactionof the requirements for the degree of

Doctor of Philosophy

in

Engineering - Electrical Engineering and Computer Sciences

in the

GRADUATE DIVISION

of the

UNIVERSITY OF CALIFORNIA, BERKELEY

Committee in charge:

Professor Jerome Feldman, ChairProfessor Bruno Olshausen

Professor Jose CarmenaProfessor Srini Narayanan

Fall 2008

Page 4: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

The dissertation of Joseph Gerard Makin is approved.

Chair Date

Date

Date

Date

University of California, Berkeley

Fall 2008

Page 5: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

A Computational Model of Human Blood Clotting: Simulation, Analysis, Control,

and Validation

Copyright c© 2008

by

Joseph Gerard Makin

Page 6: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Abstract

A Computational Model of Human Blood Clotting: Simulation, Analysis, Control,

and Validation

by

Joseph Gerard Makin

Doctor of Philosophy in Engineering - Electrical Engineering and Computer Sciences

University of California, Berkeley

Professor Jerome Feldman, Chair

Complex biological systems pose many challenges to researchers, including, inter alia, choice

of computational model, with its consequences for simulation and analysis; methods of ma-

nipulating the system exogenously (control); and model validation. I attempt to address

these issues for human blood clotting. By treating the system as comprising interacting

discrete and continuous aspects, i.e. as hybrid, the entire coagulation cascade may be

simulated: Blood proteins, elemental ions, and other state elements are modeled either

as real-valued concentrations or as binary variables (present/absent); interactions are ren-

dered either as ODEs (per their chemical equations) or as discrete events. Techniques from

nonlinear control theory are then used to devise drug therapies for diseased patients. Fi-

nally, the model is used to warp variations in the input parameters—rate constants and

initial conditions—into an output space where pathologies and healthy clotting are cleanly

separated by a semi-supervised clustering analysis. This serves to validate the model as

1

Page 7: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

well as to summarize efficiently the predicted clinical consequences of individual variations.

Professor Jerome FeldmanDissertation Committee Chair

2

Page 8: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

To S.C.M.: who predicted it.

i

Page 9: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Contents

Contents ii

List of Figures v

List of Tables viii

Acknowledgments ix

1 Introduction 1

1.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.2 Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

2 The Coagulation Cascade 8

2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2.2 Preliminaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2.3 The Cascade . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

2.3.1 Intrinsic pathway . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

2.3.2 Extrinsic pathway . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

2.3.3 Common pathway . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

2.3.4 Down-regulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

2.3.5 Completeness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

2.4 Diseases of Coagulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

2.5 Mathematical Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

3 Hybrid Systems 23

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23

3.2 Types of Hybrid-System Investigation . . . . . . . . . . . . . . . . . . . . . 25

ii

Page 10: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

3.2.1 Modeling and simulation . . . . . . . . . . . . . . . . . . . . . . . . . 26

3.2.2 Verification and decidability . . . . . . . . . . . . . . . . . . . . . . . 27

3.2.3 Controller synthesis . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

3.3 Hybrid Petri Nets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

3.4 Hybrid Systems in Biology . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

4 Simulations and Sensitivity 41

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

4.2 Model Implementation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

4.2.1 Modules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

4.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

4.3.1 Normal clotting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

4.3.2 Simulating hæmophilia A . . . . . . . . . . . . . . . . . . . . . . . . 56

4.3.3 Simulating factor-V Leiden . . . . . . . . . . . . . . . . . . . . . . . 57

4.4 Decomposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

4.5 Sensitivity Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

4.5.1 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

4.5.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

4.5.3 Sensitivity and prothrombin time . . . . . . . . . . . . . . . . . . . . 64

4.5.4 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

4.6 Discussion and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

5 Control: A First Pass 73

5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

5.2 Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

5.3 Application to the model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

5.3.1 Feedback linearization . . . . . . . . . . . . . . . . . . . . . . . . . . 78

5.3.2 Step-input control . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82

5.4 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82

5.5 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

5.6 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

5.6.1 Controllability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94

5.6.2 Non-regularity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95

5.6.3 Controller merits and demerits . . . . . . . . . . . . . . . . . . . . . 97

iii

Page 11: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

5.6.4 Significance of model assumptions . . . . . . . . . . . . . . . . . . . 98

5.7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99

6 Analysis: Stability, Model Reduction, and Control 110

6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110

6.2 Model Reduction and Bounded Tracking . . . . . . . . . . . . . . . . . . . . 111

6.2.1 Bounded tracking and minimum-phase systems . . . . . . . . . . . . 111

6.2.2 Construction of the zero dynamics . . . . . . . . . . . . . . . . . . . 112

6.2.3 Intermezzo: model reduction . . . . . . . . . . . . . . . . . . . . . . 115

6.2.4 Stability of the zero dynamics . . . . . . . . . . . . . . . . . . . . . . 118

6.3 Control: Model-Predictive and Proportional . . . . . . . . . . . . . . . . . . 119

6.3.1 State observability and feedback linearization . . . . . . . . . . . . . 119

6.3.2 Control with constraints: model prediction . . . . . . . . . . . . . . 122

6.4 Discussion and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . 126

7 Validation/Learning 135

7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135

7.2 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137

7.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139

7.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140

7.4.1 Clustering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140

7.4.2 Decision tree . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141

7.4.3 A mathematical take . . . . . . . . . . . . . . . . . . . . . . . . . . . 142

7.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144

8 Conclusions & Future Work 146

8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146

8.2 The Hybrid-System Approach . . . . . . . . . . . . . . . . . . . . . . . . . . 147

8.3 Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149

8.3.1 Modeling assumptions . . . . . . . . . . . . . . . . . . . . . . . . . . 149

8.3.2 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150

8.4 Learning/Classification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152

8.5 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153

Bibliography 158

iv

Page 12: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

List of Figures

2.1 The coagulation cascade. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

3.1 A hybrid system model of a thermostat; see text for details. . . . . . . . . . 24

3.2 A very simple Discrete Petri Net (DPN) with purely discrete components.The input and test arcs’ respective resource requirements were satisfied, butthe inhibitory arc’s requirement is not—so the transition is enabled. . . . . 30

3.3 The simple Discrete Petri Net (DPN) of Figure 3.2, advanced by one step:the input arc’s place has been depleted by its resource requirement (1 token),and the output arc’s place has been augmented by its weight (1 token). . . 31

3.4 A very simple Petri net with purely continuous components (CPN). The rateat which “token fluid” leaves m1 is the rate at which it accumulates in m2,which in this example is m2(m1 + 1) . . . . . . . . . . . . . . . . . . . . . . 35

3.5 A summary of the various types of arcs in an HPN; cf. Defs. 3.3.6 and 3.3.7. 37

4.1 The activation module, which models the activation of a zymogen (IN1) intoits active configuration (OUT) by a catalyst (IN2). . . . . . . . . . . . . . . 46

4.2 The binding module, which models the (reversible) binding of two compo-nents, IN1 and IN2, into the macromolecule OUT. . . . . . . . . . . . . . . 47

4.3 A hybrid Petri net module modeling the initiation of the intrinsic pathway. 48

4.4 An HPN module depicting the final stages of blood clotting: the activationof factors I and XIII, and the formation of a clot. . . . . . . . . . . . . . . . 50

4.5 The activation module of Figure 4.1, but here two of the reactions areswitched on (off) by the presence (absence) of a discrete variable. . . . . . . 53

4.6 A simulation of the concentration of thrombin (factor IIa) in nM as it varieswith time (in seconds) in the blood clotting of a normal patient. . . . . . . 56

4.7 Simulated results of the concentration (nM) of thrombin versus time (s) dur-ing the clotting process for a person with severe hæmophilia A (low curve)and for normal clotting (high curve). . . . . . . . . . . . . . . . . . . . . . . 57

v

Page 13: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

4.8 Simulated results of the concentration (nM) of thrombin versus time (s) dur-ing the clotting process for a person with factor-V Leiden (high curve) anda normal patient (low curve). . . . . . . . . . . . . . . . . . . . . . . . . . . 58

5.1 A conceptual rendering of feedback linearization. The arrangement can beviewed either as a controller in a feedback loop with the original control-affine system (solid lines), plus an output function; or as an equivalent outputfunction along with coupled linear and nonlinear systems (dashed lines), withthe former evolving independently of the latter, and the output dependingsolely on the former. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101

5.2 Simulated control of thrombin concentration during a clotting event in apatient with factor-V Leiden. The controller was unconstrained, and wasupdated continuously, allowing arbitrarily close tracking of the desired tra-jectory. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102

5.3 The heparin input in nM/s that generated Figure 5.2. Although this inputyields perfect tracking, it is obviously unacceptable for any realistic drug-delivery mechanism, operating as it does on a continuous time scale, andusing negative inputs (see text). . . . . . . . . . . . . . . . . . . . . . . . . . 103

5.4 Simulated control of thrombin concentration during a clotting event in apatient with factor-V Leiden, again continuously sampling. The input wasconstrained to the range 0-20 nM/s. . . . . . . . . . . . . . . . . . . . . . . 104

5.5 The input generating the output of Figure 5.4. Note that the input is con-strained to the range 0-20 nM/s. . . . . . . . . . . . . . . . . . . . . . . . . 105

5.6 Simulated control of thrombin concentration during a clotting event in apatient with factor-V Leiden, this time using a discrete controller samplingevery 0.5 s. The input was again constrained in the range between 0 and 20nM/s. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106

5.7 The discrete input, constrained to lie between 0 and 20 nM/s, that generatedFigure 5.6. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107

5.8 Simulated control of thrombin in a patient with moderate hæmophilia A (seetext), where the controller inputs the proper initial concentration of factorVIII at the first time step, then shuts off. . . . . . . . . . . . . . . . . . . . 108

5.9 Control of thrombin concentration in a hæmophiliac using a single step inputat time t = 0. The input is 6.92 pM/s. . . . . . . . . . . . . . . . . . . . . . 109

5.10 Control of thrombin concentration in a patient with factor-V Leiden by asingle step input at time t = 0. The input is 4.39 nM/s. . . . . . . . . . . . 109

6.1 Simulated proportional control of thrombin concentration during a clottingevent in a patient with factor-V Leiden, once more using a discrete controllersampling every 0.5 s (as in the previous chapter), and with input again con-fined to the range between 0 and 20 nM/s. . . . . . . . . . . . . . . . . . . . 121

6.2 The discrete input, constrained to lie between 0 and 20 nM/s, that generatedFigure 6.1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122

vi

Page 14: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

6.3 Simulated proportional control of thrombin during a clotting event in a pa-tient with factor-V Leiden, again at 2 Hz, but with input constrained to therange 0-5 nM/s. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126

6.4 The proportional controller’s discrete input for Figure 6.3, constrained to liebetween 0 and 5 nM/s . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127

6.5 Factor-V Leiden, treated by heparin via a feedback-linearization controlleroperating at 2 Hz, and within the range 0-5 nM/s. . . . . . . . . . . . . . . 128

6.6 The discrete input generated by the feedback-linearization scheme, clampedbetween 0 and 5 nM/s, that produced Figure 6.5. . . . . . . . . . . . . . . . 129

6.7 Control of thrombin using the LMPC scheme in conjunction with feedbacklinearization. Again, the uncontrolled trajectory is a result of factor-V Lei-den, and the input rate of heparin is confined to the range 0-5 nM/s. Notethe improvement over Figure 6.5. . . . . . . . . . . . . . . . . . . . . . . . 130

6.8 The discrete input (constrained to 0-5 nN/s) generated by the LMPCer andproducing Figure 6.7. Sampling rate is 0.5 Hz . . . . . . . . . . . . . . . . . 131

7.1 An output space in which hypo- (black x’s), hyper- (light gray +’s), andnormal (gray circles) coagulation are cleanly separated by k-means, withk = 4 The white hexagram, black square, and white pentagram (resp.) arethe known exemplars of the three classes. See text for an interpretation ofthe fourth class (gray diamonds). . . . . . . . . . . . . . . . . . . . . . . . . 139

7.2 The decision tree. The class labels are M = medium, L = low, H = high,and O = outliers for the four classes of coagulation levels. . . . . . . . . . . 145

vii

Page 15: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

List of Tables

2.1 Primary Coagulation Factors and their Pre-Injury Concentrations . . . . . 11

4.1 Parameter values and types used in the HPN modules of Figures 4.4 and 4.3 52

4.2 Coagulation Disorders . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

4.3 Averaged Normalized Sensitivities . . . . . . . . . . . . . . . . . . . . . . . 63

4.4 PT-Time Parameter Shifts . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

5.1 Chemical Reaction Set I . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83

5.2 Chemical Reaction Set II . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

5.3 Chemical Reaction Set III . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89

5.4 Initial Conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

viii

Page 16: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Acknowledgments

Most people probably read the acknowledgments of a thesis in search of their own names;

a few for completeness; and the rest hoping to get an inside joke. Almost everyone, then,

will be disappointed.

My first thanks are to Prof. Srini Narayanan, who conceived of the project, took me on

to work on it, and guided me along the way. But Srini has been more than an advisor: he’s

been a friend. That means I told lots of inappropriate jokes around him, but also that he

listened to my complaints about personal life, palavered about politics and the NFL (even

if he pulled for the front-runners), drank with us, and even kept some track of the number

of miles I ran in a week.

Prof. Jerry Feldman served as my “official” advisor, and many thanks are due to him

as well. Jerry was the voice of wisdom on technical matters—he suggested the learned

step-controller of Chapter 5—as well as all the practical concerns of the graduate program.

More importantly, his “neural theory of language” is what induced me to join the group.

Thanks also to my other committee members, Drs. Jose Carmena and Bruno Olshausen,

for their time—but also especially to Prof. Carmena for filling in as a reader at the last

minute and with such alacrity; and to Bruno for the hours spent in his reading group, hon-

estly the most enjoyable of my academic life at Berkeley—and for being an all-around nice

guy. I owe the other members of my qualifying-exam committee, as well: Dr. Tom Budinger

of the Lawrence-Berkeley Lab; and Prof. Shankar Sastry, who additionally provided many

useful hours of instruction in the classroom, and the authoritative textbook on nonlinear

systems.

My peers have contributed to this thesis, too, in one way or another, and deserve a shout-

out as well. Alessandro Abate ran the bio-hybrid-systems reading group, and gave me plenty

of useful advice on control theory. I owe Jeff Doyon for my tuition on chemical kinetics

and for our profitable conversations on research (mine and his both)—mostly conducted on

long runs. Matt Gedigian managed (inexplicably) to make most of my campus talks, and

read drafts of my papers. And thanks to my officemates: Steve Sinha, with whom I had

ix

Page 17: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

many enjoyable conversations about politics, and a few about technical matters; and Leon

Barrett, with whom I had many useful conversations on technical matters, and a few about

politics. The remainder of my thanks for fellow students must be divided among the rest of

our research group—Nancy Chang, John Bryant, and Eva Mok—and attributed to Friday

evenings at the bar.

Without my parents, Thomas and Donita Makin, I should not be here at all. My debt

to them is beyond evaluation.

Penultimate thanks are to the authors of the TEX template for this thesis and of the

ucthesis.cls class file! and to the indispensable Ruth Gjerde, who is solely responsible for

my having filed just about every piece of department paperwork, and with whom I always

had a pleasant chat.

Finally, my largest debt of gratitude for this dissertation is to Professor Lynne Molter

of Swarthmore College, who first taught me the rudiments of control theory, who first

employed me as a researcher, who encouraged my interest in electrical engineering and in

graduate studies, and who remains a friend.

x

Page 18: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

xi

Page 19: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Chapter 1

Introduction

1.1 Overview

It has long been a dream of science to apply the leverage of its more mathematical

disciplines to its biological, especially medical, investigations. Biology is messy, however,

and will not be shoehorned into the same kind of mathematical formalization as (say)

physics or information theory or even chemistry. Yet a mathematical approach may very

well help us better to understand a biological process, or to manipulate it, or to connect

changes in its structure to changes in its function and make predictions on that basis; and

hence to diagnose and treat diseases. This thesis, then, labors in the service of a dream,

but not a vain one.

The object of this study is human blood clotting, a process exhibiting complexity typ-

ical of biological systems: no single set of differential equations or discrete-process model

can completely describe the coagulation cascade (at least not given the current state of

knowledge); yet on the other hand it is obviously of crucial clinical importance. Disorders

are various and some not uncommon: hæmophilia A affects about one in ten thousand

individuals; protein-C deficiency one in five hundred; von Willebrand disease about one in

a hundred; and the recently (1994) identified factor-V Leiden may be as common as one in

twenty individuals (Crookston; King). (A detailed description of the blood-clotting process

1

Page 20: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

and its disorders appears in Chapter 2.) Clinicians would like to understand these diseases

better, as well as to treat them effectively.

The complexity of the coagulation cascade limits the usefulness of qualitative descrip-

tions of the roles of its proteins and reactions. Such limitations apply a fortiori to any

predictions of the effects of changes to these proteins and their reactions that are based

(solely) on a set of qualitative descriptions. Here the usefulness of a computational model

comes into view: alterations of initial protein concentrations, or of reaction rates, or of

the presence or absence of certain reaction-enabling ions (like calcium) can be simulated in

a such a model, and the consequences directly observed—at least insofar as the model is

accurate.

But how should one model a system that is not completely describable in terms of

continuous processes and states, and in which qualitative information is to play a role?

The approach of this thesis is to treat coagulation as comprising interacting discrete and

continuous aspects, i.e. as hybrid. Choosing this formalism exacts a price in analyzability,

but it pays off in the form of a complete model of the coagulation cascade, in contrast

to other models (see Chapter 2). That is, many of the mathematical tools which exist

for analyzing purely discrete or purely continuous systems (stability, location of equilibria,

controllability and observability, etc.) do not exist for hybrid systems—the theory of which

is described in detail in Chapter 3; but homogeneous models cannot (properly) embrace the

breadth of the coagulation process. To capture the entire system, then, I have constructed

a hybrid-system model, which along with its simulations appears in Chapter 4. As for

the deficiencies of analysis, we shall see that for a large class of cases, they can in fact be

circumvented.

We are interested, to repeat, in treating diseases as well as in understanding them (and

non-pathological clotting), and here too the current state of the art would benefit from

a mathematical model. So, for example, the hypercoagulatory disorder factor-V Leiden,

which serves as a principal case study in this paper, is usually treated by daily visits to the

doctor’s office, where blood is drawn and protein concentrations are measured, after which

an anti-coagulant (like heparin) is administered. If certain protein concentrations exceed

2

Page 21: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

ranges which clinical experience has found to be unsafe, the dosage is increased or decreased

as appropriate; otherwise yesterday’s dosages are re-administered.

This kind of treatment could certainly be improved upon. To begin with, the measure-

ment interval (one day) is rather large; and the method is “global” rather than local: that

is, treatment is effected by changing overall protein concentrations—throughout the entire

blood stream—rather than at the site of a thrombotic event. The difficulty here is techno-

logical, however, rather than mathematical—solution requires the appropriate sensors and

drug delivery mechanisms—which difficulty this document will ignore; in fact, this thesis

will sometimes assume that these issues are resolvable. The more interesting shortcoming of

present treatments is that they prescind away from all but the rudest qualitative knowledge

of blood clotting: too much of this protein implies a need for an anticoagulant; too little of

that protein demands a procoagulant. In Chapters 5 and 6 I shall show how a mathematical

model of coagulation can be exploited to allow much finer control of clotting, using some

tools from computer science and nonlinear control theory.

I have already alluded to the question of the model’s fidelity, which motivates the

remaining portions of this thesis. It must first be said that no complete comparison between

the data produced by the model and actual in vivo clotting is possible, since (as mentioned

above) the sensors for measuring all the protein concentrations at the site of an injury in

real time do not exist. This is not, of course, to say that the parameters (rate constants)

and structure (reactions) of this model are not the result of empirical investigation; indeed,

they are, but of countless (painstaking) experiments on the interactions of pairs or triads

or perhaps a few blood proteins. These experiments are, first of all, perforce in vitro, so

measured rate constants might vary from their in vivo cousins in virtue of (e.g.) a different

prevailing ion concentration, or of the substitution of a fluid reaction for one on a substrate.

They, secondly, depend on hypotheses about which proteins effectively interact with each

other. These considerations ensure that the model will have errors.

Since the data do not exist for a complete in vivo validation, then, how are we to

interpret the results? The simulations from the hybrid-system model (Chapter 4) should

be interpreted as demonstrating the utility of such a model, not as providing gold-standard

3

Page 22: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

simulation results; as well as providing a mode of comparison for future clinical data and

hence a method for refining the model. The control schemes of Chapters 5 and 6 are

of course not intended to provide precise numerical values for the controllers; indeed, the

feedback controllers are intended to obviate exactly the need for such model precision (since

uncertainty will anyway be introduced, even in the case of a “perfect model,” by inter-

individual differences). A fuller discussion of the effect of model errors on the control

schemes appears in Chapters 5 and 6.

What other kinds of information can profitably be reaped from an imperfect model?

One such is the sensitivity of the model to changes in its parameters, i.e. the differential

changes in state over time from the nominal (unperturbed) trajectory, introduced by (static)

changes in rate constants. Such results do not depend on exact values of rate constants; in

fact, the sensitivity of the system reveals just how much errors in rate constants matter to

overall results. Sensitivity analysis also serves inter alia a clinical purpose, revealing where

in the clotting cascade pharmaceutical intervention will have the most effect with the least

effort. A sensitivity analysis of a large portion of the clotting cascade appears in Chapter

4.

Finally, this thesis proposes a method, exploiting some well-established machine-learning

techniques, for making use of what clinical data do exist, for simultaneously validating and

refining the model. The trick is to rely on the model where correct and on “gold standard”

data where they exist. In brief, the model is interpreted as a map from an input space

into an output space, where the former consists of, e.g., initial condition values, or again

of rate-constant values; and the latter consists of (say) thrombin concentrations at various

intervals during the course of a clot event, as predicted by the model. The output data

can then be partitioned in a semi-supervised manner, the labeled data consisting of known

inputs (e.g. initial conditions) and their corresponding outputs (e.g. “hypercoagulatory”

vs. “normal” vs. “hypocoagulatory”); and the unlabelled data comprising the remaining

input/output pairs. This technique is explored in the penultimate chapter (7).

A final chapter (8) sums up the contributions of this thesis; laments the loose ends it

4

Page 23: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

failed to tie up; and recommends some of these as well as some additional threads to future

investigators.

1.2 Notation

The mathematical notation used in this thesis is standard; in fact, it is perhaps overly

punctilious, reflecting the author’s fondness for consistency and unambiguousnesss.

Lowercase letters from the end of the Latin alphabet (u - z) are used to represent

variables, with italic script for scalars and bold invariably reserved for vectors. In the

context of a control system, x, u, and y are always the state, input, and output, respectively.

Following the literature on feedback linearization, the Greek letters ξ and η are also used to

represent the state, after a change of variables. However, Greek-letter vectors and scalars

are not distinguished by font, since the difference between bold and non-bold Greek letters

is imperceptible in this typeface. The letter t is of course always reserved for time.

Real-valued functions, whether scalar- or vector-valued, are usually taken, as conven-

tionally, from the lowercase Latin letters f through h, plus r and s. Vector-valued functions

and vector fields are bolded as well, the difference between the two being indicated by the

argument font; hence f(x) and f(x), respectively. When these letters have been exhausted,

functions resort to the back of the Greek alphabet, capitalization there indicating vector

outputs. Other kinds of maps (i.e. not exclusively involving real numbers) occur in the

material on Petri nets, and there letters were chosen mostly on alliterative or mnemonic

considerations.

Integers are represented by the lowercase Latin letters from i to q—excluding l and o

for their likeness to 1 and 0, and k for its use as a generic rate constant—with n usually

reserved for the dimension of the state and q for the strict relative degree of the system.

(Constant) matrices and vectors are represented with capital and lowercase letters,

respectively, from the beginning of the Latin alphabet. Vectors are again bolded. In the

context of linear time-invariant systems, the usual conventions are respected: A is the state

5

Page 24: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

matrix and B (b) is the input matrix (vector). Constant scalars are usually drawn from the

beginning of the Greek alphabet, although feedback gains are represented (per convention)

with K and the sampling interval is always symbolized T .

Calligraphic script is reserved for sets, which use capital Latin letters. Elements of

sets are then represented with the corresponding lowercase letter. Excepted are the well-

known number sets, which are rendered in blackboard bold: N,Z,R, and C for the naturals,

integers, reals, and complex numbers, respectively. The natural numbers are taken to

include 0; restrictions to the positive or negative subsets are indicated by a superscripted

+ or −. As usual, C0 is the set of continuous functions. The calligraphic N (·) is not used

for sets but to denote the nullspace of its argument.

Subscripts denote elements of a matrix or vector: di is the ith column of D; xj is the jth

element of x. Plain numerical superscripts on the other hand may indicate exponentiation,

a recursive operation, or simply a numbering, depending on context.

Differentiation is expressed as follows. Time derivatives use Newton’s notation: one,

two, or three dots over a variable for the corresponding number of derivatives, and a paren-

thetical superscripted numeral for higher derivatives. Leibniz’s notation is used for all other

standard derivatives, including the Jacobian (total derivative): ∂f/∂x—except the gradient,

which is represented with the nabla; hence ∇xφ is the row vector [∂φ/∂x1, ..., ∂φ/∂x1] of

partial derivatives.

The Lie derivative of a (scalar-valued) function h along trajectories of the vector field

f , that is (∇xh)f(x), gets its own symbol, Lfh(x). The Lie derivative of this object, say

along trajectories of g, i.e. Lg(Lfh(x)), dispenses with parentheses and is rendered sim-

ply LgLfh(x). Iterated Lie derivatives along the same vector field are written even more

concisely as Lifh(x), where i is the number of iterations.

Square brackets are put to various uses: In the usual cases they indicate vectors or

matrices; around the name of a chemical (usually a blood protein) they denote concentration

thereof; and in the context of [·, ·], they are the Lie bracket operation; that is,

[f ,g] :=∂g∂x

f − ∂f∂x

g.

6

Page 25: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Finally:∑

i is used for summations; a superscripted T indicates the matrix transpose;

∈ denotes set inclusions; I is reserved for the identity matrix; rk(·), Sp·, and spec· are

the matrix rank, span, and spectrum (eigenvalues), respectively, of their arguments. Except

in the case of the gradient or when otherwise explicitly stated, all vectors are assumed to

be columns.

7

Page 26: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Chapter 2

The Coagulation Cascade

2.1 Introduction

This chapter provides a detailed description of the biology of blood clotting, and can-

vasses previous attempts to model it mathematically. Much of our knowledge of the clotting

process comes in the form of chemical reactions that can be translated without remainder

into differential equations, but some reactions are known only qualitatively; and still other

information is essentially bivalent, like whether or not a reaction can take place in the ab-

sence of a certain ion. The heterogeneous nature of these descriptions is stressed in what

follows, since it motivates the hybrid-system model.

2.2 Preliminaries

Human blood clotting is a complicated biochemical and (arguably) rheological control

process whose normal function is to minimize the blood loss induced by vascular trauma.

Clotting is then one aspect of hemostasis—the other aspects being widening of the rele-

vant blood vessels (vasodilation) and clot dissolution (fibrinolysis)—that is, the dynamic

maintenance of balance between excessive bleeding (hæmophilia) and excessive clotting

(thrombophilia).

8

Page 27: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Healthy blood circulates with a set of proteins called clotting or coagulation factors,

most of which are inactive enzyme or cofactor precursors. The clotting process normally

begins when a blood vessel breakage precipitates the modification of some of these proteins,

transforming them from their unactivated (zymogen) to their activated (enzymatic) forms.

As concentrations of the latter increase, these enzymes trigger the activation of other clotting

factors—sometimes requiring first the activation of their cofactors—and so on, giving rise

to a cascade-like series of activations. (As we shall see in Chapters 3 and 4, representing

these cascades perspicuously is part of the motivation for using hybrid Petri nets, in which

this structure is immediately apparent; whereas it is very difficult indeed to read this kind

of description off a list of ODEs.) Ultimately, the glycoprotein fibrinogen is converted into a

fiber-like form (aptly named “fibrin”), which binds to the site of injury along with platelets

and another clotting factor (XIIIa), producing a clot and sealing the damaged vessel. Along

the way, other proteins serve to inhibit the activation of clotting factors and still others to

dilate the blood vessel (“vasodilators”).

The coagulation factors are listed in Table 2.1. Many of these factors have a Roman-

numerical designation, their active forms indicated by an appended letter “a.” (There is

no “factor VI” because this name was historically annexed to activated factor V.) These

are all, in some sense, pro-coagulant: a deficiency in any of these factors hinders clotting.

Prekalikrein and high molecular-weight kininogen are also pro-coagulant in this sense. On

the other side are the down-regulating proteins: antithrombin (sometimes “antithrombin

III”), proteins C and S, thrombomodulin (Tm), and tissue factor pathway inhibitor.

Factors IIa, VIIa, IXa, Xa, XIa, XIIa, as well as activated protein C (APC) and

kallikrein, are serine proteases, enzymes which function by cleaving peptide bonds in other

proteins and which have serine at their active sites. Factor XIII is also an enzyme, but

a transferase rather than a serine protease—in fact, the only enzyme of the cascade that

is not a serine protease.1 Factors Va and VIIIa, along with tissue factor (TF), protein S

(PS), thrombomodulin, and high molecular-weight kininogen (HMWK), are cofactors, re-

quired for certain enzymatic reactions (see below). Antithrombin (AT) is a serine protease1A transferase is an enzyme that promotes the transfer of a functional group from one molecule to another.

9

Page 28: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

inhibitor (serpin), whereas tissue factor pathway inhibitor is a different kind of inhibitor

(Kunitz-type). The importance of this distinction for our purposes is that serpins inhibi-

tion is irreversible. This leaves so-called factor IV (this name is rarely used), which is not a

protein at all but calcium ions; and factor I, fibrinogen, a glycoprotein whose cleavage into

fibrin marks the terminal event of the coagulation cascade.

2.3 The Cascade

Under normal circumstances, the coagulation cascade is simultaneously initiated by two

different mechanisms whose resulting “pathways” meet at the activation of factor X into

factor Xa (thrombokinase). More recently, it has become clear that all three pathways

interact a good deal during the clotting process. In fact, the pathway appellations are in

some respects merely holdovers from early, more limited models of blood clotting; they

are presently retained primarily to indicate initiation mechanism. The traditional nomen-

clature also prevails because it provides a way of bracketing parts of the cascade for ease

of understanding, and because the distinctions still provide a valid and instructive way to

conceptualize the clotting cascade (i.e. in spite of their interactions). A more or less tradi-

tional description of the cascade in terms of pathways, then, follows. Its graphical depiction

appears in Figure 2.1.

2.3.1 Intrinsic pathway

The intrinsic pathway is triggered when vascular cell damage exposes blood plasma to

the negatively charged surface of the subendothelial tissue2—for which reason the intrinsic

pathway is also known as the “contact system.” Hageman factor (factor XII) binds to this

surface, which induces a conformation change in the bound proteins, transforming them

into their active (protease) state.

Historically, this was believed to be the only method of activation for the intrinsic path-

way. It is now, however, also believed that vascular injury precipitates a rise in plasma2The endothelium is the layer of cells that line the interior walls of blood vessels.

10

Page 29: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Fact

or

(ab

br.

)T

rivia

lN

am

eA

ctiv

ate

dF

orm

Init

ial

Con

c.(n

M)

Sou

rce

Ifi

bri

nogen

fib

rin

(Ia)

6000-1

3000

(Halk

ier,

1991)

IIp

roth

rom

bin

thro

mb

in(I

Ia)

1400

(Bu

ngay

etal.,

2003)

III

tiss

ue

fact

or

—0.0

05

a(B

un

gay

etal.,

2003)

IVca

lciu

mio

ns

—1.2

x10

6(G

olt

zman

,2005)

Vp

roacc

eler

inacc

eler

in(V

a)

20

(Bu

ngay

etal.,

2003)

VII

pro

conver

tin

conver

tin

(VII

a)

10/0.1

b(B

un

gay

etal.,

2003)

VII

Ianti

mop

hilia

cfa

ctor

AV

IIIa

0.7

(Bu

ngay

etal.,

2003)

IXp

lasm

ath

rom

bop

last

inco

mp

on

ent

IXa

90

(Bu

ngay

etal.,

2003)

XS

tuart

-Pro

wer

fact

or

thro

mb

okin

ase

(Xa)

170

(Bu

ngay

etal.,

2003)

XI

pla

sma

thro

mb

op

last

inante

ced

ent

XIa

30

(Bu

ngay

etal.,

2003)

XII

Hagem

an

fact

or

XII

a500

(Halk

ier,

1991)

XII

Ifi

bri

nst

ab

iliz

ing

fact

or

pla

sma

tran

sglu

tam

inase

(XII

Ia)

70

(Halk

ier,

1991)

pre

kallik

rein

(PK

)F

letc

her

fact

or

kallik

rein

500

(Halk

ier,

1991)

pro

tein

C(P

C)

—A

PC

60

(Bu

ngay

etal.

,2003)

pro

tein

S(P

S)

——

300

(Bu

ngay

etal.,

2003)

anti

thro

mb

in(A

T)

——

3400

(Bu

ngay

etal.,

2003)

thro

mb

om

od

ulin

(Tm

)—

—1

(Bu

ngay

etal.,

2003)

tiss

ue

fact

or

path

way

aka

LA

CI,

EP

I—

2.5

(Bu

ngay

etal.,

2003)

inh

ibit

or

(TF

PI)

hig

hm

ole

cula

r-w

eight

Fit

zger

ald

fact

or

—1000

(Halk

ier,

1991)

kin

inogen

(HM

WK

)

aU

pon

blo

od

ves

sel

rup

ture

;ci

rcu

lati

ng

con

centr

ati

on

is0

nM

.

bU

nact

ivate

dan

dact

ivate

dco

nce

ntr

ati

on

sre

spec

tivel

y,of

fact

or

VII

.

Tab

le2.

1:P

rim

ary

Coa

gula

tion

Fact

ors

and

thei

rP

re-I

njur

yC

once

ntra

tion

s

11

Page 30: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Figure 2.1: The coagulation cascade.

12

Page 31: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

zinc concentration (Røjkjær and Schmaier, 1999), (Shariat-Madar et al., 2002), which in

turn enables the binding of HMWK and prekallikrein, and the activation of complexed

prekallikrein into its enzymatic form. Both these events are understood as threshold medi-

ated: zinc levels must exceed about 3µM for the binding reaction and 5µM for the activation

of kallikrein. The latter reaction is enhanced by the presence of factor XIIa. Even in the

absence of zinc, however, factor XIIa can activate the uncomplexed species of prekallikrein,

albeit much more slowly than the activation of prekallikrein in the PK:HMWK complex.

The serine protease kallikrein can also activate factor XII, and thus induces a positive

feedback loop. Since this reaction takes place in plasma rather than on the endothelial wall,

it is termed the “fluid-phase” activation—the negatively charged surface inducing “solid-

phase” activation. The reaction kinetics do indeed differ, since solid-phase reactions are

highly influenced, as fluid-phase reactions are not, by the geometry of surfaces.

A sufficient quantity of factor XIIa suffices to activate factor XI. Here the intrinsic

pathway starts to interact with the other pathways, since factor XI can also be activated

by feedback from the so-called common pathway (see below). Whatever its method of

activation, factor XIa (in the presence of calcium ions) cleaves factor IX, which in turn

activates factor X. This latter reaction, however, is extremely slow in the absence of factor

IXa’s cofactor, factor VIIIa. (In fact, following (Bungay et al., 2003), the present study

neglects this reaction altogether.) But factor VIII is activated by the common pathway—so

the intrinsic pathway alone is insufficient to generate blood clots.

2.3.2 Extrinsic pathway

The extrinsic pathway is initiated when the ruptured blood vessel releases tissue factor

(TF) into the plasma (for which reason it is now sometimes called the tissue-factor pathway),

which subsequently binds with unactivated factor VII (proconvertin). Now, the TF:VII

complex is inert until activation by factor Xa, and we have just finished saying that the latter

cannot be activated by the intrinsic pathway alone. The mystery is resolved by allowing

that some (small) fraction of circulating factor VII is in enzymatic (VIIa) form (Bungay

13

Page 32: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

et al., 2003). This allowance presents no embarrassment for the non-clotting situations,

since factor VIIa is inert in the absence of its cofactor, tissue factor. Again in the presence

of calcium, the TF:VIIa complex can indeed activate factor X, as well as factor IX (reaching

over into the intrinsic pathway); and Xa in turn reciprocally activates the TF:VII complex.

This concludes the work of the extrinsic pathway.

In light of the inability of the intrinsic pathway alone to induce clotting, what might

we say about the roles of the two initiation pathways? It is now widely believed that

the extrinsic pathway serves to “kick-start” the intrinsic pathway into action via feedback

from thrombin (IIa) and thrombokinase (Xa) activating factors VIII and XI. Thus the

extrinsic pathway directly achieves minimal thrombin production but does so on the order

of seconds, whereas the intrinsic pathway generates large quantities of thrombin but on

the order of minutes (Viera Stvrtinova, 1995). (Indeed—to anticipate—this qualitative

description is congruent with the simulations from the model of this thesis: Activation by

the intrinsic pathway alone, including an equation for the [slow] activation of factor X by

factor IXa, results in a significant [i.e. normal] amount of thrombin; whereas a deficiency

of the intrinsic pathway’s factor VIII, as in hæmophilia A, results in about 1% of normal

thrombin generation.)

2.3.3 Common pathway

The common pathway begins at the activation of factor X and terminates ultimately in

the production of a fibrin clot. As we have already seen, factor Xa feeds back to activate

the TF:VII complex in the extrinsic pathway and to activate factor VIII in the intrinsic

pathway, which subsequently binds to factor IXa. These two complexes from their respective

pathways feed forward to generate more factor Xa, though again both reactions require

calcium ions. Factor X also activates factor V in the common pathway, and proceeds to

bind to it.

This complex, Xa:Va, activates the most important protein of the cascade, thrombin

(factor IIa) as follows. First, the complex binds with prothrombin (factor II). Then, in

14

Page 33: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

the presence of calcium ions, the prothrombin part of the complex is transformed into the

intermediate (between zymogen and full enzyme) protein meizothrombin (mIIa). Finally,

the complex dissociates into Xa:Va and the protease thrombin (factor IIa). (For simplicity,

meizothrombin is omitted from Figure 2.1)

Thrombin initiates the final portion of the clotting cascade, but it also participates

in three feedback loops. It activates factors VIII and XI (in the intrinsic pathway), and

factor V in the common pathway. The Xa:Va:mIIa complex can also dissociate without

transforming meizothrombin to thrombin, in which case the free mIIa proteins can also

feed back to activate factors VIII and V, albeit at a slower rate than their activation via

thrombin.

Thrombin induces clotting by cleaving the soluble protein fibrinogen (factor I) into the

insoluble protein fibrin, which then spontaneously polymerizes into a mesh. Concurrently,

thrombin also converts factor XIII into its activated form, which stabilizes the fibrin mesh

by cross-linking with it. As platelets aggregate into the cross-linked mesh, the clot, bleeding

is stopped. Both the activation of factor XIII and fibrin are known to take place at certain

thresholds of thrombin.

2.3.4 Down-regulation

In addition to all the interactions lately discussed, thrombin also participates in its own

down-regulation. In complex with the cofactor thrombomodulin, thrombin activates the

inhibitor protein C; which, along with its cofactor, protein S, irreversibly inactivates factors

Va and VIIIa by cleaving the appropriate amino acids. The serpin antithrombin also serves

as an inhibitory “sink” by irreversibly binding up free factor XIa, meizothrombin, factor

Xa, and thrombin.

The final inhibitor we consider in our model is tissue factor pathway inhibitor, which

binds up free factor Xa, neutralizing it. This complex (TFPI:Xa) has further inhibitory

powers, however, soaking up the TF:VIIa complex. Both of these bindings are, however,

15

Page 34: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

reversible, so they should be thought of not so much as sinks but as negative feedback loops,

preventing the concentrations of Xa and TF:VIIa from increasing overmuch.

2.3.5 Completeness

This concludes the elements of blood clotting that are covered by the model of this

thesis. Now, I have called my model “complete,” and it is insofar as the (time) evolution of

the coagulation cascade is concerned; but we have so far ignored (1) spatial considerations,

and (2) the ancillary systems of vasodilation and fibrinolysis. We consider the latter first.

Vasodilation is less well-understood than blood clotting, but the outlines are clear

enough. High molecular-weight kininogen, under the influence of factor XIIa and kallikrein,

releases the vasodilator bradykinin. Likewise, a low molecular-weight kininogen is cleaved

by so-called tissue kallikrein to release another kinin, kallidin. The kinins induce the relax-

ation of the local smooth muscle, dilating the blood vessel in the area. Collectively, these

proteins and their interactions are known as the kinin-kallikrein system.

The fibrinolytic system is initiated by the injury-induced release of the serine proteases

tissue-plasminogen activator (tPA) and urokinase from the vascular wall. Both (the former

more efficaciously) cleave the fibrin-bound zymogen plasminogen into plasmin, yet another

serine protease, which degrades fibrin. Urokinase and tPA are in turn down-regulated

by the serpins plasminogen activator inhibitor 1 and 2; while plasmin is inactivated by

the serpin α2-antiplasmin. Thrombin also plays a role in down-regulation by activating

the appropriately named thrombin-activatable fibrinolysis inhibitor (TAFI). It functions by

rendering fibrin a less potent cofactor for the activation of plasminogen.

These two pathways were ignored because the author did not want to needlessly com-

plicate the model; and this choice is justified by the fact that neither one feeds back to

affect the coagulation cascade, either in the form of explicit feedback loops or (as far as

is known) in the form of substrate competition. Their exclusion, then, should in no way

affect our simulations of clot formation or thrombin generation under various physiological

conditions; nor the control thereof. Nevertheless, as we shall see in the Chapter 4, the model

16

Page 35: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

was constructed in such a way as to allow the easy incorporation of addition elements; and

furthermore, the addition of these pathways could be constructed from the same primitives

as the rest of the model: a series of possibly threshold-controlled bindings and activations.

The other major feature we have ignored until now is the effect of spatial arrangements

on the reactions. Since certain reactions take place on the endothelial wall, others on the

surface of platelets, and still others extracellularly, it has been claimed that a faithful model

of blood clotting must separate these reactions from each other, modeling their interaction

through spatial-diffusion equations (Panteleev et al., 2006). Likewise, von Willebrand fac-

tor, a blood glycoprotein with which (unactivated) factor VIII circulates in complex, and

which protects the latter from degradation, binds to platelets most efficiently under high

shear stress, which is obviously a function of spatial considerations (e.g. diameter of the

vasculature at the site of injury). Or again, clot dissolution is a matter of rheology (i.e., in-

volving flows, particularly of substances with varying viscosities) as well as of biochemistry.

These “spatial effects,” as I have called them, have two unfortunate consequences for

models. The first is that accounting for diffusion requires transforming the ordinary differen-

tial equations governing the time-evolution of protein concentration into partial differential

equations governing spatial evolution, as well (by introducing partial derivatives with re-

spect to each of the three spatial dimensions). The second unfortunate consequence is that

interactions on surfaces rather than in fluids vitiates the law of mass action (see Section 2.5

below)—although the models of coagulation discussed below which do model spatial effects

seem not to have appreciated this fact—meaning that (in short) reactions may not be sim-

ply proportional to the (volumetric) concentrations of reactants. We justify the exclusion

of these effects from the present model, then, on the grounds that (1) they have generally

been neglected in the other models in the literature (again see Section 2.5 below); (2) many

of the data we use for validation come in fact from an in vitro environment, where spatial

effects do not play any significant role; (3) while not completely accurate, the fluid-based

reactions are nevertheless an approximation of the actual in vivo process; and finally (4)

although PDEs are simulatable, restricting the model to ODEs provides great traction in

analysis and control, as we shall see in Chapters 5 and 6.

17

Page 36: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Finally, and obviously, in light of the preceding, we do not mean by “complete” “free

from any error.” In particular, the rate constants (and other parameters used in the model)

are empirical, with both the negative and positive connotations of that word: They are

indeed drawn from the literature, each one the result of the detailed measurements of

some biochemical experiment or other, but there is great variance among them. Some rate

constants vary by as much as orders of magnitude from one study to the next (Panteleev

et al., 2006) (or cf. the rate constants of (Bungay et al., 2003) and (Luan et al., 2007),

and their references). Thus we stress that the model, while it matches such experimental

data as exist for the whole cascade, is underdetermined by those data. The point of this

dissertation is to show how what information we do know, about any part of the cascade,

can be incorporated into a single model; how the simulations of this model can then be used

with clinical experiments by iterating between the two to refine, seriatim, our knowledge

of coagulation and the model; and furthermore to use this information to design clinical

interventions by treating the cascade and a drug as together composing a control system.

2.4 Diseases of Coagulation

Many things can go wrong with the coagulation system and there are correspondingly

many diseases. Two in particular play a large role in this dissertation, as the subject of our

simulations and control schemes; we discuss these first. In what follows we draw primarily

on (Crookston) and (King).

The best-known blood-clotting disorder is almost certainly hæmophilia, acquaintance

with which dates from the Biblical era (Rosner, 1995), although its modern diagnosis waited

until the nineteenth century. Its most common incarnation is factor-VIII deficiency, i.e.

hæmophilia A, which affects one in about every five to ten thousand individuals, almost

all of them male since it is an X-linked trait. Treatment can be either prophylactic or in

response to an injury, but in either case involves the injection of some replacement form of

factor VIII (either isolated from serum or “recombinant,” i.e. bioengineered). Symptoms

obviously include prolonged bleeding, but also joint and muscle hemorrhage. The severity

18

Page 37: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

of the diseases varies with the extent of the deficiency of factor VIII: 5% of normal levels

has mild consequences; 1-5%, moderate; less than 1%, severe.

A less well known but in fact much more common bleeding disorder is the recently

discovered factor-V Leiden, the heterozygous form of which may affect as much as five

percent of the white population. Factor-V Leiden is a genetic disorder in which the amino

acid arginine505 of the heavy chain of factor V is replaced by glutamine505, which effectively

disables APC-mediated cleavage of factor V at this point. (There are in fact other forms of

APC resistance, but factor-V Leiden accounts for the overwhelming majority [95%] of cases.)

The consequence is hypercoagulatory behavior, increasing the odds of venous thrombosis

(clotting within a vein) by as much as a factor of ten for heterozygotes and of 80 for

homozygotes. Factor-V Leiden was first identified in 1994, and perhaps as a result is not

treated prophylactically—though of course its thromboemboli are treated, according to the

usual protocol, viz. administration of an anticoagulant like warfarin or heparin. Estrogen-

based oral contraceptives are also known to increase hypercoagulatory risk.

Hæmophilia B is less common than its aforementioned cousin, affecting one male in

about thirty thousand, and resulting from a deficiency of factor IX; its severity varies just

as hæmophilia A, mutatis mutandis for factors. Hæmophilia C, i.e. factor XI deficiency,

is even less common in the general population (one in one-hundred thousand), although

it is fairly prevalent among the Ashkenazim. The most common hemorrhagic disorder is

not any of the hæmophilias, however, but von Willebrand’s disease, a deficiency (either in

quantity or quality) of von Willebrand factor, affecting upwards of 2% of the population.

Von Willenbrand’s disease discourages clotting by hindering platelet adhesion; but also by

failing to protect the zymogen factor VIII from degradation in pre-clotting conditions, which

symptom sometimes disguises the disease as hæmophilia A.

There are of course many other coagulation diseases, for which the interested reader is

encouraged to consult (King) and especially (Crookston); but they are not relevant to this

thesis.

19

Page 38: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

2.5 Mathematical Models

Many mathematical models of blood clotting have been constructed, most in lesser detail

than the present study but a few covering aspects of the process that I have neglected. We

discuss the relevant exemplars.

Most models, like ours, are concerned primarily with the coagulation cascade (i.e. ig-

nore spatial and rheological effects). These approaches generally model the time evolution

of protein concentrations with a series of nonlinear, ordinary differential equations, deriving

from the law of mass action, the Michaelis-Menten equations, or other chemical considera-

tions. Due to the extremely complicated nature of blood clotting, these attempts usually

focus on small subsets of the entire process. So, for example, Butenas et al. (2004) model

only the interactions of coagulation factors II, IX, and X; Panteleev et al. (2002) and Adams

et al. (2002) model the extrinsic pathway; Qiao et al. (2004) model a portion of the common

pathway and a small part of the intrinsic pathway; and Xu et al. (2002) and Leipold et al.

(1995) model some of the interactions of factors II, V, VII, VIII, IX, and X.

The most complete and ambitious models of the cascade are (Bungay et al., 2003),

and (Luan et al., 2007), which use systems of 73 and 98 (respectively) coupled, nonlinear,

differential equations to describe all of the extrinsic pathway, a large portion of the (ulterior)

intrinsic pathway, and the common pathway up through thrombin production. Luan et al.

(2007) include more variables as a consequence of including the effects of platelets, but

otherwise the models cover essentially the same portions of the coagulation cascade.

Note the absence of the early intrinsic pathway—HMWK, PK, factor XII—from any of

these models, which is at least in part a consequence of its being poorly understood. Now,

Kogan et al. (2001) do build a model of the intrinsic and common pathways in the service of

matching the results of a clinical test, the so-called activated partial-thromboplastin time.

However, they cannot in fact match clinical aPTT times, nor are their choices of intrinsic-

pathway chemical reactions uncontentious: e.g., they do not include HMWK, nor is there

mention of zinc concentrations (see above, Section 2.3.1). Thus in the model of this thesis,

20

Page 39: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

as we shall see in Chapter 4, the early intrinsic pathway was modeled qualitatively as a

hybrid system.

An important distinction among these ODE models is whether they rely strictly on

the law of mass action or whether, on the contrary, they make use of the Michaelis-Menten

approximation. The former states that the rate at which (elementary) reactions take place is

proportional to the product of the chemical activities of the reactants. The proportionality

constant is empirically determined and called (appropriately) the “rate constant” for that

reaction. Now, the law of mass action is true of every (free-diffusion) reaction3 and at the

protein-concentration levels of blood clotting, we can substitute the word “concentrations”

for “chemical activities” in the formulation just given. That is not to say, however, that the

rate of product formation in, e.g., an enzymatic reaction is proportional to the product of the

concentrations of the enzyme and substrate, because such reactions involve an intermediate

reaction in which an enzyme-substrate complex forms. This is written:

E + Skon−−−−koff

E : S kcat−−→ E + P (2.1)

where E an enzyme (say, factor Xa), S is a substrate (say, factor V), E:S is the aforesaid

complex, P is the product (in this case, factor Va), and koff, kon, and kcat are the rate

constants. Nevertheless, the law of mass action holds for each elementary reaction, so here

generation of the product proceeds at a rate proportional to the concentration of the E:S

complex, rather than those of the initial reactants.

Now, on the other hand, in many enzymatic reactions, the substrate concentration

greatly exceeds the enzyme concentration (high enzyme efficacy making only a compara-

tively small amount necessary for the reaction), in which case the change in concentration of

the intermediate enzyme-substrate complex is approximately zero, and the rate of product

formation (v) can be written more simply as:

v = kcat[E]0[S]

Km + [S], (2.2)

with Km = koff+kcat

kon. These so-called Michaelis-Menten kinetics can simplify our equations,

and offer another advantage as well: chemists often do not bother to determine the on- and3In fact, the law of mass action can be derived from statistical mechanics

21

Page 40: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

off-rates separately, measuring instead the more easily acquired Km. (An instance of this

will show up in Chapter 5.) However, as we have seen, this approximation is only valid

when the substrate is in great excess over the enzyme, which we cannot guarantee over

the course of the coagulation process. Furthermore, adoption of MM kinetics violates the

conservation of mass implicit in mass-action kinetics, which conservation makes for some

nice mathematical properties about which we shall have something to say in Chapter 6.

It was mentioned above (Section 2.3.5) that the model of this dissertation ignores the

spatial effects on reactions. Recent models by two groups have explicitly included these

effects: Panteleev et al. (2006) include spatial diffusion in their model of the extrinsic

pathway and parts of the common pathway. The authors do not restrict themselves to

mass-action kinetics, using instead MM kinetics and another, similar approximation where

appropriate. The diffusion terms of course result in partial, rather than ordinary, differential

equations, as we have lately appreciated. A similar approach, modeling a more or less

coextensive portion of the cascade, appears in the thesis (Mohan, 2005), which also includes

a viscoelastic-fluid model of flow across the clot. The goal of neither of these models is to

capture the entire coagulation process, but rather to render existing, incomplete models

more realistic.

There are also computational (one hesitates to say “mathematical”) models of blood

clotting which abjure equations altogether. Mounts and Leibman (1997) use a variant of the

Petri-net formalism (as I do: see Chapter 3) to describe the coagulation cascade. Since (or-

dinary) Petri nets have a discrete state space, the tokens (again see the description of Petri

nets in Chapter 3) of the model are used to represent some number of molecules of a given

substance. The authors describe their model as “qualitative,” but in fact their discretiza-

tion is fine enough to give rather precise results. Finally, in a similar vein, Signorini and

Gruessay (2004) take an object-oriented approach, specifying the structure and dynamics of

the cascade by treating each factor as a message-passing object. The order of the message

passing is specified by an “activity diagram,” which in fact is more or less conceptually

identical to a Petri net, with mechanisms for modeling concurrency and synchronization.

22

Page 41: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Chapter 3

Hybrid Systems

3.1 Introduction

Classical control theory and system modeling have focused on systems with purely

continuous dynamics and those with purely discrete dynamics. However, many real-world

systems necessarily involve both continuous and discrete components, or are best modeled

as interacting continuous and discrete subsystems. These systems are called hybrid sys-

tems, and have motivated a great deal of research in the last two decades. What follows is

a somewhat pedantic introduction for readers completely unfamiliar with hybrid systems;

initiates should skip to Section 3.3 on hybrid Petri nets. Readers looking for more infor-

mation are encouraged to download the excellent but as-yet unpublished book on hybrid

systems, (Lygeros et al., 2001).

A system may be considered hybrid either because the state space consists of both

continuous and discrete components, or because the dynamics manifest both continuous-

time and discrete-time behaviors. Consider, for example, the by now well known example of

a thermostat. (The following treatment of the classic thermostat model was adapted with

little change from an example in (Lygeros, 2004).) Suppose the change in temperature is

governed by one differential equation when the heat is on,

x = α(85− x); (3.1)

23

Page 42: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Figure 3.1: A hybrid system model of a thermostat; see text for details.

but when the state (temperature) crosses a certain threshold, say 75F, the heat is turned

off and a new differential equation applies:

x = −αx. (3.2)

When the temperature falls below 70F, the heat goes back on and Eq. 3.1 again

applies. Thus the temperature increases exponentially toward 85F until it hits a boundary

in the state space, at which point the heat is turned off and the temperature declines, again

exponentially, toward zero degrees. The various regions where different sets of continuous-

time dynamics obtain may be modeled as different discrete states of the system; hence, the

entire network may be thought of as the first type of hybrid system, where the state is

jointly defined by a continuous variable (the temperature) and a discrete state (indicating

whether or not the heat is on).

The model is illustrated in Figure 3.1. The system is very much like a finite state

machine, but additionally a set of continuous dynamics (a differential equation) is associated

with each of the discrete states. The arrow labeled x = 72 indicates that the initial state is

(x = 72, q = off), where the pair (x, q) defines the state, with x ∈ X , the continuous state

space, and q ∈ Q, the discrete state space. The arrows connecting discrete states are, as

24

Page 43: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

in finite state machines, called “edges,” and are written (q, q′), where the edge starts at q

and ends at q′. The edges are labeled with the so-called guard conditions (x = 70, x = 75):

given the state (x, q), if x belongs to the set specified by the guard associated with the edge

(q, q′), then the system may transition to the discrete state q′. The continuous variable may

in general also be reset at a discrete transition, but all of the reset maps of the thermostat

model are the identity map.

Finally, a domain D(q) (x ≥ 70, x ≤ 75) is associated with each discrete state. When

the continuous state reaches the boundary of the domain, a discrete transition is forced—

unless no guard condition is satisfied, in which case the system is “blocked” and stops.

Alternatively, the guard may be enabled before the trajectory reaches the boundary of

the domain, in which case the discrete transition may but need not occur, and the model

becomes non-deterministic. This point is made for the sake of generality; in the thermostat

example, the guard conditions and domain have been written so as to preclude both blocking

and non-determinism. The same is true of the blood clotting model presented in this paper.

The thermostat may be modeled as a hybrid system because, again, it has discrete

states as well as continuous states and evolves in continuous time. However it should be

noted that hybrid systems may also arise in the interaction of discrete-time systems with

continuous dynamics. So, for example, we may wish to model the interaction of a purely

continuous system like a chemical batch process with a digital controller which has an

essentially continuous state space but evolves in discrete time. The blood clotting model of

this paper is of the first type.

3.2 Types of Hybrid-System Investigation

Hybrid systems arise in numerous other contexts, among them chemical batch processes

(Engell et al., 2000); road-traffic controllers (Czogalla et al., 2002), (Horowitz and Varaiya,

2000); air traffic control (Oishi et al., 2002); robotic control (Fierro et al., 2001),(Schlegl

et al., 2002); automotive applications (Pettersson and Lennarton, 2003), (Antsaklis and

Koutsoukos, 2003); embedded systems (Neuendorffer, 2004); and biological systems (Ghosh

25

Page 44: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

and Tomlin, 2001),(Chen and Hofestadt, 2003),(Matsuno et al., 2000),(Matsuno et al.,

2003), (Jong et al., 2003). In all of these examples, there is a variety of questions we may

be interested in asking, and which recent research has attempted to find ways of answering.

We provide a brief overview here in order to give the reader a flavor of the field.

3.2.1 Modeling and simulation

Perhaps the most obvious of these is modeling and simulation. For hybrid systems in

particular, formal analysis is restricted to the simplest of systems or to highly circumscribed

cases, so a powerful alternative is to construct a model of the system and simulate its

behavior rather than perform an exhaustive analysis. However, whereas a great variety of

modeling and simulation tools exists for purely continuous or purely discrete dynamics, only

recently have such tools been developed specifically for hybrid systems. The extent of the

novel modeling and simulation issues associated with hybrid systems are beyond the scope

of this paper, but three should be mentioned because of their relevance and their ubiquity.

The first is choice of representation. Whereas purely discrete and purely continuous

dynamics have fairly well-established representations from the computer science and control

theory disciplines (respectively), a standard framework for modeling hybrid systems has

not yet arisen. Hybrid systems also come in a host of flavors, and ideally a simulation

or modeling program should be able to accommodate all of these. Lygeros (2004) has

emphasized that a hybrid-system modeling language should be descriptive, in the sense of

being able to model a wide range of continuous and discrete dynamics and their interactions,

and to accommodate stochasticity in a variety of contexts; composable from smaller units

into larger networks; and abstractable, in the sense of being able to cash out composite

model specifications in terms of component specifications as well as determine composite-

level behavior via knowledge of component behavior.

The second simulation concern is the development of accurate and efficient numerical

integration techniques. In a familiar problem from the hybrid-systems literature, imprecise

numerical integration triggers a discrete event—and hence, perhaps, a new set of differential

26

Page 45: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

(state) equations—in the simulation, where no such event occurs in the actual system. Pu-

tative solutions which simply shrink the step size, however, can greatly increase simulation

time, and moreover are not per se a guarantee of eliminating this type of simulation error.

This issue will arise again in the context of the present study (see Section 8.2).

The third and final simulation issue to be discussed here is the problem of so-called

“Zeno” systems, in which trajectories of the system take an infinite number of discrete

transitions in finite time. In such cases, simulation time “stops,” and the simulation termi-

nates only when the system hangs. The well-known bouncing-ball hybrid system exhibits

this type of behavior: the “moving up” and “moving down” state alternate increasingly

faster, according to a geometric series that converges in the limit but will never converge

in simulation. Even in non-Zeno systems, existence and uniqueness of solutions is not in

general guaranteed for hybrid systems, and special care must be taken in their simulation.

3.2.2 Verification and decidability

A second type of question we may be interested in asking about a particular hybrid

system is known as verification. Verification is the formal proving of certain (interesting)

system properties, given the system and a range of inputs. There are two variations: algo-

rithmic verification (“model checking”) and deductive verification, where the former uses

search techniques and the latter involves the construction of a formal proof (Kowalewski,

2002). In both cases, the property of interest is usually reachability. In general, however,

reachability analyses on hybrid systems are prohibitively difficult. Results have been con-

fined to small classes of highly circumscribed systems (e.g., timed automata and rectangular

automata (Henzinger et al., 1998)). This problem, too, will return to haunt our model. Ver-

ification is also sometimes performed vis-a-vis the stability of the system; we may want to

ask, say, if the closed-loop system is asymptotically stable (Lygeros, 2004).

It has also proven useful for analysis techniques, particularly reachability, to perform

abstractions on systems; i.e., to replace a complicated hybrid system model with a less

complicated one in which properties previously difficult or impossible to prove are rendered

27

Page 46: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

tractable. The general procedure is to construct a simplified model which contains the

behavior of the original system as well as some new behavior, an artifact of the abstraction.

This system is then tested for some appropriately abstracted version of the original property

to be tested (e.g. reachability); for example, if a state can be shown to be unreachable in

the abstracted system, then it has been shown to be unreachable as well in the original

system (Kowalewski, 2002),(Antsaklis and Koutsoukos, 2003).

A related question is that of decidability: whether a problem can be answered, affirma-

tively or negatively, by some algorithm in finite time. Specifically, in the present case, the

question is whether or not a system can be verified in finite time; and more specifically,

whether or not the reachability of some state(s) or region in the state space can be com-

puted in finite time. It should be noted that reachability analyses can be fruitfully pursued

even for undecidable systems, since the particular (say) state about which reachability is to

be determined may be decidable, even though reachability in general is not (Kowalewski,

2002),(Henzinger et al., 1998).

3.2.3 Controller synthesis

A third type of question that can be asked about hybrid systems concerns controller

synthesis. Controller specifications in hybrid systems are often given in terms of temporal

logic. For instance, we may require that all trajectories of a system remain within a set of

states F ⊆ X×Q, where X and Q are the continuous and discrete state spaces, respectively.

The condition is written as

((q, x) ∈ F), (3.3)

where q ∈ Q and x ∈ X are the discrete and continuous state, respectively. Or we may

insist that the trajectory eventually reach some set of states F , written

♦((q, x) ∈ F). (3.4)

Designing a controller then amounts to picking a set of inputs from the input space for each

state of the system such that the specification of interest is satisfied. There are various

techniques for performing this task. For example, both the theory of optimal control and

28

Page 47: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

game-theoretic approaches can be used to derive the Hamilton-Jacobi partial differential

equations whose solutions are the boundaries of the reachable sets. These can then be

solved approximately, providing (real-time) feedback-control laws which provably satisfy

the specifications (Lygeros et al., 2001).

3.3 Hybrid Petri Nets

There are numerous hybrid system modeling languages (see, for example, (Alur

et al.),(Chutinan et al.),(Dang and Maler),(Henzinger et al.),(Miyano et al.), (Lee

et al.),(Mat),(Daws et al.),(de Moura et al.),(Deshpande et al.), (Drath, 2002), and (Larsen

et al.)) and in choosing among them the following constraints were considered. First and

foremost, of the different types of hybrid-system investigations canvassed above, our present

focus is primarily simulation. (Why are we unconcerned with the control problem, the im-

portance of which, in connection with disease treatment, we saw in Chapter 1? We discuss

this in Chapters 5 and 6, but to anticipate: the high dimension of the state space precludes

the standard hybrid-systems control techniques, so an alternative based on decomposing

the system is pursued instead.)

A second major design focus was ensuring that the model be easily and intuitively

modifiable, not just in parameter settings like concentrations of proteins, but in structure

as well, so that new (say) proteins and reactions can be incorporated painlessly into it. An

ideal model will also provide a perspicuous representation of the system of interest. This

is especially significant in the present case since is intended in part for use in a clinical

setting by biologists who may have little or no familiarity with programming languages.

This constraint obviously militates in favor of a graphical modeling framework.

A third and final consideration is that the clotting cascade consists in large part of

numerous similar reactions among different clotting factors. The reactions often involve

reactants which play no roles in any other reactions or events. These two facts suggest an

object-oriented approach to modeling, since this affords methods of abstraction, reuse, in-

29

Page 48: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Figure 3.2: A very simple Discrete Petri Net (DPN) with purely discrete components. The

input and test arcs’ respective resource requirements were satisfied, but the inhibitory arc’s

requirement is not—so the transition is enabled.

formation hiding, and inheritance. (For another object-oriented approach to blood clotting,

see (Signorini and Gruessay, 2004).)

A modeling language which meets all of these constraints is the hybrid Petri net (HPN).

Petri nets tout court are a graphical modeling formalism for distributed discrete-event sys-

tems which generalize automata theory to allow notions of concurrency and resource con-

sumption; their hybrid cousins are a further generalization to include notions of continuous

state and continuous time. We begin with an informal description of Petri nets, followed

by a formal exposition of the network semantics, and then move on to HPNs.

An ordinary Petri net comprises three kinds of components: transitions, places, and

directed arcs (see Figure 3.2). Places (drawn as circles) represent discrete quantities in

virtue of the number of “tokens” (n) they carry (n ∈ N), and events are modeled as firings

of transitions (drawn as black rectangles). In the figure, tokens are drawn as little black

circles which live in places. The distribution of tokens over all the places in the net is called

the marking of the net. The marking changes when transitions fire and tokens are consumed

from the input places and produced at the output places.

A transition is enabled if and only if each place connected via an input arc contains as

30

Page 49: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Figure 3.3: The simple Discrete Petri Net (DPN) of Figure 3.2, advanced by one step: the

input arc’s place has been depleted by its resource requirement (1 token), and the output

arc’s place has been augmented by its weight (1 token).

many tokens as the “resource requirement” of its corresponding arc. If an arc is unlabeled,

its resource requirement is assumed to be unity. When a transition is enabled, it “fires,”

meaning that tokens are consumed from input places according to the resource requirement,

and produced in the output place(s) according to the weight(s) associated with the outgoing

arc(s); since the outgoing arc is unlabeled in the figure, the output is assumed to be a single

token.

Petri Nets are often outfitted additionally with “test” input arcs, which function ex-

actly like normal arcs except that tokens are not consumed when a transition fires; and

“inhibitory” input arcs, which enable transitions just in the case that the input place has

fewer tokens than the resource requirements (again with no corresponding token consump-

tion). These also appear in Figure 3.2. The test arc is drawn as a dashed arrow; since the

preceding place contains two tokens, it satisfies the resource requirement given by the test

arc’s weight. Similarly, the input place to the inhibitory arc, drawn as an arrow ending in

an open circle, contains fewer than three tokens, so it does not inhibit the transition. The

firing is determined by the conjunction of these conditions; the Petri net in Figure 3.2 will

thus transition to the one shown in Figure 3.3.

31

Page 50: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Definition 3.3.1. Discrete Petri Net (DPN). A DPN is a tuple (P,S,A,W,M0),

where:

• P is a set of discrete places.

• S is a set of discrete transitions.

• A is a set of weighted directed arcs which connect places to transitions and transitions

to places; i.e. A ⊆ (P × S) ∪ (S × P).

• W : A 7→ N+ maps each arc a ∈ A to a weight w from the positive integers.

• M0: P 7→ N is the initial marking of the network, which gives the original token

distribution in the network via a map from each place p ∈ P in the net to the natural

numbers.

The meaning of an arc differs according to whether it connects places to transitions or

vice versa, and on which of three flavors a member of the former set comes in: test arcs E ,

inhibitory arcs I, or resource arcs R.

Definition 3.3.2. DPN Arcs. A = ∗A ∪ A∗, where

• ∗A ⊆ (P × S) are input arcs, and

• A∗ ⊆ (S × P ) are output arcs.

Furthermore, ∗A = E ∪ I ∪ R

DPNs have a well specified real-time execution semantics where the next state function

is specified by the firing rule. In order to simulate the dynamic behavior of a system, a

marking of the DPN is changed according to the following firing rule:

Definition 3.3.3. DPN Execution Semantics.

• A transition s ∈ S is said to be enabled if:

32

Page 51: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

1. the source place p of each inhibitory arc i ∈ I of s has (strictly) fewer tokens

than wps, and

2. the source place of each test and resource arc (e ∈ E and r ∈ R, respectively)

contains at least wps tokens;

where in each case wps is the weight of the input arc from each source place p to s.

• The firing of an enabled transition, s, removes wps tokens from the source, p, of each

resource arc, and places wsp tokens in each output place p.

Notice that the semantics associated with the enable condition make reference only to

the input arcs ∗A, whereas the firing semantics invoke both input and output arcs. It should

also be pointed out that the definition implies that transitions which have no input arcs are

always enabled.

Transition firings take place as soon as their resource requirements are satisfied; i.e.,

time does not pass, even through successive transitions; or we might say, more accurately,

that so far the DPN has no time concept. Of course, we may want to model not simply

the sequencing of events but the time it takes for the events to transpire. In this case we

can assign delay times to the transitions: a transition with a delay of τ seconds will fire

exactly τ seconds after its resource requirements have been met. If in the meantime the

resources are depleted and the requirement is no longer fulfilled, then the transition will not

fire. Time flows whenever none of the transitions is firing, which means that each transition

is either not enabled or is enabled but experiencing a delay. More formally, we need to

augment the execution semantics as follows:

Definition 3.3.4. Time. A DPN may be augmented with a time concept, where time

t ∈ R. Time stops running (i.e. increasing from t0 = 0) whenever a (discrete) transition

fires.

Definition 3.3.5. Transition delay. Associate to each transition sj ∈ S a delay τj ∈ R.

The firing of an enabled transition takes place at time t∗ + τ , where the transition was

enabled at time t∗ and remained enabled throughout the interval [t∗, t∗ + τ ].

33

Page 52: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Assigning a delay of zero seconds to a transition restores the original execution seman-

tics, i.e. makes firing take place as soon as the transition is enabled.

In the present study, the Petri net language was additionally required to represent

continuous-time events and a continuous state space. Hybrid Petri nets meet these re-

quirements by providing, respectively, continuous transitions and continuous places with

real-valued “token fluid.”

The definition deserves some preparatory remarks: (1) In contrast to a DPN, an HPN

marks its continuous places with real numbers, in addition to assigning integers to the

discrete places. (2) No weights are assigned to arcs which link continuous places to con-

tinuous transitions, or which link continuous transitions to continuous places (since input

and output are governed by the differential equation associated with the transition). Fur-

thermore, the weight maps assign real or natural numbers where appropriate. (3) Neither

resource arcs nor output arcs can join discrete places with continuous transitions (since the

places’ requirement of discrete state is incompatible with the transitions’ requirement of

continuous state change). In contrast, continuous places and transitions can only be joined

by resource or output arcs. Figure 3.5 summarizes these constraints. (4) The meanings

of the arc-starring convention and of E , I, and R are the same as above. Notice that the

continuous input arcs of ∗Ac are not members of any of these sets.

Formally:

Definition 3.3.6. Hybrid Petri Net (HPN). A hybrid Petri net is a tuple

(P,S,A,Wc,Wd,Mc0,Md0, T, F ), where:

• P = Pc ∪ Pd is a set of continuous places and discrete places.

• S = Sc ∪ Sd is a set of continuous transitions and discrete transitions.

• A ⊆ (P × S) ∪ (S × P ) is a set of weighted directed arcs. Furthermore,

– Ad = a ∈ A|a ∈ (Pd × Sd) ∪ (Sd × Pd) ⊆ (E ∪ I ∪ R ∪Ad∗);

– Adc = a ∈ A|a ∈ (Pd × Sc) ∪ (Sc × Pd) ⊆ (E ∪ I);

34

Page 53: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Figure 3.4: A very simple Petri net with purely continuous components (CPN). The rate

at which “token fluid” leaves m1 is the rate at which it accumulates in m2, which in this

example is m2(m1 + 1)

– Acd = a ∈ A|a ∈ (Pc × Sd) ∪ (Sd × Pc) ⊆ (E ∪ I ∪ R ∪Acd∗);

– Ac = a ∈ A|a ∈ (Pc × Sc) ∪ (Sc × Pc); and

– A = Ad ∪ Adc ∪ Acd ∪ Ac.

• Wc : Acd 7→ R+ is the continuous weight map.

• Wd : (Ad ∪ Adc) 7→ N+ is the discrete weight map.

• Mc0: Pc 7→ R is the initial token-fluid marking of the network.

• Md0: Pd 7→ N is the initial token marking of the network.

• T : Sd 7→ (R+∪0) maps each discrete transition s ∈ Sd to a non-negative real-valued

delay τ .

• F : Sc 7→ C0 maps each continuous transition s ∈ Sc to a function f(m1, ...,mn) of

the network markings, from the space of continuous functions.

The fundamental addition to the execution semantics is continuous token-fluid flow. As

shown in Figure 3.4, a continuous transition fires continuously according to a “firing speed”

(i.e. a differential equation) which defines the speed of consumption and production of fluid

from the various input and output places, respectively, associated with it. Token fluid leaves

the input places at this rate and enters the output places at the same rate.

Definition 3.3.7. HPN Execution Semantics.

• The enabling of an HPN is identical to that of a DPN, given in Def. 3.3.3.

35

Page 54: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

• The firing for all arcs asp ∈ (Ad ∪ Adc ∪ Acd) is identical to that of a DPN, given in

Def. 3.3.3.

• The firing for all continuous input arcs ∗aps ∈ ∗Ac from place p to transition s is

given by the equation

dmp(t)dt

= −fs(m1(t), ...,mn(t)), (3.5)

where mp(t) ∈ Mc is the marking associated with the input place p, and fs is the

function associated with the transition s.

• The firing for all continuous output arcs a∗sp ∈ Ac∗ is given by the equation

dmp(t)dt

= fs(m1(t), ...,mn(t)), (3.6)

where mp(t) ∈ Mc is the marking associated with the output place p, and fs is the

function associated with the transition s.

The reader should take care to note that the enable semantics from Def. 3.3.3 apply

only to arcs ∗a ∈ E ∪ I ∪ R = (∗Ad ∪ ∗Adc ∪ ∗Acd), and that as a consequence, HPN

transitions which are fed only by continuous input arcs ∗a ∈ ∗Ac or by no arcs at all are

always enabled. Additionally, Def. 3.3.3 refers only to “tokens,” but in Def. 3.3.7 it is

assumed that this be taken as either tokens or token fluid, as the case may be.

3.4 Hybrid Systems in Biology

We conclude with a brief survey of some (related) recent work in applying hybrid models

to biological systems.

In (Ghosh and Tomlin, 2001), a model of cell differentiation is constructed. The authors

represent the spatiotemporal evolution of protein concentrations within a cell by ordinary

differential equations, derived from a simplified form of the chemical kinetics. These dynam-

ics are turned on and off by discrete switches, which are triggered by protein concentrations

reaching set thresholds. In both these respects, this model is like the blood-clotting model

36

Page 55: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Figure 3.5: A summary of the various types of arcs in an HPN; cf. Defs. 3.3.6 and 3.3.7.

37

Page 56: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

of this thesis. On the other hand, the model assumes that cells interact only with their

six neighbors in their planar hexagonal array. The authors produce various simulations for

reasonably sized grids, for comparison with biological data; and prove the existence and

attraction of certain equilibria; but the latter results are restricted to a two-cell network.

This is a consequence of the restriction, alluded to above, that high state dimensionality

imposes on reachability results. In fact (and interestingly), the authors claim that this

restriction can be overcome with a model checker (see Section 3.2.2 above), since the con-

tinuous dynamics of their model admit analytic solution. This is certainly not the case in

the coagulation model; i.e. no analytic solutions to its ODEs are possible.

Chen and Hofestadt (2003) present a methodology for the modeling of intracellular

metabolic networks and their regulation by genes, including a case study on the urea cycle.

The paper advocates a hybrid model for these metabolic networks in virtue of its ability

to incorporate qualitative and quantitative data into a single framework—a consideration

which (among others), as discussed in the introduction to this thesis, motived this author’s

adoption of a hybrid model. Here, once again, the continuous dynamics are generally a

consequence of chemical kinetics, with continuous state variables representing species (not,

however, necessarily protein) concentrations. The authors also consider spatial dynamics,

however, including diffusion transportation, which is neglected in the coagulation model

(for reasons that will be explained in the next chapter, 4). The discrete aspects coincide

with the gene regulation, since (1) these processes are less well understood, and (2) many

of them anyway very closely resemble switches. Finally, and again for very similar reasons,

Chen and Hofestadt (2003) choose hybrid Petri nets as their modeling language.

Similar approaches are taken in the work of Matsuno and friends, (Matsuno et al.,

2000) and (Matsuno et al., 2003). In (Matsuno et al., 2000), genetic regulatory networks are

considered, in particular the genetic switching of the lambda phage, a virus particle, between

lysis and lysogeny (two different methods of virus reproduction). Again HPNs are employed.

Here, the authors additionally stress the ability of HPNs easily to model stochastic behavior

and to be constructed hierarchically. In (Matsuno et al., 2003), the scope is broadened to

biopathways in general; case studies of circadian-rhythm regulation and Fas-ligand-induced

38

Page 57: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

apoptosis (programmed cell death) are modeled, again with a variant on the hybrid Petri

net. Here the authors also stress the advantage in terms of perspicuousness of an HPN over

ODEs for representing large networks of cascades—another consideration that influenced

the present blood-clotting model, coagulation being traditionally described precisely as a

series of cascades. In these papers, as in (Chen and Hofestadt, 2003), large systems are

modeled, and so the inevitably large state space restricts the researchers to simulations, i.e.

away from any analysis.

A very different approach to modeling genetic regulatory networks as hybrid systems

appears in (Jong et al., 2003). There, the authors again consider the threshold-controlled

time evolution of protein concentrations according to differential equations, but assume

that only the orderings of the thresholds are known. Similarly, the rate constants governing

degradation and accumulation of proteins are specified only by inequalities. Specifically,

since the ODEs within each threshold-bounded domain are linear and uncoupled, the equi-

libria within each domain can be expressed explicitly in terms of the rate constants; which

equilibria are then specified only to lie between some pair of the thresholds. This turns

out to be enough information for an algorithm to generate the “qualitative states” of the

system (equilibria and cycles), and the transitions between them.

The theory of hybrid systems can be exploited for the control of biological systems, as

well. The chemical kinetics of many biological systems can be translated into differential

equations which are, in general, multi-affine: i.e., polynomial in their state variables, with

the proviso that the degree of any of those variables is no greater than one.1 A recent set

of papers (Belta et al., 2002, 2004) supplies necessary and sufficient conditions for steering

of multi-affine systems on N -dimensional rectangles; and a recipe for deriving the control

law to effect this steering. The several dynamics in their several rectangles, then, together

compose a hybrid system.

This clever scheme was considered for control of the blood-clotting model, so more

details are in order. The control task of steering the system through a particular facet1The idea is that, picking for a moment a single variable xi and holding all the others constant, the

system is affine in that variable: f(c1, c2, ..., xi, ci+1, ...cn) = mxi + b.

39

Page 58: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

(an [n− 1]-dimensional face), F ∗, of an n-dimensional polytope is equivalent to steering it

through the corresponding facet of the n-dimensional hypercube (since the polytope can

be transformed into a hypercube by an affine transformation of the original state variables,

which transformation will evidently result in another multi-affine system); and this control

task in turn can be be specified in terms of necessary and sufficient conditions on the

differential equations f(x) + g(x)u merely at the vertices of the polytope (rather than

anywhere within the rectangle). In particular, if the vector fields can be shown to point in

the “correct” direction at each vertex—i.e., in the direction of the outward-pointing vector

normal to F ∗ and away from (or parallel to) all of the other facets—then the trajectory

of the state vector, starting from any point within the hypercube, will exit the rectangle

through F ∗. The goal is to choose the input u so as to guarantee that the state remains in

“safe” or acceptable regions. (This scheme is inappropriate for our model, unfortunately;

see Chapter 8.)

Finally, although the present study does not include stochastic processes, we mention

here for completeness two hybrid-system models that do. In (Hu et al., 2004), the produc-

tion of the antibiotic subtilin by the bacterium Bacillus subtilis is modeled via differential

equations, each particular to the current state of a Markov chain. Alternatively, stochas-

ticity enters the model of (Joshi et al., 2004) in the following way: Chemical kinetics, as we

have been stressing, are translatable into differential equations describing the time evolu-

tion of concentrations. This translation, however, assumes that there are enough particles

(proteins, etc.) that the stochastic effects are “washed out”; whereas we may be concerned

with very small quantities, in which case accuracy requires treating the reaction as a se-

ries of discrete, random events. Joshi et al. (2004) proposes a method for dynamically

partitioning the reactions and their corresponding reactants into discrete and continuous

sets, the evolution of the latter being governed by deterministic ODEs and the former by

a continuous-time Markov chain. This method allows accurate and efficient simulation of

chemical reactions with simultaneous fast and slow dynamics. The blood clotting model

of this dissertation, however, will be concerned only with (relatively) large quantities of

proteins, obviating the need for such a technique.

40

Page 59: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Chapter 4

Simulations and Sensitivity

4.1 Introduction

As we saw in Chapter 2, the process of blood coagulation in mammals is complicated,

and involves the interaction of more than a dozen coagulation factors as well as a number of

proteins from the kinin-kallikrein system and protein inhibitors. Attempts to model coagula-

tion mathematically therefore usually focus on a smaller subset of interactions, perhaps one

of the so-called pathways or just a portion of one of them (cf. (Panteleev et al., 2002),(Ko-

gan et al., 2001),(Bungay et al., 2003), (Leipold et al., 1995),(Qiao et al., 2004),(Butenas

et al., 2004) and (Pohl et al., 1994)). Such models generally consist of a set of coupled, usu-

ally nonlinear, differential equations governing the time evolution of protein concentrations.

Although on one level of analysis all of the processes in blood clotting comprise discrete

events, like cleavage of chemical bonds, formation of bonds, and the like; nevertheless, at

the scale of interest it is concentrations that matter, and these exhibit continuous dynamics.

There are, however, at least two reasons why this methodology is inadequate for mod-

eling the entire coagulation cascade. First, there is reason to believe (Halkier, 1991) that

certain events in the cascade are better modeled as “switched.” So, e.g., many of the chem-

ical reactions of the cascade require the presence of calcium ions; that is, not the speed of

41

Page 60: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

these reactions but whether or not they take place at all is controlled by calcium.1 Simi-

larly, concentrations of free zinc ions are thought to toggle the activation of several of the

proteins of the contact activation portion of the clotting cascade (Røjkjær and Schmaier,

1999),(Shariat-Madar et al., 2002); that is, the reactions can take place only after a thresh-

old concentration of zinc has been exceeded. Relatedly, the reactions of the coagulation

cascade take place on very different time scales (which fact, inter alia, causes difficulties

for simulation: it damns the HPN simulator to operate at the granularity of the fastest

reaction, i.e. the smallest time step). Although we have not systematically purged fast

reactions from the model by replacing them with switches, some fast reactions have been

modeled thus for that reason2; and furthermore the hybrid framework pioneered here shows

how such a purge might be accomplished.

The second reason for including discrete events in the model is that the current state

of clinical knowledge is not sufficient to provide a precise description of the various param-

eters and their interactions in terms of differential equations. So, for example: the rate

constants for the various steps in factor XIII activation are not known—even the form of

the equation is unclear; the role of high molecular-weight kininogen in enabling activation

of various members of the intrinsic pathway has not been formalized; and the individual

on- and off- rates for substrate-enzyme interactions has not been determined for various

proteins.3 Certainly there is no closed-form set of differential equations for the system as a

whole. On the other hand, more coarse-grained information is available—e.g., the thrombin

concentration that is required for appreciable conversion of factor XIII into its activated

form. This information can often be incorporated in the form of discrete events toggled

(possibly) by thresholds.

Thus the alternative pursued here is to model the coagulation cascade as a hybrid system1It is perhaps possible that the speed of some of these reactions does indeed vary continuously rather

than discretely with calcium concentration—though to my knowledge this has not been reported—but thatin any case it makes no matter since either (1) intermediate levels of calcium never occur physiologically;or (2) (more likely) the change from “off” to “on” occurs too steeply, i.e. over too small a range of calciumconcentrations, for a continuous variation to be either measurable or meaningful. This latter consideration isakin to the (somewhat standard) modeling practice of replacing steep sigmoid functions with step functionswhere this simplifies things.

2See the previous footnote3These last, however, can at least be approximated by differential equations with known parameters, i.e.

using Michaelis-Mentin rather than mass-action kinetics; see the discussion in Chapter 2.

42

Page 61: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

(HS), i.e. one consisting of interacting continuous and discrete dynamics. Hybrid systems

were discussed at length in Chapter 3; in the present case, our aim was to construct a robust

and faithful model of the coagulation process: faithful in the sense of accurately modeling

human (or generally, mammalian) blood clotting, and robust in the sense of doing so over a

wide range of parameter settings. We additionally required that our model be perspicuous

(with the biologist in mind), and easily modifiable. In light of these constraints, the model

was implemented using hybrid Petri nets (HPNs), described in Chapter 3 in excruciating

detail.4

The model serves both a specific and a more general purpose. Specifically, by accurately

simulating the blood-clotting process, the model serves as a basis for predictions: the effect

of alterations in coagulation-factor concentrations, on both the overall clotting time and on

the concentration of other factors, can be simulated effectively. These simulations can serve

as the basis for predictions about the effect of pharmacological intervention; for understand-

ing the nature of certain blood-clotting disease pathologies (e.g. the various hæmophilias,

factor-V Leiden, etc.); and for refinement of our understanding of blood clotting in general.

More generally, the model demonstrates the utility of the design methodology, viz. using

hybrid systems, and in particular hybrid Petri nets, to model cascade-like biological pro-

cesses where both discrete and continuous dynamics play a role. In virtue of its ability to

incorporate both types of dynamics, the model is able to support robust analysis and pre-

diction in cases where parts of a complex process may be known precisely (with differential

equations) while other aspects may have qualitative descriptions only (through punctuated

phase changes, discrete transitions, and threshold behaviors). This ability to reason ef-

fectively with representations of multiple granularities addresses a central requirement in

modeling complex biological processes.

The model was tested by simulating normal blood clotting, as well as various blood

clotting disorders. The resulting time to clotting and the time course of blood protein

concentrations were compared against the clinical literature, and gave consistent results.4Our current implementation is based on the Visual Object Net++ platform, a dedicated HPN mod-

eling and simulation environment (Drath, 2002), which includes a graphical language that offers a suite ofobject-oriented programming (OOP) features: hierarchical organization, inheritance and object reuse.

43

Page 62: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

4.2 Model Implementation

The cascade consists largely of a series of very similar reaction types, as we shall see, for

which reason I used an object-oriented HPN modeling language; that is, one with constructs

for information hiding and object reuse.

The HPN version of the coagulation cascade is depicted in Figure 4.9, at the end of

the chapter. (A more readable color version of this figure—and the rest of the thesis—

can be found online(Makin, 2008).) Places at the left hand side represent unactivated

blood factors—zymogens of serine proteases, their cofactors, and various ancillary proteins,

regulatory and otherwise—at pre-injury in vivo concentrations. The figure refers to the

factors by their Roman numeral designations (if such they have); alternative names, along

with initial concentrations and their sources in the literature, appear in Table 2.1 in Chapter

2.

The boxes in the figure are obviously neither places nor transitions; they are rather

objects which themselves contain Petri nets. The majority of these objects are in fact

three continuous-time/state modules; the remaining two, the modules Init-Intrinsic and

Fibrin, are hybrid Petri nets.

4.2.1 Modules

Figure 4.9 shows the entire clotting cascade. The overall model comprises multiple

instances of four basic modules (along with a few other places and transitions), with different

parameters for different pathways, factors, and enzymes. There are two modules with purely

continuous dynamics and another two with hybrid dynamics. The two continuous modules

involve (a) blood factor activation and (b) factor-factor binding; the two hybrid modules

involve (a) the initiation of blood clotting in the intrinsic pathway and (b) the formation

of a fibrin clot. These modules are described below.

44

Page 63: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Blood Factor Activation

Figure 4.1 depicts one of the basic aspects of the coagulation cascade, the enzyme-

induced transformation of a blood factor (which may be either a serine protease or a glyco-

protein) from its inactive (zymogen) form to its active configuration. In fact, the coagulation

cascade consists largely of a series of such activations, where the newly activated coagu-

lation factor proceeds to activate another factor (again see Figure 2.1). The places IN1,

IN2 and OUT represent the concentration (in nanomolarity [nM], i.e. nanomoles per liter,

though the units are not conceptually relevant) of various blood factors: IN1 is the zymogen

(substrate) and OUT is its activated form; IN2 is the catalyzing enzyme; IN1:IN2 is an in-

termediate macromolecule. The places labeled with k’s are the constants of classic enzyme

kinetics: on-rates, off-rates, and catalytic rates. This reaction can also be written as a set

of differential equations; square brackets are used to remind the reader that concentrations

are being indicated:d[IN1]dt

= koff[IN1:IN2]− kon[IN1][IN2] (4.1)

d[IN2]dt

= koff[IN1:IN2]− kon[IN1][IN2]

+kcat[IN1:IN2] (4.2)

d[IN1:IN2]dt

= kon[IN1][IN2]− koff[IN1:IN2]

−kcat[IN1:IN2] (4.3)

d[OUT]dt

= kcat[IN1:IN2] (4.4)

Of course, since the variables IN1, IN2, and OUT participate in other reactions, Eqs. 4.1,

4.2, and 4.4 do not completely define the dynamics of any of these variables; the complete

governing equations may contain additional additive terms from other reactions. That is

why these three places have shaped outer rings in Figure 4.1: it indicates that they are

“published,” i.e. available to interact with other objects and hence other reactions. (The

rate constants are also published, but this is rather so that the user can easily change them

without having to open up the object of interest.)

45

Page 64: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Figure 4.1: The activation module, which models the activation of a zymogen (IN1) into its

active configuration (OUT) by a catalyst (IN2).

Factor-Factor Binding

The second recurring reaction is the binding of two blood factors, depicted in Figure 4.2.

Note that, as in the activation reaction (Figure 4.1), the rates constants ki are connected

to transitions via test arcs; this reflects the fact that these quantities are unchanged by the

reaction. In the present case, the governing equations are simply:

d[IN1]dt

= koff[OUT]− kon[IN1][IN2] (4.5)

d[IN2]dt

= koff[OUT]− kon[IN1][IN2] (4.6)

d[OUT]dt

= kon[IN1][IN2]− koff[OUT] (4.7)

Both of these modules appears in many instantiations throughout the model, differing

from each other only in their rate constants and their interconnections with the rest of the

network. The differential equations that they model were drawn from (Bungay et al., 2003).

46

Page 65: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Figure 4.2: The binding module, which models the (reversible) binding of two components,

IN1 and IN2, into the macromolecule OUT.

The Intrinsic Pathway

The Init-Intrinsic module, shown in Figure 4.3, models the initiation of blood clotting

via the intrinsic pathway. (Details of this process were drawn from (Shariat-Madar et al.,

2002) and (Røjkjær and Schmaier, 1999).) The continuous transition labeled “zinc flow

rate” models the increase in zinc concentration according to a simple first-order differential

equation (exponential growth up to an asymptote). At the same time, exposure of a nega-

tively charged surface (modeled as the binary variable m1) allows solid-phase activation of

factor XII (Hageman factor).

When zinc-ion concentration exceeds 0.3 µM, high-molecular-weight kininogen

(HMWK) binds with the plasma protein prekallikrein. The ambient zinc concentra-

tion continues to rise, meanwhile, and when it exceeds 5 µM the activation of prekallikrein

to kallikrein is enabled. This process is greatly enhanced by the presence of activated factor

XII, sufficient quantities of which enable activation via the “fast” transition.

Once activated, kallikrein participates in a feedback loop by enabling the fluid-phase

activation of factor XII, which in turn activates more kallikrein. Notice that factor XIIa can

47

Page 66: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Fig

ure

4.3:

Ahy

brid

Pet

rine

tm

odul

em

odel

ing

the

init

iati

onof

the

intr

insi

cpa

thw

ay.

48

Page 67: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

activate kallikrein through either a “slow” or “fast” transition, where the former corresponds

to activation of free prekallikrein and the latter to activation of prekallikrein bound to the

surface of HMWK. The speeds of these reactions, fast and slow, are modeled by assigning

appropriate time delays to the discrete transitions.

Eventually the concentration of factor XIIa crosses the threshold for the activation of

factor XI (plasma thromboplastin antecedent, or PTA), generating a discrete quantity of

factor XIa (given by the output arc weight). A sufficient quantity of factor XIa prevents

further activation by factor XIIa, hence the inhibitor arc returning from the place XIa to the

activation and binding transition. Factor XIIa production is itself inhibited by the serpin

C1-inhibitor, which is activated by sufficient quantities of fXIIa.

The Fibrin Module

Figure 4.4 shows the Fibrin module, which models the final portion of the blood clotting

pathway: the activation of factors XIII and I, and the formation of a fibrin clot. (Data for

this module were drawn from (Muszbek et al., 1999),(Greenberg et al., 1985),(Standeven

et al., 2005),(Lewis et al., 1997), (Brummel et al., 2002), and (Mosesson, 2000); see Table

4.1 for details.) Factor XIII (fibrin stabilizing factor) normally circulates in plasma bound

to fibrinogen (factor I). When thrombin concentration exceeds a threshold, the Arg37-Gly38

bond of factor XIII is cleaved, producing factor XIIIa′. There is again a “fast” and “slow”

cleavage, modeled by transitions with identical delays but which are enabled by differently

weighted test arcs from the place IIa. The fast cleavage requires lower levels of thrombin

but additionally the presence of fibrin or fibrinogen. (The reader may convince himself that

the boxed elements labeled “OR” in Figure 4.4 do in fact enforce an or-gate of sorts.) Factor

XIII can also be cleaved at the Lys513-Ser514 bond, which renders it useless with respect to

clotting. This cleavage is however entirely inhibited by the presence of calcium ions (Ca2+),

hence the inhibitor arc from the calcium place to this transition. Meanwhile, thrombin

levels rise, and when they exceed 2.48 nM, fibrinopeptide A is released from fibrinogen

thereby activating it to fibrin.

49

Page 68: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Fig

ure

4.4:

An

HP

Nm

odul

ede

pict

ing

the

final

stag

esof

bloo

dcl

otti

ng:

the

acti

vati

onof

fact

ors

Ian

dX

III,

and

the

form

atio

nof

acl

ot.

50

Page 69: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

In the presence either of calcium or fibrin (or both), the A′ and B subunits of factor

XIIIa′ dissociate. Next, again only if calcium ions are present, the active site on the A′

subunits is unmasked, resulting in the transglutaminase FXIIIa*. Finally, FXIIIa* and the

fibrin polymers cross-link to form a clot.

The place labeled m1 in the figure represents the percentage of cross-linking accom-

plished, where each token corresponds to 10%. Thus when m1 acquires ten tokens, cross-

linking is complete. Now, cross-linking is self-regulating in that it inhibits the upstream

promoter effects of fibrin and fibrinogen on factor XIII activation once about 40% of the

cross-linking has been accomplished (Muszbek et al., 1999); hence the inhibitor arc from

m1 to the “fast” transition.

Both of these HPNs made use of various thresholds, arcweights, timing delays, and

binary dependencies (i.e. switches). Table 4.1 lists all of these parameters, and their sources

in the literature, if such there be. Parameters which do not have literature sources were

either interpolated from other relevant data or, in the limit, are informed guesses. As for

the continuous portions of the network, its parameters are all rate constants; they appear

in the Chapter 5.

Finally, several of the activation reactions of the clotting cascade require free calcium

ions. Figure 4.5 reprises Figure 4.1, except that the discrete variable sw switches the binding

and catalytic reactions on and off; if the sw place holds a token (or more than one), then

the reactions may take place, otherwise they may not.

The Complete Cascade

With the component modules explained, we can now turn to the entire coagulation

cascade, referring to Figures 4.9 and 2.1 throughout. To reiterate, the cascade is initiated

through both the intrinsic and extrinsic pathways. The initiation of the former has been

explained; the latter pathway (see the lower left corner of Figure 4.9) begins with the binding

of tissue factor (TF) to lipid-bound factor VII, in both its activated and unactivated forms.

(For the sake of brevity, the many lipid binding reactions will not be mentioned explicitly).

51

Page 70: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Para

met

erT

yp

eD

escr

ipti

on

Valu

eSourc

e

[Zn

2+

]arc

wei

ght

thre

shold

for

HM

WK

:PK

bin

din

g0.3µ

M(S

hari

at-

Madar

etal.,

2002)

[Zn

2+

]arc

wei

ght

thre

shold

for

PK

act

ivati

on

on

cells

M(R

øjk

jær

and

Sch

maie

r,1999)

[Zn

2+

]arc

wei

ght

thre

shold

inhib

itio

nof

HM

WK

:PK

bin

din

g10µ

M(S

hari

at-

Madar

etal.,

2002),

(Røjk

jær

and

Sch

maie

r,1999)

[Zn

2+

]arc

wei

ght

thre

shold

for

FX

IIfluid

-phase

act

ivati

on

10µ

M(S

hari

at-

Madar

etal.,

2002)

[Zn

2+

]arc

wei

ght

thre

shold

for

FX

I:H

MW

Kbin

din

g10µ

M(S

hari

at-

Madar

etal.,

2002)

Zin

cflow

rate

conti

nuous

transi

tion

rate

of

zinc

ion

acc

um

ula

tion

˙[Zn

]=

20−

[Zn

]ex

trap

ola

ted

from

intr

in.

(diff

eren

tial

equati

on)

path

way

tim

edata

[IIa

]arc

wei

ght

thre

shold

for

fast

clea

vage

of

Arg

37-G

ly38

bond

1.5

6nM

inte

rpola

ted

from

(Bru

mm

elet

al.,

2002)

[IIa

]arc

wei

ght

thre

shold

for

slow

clea

vage

of

Arg

37-G

ly38

bond

90nM

inte

rpola

ted

from

(Bru

mm

elet

al.,

2002)

[IIa

]arc

wei

ght

thre

shold

for

rele

ase

of

fibri

nop

epti

de

A2.4

8nM

inte

rpola

ted

from

(Bru

mm

elet

al.,

2002)

[IIa

]arc

wei

ght

thre

shold

for

rele

ase

of

fibri

nop

epti

de

B3.2

8nM

inte

rpola

ted

from

(Bru

mm

elet

al.,

2002)

[XII

I]arc

wei

ght/

del

ayam

ount

of

fXII

Icl

eaved

per

unit

tim

e:

fast

clea

vage

1nM

/3s

inte

rpola

ted

from

(Bru

mm

elet

al.,

2002)

slow

clea

vage

0.3

3nM

/3s

inte

rpola

ted

from

(Bru

mm

elet

al.,

2002)

cross

-lin

kin

garc

wei

ght

per

centa

ge

of

cross

-lin

kin

gat

whic

hpro

mote

reff

ect

of

fibri

non

Arg

37-G

ly38

clea

vage

isin

hib

ited

40%

(Musz

bek

etal.,

1999)

Tab

le4.

1:P

aram

eter

valu

esan

dty

pes

used

inth

eH

PN

mod

ules

ofF

igur

es4.

4an

d4.

3

52

Page 71: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Figure 4.5: The activation module of Figure 4.1, but here two of the reactions are switched

on (off) by the presence (absence) of a discrete variable.

Now, factor VII may be activated by factor Xa, but this requires completion of either the

intrinsic or extrinsic pathways, both of which, as we shall see, require the activation of factor

VII. Thus it is generally accepted (Bungay et al., 2003) that some very small quantity of

factor VIIa (0.1 nM) exists in the blood stream prior to vascular injury. This does not

imply spontaneous activation of the coagulation system since without tissue factor the rest

of the pathway cannot proceed.

The TF:VIIa complex proceeds to bind with lipid-bound factor X (lower right corner of

Figure 4.9), which complex in turn transforms (see the subsequent continuous transition)

into TF:VIIa:Xa via cleavage of a bond in the factor X component. The complex then dis-

sociates (the subsequent transition) into factor Xa (i.e., its activated form, thrombokinase)

and the TF:VIIa complex. Factor Xa then feeds back to activate lipid-bound factor VII

into VIIa as well as the TF:VII complex into TF:VIIa. Factor Xa also binds with a protein

called tissue factor pathway inhibitor (TFPI) (bottom right of Figure 4.9), which inhibits

the extrinsic pathway by binding to free TF:VIIa complex and removing it from further

reactions. Activation of factor X demarcates the traditional ending point of the extrinsic

pathway.

As we have already seen in the Init-Intrinsic module, the exposure of HMWK,

kallikrein, and factor XII to an electro-negative surface results ultimately in the activa-

53

Page 72: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

tion of factor XI to XIa. Factor XIa can then activate factor IX to IXa (middle right of the

figure), but the cascade can proceed no further until factor Xa resulting from the action of

the extrinsic pathway activates lipid-bound factor VIII. This comports with the accepted

theory that the extrinsic pathway serves to kick-start the intrinsic pathway into action (see

above). The activated form of factor VIII (VIIIa) forms a complex with factor IXa, which

in turn activates more (lipid-bound) factor X. This marks the end of the intrinsic pathway.

The common pathway of the coagulation cascade involves factors V, II, and X; plus

the proteins antithrombin, thrombomodulin, protein S, and protein C (see the upper half

of the figure). Lipid-bound factor V is activated by factor Xa. The activated form, Va,

then forms a complex with Xa, which in turn binds to lipid-bound factor II. This Xa:Va:II

complex spontaneously converts (via the continuous transition) to Xa:Va:mIIa, where mIIa

is meizothrombin, an intermediate form of factor II. This complex then dissociates either

into Xa:Va and the activated form of factor II, thrombin (via the continuous transition), or

into Xa:Va and mIIa (via the Va:Xa:mIIa reversible binding module). Both meizothrombin

and thrombin feed back to activate more factor V. Thrombin also feeds into the intrinsic

pathway to activate factor VIII, as does meizothrombin; and to activate factor XI.

The common pathway is inhibited by antithrombin, thrombomodulin, and proteins C

and S. Thrombin binds to thrombomodulin (upper right corner), effectively removing it

from further catalytic reactions. This complex has a further inhibitory role, however, in

activating lipid-bound protein C. Activated protein C (APC) binds with lipid-bound protein

S, and the APC:PS complex feeds back to inactivate both VIIIa in the intrinsic pathway

and Va in the common pathway. Factors Xa, mIIa, and IIa (thrombin), meanwhile, are

irreversibly bound by antithrombin, removing them completely from further reactions.

Thrombin’s final role, of course, is to activate factors I and XIII, but this has already

been detailed above in the description of the Fibrin object.

54

Page 73: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

4.3 Results

We now present the results of various simulations. We start with the normal clotting

pathway, and then simulate some disorders of blood clotting as changes to either the relevant

coagulation factors and associated protein complexes—i.e. the initial values of places in the

network—or to the structure of the network. The results shown for the abnormal cases are

for thrombin levels, but of course simulations produce time-course results for all the places

in the network. Thus the impact of specific diseases or combinations of diseases on the

entire clotting cascade may be examined. This includes both qualitative and quantitative

aspects, like coagulation with versus without therapeutic intervention, and like the effect of

specific dosage levels of an intervention, respectively.

4.3.1 Normal clotting

We choose to examine thrombin because it is the most important enzyme product of

the coagulation cascade: it participates in far more reactions than any of the other factors,

including both feedforward and feedback regulation, and is essential for normal blood clot-

ting. Time to thrombin activation is consequently one of the major parameters measured

in clinical tests. To evaluate the baseline performance of our model, we set the initial con-

centrations as shown in Table 2.1 and compared the time course of thrombin production in

our computational simulation with results reported in the clinical literature (Halkier, 1991;

Bungay et al., 2003). Figure 4.6 shows the concentration of thrombin produced under nor-

mal conditions upon triggering of the clotting cascade. The value of thrombin is shown in

nM, as a function of time, in seconds. Consistent with previously reported clinical studies,

thrombin concentration using the model peaks at about 160 nM around 100 seconds (cf.

(Undas et al., 2001)). A complete clot (i.e. 100% cross-linking) occurs after 105 seconds,

which is again a reasonable figure.

55

Page 74: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Figure 4.6: A simulation of the concentration of thrombin (factor IIa) in nM as it varies

with time (in seconds) in the blood clotting of a normal patient.

4.3.2 Simulating hæmophilia A

Now consider a simulation of hæmophilia A, a disease of the clotting system which

results in excessive bleeding. The cause of the pathology is a deficiency of coagulation

factor VIII, which can range from mild to severe (see Table 4.2).

Table 4.2: Coagulation Disorders

Condition Cause Result Notes

Hæmophilia A deficiency excessive severity:

of fVIII bleeding > 5% (mild)

1− 5% (moderate)

< 1% (severe)

Factor V APCR in fV hypercoag. APC resistance

Leiden ≈ 5% of population

The coagulation disorders simulated in this chapter.

56

Page 75: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Figure 4.7: Simulated results of the concentration (nM) of thrombin versus time (s) during

the clotting process for a person with severe hæmophilia A (low curve) and for normal

clotting (high curve).

To simulate the clotting disorder, therefore, the continuous HPN place VIII, representing

the initial (pre-injury) plasma concentration of clotting factor VIII, is set to 0.035 nM, which

corresponds to the borderline between mild and moderate hæmophilia (cf. Tables 2.1 and

4.2), and the simulation is re-run. The result is shown in Figure 4.7: thrombin concentration

peaks later (at 120 seconds) and much lower (at just over 6 nM), which is congruent with

clinical observations. The maximum cross-linking achieved is 40%, which again is consistent

with the disease pathology.

4.3.3 Simulating factor-V Leiden

Factor-V Leiden is a coagulation disorder characterized by a condition called activated

protein C resistance (APCR), in which a genetic mutation in the factor-V gene renders the

resulting factor-V protein resistant to inactivation by activated protein C (APC). (Recall,

following Figure 2.1, that the function of APC is to inactivate factor Va and factor VIIIa.)

Factor V is a procoagulant, so the consequence of its slower rate of inactivation is generally

a thrombophilic (propensity to clot) state. The disorder has in fact only recently been doc-

57

Page 76: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Figure 4.8: Simulated results of the concentration (nM) of thrombin versus time (s) during

the clotting process for a person with factor-V Leiden (high curve) and a normal patient

(low curve).

umented: APCR was first described in 1993; factor-V Leiden was subsequently discovered

in 1994.

More sophisticated ways of simulating factor-V Leiden are available within the present

model—and in fact I shall demonstrate one such in the context of the control schemes in

the next chapter—but for now, we choose simply to remove the module by which APC

inactivates factor V (see Section 4.2.1). As expected, the result (Figure 4.8) is an increase

in the amount of thrombin produced, congruent with clinical observations (cf., e.g., (van’t

Veer et al., 1997)). This simulation of factor-V Leiden does not predict a shortened clotting,

and this, too, is in agreement with clinical results.

4.4 Decomposition

We noted at the outset that the primary goal of the HS model was simulation: hybrid-

system theory affords far fewer and less powerful analytical tools than the complementary

methods for purely discrete and purely continuous systems. We should like to exploit some

58

Page 77: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

of those methods in the remainder of this thesis—for sensitivity analysis, control, and model

reduction. Therefore, we now propose a decomposition of the system into interacting sub-

systems, with the aim of applying our control techniques to the purely continuous portions,

and then proving the efficacy of these techniques on the overall process by applying a

reachability analysis to the other subsystems.

From Figure 4.9, it is apparent that the Init-Intrinsic module interacts with the other

(continuous) modules only via factor XIa; and that the Fibrin module interacts with these

modules only via thrombin. In the latter, moreover, interaction is only in one direction:

because thrombin affects the Fibrin module only by activating certain (discrete) transitions,

no thrombin is actually consumed by the fibrin module. Now, it must be noted that this is

not strictly correct, since some amount of thrombin complexes with both factors I and XIII;

however, we assume that very little of either complex forms, as in (Khanin et al., 1998). The

Init-Intrinsic module, on the other hand, affects and is affected by the purely continuous

modules (through the evolution of factor XIa). However, deficiencies of the clotting factors

of this early portion of the intrinsic pathway do not have hemorrhagic consequences(Adcock

et al., 2002), (which is in agreement with the present model).

Armed with these considerations, we restrict our attention from here on to the continu-

ous subsystem of Figure 4.9 which comprises all those parts of the cascade not included in

the Init-Intrinsic or Fibrin modules. Now, to prove that control of these ODEs suffices

to induce healthy clotting, we should at least have to show that the amount of factor XIa

produced by the Init-Intrinsic module is negligible compared to the amount produced by

reciprocal activation by factors Xa and IIa. In fact we might be interested in showing for

what range of initial conditions in the Init-Intrinsic module this result obtains. Similarly,

a proof would require us to show that the thrombin trajectories we control are sufficient

to produce a clot, and within a reasonable amount of time. This question, too, might be

posed in terms of ranges: what range of thrombin is sufficient to effect this end?

Both of these questions are questions of reachability. As we have alluded to previously,

and shall spell out in more detail in the next chapter, reachability on large hybrid systems

is very expensive in computational terms, often prohibitively so. A more feasible approach

59

Page 78: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

in the present case is to discretize the state space of these two HPN modules and perform a

discrete reachability analysis. In fact, reachability in (discrete) Petri nets often (depending

on the structure of the network) has a very simple solution (Bause and Kritzinger, 1996).

Whether or not such a discretization is feasible, and whether the resulting network is indeed

amenable to the normal reachability algorithms, are questions we leave to the final chapter,

8.

The idea, then, is to steer model thrombin concentrations, with the hope that it can be

proven that these trajectories suffice to induce proper clotting. Lacking such a proof (and

not wanting to rely on hope), we will throughout aim at the tightest control possible: that

is, we will attempt to force thrombin concentrations in diseased-patient models to match

the trajectory of thrombin during a healthy clotting event. This may be overkill in the

sense that looser criteria might be adduced, via the reachability analysis lately outlined,

to guarantee proper clotting, but it is certainly sufficient (at least insofar as the model is

correct).

Finally, if our controller is going to neglect these two HPN modules, then certainly the

input must not interact with them. Fortunately, the major drugs for the major clotting

problems do not. In the following chapters, we focus on the two most common bleeding dis-

orders, factor-V Leiden and hæmophilia A. The latter is treated with (recombinant) factor

VIII, which lies in the purely continuous subsystem; and the former is treated with heparin,

which—as we shall see—increases the efficacy of antithrombin in inhibiting thrombin and

factor Xa, all of which proteins live in the purely continuous portion of the model.

4.5 Sensitivity Analysis

What if our model is wrong? In particular, what if the rate constants in the ODEs, which

after all were determined empirically, are in error? How much will such an error affect the

model? Alternatively, suppose we wanted to change the course of coagulation by changing

a reaction rate; where shall we get the most bang for our buck? Or again, if a blood-clotting

disease alters certain rate constants (as e.g. factor-V Leiden does), how dramatic will the

60

Page 79: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

effect be on the concentrations of blood proteins? These questions are all variations on the

same theme, namely finding how sensitive a model is to its parameters; and we should like

to compute these sensitivities not simply for this or that parameter, but for the model as a

whole, i.e. for all the rate constants. In what follows we restrict ourselves to a large portion

of the cascade which can be described entirely by a set of ordinary differential equations;

however, the sensitivity analysis can (in principle) and should be extended to the HPNs,

as well (cf. for example (Hiskens and Pai, 2000)). The ODEs analyzed here are, then, the

completely continuous subsystem discussed in the previous section.

4.5.1 Methods

We follow the technique employed on another model of blood clotting, Luan et al. (2007),

which we recapitulate here for completeness. The raw sensitivity σij(t) is the change in state

i (blood-protein concentration) with a change in the jth parameter value (rate constant):

σij(t) :=∂xi∂kj

∣∣∣∣∣t

(4.8)

Note that sensitivities are time-varying. We can find the ODE governing the raw sensitivities

by notingd

dt

∂x∂kj

=∂

∂kjx =

∂f∂kj

+∂f∂x

∂x∂kj

.

Thus the vector of sensitivities σj(t) = [∂x/∂kj ]|t obeys the differential equation

σj = A(x(t),k)σj + bj(x(t),k), (4.9)

where

A(x(t),k) =∂f∂x

∣∣∣∣∣x∗(t),k∗

bj(x(t),k) =∂f∂kj

∣∣∣∣∣x∗(t),k∗

,

i.e. the derivatives evaluated along the nominal trajectory of x and for the numerical

value of k. The matrix A and vector b can be determined analytically (although they

are only estimated in (Luan et al., 2007)), but their value at any given point in time

must be evaluated based on the numerical simulation of the governing ODE x = f(x).

Additionally, the differential equation (4.9) admits no analytic solution, so the two systems

are simulated in tandem. Now, the differentially perturbed trajectory must have the same

61

Page 80: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

initial conditions x0, which is tantamount to insisting that [∂x/∂kj ]|t=0 = 0; so Eq. 4.9 is

simulated with zeroed initial conditions. This procedure is then repeated for all |k| rate

constants.

Having computed σij(t) for each parameter, state, and time sample, we proceed to

combine these values into |k| overall sensitivities, one for each parameter, taking a Euclidean

norm across both time and state, and scaling the contribution of each state variable by (the

inverse of) its magnitude. The result is then scaled by the nominal parameter value, under

the assumption that variations in them will be proportional to their magnitude. Thus the

overall sensitivity for parameter kj is the unitless quantity

σj = kj

[ tf/T∑m=1

n∑i=1

(1x∗i

∂xi∂kj

)∣∣∣∣2t=mT

]1/2

, (4.10)

where T is the sampling interval, tf is the final time of the simulation, and n is (as ever)

the dimension of the state space.

Now, we are interested in the sensitivity of the system not only along the orthodox

trajectory predicted by our model, but for nearby trajectories as well; after all, our rate

constants may be wrong. So, again following (Luan et al., 2007), we simulate Eq. 4.9 for

a range of different parameter values. On each of 200 trials we independently draw each

parameter kj from a uniform distribution stretching from 50% to 150% of its orthodox value

and re-solve (numerically) the two systems of differential equations. The resulting overall

sensitivity vectors σ are normalized at each trial into a [0, 1] range, and averaged over trials.

4.5.2 Results

Table 4.3 shows the 25 most sensitive rate constants, ranked according to average nor-

malized sensitivity σj . We note that 200 trials appears to be sufficient for the σj to converge.

62

Page 81: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Table 4.3: Averaged Normalized Sensitivities

Ranking Reaction Avg. sensitivity ± variance

1 Xa:Va:II Xa:Va:mIIa 0.87 ± 0.028

2 Xa:Va + IIL Xa:Va:II 0.72 ± 0.040

3 TF:VIIa:X TF:VIIa:Xa 0.71 ± 0.046

4 TF + VIIaL TF:VIIa 0.65 ± 0.040

5 XaL + VaL Xa:Va 0.65 ± 0.039

6 Xa:Va + IIL Xa:Va:II 0.62 ± 0.037

7 V:mIIa VaL + mIIaL 0.56 ± 0.030

8 XaL + VaL Xa:Va 0.53 ± 0.026

9 XL + TF:VIIa TF:VIIa:X 0.52 ± 0.031

10 XL + TF:VIIa TF:VIIa:X 0.51 ± 0.029

11 XIa:IX IXaL + XIa 0.48 ± 0.023

12 PCL + IIa:Tm IIa:Tm:PC 0.47 ± 0.019

13 XI:IIa XIa + IIa 0.46 ± 0.022

14 IXaL + VIIIaL IXa:VIIIa 0.46 ± 0.021

15 XI + IIa XI:IIa 0.46 ± 0.022

16 VL + mIIaL V:mIIa 0.43 ± 0.018

17 XL + IXa:VIIIa IXa:VIIIa:X 0.43 ± 0.017

18 VIIIaL + APC:PS APC:PS:VIIIa 0.41 ± 0.017

19 IIa:Tm:PC APCL + IIa:Tm 0.40 ± 0.017

20 VIII:mIIa VIIIaL + mIIaL 0.39 ± 0.014

21 XI + IIa XI:IIa 0.38 ± 0.017

22 VL + mIIaL V:mIIa 0.38 ± 0.014

23 IXL + XIa XIa:IX 0.37 ± 0.012

24 IXa:VIIIa:X XaL + IXa:VIIIa 0.36 ± 0.014

25 APC:PS:VIIIa VIIIaiL + APC:PS 0.35 ± 0.011

Averaged, normalized sensitivities (see Methods) for the 25 most influential rate constants. The rate

constant is given by its associated chemical reaction.

63

Page 82: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

How does this compare to the sensitivity analysis of (Luan et al., 2007), which models

more or less the same portions of the cascade, but including a few more equations for

platelet binding? First of all, these platelet reactions account for about half their list of

most “fragile” (sensitive) reactions—which is somewhat curious, since platelet disorders of

coagulation are not common. Second, the reactions involved in Xa:Va activation of thrombin

are on average more sensitive in our model than the reactions for activation of factor X by

the TF:VIIa complex; whereas the converse is true in the Luan et al. (2007) model. In

both, however, these reactions (and their related platelet reactions in (Luan et al., 2007))

dominate the list.

4.5.3 Sensitivity and prothrombin time

One of the (two) most common clinical tests for a clotting problem is the prothrombin-

time (PT) test, a measure of the time required for TF-initiated coagulation of blood plasma.

The test is performed by first drawing blood into a test tube containing citrate, which soaks

up the free calcium ions and thus prevents the main clotting reactions. The blood is then

centrifuged and the plasma separated from the blood cells, after which an excess of calcium

and tissue factor are added. Clotting time is measured from the addition of TF to the

first visible signs of clot formation in the tube, i.e. when about 10 nM thrombin has been

produced (Mann et al., 2003). The healthy range for this time is considered to be 12-15

seconds.

Now, this set-up is very close to the simulations of this chapter, but in the PT test,

clotting is initialized by a saturating amount of tissue factor (20 nM, vs. 0.005 nM in

the simulations above). However, in our model, this results in a clot time around 30

seconds. Moreover, the differences of the present study with the PT test all suggest a shorter

rather than a longer clot time: our model includes phospholipids (recall that the PT test is

performed on purified plasma), which enhance reactions rates; as well as another initiation

mechanism for clotting, the intrinsic pathway. Why do we find this large disparity between

clot times, then? If we assume the structure of the model is correct, then the difference must

be caused by parameter errors. And this seems rather plausible, since literature values for

64

Page 83: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

rate constants vary by upwards of orders of magnitude (Panteleev et al., 2006) (or cf. the

rate constants of (Bungay et al., 2003) and (Luan et al., 2007), and their references).

How, then, might we change the parameters as little as possible to achieve a PT of

(say) 13 seconds? A principled approach will exploit the sensitivities of the system; in

particular, we might try to compute the sensitivity of thrombin over the first p seconds to

changes in the rate constants, i.e. ∂z/∂kj , where z is the concentration of thrombin; and

then change only the most sensitive parameters, moving them in the direction of greater

thrombin. Without an explicit solution to this maximization problem, this would amount

to a kind of gradient ascent, changing parameters incrementally in the direction of greatest

thrombin increase.

Unfortunately, true gradient ascent would require the recalculation of the thrombin

sensitivity ∂z/∂kj at every step through the ascent, and as we have seen, Eq. 4.9 has no

analytic solution; and furthermore the numerical solution is quite expensive, computation-

ally, since it requires simulation of a (|k|+ |x|)-dimensional system, where (|k|+ |x|) ≈ 200.5

Rather than repeatedly recalculate the gradient, then, we employ a similar method to the

one lately described, viz. averaging the sensitivity over a range of different parameter values,

each randomly perturbed from the “correct” parameters (see above). This time, however,

we care about the sign of the sensitivity (it tells us which whether to increase or decrease

the parameter), so we dispense with the squaring in Eq. 4.10 and simply average the results

of the several unnormalized trials:

s[IIa]j = kj

(1x∗i

∂xi∂kj

)∣∣∣∣t=t∗

, (4.11)

where t∗ is the desired clot time, 13 seconds. Now we can use this thrombin-sensitivity vector

s[IIa] as an approximation of the gradient throughout. Also note that although we have

again neglected the discrete parts of the network, we can do so with some confidence, since

(1) the PT test ignores the intrinsic pathway, and (2) we can estimate clotting time from

thrombin alone as long as we bear in mind that this estimate assumes healthy functioning5Now, Eq. 4.9 is a linear, time-varying differential equation, for which an analytic solution exists in

the form of a matrix power series (Blair, 1971). However, the early terms of the series are so large as tointroduce rounding errors into the low (and ultimately most meaningful) bits of the sum, rendering thetechnique useless for the present ODE.

65

Page 84: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

of fibrin and factor XIII (but after all, in these simulations we are assuming that the entire

clotting process is healthy). We therefore follow (Mann et al., 2003) and take 10 nM of

thrombin to correspond to the formation of a visible clot in the test tube.

4.5.4 Results

In all experiments, the initiating dose of TF is 20 nM, as in the actual PT test (Mann

et al., 2003). Now, with our estimate of the thrombin gradient in hand, one question we

might ask is: How few parameters can we alter and still effect the desired PT, given an

upper limit on those parameter shifts? Table Table 4.4 shows the results for three different

upper limits.

Table 4.4: PT-Time Parameter Shifts

Max Shift (%) Min Params

60 12

50 19

40 70

Minimum number of parameters required to shift in order to achieve a PT time of 13 seconds, for

three different limits on the maximum parameter shift.

Two interesting features of these results should be noted. First, entertaining parameter

variances of the reasonable (in light of the dispersion of literature values) figure of 60%

enables the model to match clinical PT times with the shifting of just 12 (out of over 100)

parameters. This shows that very few parameters can do a lot of work. Contrariwise, forcing

all rate constants to live within 40% of their nominal values precipitously increases this

minimum number of rate constants, up to about two-thirds of them. Thus the relationship

between the maximum shift and minimum number of parameters is evidently nonlinear;

and in particular the latter moves from a small minority of the total number of parameters

to a sizable majority as the maximum shift is drawn down under the 50% mark.

66

Page 85: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Before we leave this small investigation, we reverse the question and ask how small we

can make the maximum parameter shift if we let, say, all of the parameters vary. The

answer is 38%—which, considering that 40% shifts require the movement of about 40 fewer

rate constants, and 50% shifts require alteration of 90 fewer, indicates how very little effect

these remaining constants have on thrombin levels—at least at t = 13 seconds.

4.6 Discussion and Conclusions

The modeling framework and software program described provides a robust, faithful,

interactive, and graphical computer simulation of the entire coagulation process. The frame-

work is faithful in that it accurately models mammalian blood clotting; robust in the sense

of doing so over a wide range of parameter settings; interactive in that the system operation

and parameter settings can be interactively changed while the software program is execut-

ing, and hypothetical “what-if” simulations performed; and graphical in that the model is

a formal graphical structure that supports visualization of the clotting process as well as

exact quantitative analysis.

The simulations of factor-V Leiden and hæmophilia A were consistent with clinical

results, and to that extent vindicate both the present model and its parameters as well

as the methodology, i.e. a HS approach. Of course, consistency is not tantamount to

quantitative identity, but here the obstacle lies not with the model so much as the state of

the clinical data. As Wagenvoord et al. (2006) has pointed out, most thrombin curves can

be fit by an equation with a mere four parameters, so these data vastly underdetermine the

parameters of the model. Including clot time does not substantially decrease the number

of free parameters.

We interpret our simulations, then, as having demonstrated a method for incorporating

qualitative or otherwise discrete information into a single model, which model is consistent

with clinical findings; but I do not claim the model to be error-free. On the other hand, the

model also provides a platform for testing clinical hypotheses, quantitative or qualitative,

about the coagulation cascade, since consistency with known data is a necessary (if not

67

Page 86: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

sufficient) criterion for the correctness of those hypotheses. This provides a method for

model refinement, as well. Furthermore, inasmuch as this kind of model can indeed capture

all the relevant details of coagulation, the analysis techniques demonstrated in this chapter

and those to come will apply equally well to a “final,” error-free model.

In anticipation of the control techniques of the next two chapters, we discussed how

two hybrid subsystems could be decomposed from the rest of the model, leaving a large

system of ODEs—which will be augmented in the next chapter by additional differential

equations for the interaction of heparin with the system and for a more accurate simulation

of factor-V leiden. This decomposition allows us to treat our design of clinical interventions

as a problem in nonlinear control theory.

Finally, we saw that (restricting our view to the continuous subsystem) relative changes

in protein concentrations are most sensitive to changes in the rate constants involving

activation of thrombin by the Xa:Va complex and activation of Xa by the TF:VIIa complex.

We also saw that these sensitivities could be exploited to change the model in the most

parsimonious way in order to match the results of the PT test: shifting a mere 12 rate

constants by as little as 60% sufficed to achieve a reasonable 15 seconds for this test.

On the other hand, it turns out that using these parameters in the original clotting

simulations of this chapter vitiates their agreement with clinical results. That suggests that

something more subtle than rate constant alterations is responsible for the discrepancy in

PT-test results.

68

Page 87: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Figure 4.9: The HPN model

69

Page 88: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Figure 4.9: The HPN model

70

Page 89: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Figure 4.9: The HPN model

71

Page 90: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Figure 4.9: The HPN model

72

Page 91: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Chapter 5

Control: A First Pass

5.1 Introduction

The complexity of the coagulation process that has been stressed so far renders treat-

ment of its disorders difficult. Indeed, current treatments make no attempt to manipulate

coagulation in real time; rather, periodic (on the order of one day to one week) measure-

ments and interventions are made with the aim of keeping certain known risk factors within

safe ranges. This limitation is a consequence of two facts: first, the state of current sensor

technology, which precludes real-time measurements of blood proteins (and hence real-time

feedback control); and second, the lack of theoretical techniques for such control. This chap-

ter addresses the second of these limitations by applying techniques from computer science

and from mathematical control theory and demonstrating their validity—and limitations—

both mathematically and in a series of simulations. By providing part of the solution to

the theoretical problem of controlling blood clotting, I hope to provide an impetus for the

development of the relevant sensor technologies.

The simulations of Chapter 4 demonstrate the utility of the model for certain purposes,

namely prediction, theoretical investigation, and sensitivity analysis. However, ultimately

we should like to use it as a basis for the real-time control of blood-clotting. Now, as

we saw in Chapter 3, the standard technique for control of hybrid systems is numerical

73

Page 92: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

solution of a set of partial differential equations—the Hamilton-Jacobi equations. However

it turns out that for systems of dimension greater than about five, solution of the equations

is computationally infeasible (a consequence of the curse of dimensionality) (Mitchell et al.,

2001). Unfortunately, our system has upwards of 100 state variables, so this technique is

patently unworkable.

The approach taken here is to consider only a subsystem of the model, a set of nonlinear

ODEs which were originally drawn largely from (Bungay et al., 2003) and (Hockin et al.,

1999). In fact, the model can be decomposed into this (purely continuous) subsystem and

two hybrid systems, the three interacting only through two state variables, as we have seen

in Chapter 4, so that control of clot formation in the complete model can be approached

as a problem of continuous (nonlinear) control. Thus in the present study we content

ourselves with deriving and simulating the control techniques for the system of ODEs alone.

Specifically, the control task is to steer the concentration of thrombin (activated blood

factor II) along a desired trajectory during a clotting event, by controlled rate of injection

of an exogenous pharmaceutical (e.g. heparin). I demonstrate two different techniques

to effect this control: the first, following (Sastry, 1999), using nonlinear feedback and a

change of variables to partially linearize the ODEs, and then controlling the linearized,

single-input/single-output (SISO) system using standard methods; and the second based

on step-input control. In the simulations, the input is either the anti-coagulant heparin or

(recombinant) factor VIII, and the output is thrombin, the most important enzyme product

of the coagulation cascade (see below).

We present the theory of the control techniques first, along with some preparatory results

on alternative approaches, before presenting a series of simulations. We then simulate the

clotting process in a patient with the pro-coagulatory disorder factor-V Leiden and in a

patient with moderate hæmophilia A, and then repeat the simulation under application of

the control techniques. Finally, we discuss both the relevance of this control task to the

overall task of controlling blood clotting, as well as implementation issues, and then propose

an alternative technique to remedy some of the defects of the present approach.

74

Page 93: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

5.2 Theory

The model to be controlled is a system of about 100 coupled, nonlinear ODEs, of the

most general form,

x = f(x, u) (5.1)

where u is the (single) control variable (say, an anti- or pro-coagulant). The control task

is to force one of the state variables (thrombin) to track a desired trajectory. In fact, the

ODEs can be expressed, as we shall see, in a less general form; and the approach described

in this chapter will be to exploit some of the peculiarities of the system which distinguish

it from the most general case.

Now, control of linear systems is comparatively easy, so a standard approach is to design

the controller around a linear approximation to the true system, found by considering only

the first term of the Taylor expansion, the Jacobian [∂f/∂x](x0), near an equilibrium point

x0. Unfortunately, the Jacobian in our model is singular, so (by the Hartman-Grobman

theorem) the linearization is not guaranteed to approximate the true system.

Alternatively, the system may be exactly linearized (as opposed to merely approximated

by a linear system) by choosing the appropriate nonlinear feedback u = ψ(h(x)) and looking

at the system through a change of variables ξ = Φ(x):

ξ = Aξ + bv,

y = ξ1, (5.2)

where in fact (A,b) are in controllable canonical form, so the system is completely control-

lable. Here h(x) is an output function which reflects our observations of the state variables,

and v is a synthetic input related to the true input u by a known function. From this per-

spective we can ask whether or not, given the dynamics in Eq. 5.1, there exists an output

function rich enough to support the linearization. The answer is particularly straightforward

if, as in our case, the ODEs are affine in the control, i.e. can be written as:

x = f(x) + g(x)u, (5.3)

75

Page 94: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

in which case necessary and sufficient conditions for the linearization can be given in the

form of conditions on the vector fields f(x) and g(x) and iterated Lie brackets thereof.

Specifically, the matrix of vector fields

[g(x), adfg(x), ..., adn−2f g(x), adn−1

f g(x)], (5.4)

often referred to as the strong-accessibility distribution, must have full rank (n) in the region

of interest; and the set of vector fields

g(x), adfg(x), ..., adn−3f g(x), adn−2

f g(x) (5.5)

must be involutive 1 in the region. Here, by recursive definition,

adkf g := [f , adk−1f g], k > 0

adkf g := g, k = 0.

If these conditions are not met and full-state linearization is not possible, one may

attempt to partially linearize the system. Here, the linearization is carried out with respect

to some given output function, which will admit the formulation of some number q of new

state variables with linear dynamics. Of course, if full-state linearization is not possible,

then q, called the relative degree of the affine-control system, is strictly less than n, the

dimension of the state space.

Consider again the differential equations of our system as the affine-control system of

Eq. 5.3, where this time h(x) = y, the variable we wish to control. Differentiating the

output with respect to time yields:

y = (∇xh)x (5.6)

= (∇xh)f(x) + (∇xh)g(x)u

= Lfh(x) + Lgh(x)u,

where Lfh(x) := (∇xh)f(x), the Lie derivative of the function h along the vector field given

by f . Now if Lgh(x) is nonzero for all x, then the system is said to have a strict relative

1A set of vector fields is involutive if the Lie bracket of any two of those vector fields is within the spanof the original set, where the Lie bracket [f ,g] of two vector fields is defined as [∂g/∂x]f − [∂f/∂x]g.

76

Page 95: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

degree of one, and we can force it to track a desired trajectory yd by choosing:

ud =1

Lgh(x)(yd − Lfh(x)), (5.7)

and making sure that the initial conditions match (y(0) = yd(0), y(0) = yd(0)). Changing

variables according to ξ = Φ(x) := h(x) and defining for the nonce yd =: v, a synthetic

input, we see a one-dimensional linear system ξ = 0 · ξ + v and an (n − 1)-dimensional

nonlinear system η = λ(ξ, η). (The variable v is called a synthetic input in virtue of its

role as the input to this linear system.) Notice that the linear system is decoupled from the

nonlinear one, in the sense that ξ is not a function of η.

If, on the other hand, Lgh(x) is zero for all x, then we differentiate y a second time to

get:

y = L2fh(x) + LgLfh(x)u, (5.8)

and, if for all x, LgLfh(x) 6= 0 (i.e. the system has strict relative degree two), choose:

ud =1

LgLfh(x)(yd − L2

fh(x)), (5.9)

this time making sure that y(0) = yd(0) as well. Choosing ξ = Φ(x) := [h(x), Lfh(x)]T

again yields a linear system, this time two-dimensional, which is decoupled from the re-

maining (n− 2)-dimensional nonlinear system.

In general, for a system of strict relative degree q, the control law is

ud =1

LgLq−1f h(x)

(v − Lqfh(x)), (5.10)

where we choose the synthetic input v = y(q)d and set all the initial conditions

yd(0), yd(0), ..., y(q)d (0) appropriately. Note that, if q = n, the dimension of the state space,

then

Φ(x) := [h(x), Lfh(x), ..., Ln−1f h(x)]T (5.11)

is a valid change of coordinates (locally diffeomorphic), which in fact puts the system into

controllable canonical form—i.e., (Eq. 5.3) is fully linearizable by state feedback. The

procedure is illustrated in Figure 5.1.

77

Page 96: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Finally: we have so far ignored the case where LgLq−1f h(x) is zero for some values of x,

and nonzero for others. In this case the relative degree of the system is not well-defined,

and complications ensue. (We shall see in the sequel that this is indeed the case in the

present system.)

5.3 Application to the model

5.3.1 Feedback linearization

Full-state linearization as lately described faces a significant issue when applied to the

present model. Although verifying the two conditions associated with (5.4) and (5.5) is

mathematically straightforward, it is computationally intensive. Computing the vector

fields requires the computation of two Jacobians for every Lie derivative, although one of

them, ∂f/∂x, need only be computed once. However, each of the approximately 100 vector

fields of (5.4) needs to be computed, and computation of each of the associated Jacobian

matrices entails 1002 derivatives of polynomials. Even neglecting the multiplication and

addition operations, this brings the total to a million computations—each of which is a

(symbolic) derivative of a polynomial, increasing the total number of computations yet

more. And most unfortunately of all, each derivative (at least potentially) generates more

variables, via the chain rule of differentiation, so the polynomials have increasingly more

terms in later vector fields.

Apart from computational difficulties, it remains to construct the output function, h(x).

No mechanical procedure for this construction exists.

In fact, we can circumvent the computational obstacles: Bastin and Levine (1993) have

shown that, in the case of reaction systems like the present one, the strong-accessibility

rank is bounded from above by the rank of a matrix D (the “accessibility matrix”) which

is independent of x, and is computed very simply. The existence of such a matrix depends

once again on being able to express the governing ODEs in an even more specific form. In

78

Page 97: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

particular, Eq. 5.3 can be written

x = Cr(x) + bu, (5.12)

where each element in the vector r(x) is a monomial corresponding to one of the chemical

reactions of the system (i.e one of the arrows in the chemical formulae); and b is a constant

vector, i.e. does not depend on the state. The present case is a special case even of the one

discussed in (Bastin and Levine, 1993), since our model includes neither inflow nor outflow

of reactants. For this system, the accessibility matrix is the augmented matrix D := [b, C],

which we now show.

Consider a vector field s(x) in the span of D, that is

s(x) =∑i

diφi(x), (5.13)

for arbitrary functions of x, φi, and where di are the columns of D as defined above. Then

the Lie bracket of the drift vector field f with s is

[f , s] =∂s∂x

f(x)− ∂f∂x

s(x)

=∑i

di(∇xφi)f(x)− C ∂r∂x

s(x). (5.14)

Now, since (∇xφi)f(x) is just another arbitrary function of x, the first term is within the

span of D (compare this term with Eq. 5.13). Similarly, the second term lies within the

span of C and thus a fortiori lies within the span of D. Thus we see that the Lie bracket

of f , the drift term of our ODE, with any vector field in the span of D, yields a vector field

which is itself in the span of D. Obviously g ∈ SpD, and hence ad1f g = [f ,g] is as well,

by the fact just noted. And now since ad1f g ∈ SpD, then so is ad2

f g = [f , ad1f g]—and so

on, extending the claim to all the vector fields in the strong-accessibility distribution. Thus

we can say:

∀x, rk(D) ≥ rk([g(x), adfg(x), ..., adn−2f g(x), adn−1

f g(x)]). (5.15)

Crucially, the rank of D is easy to compute, being but a constant matrix, and we avoid

crunching through the iterated Lie derivatives of the right-hand side.

79

Page 98: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

For a hæmophiliac patient, we use factor VIII as the single input, giving rk(D) = 83 <

n = 98—so the system is not fully linearizable by state feedback; i.e. there is no single-

variable output function h(x) that can be constructed to support such feedback. In the case

of factor-V Leiden, several reactions corresponding to the inactivation of factor Va, as well

as the associated state variables (namely, the inactivated forms of FVa), are removed from

the model. (The modeling of factor-V Leiden is explained at length below.) Treatment

with heparin in turn requires the addition of several reactions and state variables. However,

none of these changes, nor the different control vector field b—with its single nonzero entry

in a different row—suffice to eliminate the deficit between the rank of D and the dimension

of the state: rk(D) = 73 < n = 89. The meaning of this deficit we shall discuss later.

So the system is not completely controllable. However, there is reason to believe that

clotting can be controlled just by controlling the concentration of thrombin (activated factor

II): Clotting occurs downstream in the coagulation cascade of the ODEs of this study, and

interacts with them only via thrombin. (See the previous chapter for the full justification.)

Thus we attempt to partially linearize the system of Eq. 5.3 with thrombin the output

y = h(x). Indeed, using heparin as the control variable and augmenting the system with

the appropriate terms from the heparin chemical reactions, the strict relative degree of the

system is two.2 Alternatively, using factor VIII as the input (for hæmophilia) yields a strict

relative degree of three.

However, the relative degree for neither of these systems is well-defined; in particular, the

relevant Lie derivative is zero at the initial condition, x0. We circumvent this difficulty by

allowing the plant to run uncontrolled until the concentration of thrombin is some distance

δ from the origin, i.e. until the state has drifted away from the singularity at which the

relative degree changes, before turning the controller on. The parameter δ is tuned manually2Lacking complete controllability, is there something the system has gained? I.e., is there a trade-off in

the area? An affirmative answer can be given if one considers the computational complexity: In order tocompute the feedback law, we need only calculate (two) twice-iterated Lie brackets; whereas if the systemwere completely controllable, the two vector fields would require the calculation of n-iterated Lie brackets.In a system as complicated as ours, this would be a serious problem, even given a simple output function.Additionally, we need only specify two, rather than n, initial conditions. This question is even more relevantwhen an output function which will provide strict relative degree n is available.

80

Page 99: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

for ideal tracking. (The non-regularity of the system and its consequences are discussed in

more detail in Section 5.6.2 below.)

Error Correction

Now, in practice, small numerical discrepancies arise—e.g. from a mismatch between the

plant and the model, or from numerical approximations—so that the trajectory produced

by the control law of Eq. 5.10 deviates from the desired trajectory. Compensation is made

in the form of additional elements in the feedback loop, usually some variant on the PID

controller. A proportional and an integral term are generally used (Henson and Seborg,

1996), but here we content ourselves with proportional and derivative terms (as many as

the strict relative degree of the system) since (1) the integral term did not seem to improve

tracking in our simulations; and (2) it can be shown (Sastry, 1999) that, if the system of

Eq. 5.3 is globally exponentially minimum-phase, then this control law guarantees bounded

trajectory tracking, i.e. the tracking error and its derivatives tend asymptotically to zero.

That is, we replace the synthetic input v in Eq. 5.10 with:

v(t) = y(q)d (t) + [Kp K

1d · · · K

q−1d ][yd − ξ], (5.16)

where yd is the vector of the desired output and its derivatives:

yd = [yd yd · · · y(q−1)]T,

and ξ is the linearized state, which in our system corresponds to the actual output y and

its derivatives. Determining whether the system is in fact globally exponentially minimum

phase, however, requires some discussion, which we defer to Chapter 6.

The gains were calculated by treating the system as a linear-quadratic regulator, with—

unless otherwise noted—state-penalty matrix set to the identity matrix and the input

penalty set to unity. However, control of the nonlinear system was observed to be fairly

insensitive to the choice of cost parameters.

81

Page 100: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Discretization

Finally, practical application will also require discretization of the controller. Now, the

sampling rate of the controller affects the controllability of the system: under a sufficiently

low rate the theory lately outlined will fail to achieve exact output tracking. We demonstrate

a working discrete controller in the simulations below.

5.3.2 Step-input control

A much simpler control technique was also applied; the rationale for it is discussed below.

This strategy is predicated on the assumption that a single step input might suffice to force

the system to track the desired thrombin trajectory “reasonably well.” More precisely, the

assumption is that the step input that matches both the desired peak concentration of

thrombin and the occurrence of this peak will result in a thrombin trajectory that deviates

very little over its entirety from its desired counterpart. Therefore this technique was

implemented by performing a parameter search over repeated trials for that step input

which would minimize the peak-concentration and peak-time discrepancies. Thus the input

u was modified on successive trials according to:

u = u+ α(maxty(t) −max

tyd(t)) + β(argmax

ty(t) − argmax

tyd(t)), (5.17)

where α and β—positive for anti-coagulant inputs and negative for pro-coagulants—scale

the relative contribution of each term, and were adjusted by hand.

5.4 Methods

We draw our chemical reactions from three main sources, the first two of which collect

their own data from a number of primary sources (the last is itself a primary source):

(Bungay et al., 2003), (Hockin et al., 1999), and (Olson, 1988). A description of each

follows.

Forty-eight chemical reactions for normal blood clotting, with no exogenous interven-

tion, were taken directly from (Bungay et al., 2003), from which also the numerical values

82

Page 101: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

of the rate constants were supplied. They are reproduced here in Table 5.1. The blood

factors are referred to by their Roman-numeral designations, a lowercase “a” denoting the

activated form. Other abbreviations include: mIIa for meizothrombin, LBS for the concen-

tration of lipid binding sites, PS for protein S, PC for protein C, APC for its activated form,

TFPI for tissue-factor-pathway inhibitor, Tm for thrombomodulin, AT for antithrombin,

and TF for tissue factor. A subscripted “i” indicates the inactivated form of an enzyme,

and a subscripted “L” indicates the lipid-bound form.

Table 5.1: Chemical Reaction Set I

No. Reaction kon (nM−1 s−1) koff (s−1) kcat (s−1)

1 II + LBS IIL 0.0043 1 -

2 mIIa + LBS mIIaL 0.05 0.475 -

3 V + LBS VL 0.05 0.145 -

4 Va + LBS VaL 0.057 0.17 -

5 VII + LBS VIIL 0.05 0.66 -

6 VIIa + LBS VIIaL 0.05 0.227 -

7 VIII + LBS VIIIL 0.05 0.1 -

8 VIIIa + LBS VIIIaL 0.05 0.335 -

9 IX + LBS IXL 0.05 0.115 -

10 IXa + LBS IXaL 0.05 0.115 -

11 X + LBS XL 0.01 1.9 -

12 Xa + LBS XaL 0.029 3.3 -

13 APC + LBS APCL 0.05 3.5 -

14 PS + LBS PSL 0.05 0.2 -

15 VIIIai + LBS VIIIai,L 0.05 0.335 -

16 PC + LBS PCL 0.05 11.5 -

17 TFL + VIIaL TF:VIIaL 0.5 0.005 -

18 TFL + VIIL TF:VIIL 0.005 0.005 -

19 TF:VIIaL + IXL TF:VIIa:IXL

→ TF:VIIaL + IXaL 0.01 2.09 0.34

20 TF:VIIaL + XL TF:VIIa:XL → TF:VIIa:XaL 0.1 32.5 1.5

21 TF:VIIa:XaL → TF:VIIaL + XaL - 1 -

22 TF:VIIL + XaL TF:VII:XaL → TF:VIIaL + XaL 0.05 44.8 15.2

(continued on next page)

83

Page 102: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Table 5.1 – continued

No. Reaction kon (nM−1 s−1) koff (s−1) kcat (s−1)

23 IXaL + VIIIaL IXa:VIIIaL 0.1 0.2 -

24 XaL + VaL Xa:VaL 1 1 -

25 IXa:VIIIaL + XL IXa:VIIIa:XL

→ IXa:VIIIaL + XaL 0.1 10.7 8.3

26 VL + XaL V:XaL → VaL + XaL 0.1 1 0.043

27 VIIIL + XaL VIII:XaL → VIIIaL + XaL 0.1 2.1 0.023

28 VL + IIa V:IIaL → VaL + IIa 0.1 6.94 0.23

29 VIIIL + IIa VIII:IIaL → VIIIaL + IIa 0.1 13.8 0.9

30 Xa:VaL + IIL Xa:Va:IIL 0.1 100 -

31 Xa:VaL + mIIaL Xa:Va:mIIaL 0.1 66 -

32 Xa:Va:IIL → Xa:Va:mIIaL - - 13

33 Xa:Va:mIIaL → Xa:VaL + IIa + LBS - - 15

34 VIIL + XaL VII:XaL → VIIaL + XaL 0.05 44.8 15.2

35 XI + IIa XI:IIa → XIa + IIa 0.1 10 1.43

36 APC:PSL + VIIIaL APC:PS:VIIIaL

→ APC:PSL + VIIIai,L 0.1 1.6 0.4

37 TFPI + Xa TFPI:Xa 0.016 0.000333 -

38 TFPI:Xa + TF:VIIaL TFPI:Xa:TF:VIIaL 0.01 0.0011 -

39 IXa + AT → IXa:AT 4.9 ×10−7 - -

40 Xa + AT → Xa:AT 2.3 ×10−6 - -

41 IIa + AT → IIa:AT 6.83 ×10−5 - -

42 VL + mIIaL V:mIIaL → VaL + mIIaL 0.1 6.94 1.035

43 VIIIL + mIIaL VIII:mIIaL → VIIIaL + mIIaL 0.1 13.8 0.9

44 IIa + TmL IIa:TmL 1 0.5 -

45 IIa:TmL + PCL IIa:Tm:PCL → IIa:TmL + APCL 0.1 6.4 3.6

46 mIIa + AT → mIIa:AT 6.83 ×10−6 - -

47 APCL + PSL APC:PSL 0.1 0.5 -

48 XIa + IXL XIa:IXL → XIa + IXaL 0.01 1.417 0.183

The chemical reactions and rate constants drawn from (Bungay et al., 2003). They govern all the

reactions except factor Va inactivation and the interaction of heparin with the system.

The simple mechanism of factor Va inactivation of the Bungay et al. (2003) model

84

Page 103: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

consisted of a single chemical equation. We substituted for this equation 3 the model of

factor-Va inactivation of (Hockin et al., 1999), which was designed specifically to match data

for factor-V Leiden. These 34 chemical equations are listed in Table 5.2. As discussed in

Chapter 2, factor-V Leiden is a genetic disorder in which the amino acid Arg505 of the heavy

chain of factor V is replaced by Gln505, which effectively disables APC-mediated cleavage

of factor V at this point. Hockin et al. (1999) were able to match data for inactivation

of FVaLEIDEN by assuming that the form of factor V cleaved at Arg505, Va5, has no effect

on clotting, i.e. is “inactivated.” (This is so even despite its reported 60% efficacy in

complex with factor Xa as prothrombinase. Since it has no overall effect, and for the sake

of simplicity, we treated Va5L as inert.)

We made a further simplification in the incorporation of this model: Hockin et al. (1999)

model the inactivation of factor Va by APC, rather than the more efficacious APC:PS com-

plex, as in (Bungay et al., 2003). Now in fact, uncomplexed APC competes with APC:PS

for binding with factor Va, but Bungay et al. (2003) leaves out this competition—i.e., only

APC:PS is modeled as binding with Va—presumably because, under normal conditions,

very little free APC circulates: the vast majority of it is immediately bound up by protein

S. We follow this expedient in appropriating the Hockin et al. (1999) model, replacing APC

with APC:PS in all the relevant chemical equations. Doesn’t this substitution require an

alteration of the rate constants? After all, APC:PS is supposed to be more efficacious.

Yes: and here we follow the results of (Egan et al., 1997), who found that the increased

efficacy of the complex results from a factor of three increase in the cleavage rate of Va at

Arg306.4 In the interest of perspicuity, the factor of three and the original rate constant

from (Hockin et al., 1999) are listed explicitly in Table 5.2. Of course, if we wanted to

model (e.g.) protein-S deficiency, we should need to put the equations for (uncomplexed)

APC inactivation of FVa into the model.3For completeness: in fact, we used all of the chemical equations from Bungay et al. (2003) except two:

the deactivation of FVaL by APC; and the unbinding of FVaiL, the lipid-bound, inactivated form of factorVa, from its phospholipid surface—since, after all, FVaiL never shows up in our model, being replaced byvarious cleaved forms of factor Va.

4Previously it had been reported that the factor was twenty rather than three, but as Egan et al. (1997)notes, this would abolish the thrombophilic effect of factor-V Leiden: In the presence of protein C, thecleavage at Arg306 would dominate the cleavage at Arg505, so that the disabling of the latter in FVL wouldhave no overall effect on clotting.

85

Page 104: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Does this complicated replacement for the single equation of (Bungay et al., 2003) yield

similar results under normal (healthy) conditions? In fact, there is a significant change:

under the more complicated model, thrombin peaks during normal clotting around 90 nM,

whereas under Bungay et al. (2003)’s model, it peaked at almost 40 nM (though, curiously,

the time of the peak is the same in both models). Now Bungay et al. (2003) verifies their

model by comparing it to clinical observations of thrombin curves under various physio-

logical conditions, viz. Figure 5 of (Butenas et al., 1999). However, oddly enough, these

conditions assume no protein C or protein S, so Bungay et al. (2003)’s simulations do not

actually verify this part of their model. In point of fact, Bungay et al. (2003)’s “verifi-

cations” are anyway qualitative: the thrombin curves from (Butenas et al., 1999), while

peaking at similar concentrations, are somewhat delayed compared with those of (Bungay

et al., 2003). Resolution of this issue, too, we defer to the discussion section below. For

now it should be said that the more detailed model of (Hockin et al., 1999) is probably

more accurate than the single equation of (Bungay et al., 2003), so we proceed with some

confidence.

The heretofore unexplained abbreviations appearing in the table are as follows. The

numbers 3,5, and 6 denote the site or sites at which the species of factor Va has been cleaved,

respectively Arg306, Arg505, and Arg662. Thus Va53 is factor Va cleaved at both Arg306 and

Arg505; and VaA3 is factor V’s A2 peptide chain, cleaved at Arg306. LC denotes the light

peptide chain of factor Va, and HC its heavy chain; and VaLC is the light chain and A1

peptide chain. The subscripted L again denotes the lipid-bound species.

Table 5.2: Chemical Reaction Set II

No. Reaction kon (nM−1 s−1) koff (s−1) kcat (s−1)

49 LCL + HC VaL 2.63× 10−6 1.72× 10−5 -

50 LCL + HC5 Va5L 2.63× 10−6 1.72× 10−5 -

51 LCL + HC3 Va3L 2.63× 10−6 1.72× 10−5 -

52 LCL + HC53 Va53L 2.63× 10−6 1.72× 10−5 -

53 LCL + HC36 Va36L 2.63× 10−6 1.72× 10−5 -

54 LCL + HC56 Va56L 2.63× 10−6 1.72× 10−5 -

(continued on next page)

86

Page 105: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Table 5.2 – continued

No. Reaction kon (nM−1 s−1) koff (s−1) kcat (s−1)

55 LCL + HC536 Va536L 2.63× 10−6 1.72× 10−5 -

56 LCL + APC:PS L LC:APC:PSL 2.63× 10−6 1.72× 10−5 -

57 VaL + APC:PSL Va:APC:PSL 0.1 0.7 -

58 Va5L + APC:PSL Va5:APC:PSL 0.1 0.7 -

59 Va3L + APC:PSL Va3:APC:PSL 0.1 0.7 -

60 Va53L + APC:PSL Va53:APC:PSL 0.1 0.7 -

61 Va36L + APC:PSL Va36:APC:PSL 0.1 0.7 -

62 Va56L + APC:PSL Va56:APC:PSL 0.1 0.7 -

63 Va536L + APC:PSL Va536:APC:PSL 0.1 0.7 -

64 VaLCL + APC:PSL VaLC:APC:PSL 0.1 0.7 -

65 VaLCL + VaA3 Va3L 2.57× 10−6 0.028 -

66 VaLCL +VaA53 Va53L 2.57× 10−6 0.028 -

67 VaLCL +VaA36 Va36L 2.57× 10−6 0.028 -

68 VaLCL +VaA356 Va536L 2.57× 10−6 0.028 -

69 VaLC:APC:PSL + VaA3 Va3:APC:PSL 2.57× 10−6 0.028 -

70 VaLC:APC:PSL + VaA53 Va53:APC:PSL 2.57× 10−6 0.028 -

71 VaLC:APC:PSL + VaA36 Va36:APC:PSL 2.57× 10−6 0.028 -

72 VaLC:APC:PSL + VaA356 Va536:APC:PSL 2.57× 10−6 0.028 -

73 Va:APC:PSL Va3:APC:PSL - - 3*0.064

74 Va5:APC:PSL Va53:APC:PSL - - 3*0.064

75 Va5:APC:PSL Va56:APC:PSL - - 5.2× 10−4

76 Va3:APC:PSL Va36:APC:PSL - - 5.2× 10−4

77 Va53:APC:PSL Va536:APC:PSL - - 5.2× 10−4

78 Va56:APC:PSL Va536:APC:PSL - - 3*0.064

79 Va:APC:PSL Va5:APC:PSL - - 1

80 Va3:APC:PSL Va53:APC:PSL - - 1

81 Va36:APC:PSL Va536:APC:PSL - - 1

The chemical reactions and rate constants of the Hockin et al. (1999) model of factor-Va deactivation.

Factor-V Leiden is modeled as knocking out the reactions which cleave factor V at the Arg505 site.

Proteins cleaved at this site are marked by a 5. See text for abbreviations.

A word on the lipid-binding: The reader may notice that not all of the proteins sub-

87

Page 106: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

scripted with an L have equations for dissociating from their lipid substrates. In this we

are for the most part following (Bungay et al., 2003), in which only the simple proteins,

but not the compounds, are allowed to dissociate. For the remainder, we have followed

Hockin et al. (1999) in not allowing the uncomplexed proteins LCL, Va5L, Va3L, Va53L,

Va36L, Va56L, Va536L, and VaLCL to dissociate from their substrates. Now, in fact, the

experiments in Hockin et al. (1999) assume a saturating amount of lipids; whereas in our

model, the phospholipid concentration is something less than saturating. Technically, then,

these proteins should be allowed to dissociate. However, assuming on- and off-rates equal

to that for (uncleaved) factor Va, these dissociations have an entirely negligible effect even

on the concentrations of the FVa-variants themselves, let alone on thrombin levels—which

is to be expected, since phospholipids never fall below about 1500 nM throughout the sim-

ulations, at which level binding forces vastly predominate over unbinding forces (cf. the

rate constants of chemical reaction 4). Notice, also, that the heavy-chain variants of factor

Va are not lipid bound: the binding site for lipids lies in the light chain of factor V Hockin

et al. (1999).

Finally, on this head, a correction was required: Bungay et al. (2003) claim that throm-

bin is released from the lipid surface when it is activated. However, conservation of mass

requires these lipid binding sites to be returned to the pool of uncomplexed LBS; thus in

chemical reaction 33 I have added an LBS term that was missing from (Bungay et al., 2003).

Of course, as I have just finished saying, the concentration of binding sites is in excess over

the other proteins, so this addition makes no discernible difference to the simulations.

The form of the heparin equations comes from (Olson, 1988) and appear in Table 5.3.

The abbreviated proteins are heparin, thrombin (activated factor II), and antithrombin;

(IIa-AT) is a stable complex which can no longer dissociate into inhibitor and protease.

Such are the forms of the heparin reactions; however, to my knowledge, the exact values

of the on- and off-rates under physiological conditions have never been determined. Olson

(1988) does, however, give estimates of the ratios of off- to on-rates; the following expedient

was therefore adopted: On-rates were set to an intermediate value within the biologically

normal range for enzyme-substrate reactions (see for example Table 4.4 in (Fersht, 1999),

88

Page 107: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

but also (Bungay et al., 2003)), viz. 0.1 s−1 nM−1. The corresponding off-rates were then

computed by multiplying the aforementioned ratios by the on-rates, i.e. 0.1. Again, this

is represented explicitly in the table. This approximation does not vitiate the theoretical

apparatus, but it does have numerical consequences; we take up this question again in the

discussion below.

The final rate constant is a catalytic rate, and as such is taken directly from (Olson,

1988). However, we also deviate in one other particular from that paper: Whereas Bungay

et al. (2003) model AT binding to IIa as irreversible, with a very slow on-rate, Olson (1988)

considers the reaction to be reversible, with a very high off-to-on ratio. These mechanisms

are certainly not equivalent, and we use the more recent mechanism from (Bungay et al.,

2003). However, in any case, this reaction will be dominated by the much higher affinity

reaction between AT and heparin, the primary path to the Hep:AT:IIa complex, so the

choice has little practical consequence.

Finally: heparin is also thought to facilitate the inactivation of factor Xa, but once

again for simplicity these interactions have been neglected.

Table 5.3: Chemical Reaction Set III

No. Reaction kon (nM−1 s−1) koff (s−1) kcleavage (s−1)

82 Hep + IIa Hep:IIa 0.1 0.1 ∗ 3× 104 -

83 Hep + AT Hep:AT 0.1 0.1 ∗ 2× 102 -

84 Hep:IIa + AT Hep:AT:IIa 0.1 0.1 ∗ 20 -

85 Hep:AT + IIa Hep:AT:IIa 0.1 0.1 ∗ 3× 103 -

86 IIa:AT + AT Hep:AT:IIa 0.1 0.1 ∗ 0.6 -

87 Hep:AT:IIa → IIa-AT - - 5

The chemical reactions of the heparin reactions. The equations were drawn from (Olson, 1988) but

the rate constants were estimated as in the text.

Table 5.4 lists all the non-zero initial concentrations. The concentration of lipid binding

89

Page 108: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

sites and the initial activating amount of tissue factor were chosen to match (Butenas et al.,

1999), as in (Bungay et al., 2003). All other concentrations are normal physiological values.

Table 5.4: Initial Conditions

Species Conc. (nM) Species Conc. (nM) Species Conc. (nM)

Tissue Factor 0.005 Factor VIII 0.7 TFPI 2.5

Factor II 1400 Factor IX 90 Antithrombin 3400

Factor V 20 Factor XI 30 Protein C 60

Factor VII 10 Factor X 170 Protein S 300

Factor VIIa 0.1 Thrombomodulin 1 LBS 3396

The non-zero initial conditions. All other proteins are initialized at zero.

The chemical reactions of Tables 5.1, 5.2, and 5.3 are equivalent to a set of coupled,

nonlinear, ordinary differential equations. For completeness we illustrate the transformation

for the first chemical equation, II + LBS IIL. Then

d[II]dt

= −kon[II][LBS] + koff[IIL]

d[LBS]dt

= −kon[II][LBS] + koff[IIL]

d[IIL]dt

= kon[II][LBS]− koff[IIL],

where square brackets denote (time-varying) concentration. More succinctly, as in Eq. 5.12,

we may write:

d

dt

[II]

[LBS]

[IIL]

=

−1 1

−1 1

1 −1

kon[II][LBS]

koff[IIL]

. (5.18)

Incorporating the other chemical reactions amounts to adding monomials to each ODE as

necessary, and of course adding an ODE for each new chemical introduced; or, alternatively,

to adding new columns to the numerical matrix, C, for each new rate constant, and adding

a new row for each new chemical.

All simulations were performed in Matlab (Mat) using the stiff ODE solver ode15s.

90

Page 109: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Computation of the strict relative degree of the system were also performed in Matlab,

using the symbolic toolbox.

5.5 Results

We first simulate coagulation in a patient with the hypercoagulatory disorder factor-V

Leiden, with and without intervention by the anti-coagulant heparin, as well as normal

(nonpathological) clotting. Initiation of the clotting event is assumed to take place via the

intrinsic pathway and is therefore modeled by initializing tissue factor at a concentration

of 5 pM (following (Bungay et al., 2003)).

Figure 5.2 shows the thrombin profile during normal clotting (dark blue), clotting in a

“patient” with factor-V Leiden (light blue), and that clotting in the same patient but with

the feedback-linearizing controller (green). Exact output tracking has been achieved; how-

ever, several defects are immediately obvious. First, the controller operates continuously,

i.e. updating as often as the numerical simulation of the ODEs does, whereas any practical

controller must be digital. The input, appearing in Figure 5.3, is obviously not feasible for

any imaginable drug-delivery mechanism. In fact, to avoid numerical errors, the input was

clamped at 100 nM/s. (I explain this issue in the discussion section below.) Finally, and

relatedly, a realistic controller will have a maximum and minimum input rate; certainly, the

rate cannot be negative, since this implies withdrawal of heparin from the site of the injury.

The present controller was allowed to do just that.

We address the final defect first. Note, however, that the theory outlined above does

not guarantee the success of any these remedies. As lately noted, the input cannot be a

negative number, nor can it be much higher than a single-digit nanomolarity per second.5

We impose an upper bound (somewhat arbitrarily) at 20 nM/s and a lower bound of zero

by “squashing” inputs outside this range, i.e. setting any u calculated from Eq. 5.10 which5Continuous intravenous infusions over a 24-hour period total between 20,000 and 40,000 IU (International

units), which comes to about 10 pM/s Since our treatment lasts only over the course of a few minutes, wecan presumably use higher doses over this short interval without overmuch hæmophilic risk downstream.

91

Page 110: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

exceeds one of these bounds to the value of the bound. The result appears in Figure 5.4,

and the counterpart input in Figure 5.5.

There is indeed some degradation from the unconstrained controller; in particular, the

controller overshoots the peak trajectory somewhat, and is unable to compensate for the

final undershoot. These defects are respectively the consequences of the maximum and

minimum constraints; but the latter at least might be remedied by some form of anticipation;

that is, if the controller could predict the final undershoot, it might ease off the input earlier.

This possibility is explored in the discussion below. As for the overshoot, it can be eliminated

by raising the maximum input to 50 nM/s (graph not shown). However, in the absence of

an established maximum, it might be safer to try to address this issue with an anticipatory

control as well.

The corresponding input in Figure 5.5 still exhibits some chatter, naturally, since it is

allowed to vary continuously. That shortcoming is rectified by discretizing the controller,

which results in the output and input of Figures 5.6 and 5.7, respectively. Here the out-

put again overshoots the desired thrombin trajectory, but also undershoots the pre-peak

trajectory, evidently because it was unable to recover from the initial input; and, for sim-

ilar reasons, undershoots the post-peak thrombin curve slightly more than its counterpart

continuous controller (Figure 5.4).

A smaller controller step yields, as expected, superior results (not shown); and raising

the input maximum again eliminates the overshoot. In the limit, of course, we should be

able to reproduce the results of the continuous controller. However, and for that very reason,

the smaller the step, the more unrealistic the controller. And then, in any case, we shall

presumably not be able to eliminate the undershoot that the continuous controller exhibits,

since that results from the input minimum; nor, if we take seriously the input maximum of

20 nM/s, shall we be able to eliminate the overshoot.

Whether the controlled trajectory of Figure 5.6 presents a hypercoagulatory risk is

something of an open question: Rapid product formation “downstream” in the cascade—i.e.

the formation of fibrin from fibrinogen and activation of factor XIII—requires concentrations

92

Page 111: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

of thrombin less than 2 nM (Brummel et al., 2002), so it is not clear how important the

exact trajectory of the thrombin curve is. (Nevertheless, it may be important, and so we

propose a more sophisticated approach below.)

We turn to “moderate” hæmophilia A, at which native levels of the zymogen factor VIII

are about 2.5% of normal levels. Now, if indeed we have a controller that can be switched

on precisely at the onset of a clotting event (perhaps by the release of tissue factor), and can

apply the control locally—assumptions under which we have been operating so far—then

feedback linearization is overkill: We can simply dump in the healthy initial concentration

of factor VIII in the first controller time step, and then turn the controller off. Using again

a control sampling rate of 2 Hz, this technique yields the near-perfect results of Figure

5.8. Here the nonzero input at the first time step (not shown) is x(0)/T , where x(0) is the

desired initial concentration of factor VIII, 0.7 nM, and T is the controller time step, 0.5s.

An added boon of this technique is that it requires no sampling at all: the controller would

not require sensors.

Now, the simple, step controller discussed in Section 5.3.2 above operates on similar

principles, although it makes no assumptions as to the controller operating frequency. Ap-

plying the simple learning algorithm of Eq. 5.17 with α = 1 × 10−4 and β = 1 × 10−5

yields the trajectory in Figure 5.9. The constant input (not shown) of factor VIII is the

very low rate of 6.92 pM/s. The price we pay for the simplicity of this control scheme is

the overshoot on the back half of the trajectory.

Finally, we ask how well we can manage factor-V Leiden with the step controller. In

fact, we can do quite well. Choosing α = 0.05 and β = 0.01 in Eq. 5.17, the input will

settle on u = 4.39 nM/s, at which rate the peak concentration and its occurrence can be

very nearly matched (Figure 5.10). The only question that remains is how significant the

undershoot is from a clinical perspective.

93

Page 112: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

5.6 Discussion

5.6.1 Controllability

We have shown via an algebraic criterion that our system of ODEs is not full-state

linearizable. As a matter of fact, we can say more. It follows from the method of trans-

forming chemical equations into ODEs that every reversible reaction will introduce into C

two columns that are additive inverses of each other. Thus, while the rank of C is obviously

upper-bounded by the number of unidirectional reactions—i.e. the number of arrows in all

our chemical equations, or again the number of rate constants—since this is the number

of columns in C, it is also upper-bounded by the (generally smaller) number of uni- or

bidirectional reactions—i.e. counting each pair of arrows as just one reaction; henceforth

simply called “reactions.” If this number falls short of the dimension of the state space, then

full-state linearization requires the deficit to be covered by the control vector fields—which,

in the case of a single-input, can provide additional rank of at most one. So in chemical

systems of mass-action kinetics with no outflow, a single output, and a single input, a nec-

essary (though insufficient) condition for full-state linearization is that there be at least as

many reactions as state variables.

We can translate this result into a statement about the controllability of the system.

The rank of the strong-accessibility distribution (at some point x0) is the dimension of the

locally accessible manifold (from x0); or alternatively, the difference between this rank and

the dimension of the state, n, is the number of uncontrollable modes of the system. So

the dimension of the locally accessible manifold is upper-bounded by the total number of

reactions (again, counting bidirectional reactions just once).

Considering the present system, we find that without the heparin reactions there are

96 reactions and 98 state variables. Assuming factor-V Leiden knocks out some reactions

and some state variables, leaving 77 and 84, respectively. Now, we are obviously interested

in pharmaceutical interventions in addition to heparin (a whole host of pro-coagulants,

for instance). But it can now be claimed that in order to even stand a chance of fully

linearizing the system with one of these interventions as the (sole) input, or equivalently of

94

Page 113: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

fully controlling the system, the drug must interact with the system via at least six more

reactions than the number of variables that these reactions introduce. (We say six because

the control vector field may not be in the span of C and hence cover the remaining dimension;

recall Eq. 5.15.) And indeed, heparin introduces only six reactions for its five new state

variables, yielding 83 and 89, respectively. So even though there are more unidirectional

reactions (rate constants, columns of C) than state variables (proteins, rows of C), we know

that there are uncontrollable modes even without constructing C simply by noting that the

number of reactions falls short of the number of state variables.

Finally on the topic of full-state linearization, it can be shown that outflow of the blood

factors—in particular, heterogeneous outflow—can increase the rank of the accessibility

matrix, and hence of the strong-accessibility distribution (Bastin and Levine, 1993). For

simplicity, the model as it stands neglects outflow, but in fact in vivo coagulation will

perforce have some flow (both in and out) of clotting factors, and if it be not negligible over

the time scale of interest, this may provide for greater controllability of the system.

5.6.2 Non-regularity

The coagulation system under application of heparin is, we have been told, non-regular,

i.e. the relative degree is not constant across the state space. What consequences ensue for

control?

It appears that the technique of allowing the system to drift away from the singularity

has no adverse effects on control. We describe now in detail why this should be so. The

feedback nonlinearity that shows up in the denominator of the control law, Eq. 5.10, is in

the case (q = 2):

LgLfh(x) = −kon[IIa](t), (5.19)

inducing a singularity precisely when concentrations of thrombin (factor IIa) are zero. How-

ever, during a clotting event, thrombin concentrations will always be greater than zero, so

the state is safely bounded away from the singularity. We could make this bound precise

by choosing some δ in thrombin-concentration space below which we don’t care to regulate

95

Page 114: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

the thrombin trajectory. And the simulations have demonstrated that in the concentration

range of interest—viz., on the order of nanomolarity—the state is not near enough to the

singularity to introduce numerical errors into the control.

Now, on the other hand, some (too large) choice of δ would violate the initial-condition

criteria too grossly for the controller to be able to recover. That it does recover at all is

a consequence (presumably) of the error correction terms added to the control law in Eq.

5.16. Again, the simulations have demonstrated that the controller can indeed recover as

late as 60 seconds into the clotting event.

We saw in the Results section above that the unconstrained (continuous) controller led

to numerical errors in the ODE simulation; which is why the data shown in Figures 5.2 and

5.3 were produced with input clamped at 100 nM/s. These errors are indeed the consequence

of the non-regularity of the system, but a peculiar consequence that could not arise in a

realistic controller: Since the heparin input rate was allowed to assume negative values, the

concentration of heparin itself as well as other state variables could be driven below zero.

And of course pushing thrombin levels through zero means operating the controller in the

region of the singularity. Lower-bounding the input at zero, as we must in any realistic

controller and as we do in the other simulations, eliminates this threat to the numerical

stability of the controller.

Are there not more theoretically sound, less heuristical, bases for avoiding the singular-

ity? There are, but these (Tomlin and Sastry, 1997),(Hauser et al., 1989) involve variations

on the following theme: compute the relative degree of the system at the singularity, and use

the corresponding control law (again Eq. 5.10, but with a higher strict relative degree, q) in

the region of the singularity. The problem with this type of technique for our model is that

the relative degree at the singularity is higher than the ability of our machine to compute

iterated Lie derivatives; that is, we are back to the same, apparently insuperable, compu-

tational obstacle that prevented direct computation of the rank of the strong-accessibility

distribution. Fortunately, the simulations presented here evidently demonstrate the super-

fluity of these more sophisticated methods.

96

Page 115: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

5.6.3 Controller merits and demerits

How precise does our control need to be, after all? The answer is not known, but there

is some suggestion that the required amount of thrombin for clotting is much less than

the actual peak concentration: Brummel et al. have demonstrated in an in vitro study

that less than 2 nM of thrombin is required for rapid product formation downstream in

the cascade(Brummel et al., 2002). Certainly this suggests that step inputs suffice to steer

hæmophilia-A and factor-V-Leiden patients through safe clotting, given the qualitative

matches of Figures 5.10 and 5.9. In case these matches are not sufficient, however, or

more critically in case the upper bound on the input needs to be lowered, we propose in

the following section a model-predictive approach to remedy the defects of the feedback-

linearizing controller.

However, the step controllers have another enormous advantage over the feedback-

linearizing control: they require no sampling, only initiation at the inception of a clotting

event. There is not currently a method for measuring blood-protein concentrations in vivo in

real time—certainly not at the required 2 Hz sampling rate. (On the other hand, thrombin

can be measured in real-time ex vivo; see (Hemker et al., 2002).) We interpret our results

on feedback linearization, then, as showing that if indeed a greater degree of accuracy is

required for thrombin tracking, then it would be fruitful to devise the appropriate sensors

for real-time sampling, since they would make possible this suitable technique. Now, naıvely

constraining the input vitiates the accuracy of this technique (Figures 5.4 and 5.6), hence

the need for the model-predictive controller.

On the other hand, two significant advantages of the feedback controllers must be

stressed. First of all, they require no training. Patient-to-patient variation and model errors

(e.g. in rate constants, which vary fairly widely from study to study) make pre-training on

a model insufficient for the development of actual treatments, though it certainly provides

a good starting point. So in the limit, the training of the step-controller approaches the

actual (barbarous) state of the art, namely testing the patient once a day and modifying

doses accordingly. Second, and perhaps more significantly, the feedback controllers operate

97

Page 116: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

fairly robustly over a wide range of delays in turning on. So, e.g., in the case of constrained,

discrete control (the most plausible scenario), turning on the controller a full minute after

the release of tissue factor increases the overshoot by less than 5%, with the rest of the

trajectory tracking essentially identical (data not shown). The issue here is how similar

the state at (e.g.) t = 60s is to the desired initial conditions—the closer, the more easily

the controller can recover—and as is clear from the plots above, thrombin concentrations

(presumably, inter alia) do not increase significantly until about 85 seconds.

One envisages, then, the following real-world scenario in which this type of controller

would be particularly useful: An individual suffering from factor-V Leiden cuts herself,

and—within the first minute—applies to the site of injury a device which injects heparin,

in response to the local concentration of all (or some subset of) the other blood proteins.

5.6.4 Significance of model assumptions

Three model assumptions require elaboration. First, both the feedback and step con-

trollers show some sensitivity to the choice of rate constants. An earlier version of the model,

for instance, estimated both on- and off-rates for the heparin reactions, again setting the

former to 0.1 nM/s but also fixing the latter at 10s−1, and achieved essentially perfect

tracking with the same input limits and sampling rate. As I have said repeatedly, reported

rate constants in the blood clotting literature vary somewhat widely, so this consideration

is quite relevant. However, even if we assume the affinities in (Olson, 1988) are correct, as

in the present model, there is still the matter of our assumption of the on-rates (recall that

the unknown heparin on- and off-rates were estimated from their ratio, fixing the on-rate

in the middle of the reasonable physiological range at 0.1 nM/s and then computing the

off-rate from their known ratio; see Section 5.4).

The effect on the controller turns out to be quite small. Similar tracking results (not

shown) can be achieved by both controllers for rate-constant values which span three orders

of magnitude, with kon from 0.01−1 nM/s and koff set to maintain the desired ratio. Values

98

Page 117: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

outside this range are unlikely, but there the step controller can match the thrombin peak’s

timing and height only on pain of greatly distorting the subsequent trajectory.

What about model errors in the structure, i.e. the chemical reactions, as opposed to the

parameters of the model? Discussion of this aspect is more involved and will be deferred to

the following chapter (and there discharged the remaining promissory note on model errors

issued in the Introduction to this thesis, Chapter 1).

The final model assumption adverted to at the beginning of this section is that of locality:

the supply of unactivated (zymogen) blood factors was treated as limited. This choice was

made to conform to the model of (Bungay et al., 2003) as well as to the in vitro results of

(Butenas et al., 1999). In reality, however, zymogens are replenished by circulating blood,

and activated proteins are similarly removed. On the other hand, the limited amount of

lipids also restricts the amount of each zymogen which can ever take part in the reaction,

since nearly all of the clotting reactions take place on a phospholipid surface. Congruently,

simulations (not shown) indicate that even if the zymogens are modeled as inexhaustible,

the thrombin curve remains qualitatively the same. We proceeded then with the locality

assumption, in order to avoid the complexities of an added circulation model.

5.7 Conclusions

Regulation of many clotting disorders requires daily visits to a doctor both for the ad-

ministration of blood thickeners or thinners, and for assaying blood-protein concentrations

to determine dosages. In exceptional circumstances, continuous intravenous administration

is necessary. Neither of these alternatives is desirable, and both are in some sense a conse-

quence of taking into account only the most primitive knowledge about human coagulation:

certain drugs encourage clotting, while others discourage it.

This chapter has proposed and demonstrated two control techniques that are much more

powerful in virtue of exploiting a mathematical model of a large portion of the coagulation

cascade and the effects of the relevant pharmaceuticals. Both controllers assume local

application and the ability to detect a clotting event (release of tissue factor); under which

99

Page 118: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

assumptions both have been shown capable of regulating the time-varying concentrations

of thrombin, the key blood protein which alone of the proteins in this model determines

clot formation.

Each controller has its merits and demerits: The step controller is more critically depen-

dent on the accuracy of the model but requires no sensors; whereas the feedback controller is

robust to model changes but requires real-time (on the order of 2 Hz) sampling of a number

of protein concentrations (viz., all those involved in the feedback terms). Feedback con-

trol is also (slightly) degraded by constraining the input rate to lie within a feasible range.

We shall address both of these deficiencies of feedback linearization in the next chapter

(although not, unfortunately, at the same time), demonstrating a feedback controller that

requires observation of only one state variable, and a different (model-predictive) controller

which overcomes some of the constraint-induced tracking degradation.

The significance of both the controllers presented in this paper for clinical application

would be greatly advanced by various improvements to the model. First, as lately noted, the

rate constants for heparin interactions with the clotting process are not known; if a precise

controller is to be constructed along the lines proposed here, these must be determined.

Second, a circulation model would dispense with the somewhat dubious locality assumption,

as well as possibly afford more powerful control, since inflow and outflow can increase the

strict relative degree of the system (and hence the number of linearized variables which can

be controlled).

Finally, this chapter has made some useful general observations about chemical kinetics

and controllability. In particular, I have shown that the dimension of the locally accessible

manifold is upper-bounded by the number of reactions in a chemical system, where bi-

directional reactions are counted just once. The author has not seen this result elsewhere.

We can say then, e.g., that since full-state linearization requires that the dimension of the

locally accessibly manifold be equal to the dimension of the state, in chemical systems it

requires that the number of chemicals not exceed the number of reactions.

100

Page 119: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Figure 5.1: A conceptual rendering of feedback linearization. The arrangement can be

viewed either as a controller in a feedback loop with the original control-affine system (solid

lines), plus an output function; or as an equivalent output function along with coupled

linear and nonlinear systems (dashed lines), with the former evolving independently of the

latter, and the output depending solely on the former.

101

Page 120: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

0 50 100 150 200 250 3000

10

20

30

40

50

60

70

80

90continuous control

time (s)

conc

entra

tion

(nM

)

[IIa] Desired[IIa] Uncontrolled[IIa] Controlled

Figure 5.2: Simulated control of thrombin concentration during a clotting event in a patient

with factor-V Leiden. The controller was unconstrained, and was updated continuously,

allowing arbitrarily close tracking of the desired trajectory.

102

Page 121: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

0 50 100 150 200 250 300−80

−60

−40

−20

0

20

40

60

80

100Input

time (s)

conc

entra

tion

rate

(nM

/s)

Figure 5.3: The heparin input in nM/s that generated Figure 5.2. Although this input yields

perfect tracking, it is obviously unacceptable for any realistic drug-delivery mechanism,

operating as it does on a continuous time scale, and using negative inputs (see text).

103

Page 122: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

0 50 100 150 200 250 3000

10

20

30

40

50

60

70

80

90continuous control

time (s)

conc

entra

tion

(nM

)

[IIa] Desired[IIa] Uncontrolled[IIa] Controlled

Figure 5.4: Simulated control of thrombin concentration during a clotting event in a patient

with factor-V Leiden, again continuously sampling. The input was constrained to the range

0-20 nM/s.

104

Page 123: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

0 50 100 150 200 250 3000

2

4

6

8

10

12

14

16

18

20Input

time (s)

conc

entra

tion

rate

(nM

/s)

Figure 5.5: The input generating the output of Figure 5.4. Note that the input is constrained

to the range 0-20 nM/s.

105

Page 124: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

0 50 100 150 200 250 3000

10

20

30

40

50

60

70

80

90naive constraints

time (s)

conc

entra

tion

(nM

)

[IIa] Desired[IIa] Uncontrolled[IIa] Controlled

Figure 5.6: Simulated control of thrombin concentration during a clotting event in a patient

with factor-V Leiden, this time using a discrete controller sampling every 0.5 s. The input

was again constrained in the range between 0 and 20 nM/s.

106

Page 125: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

0 50 100 150 200 250 3000

2

4

6

8

10

12

14

16

18

20Input

time (s)

conc

entra

tion

rate

(nM

/s)

Figure 5.7: The discrete input, constrained to lie between 0 and 20 nM/s, that generated

Figure 5.6.

107

Page 126: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

0 50 100 150 200 250 3000

5

10

15

20

25one−shot control

time (s)

conc

entra

tion

(nM

)

[IIa] Desired[IIa] Uncontrolled[IIa] Controlled

Figure 5.8: Simulated control of thrombin in a patient with moderate hæmophilia A (see

text), where the controller inputs the proper initial concentration of factor VIII at the first

time step, then shuts off.

108

Page 127: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

0 50 100 150 200 250 3000

5

10

15

20

25step control/parameter learning

time (s)

conc

entra

tion

(nM

)

[IIa] Desired[IIa] Uncontrolled[IIa] Controlled

Figure 5.9: Control of thrombin concentration in a hæmophiliac using a single step input

at time t = 0. The input is 6.92 pM/s.

0 50 100 150 200 250 3000

10

20

30

40

50

60

70

80

90step control/parameter learning

time (s)

conc

entra

tion

(nM

)

[IIa] Desired[IIa] Uncontrolled[IIa] Controlled

Figure 5.10: Control of thrombin concentration in a patient with factor-V Leiden by a single

step input at time t = 0. The input is 4.39 nM/s.

109

Page 128: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Chapter 6

Analysis: Stability, Model

Reduction, and Control

6.1 Introduction

Two threads were left dangling at the end of the previous chapter. In the context of the

hypercoagulatory disorder factor-V Leiden, we explored two different types of controllers,

one based on feedback linearization; and the other applying a simple step input, whose

value was learned over repeated trials so as to match the timing and concentration of the

thrombin peak. Each has its advantages and disadvantages (see the Discussion section of

the previous chapter), but neither was able to track perfectly the thrombin trajectory, at

least with the most likely choice of input on-rates (i.e. rates at which heparin binds to

blood proteins), and especially as the constraints on the input were made more stringent.

In this chapter we propose and demonstrate a modification of the feedback-linearization

scheme to overcome this problem. While exploring that terrain we also find a very simple

controller that operates with minimal error, at least when the input upper bound is relaxed,

and moreover obviates the need for observation of almost all of the state variables (all but

one).

The second residual issue is whether the system is globally, exponentially “minimum

110

Page 129: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

phase,” which in turn depends on the stability of the so-called zero dynamics. If this and

another, technical condition are satisfied, then the control law (5.10) guarantees bounded

tracking.1 We shall see in the sequel that this is in fact not the case, and that the success

(such as it was) of our control law, curiously, was not guaranteed—at least not by any

theorem the author knows. Along the way we adduce some nice additional facts about the

controllable subspace of the system which we exploit to effect a model reduction.

6.2 Model Reduction and Bounded Tracking

6.2.1 Bounded tracking and minimum-phase systems

In what follows we rely on (Sastry, 1999). Recall that feedback linearization separates

the system into a set of completely controllable, linear dynamics:

ξ = Aξ + bv,

y = ξ1, (6.1)

on the one hand; and a nonlinear system:

η = λ(ξ, η) (6.2)

on the other. (See again Figure 5.1 for a conceptual rendering.) Now, the nonlinear system

(6.2), when confined to the manifold where the linear dynamics (6.1) are identically zero,

is known as the zero dynamics of the system. That is, if we insist that the output y(t) is

identically zero, which implies that all the linearized states ξi are as well (since ξi is simply

the integral of ξi+1), then the remaining variables evolve according to the so-called zero

dynamics,

η = λ(0, η), (6.3)

on an (n− q)-dimensional manifold, where again q is the relative degree of the system.

It turns out that the stability of an equilibrium point of these dynamics tells us some-

thing about the stability of certain equilibrium points of the original system. In particular,1Recall that “bounded tracking” means that the error between the desired and actual output and its

derivatives tends asymptotically to zero, and that the state is bounded.

111

Page 130: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

suppose x0 is an equilibrium of Eq. 5.3 that also produces zero output, i.e. h(x0) = 0.

Then if we define (ξ0, η0) := Φ(x0), then ξ0 = 0 for any change of coordinates (ξ, η) = Φ(x).

(This can be seen by setting the output to zero in Eq. 6.1.) Thus η0 is an equilibrium of

the zero dynamics, and furthermore by appropriate definition of Φ(x) we make sure η0 = 0.

We are now in a position to state the minimum-phase property of an affine-control

system. The latter is called exponentially minimum phase at x0 if the equilibrium point

η = 0 of the zero dynamics is exponentially stable. This is in turn equivalent to

spec∂λ∂η

(0, 0)< 0, (6.4)

that is the spectrum (eigenvalues) of the zero dynamics at the equilibrium point η = 0 are

all in the left half of the complex plane.

Now, we care about these dynamics because a theorem exists (Sastry, 1999) that the

control law (5.10) of the previous chapter produces bounded tracking, in the sense lately

defined, if the zero dynamics are defined everywhere that our trajectory tracking requires

the state to travel, and if furthermore they are indeed exponentially stable over that same

region—that is, if the affine-control system is exponentially minimum phase.2

6.2.2 Construction of the zero dynamics

So we have at hand a simple mechanical procedure for adducing a sufficient condition for

bounded tracking and furthermore giving us some insight into the dynamics of the system.

In order to apply this procedure, however, we must first construct the nonlinear subsystem

(6.2). In particular, we find n− q solutions to the PDE

(∇xηi)g(x) = 0, (6.5)

and make sure that they are independent of the ξi. Why is this condition necessary? The

nonlinear dynamics of η are defined to be independent of the input, so these functions must

be orthogonal to the input vector field, g(x). Why does this suffice? The Frobenius theorem2The zero dynamics are also required to be Lipschitz continuous, which they are, though we do not

belabor the point here. Briefly: since the state variables are confined to finite quantities, the steepness ofthe function is concomitantly limited.

112

Page 131: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

tells us that a distribution of p vector fields in an n-dimensional space is integrable—i.e.,

admits of n − p solutions η to the above equation—if and only if it is involutive (for the

definition of which, see the footnote on page 76in previous chapter). Now g(x) alone

constitutes a one-dimensional distribution, which is trivially involutive, so we know that

there are indeed n− 1 solutions. We shall pick the n− q of them that are also independent

of the ξi.

It is not normally easy to find solutions to a large PDE of the form (6.5), but the present

situation is an exception. In the case we have been considering at length, i.e. a heparin

input to a patient with factor-V Leiden, the vector field g(x) is a constant vector of all

zeros and a single one at the index of heparin—call it r. That means that the solutions all

have the very simple form∂ηi∂xr

= 0, (6.6)

where xr is heparin. Notice that ηi = xj is a solution to this equation for all j 6= r. Now, to

ensure that (ξ, η) = Φ(x) is a valid change of coordinates (i.e., that ∂Φ/∂x has full rank),

we obviously cannot let ηi = ξi; and in particular since ξ1 is the output we wish to control,

viz. thrombin, we don’t assign thrombin to any of the nonlinear variables either. This

leaves exactly n− 2 variables xi remaining for the variables ηin−2i ; all that remains is to

make sure that they are independent of ξ2 as well.

Thus by construction, the Jacobian ∂Φ/∂x of the change of coordinates is a permuted

identity matrix, except for the second row, corresponding to ∂ξ2/∂x. For the Jacobian to

have full rank, this row must be nonzero at the rth column, where again r is the index

corresponding to heparin. And indeed,

∂ξ2

∂xr= −k[IIa], (6.7)

that is, some rate constant k times the concentration of thrombin. Thus the matrix loses

rank only at the point [IIa] = 0—but that is, unfortunately, precisely the point we care

about, since we are trying to construct the zero dynamics (which means ξi(t) = 0 ∀i, t).

However, we can avoid this singularity in the zero dynamics by shifting the original state

variables. If, that is, xm = [IIa] in the original system, then we define xm = [IIa]− δ, where

113

Page 132: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

δ > 0. Then∂ξ2

∂xr= −k(xm + δ), (6.8)

which is zero only when xm = −δ. The zero dynamics are globally defined, then, which

is a requirement for the bounded-tracking theorem lately reviewed. And the coordinate

change (ξ, η) = Φ(x) is defined almost everywhere: everywhere except along the manifold

where the concentration of thrombin is precisely zero. This irregularity corresponds to

the non-regularity we noted in the last chapter, i.e. the failure of the relative degree to

be constant across the entire state space; it, too, failed precisely at zero concentration of

thrombin. Recall from that chapter that our solution was to apply our control scheme

only after thrombin had drifted some (small) distance δ away from the singularity, so that

our control scheme was unaffected by it. Correspondingly, we only consider the change of

variables Φ(x) away from this manifold. (We might think of Eq. 6.3 as defining the “δ-

dynamics” of the system, then: the evolution of the nonlinear subsystem with the output

we really care about, thrombin, held at a concentration of δ.)

We have one more requirement, stipulated at the outset, and that is to make sure that

the equilibrium point η0 of the zero dynamics is 0. This can be accomplished by shifting the

η variables by the equilibrium point of the original system. (And of course adding constants

to η does not change its status as a solution to Eq. 6.5.)

Following the construction for a change of variables just outlined, then, our nonlinear

dynamics are:

η = λ(ξ, η) = f(Φ−1(ξ, η))− x0, (6.9)

where f(x) is just f(x) without the elements corresponding to heparin and thrombin; likewise

for x0. So we can now ask: are indeed the eigenvalues of the linearization of the zero

dynamics all strictly negative? that is, does Eq. 6.4 hold? Well, we can see without any

calculations that the answer must be “no.” We saw in the previous chapter that we can

express our ODE as:

x = Cr(x) + bu, (6.10)

where rk[b, C] < n. That means that the linearization ∂f/∂x = C∂r/∂x cannot have full

rank—and now some of these singularities carry over directly into ∂λ/∂x = ∂ f/∂x.

114

Page 133: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

6.2.3 Intermezzo: model reduction

Is this the end of the line? No: these singularities are, as we shall see, a consequence

of constraints that in fact allow us to rewrite the dynamics in terms of a reduced system

evolving on a lower-dimensional manifold. We can then re-derive the nonlinear dynamics

(6.2) for this reduced system, which will no longer contain the singularities induced by the

rank deficiency of D = [b, C].

We begin by noting that

rk(D) =: p < n

and therefore

rk(N (DT )) = n− p,

i.e. the nullity (rank of the nullspace) of DT is n − p. Thus, there exist n − p linearly

independent (constant) vectors zin−pi such that

zTi (Cr(x) + bu) = 0

for any u and for all time. Alternatively, we write:

0 = zTi x =d

dt(zTi x),

so that integrating with respect to time we find that zTi x = ci for some constants cin−pi ;

or more specifically:

zTi x = zTi x0. (6.11)

Thus the n−p rank deficiency of the matrix D = [b, C] gives rise to exactly n−p linearly

independent algebraic constraints of the form (6.11)3, where the vectors zi form a basis for

the nullspace of DT . Each constraint can therefore be solved for one of the state variables

and substituted into Eq. 6.10, reducing the dimension of the system to p. Rather than

solve haphazardly for this nullspace4, however, we consider their physical meaning. These

constraints represent the conservation of mass inherent in the original differential equations,

enforced by using mass-action kinetics throughout (recall the discussion in Chapter 2).3In robotics, these constraints are called holonomic.4Matlab provides a command to do this in one step: Z = null(D’).

115

Page 134: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

The idea behind conservation of mass in a chemical system is that the total amount

of (e.g.) factor XI in the system must be constant—in fact, equal to the initial amount

of factor XI—although how much of that protein is (e.g.) in complex with thrombin, or

in its activated form, or etc. may vary. Thus we could write the conservation equations a

priori—in the example just given:

[XI] + [XIa] + [XI:IIa] + [XIa:IX] = [XI]0; (6.12)

however that process is extremely tedious (and error-prone), especially for a very large

system like the present one. We present instead a mechanical procedure for recovering the

equations from D, requiring only an enumeration of the pure (i.e. uncomplexed) proteins.

Without loss of generality, we assume that the first n − p elements of x, i.e.

x1, x2, ..., xn−p, are these uncomplexed proteins. Each of these will show up in exactly

one conservation equation apiece since, after all, each contains only one type of protein.

We define Z = [z1, z2, ..., zn−p], a basis for the nullspace of DT (i.e. the columns of Z are

the vectors zi above), with each column corresponding to a conservation equation. So, for

example, the vector zi corresponding to the conservation equation (6.12) above has ones

in those rows corresponding to XI, XIa, XIa:IIa, and XIa:IX, and zeros elsewhere. In fact,

the only column with a one in the row corresponding to XI is this one: again, XI shows up

in no other conservation equation. Thus, given the ordering lately defined, with the pure

proteins listed first in the state vector, we can partition the matrix Z as:

Z =

In−pZ2

(6.13)

where In−p is an (n−p)-dimensional identity matrix. Partitioning D comformably, we have

DT = [DT1 D

T2 ], (6.14)

so that instead of writing DTZ = 0 we may say DT1 In−p = −DT

2 Z2, or

Z2 = −(DT2 )+DT

1 , (6.15)

with the “+” denoting the pseudoinverse. Now, Eq. 6.15 can then solved in conjunction

116

Page 135: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

with Eq. 6.11 for the uncomplexed proteins, and these equations substituted back into the

final p equations of Eq. 6.10, giving an equivalent, reduced (p-dimensional) system.

For our system, this procedure yields the following conservation equations:

TF:VIIa + TF + TF:VII + TF:VIIa:IX + TF:VIIa:X + TF:VIIa:Xa + TF:VII:Xa

+TFPI:Xa:TF:VIIa = 0.005

IIa:AT + IIL + mIIa + mIIaL + IIa + II + VIII:IIa + Xa:Va:II + Xa:Va:mIIa + VIII:mIIa

+IIa:Tm:PC + mIIa:AT + Hep:IIa + Hep:AT:IIa + V:IIa + XI:IIa + V:mIIa

+IIa:Tm + TAT = 1400

Va36:APC:PS + VaLC + VaLC:APC:PS + LCL + Xa:Va + Va3 + VL + VaL + V

+Va + Xa:Va:II + Xa:Va:mIIa + Va:APC:PS + Va3:APC:PS + LC:APC:PS + V:Xa + V:IIa

+V:mIIa + Va36 = 20

TF:VIIa + VII + VIIL + VIIa + VIIaL + TF:VII + TF:VIIa:IX + TF:VIIa:X

+TF:VIIa:Xa + TF:VII:Xa + VII:Xa + TFPI:Xa:TF:VIIa = 10.1

APC:PS:VIIIa + IXa:VIIIa:X + VIIIaL + VIIIaiL + VIII + VIIIL + VIIIa + VIIIai

+IXa:VIIIa + VIII:Xa + VIII:IIa + VIII:mIIa = 0.7

IXa:VIIIa:X + IXL + IXa + IXaL + IX + TF:VIIa:IX + IXa:VIIIa + IXa:AT + XIa:IX = 90

IXa:VIIIa:X + Xa:Va + XL + XaL + X + Xa + TF:VIIa:X + TF:VIIa:Xa + TF:VII:Xa

+VIII:Xa + Xa:Va:II + Xa:Va:mIIa + TFPI:Xa + V:Xa + VII:Xa

+TFPI:Xa:TF:VIIa + Xa:AT = 170

XIa + XI + XI:IIa + XIa:IX = 30

APC:PS:VIIIa + Va36:APC:PS + VaLC:APC:PS + PCL + APC + APCL + PC + IIa:Tm:PC

+Va:APC:PS + Va3:APC:PS + LC:APC:PS + APC:PS = 60

APC:PS:VIIIa + Va36:APC:PS + VaLC:APC:PS + PSL + PS + Va:APC:PS + Va3:APC:PS

+LC:APC:PS + APC:PS = 300

117

Page 136: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Tm + IIa:Tm:PC + IIa:Tm = 1

IIa:AT + AT + mIIa:AT + Hep:AT:IIa + IXa:AT + Xa:AT + Hep:AT + TAT = 3400

TFPI:Xa + TFPI + TFPI:Xa:TF:VIIa = 2.5

3 APC:PS:VIIIa + 3 VaLC:APC:PS + 3 Va36:APC:PS + PCL + 3 IXa:VIIIa:X + XIa:IX + VIIIaL

+VIIIaiL + TF:VIIa + IIL + mIIaL + VL + VaL + VIIL + VIIaL + VIIIL + IXL + IXaL + XL + XaL

+APCL + PSL + TF:VII + 2 Xa:Va + 2 V:Xa + V:IIa + 2 VII:Xa + 2 TF:VIIa:IX + 2 TF:VIIa:X

+2 TF:VIIa:Xa + 2 TF:VII:Xa + 2 IXa:VIIIa + 2 VIII:Xa + VIII:IIa + 3 Xa:Va:II + 3 Xa:Va:mIIa

+2 VIII:mIIa + IIa:Tm:PC + 3 Va:APC:PS + 3 Va3:APC:PS + 3 LC:APC:PS + 2 APC:PS + 2 V:mIIa

+TFPI:Xa:TF:VIIa + LBS + Va3 + Va36 + VaLC + LCL = 3396

HC3 + HC36 + HC = LCL + LC:APC:PS

VaA3 + VaA36 = VaLC + VaLC:APC:PS

6.2.4 Stability of the zero dynamics

Recall that by construction, each monomial in the original differential equation x =

f(x) + g(x)u is either a single state variable scaled by a rate constant (kxi), or a product of

(different) state variables scaled by a rate constant (kxixj , i 6= j). That is, the dynamics are

multi-affine, meaning that if we hold all the variables constant except one, xi, each equation

is affine in xi. In fact, since the monomials never have degree higher than two, we might call

these equations “bi-affine.” This fact, along with the nature of the conservation equations—

viz., weighted sums of state variables and a constant—implies that our substitutions will

not change the form of the equations: they will still be bi-affine, post-reduction. We can

therefore treat the reduced system exactly like the original system—and in fact, from here

on we shall take Eqs. 6.1 and 6.2 to refer to the linear and nonlinear decompositions

(respectively) of the reduced system. The new system likewise (necessarily) has the same

relative degree, and the same construction described in Section 6.2.2 yields an everywhere-

defined zero dynamics.

118

Page 137: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

So where are the eigenvalues of its linearization? That depends on where we evaluate it.

The trouble is that the complexity of f makes it hard to solve analytically for the equilibrium

point of interest. (We are not interested in the easily calculated equilibrium at x = 0, since

in fact the system will obviously never reach that point from any activated state: at the

very least we know that the irreversible binding of AT with the proteases it inhibits will

produce nonzero quantities of those bound complexes. Nor are we interested in the stable

quiescent state, i.e. the physiological steady state prior to the introduction of tissue factor,

for similar reasons.) We approach the problem numerically, then, evaluating Eq. 6.4 at

points along the controlled trajectory. And indeed, along this trajectory the eigenvalues of

the linearization lie in the strictly negative half of the complex plane. However, as the state

approaches an equilibrium, the rank of the Jacobian collapses, i.e. some of the eigenvalues

migrate rightward to (within the numerical precision of the machine) the jω axis.

We conclude, then, that system is not exponentially minimum phase; the convergence

to the equilibrium of interest is at best asymptotic. That means that we cannot guarantee

the boundedness of our tracking scheme—at least not with the theorem at hand.

6.3 Control: Model-Predictive and Proportional

The feedback controller of the previous chapter suffered from two major infirmities:

(1) the need to observe the state more or less continuously, which in turn requires a great

deal of technology in the way of protein-concentration sensors; and (2) an ad hoc approach

to constraint handling. We address both these problems in this section—although not,

unfortunately, at the same time.

6.3.1 State observability and feedback linearization

In the discussion at the end of the previous chapter, we lamented an unfortunate re-

quirement of the feedback-linearization scheme: it may in general demand observation of

the entire state (see Section 5.6.3). We explore that requirement in more detail presently.

119

Page 138: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

As throughout this chapter, we restrict our focus to the case of treating factor-V Leiden

with heparin.

Now we saw in Chapter 5 that a feedback-control law of the form:

ud =1

LgLq−1f h(x)

(v − Lqfh(x)) (6.16)

could be used to exactly linearize the system—if, that is, we can calculate the Lie derivatives

LgLq−1f =: a(x) and Lqfh(x) =: b(x), where in the present case q = 2. Since these are

functions of x, they require observation of the state to calculate. For practical reasons,

the fewer states the controller must measure, the better. How few can it get away with

observing?

To begin with, a(x) and b(x) are functions, in the present case, of only 32 of the 89

proteins involved in the system (or anyway our model of it)—a reduction of 57 dimensions

for free. As a matter of fact, a(x) is a function of thrombin only—the most useful and most

easily measured protein—so from here on we concern ourselves only with b(x). Now certainly

not all of these 32 proteins contribute significantly to the two terms we are belaboring—so we

now apply a very simplistic heuristic to eliminate the (practically) irrelevant ones: If setting

a state xi = 0 in the equations for a and b does not appreciably change their trajectory,

then we add xi to a list of possibly irrelevant proteins. “Appreciably” is intended here in a

least-squares sense:

tf/T∑m=1

[b(x(mT ))− b(x(mT ))∣∣xi=0

]2 < θ

tf/T∑m=1

b(x(t))2, (6.17)

for some (small) threshold θ. After running through all 32 variables, we check if zeroing all

the elements on the list at once has any appreciable effect on the model trajectories, and

lo! it does not (otherwise we should have been forced to think harder about this heuristic);

this leaves only 21 relevant proteins. Notice, moreover, that we could have chosen smarter

heuristics with a little more effort (and time): perhaps considering (e.g.) the effect on b(x)

of holding xi = ci, where ci is the initial value of xi, or some constant intermediate between

the initial and final values. We content ourselves with showing that at most 21 protein

concentrations need be measured for feedback linearization with heparin, in the case of

factor-V Leiden.

120

Page 139: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Simulation of the control scheme with this “reduced” control law—i.e. Eq. 6.16 but

with the 11 ineffectual variables zeroed—yields, unsurprisingly, identical figures to those in

the previous chapter. Still, we ask: can we do better? As an extreme case, we consider

measuring only the concentration of thrombin, which evidently involves forsaking feedback

linearization (at least for the time being). In particular, we shall restrict ourselves to a

strictly proportional feedback, with the gain term manually adjusted (as usual). That is:

u = Kpe(t) = Kp(yd − y). (6.18)

Quite remarkably, this control law yields excellent results. Compare Figure 6.1 with the

corresponding results for feedback linearization in Figure 5.6: if there is a difference to be

appreciated, it favors the proportional controller!

0 50 100 150 200 250 3000

10

20

30

40

50

60

70

80

90naive constraints

time (s)

conc

entra

tion

(nM

)

[IIa] Desired[IIa] Uncontrolled[IIa] Controlled

Figure 6.1: Simulated proportional control of thrombin concentration during a clotting event

in a patient with factor-V Leiden, once more using a discrete controller sampling every 0.5

s (as in the previous chapter), and with input again confined to the range between 0 and

20 nM/s.

So: at least in our favorite test case, factor-V Leiden under treatment with heparin,

121

Page 140: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

0 50 100 150 200 250 3000

2

4

6

8

10

12

14

16

18

20Input

time (s)

conc

entra

tion

rate

(nM

/s)

Figure 6.2: The discrete input, constrained to lie between 0 and 20 nM/s, that generated

Figure 6.1.

and with “reasonable” constraints on the input rate, a very simple proportional controller

operating at 2 Hz can achieve excellent tracking results.

6.3.2 Control with constraints: model prediction

Here we pursue a different objective from that of the previous section. The controller

exhibited in the last chapter was degraded by the imposition of constraints, so we implement

a model-predictive control scheme which explicitly accounts for these constraints. Such a

scheme is considerably simplified, however, when applied to linear dynamics. Fortunately

we have at hand a technique for rendering the control problem linear, namely: feedback

linearization. So we shall employ a linear model-predictive controller (LMPCer), operating

on the linearized subsystem 6.1. I say this approach is at odds with the attempt to minimize

the number of observed variables, then, because it will require the full power of feedback

linearization.

122

Page 141: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Theory

Our method is adapted closely from (Kurtz and Henson, 1997) and (Henson and Seborg,

1996), so the exposition shall be brief. The basic idea of model-predictive control is to find,

not the “right” control input for the current time step, but the best series of control inputs

u = [u(t), u(t + T ), ..., u(t + (m − 1)T )] for a time horizon of mT seconds, where T is

the sampling interval and m is some integer. As usual, “best” means minimal error in a

least squares sense between the actual and desired outputs. This minimization is especially

felicitous in the (time-discretized) linear, time-invariant setting, where a single, quadratic

cost function can be written for the entire control horizon. Minimizing this cost subject

to the constraints on the input amounts to a quadratic-programming problem (QP). The

control is applied by using only the first element of the control vector u; at the next time

step, the QP is re-solved for a new control vector, this one evidently stretching one step

further than its predecessor.

More explicitly, at each time step i we wish to find:

v∗ = argminv

m−1∑j=1

[ξj − ξd((i+ j)T )]TQ[ξj − ξd((i+ j)T )] + α[vj − vd((i+ j)T )]2

(6.19)

where v = [v1, v2, ..., vm−1], a vector of synthetic inputs for the next m time steps; the

vectors [ξ1, ξ2, ..., ξm−1] are the linearized states for the next m steps that would be produced

by using these inputs; and vd(iT ) and ξd(iT ) are respectively the desired input and state at

time step i. (Those circumflexes are here meant to indicate that these variables aren’t “real,”

i.e. they may never actually be realized in the model.) That is, we penalize deviations from

the desired state with the matrix Q and deviations from the desired input with the scalar

α. (In general, one may penalize fast controller moves, (vj − vj−1)2, as well, but this turns

out not to improve performance, so we leave Eq. 6.19 uncluttered.)

Now, vd is again just the qth derivative of the desired output (where again q = 2 is the

relative degree of the system and hence the dimension of the linearized state), as with vanilla

feedback linearization; and likewise the vector ξd is the desired output itself and its first q−1

derivatives (recall that the linearized system of Eq. 6.1 consists simply of q integrators of

the input). That leaves the anticipated future states ξi: these are transformed, via the well

123

Page 142: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

known solution to the discrete-time state-space equations, into functions of their initial state

(i.e., the state of the linearized system at time iT ), of the input v, and of the (discretized)

state and input matrices, A and b. Using this formula and some algebra, we can transform

Eq. 6.19 into the more useful form5

v∗ = argminv

12vTHv + dTv

, (6.20)

where the matrix H is computed from the parameters Q, A, b, and α; and the vector

d is calculated from these as well as ξd, vd, and the current linearized state ξi. Notice,

incidentally, that d must be recalculated at every step, whereas H need not be.

This accounts for the cost function: now to the constraints. We impose a hard constraint

on the (linearized) state, namely that ξ1 not differ from the desired output by more that

δ at any point. Notice that this is not redundant with the cost function, since its overall

minimization might very well call for a gross output error at some time steps. Finally,

the constraints on the input u must be transformed into constraints on the synthetic input

v, and here a complication arises. At the current (ith) time step the transformation is

straightforward: we simply invert the control law, Eq. 6.16—although since it may swap

the maximum and minimum, we write:

vmin = minu∈umin,umax

ua(x(iT )) + b(x(iT ))

(6.21)

vmax = maxu∈umin,umax

ua(x(iT )) + b(x(iT ))

,

with a and b as defined above. However, the constraint v will (in general) have moved by

the next time step, since a and b will have moved. Kurtz and Henson (1997) suggest two

possible solutions. The first is to pretend that the constraints don’t move, i.e. to use the

present vmin and vmax for the entire control horizon. This turns out, as a matter of fact, not

to work so well. Alternatively, but much more expensively, we can simulate the nonlinear

subsystem over the control horizon, using the previous step’s synthetic input vector v∗—and

this achieves reasonable results, as we shall see.

Recall that the linear subsystem we are trying to control is in controllable canonical form,5Our constant companion Matlab has a built-in function to solve quadratic programming problems with

cost functions of this form.

124

Page 143: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

with the bottom row all zeros. Thus it is marginally stable (in fact, all of its eigenvalues are

located at zero), so to ensure the stability of the controlled system we again add proportional

and derivative terms to yqd in reckoning vd, as in Eq. 5.16.

Finally: if the QP proves infeasible, we scale back the control horizon by one time

step. If no control horizon suffices to render the problem feasible—and this can and does

happen—the input from the previous time step is used.

Results

In all cases we consider a patient with factor-V Leiden being treated by heparin. The

output to be controlled is, as ever, thrombin concentration. The controller operates at

2 Hz for the proportional and feedback-linearized (FL) controllers, and at 0.5 Hz for the

LMPCer. In the previous chapter, we suggested that heparin should not be injected at

a rate “much higher than a single-digit nanomolarity per second,” and then proceeded to

cap it “somewhat arbitrarily” at 20 nM/s. Here we explore what happens if the maximum

input is lowered to 5 nM/s. As before, the controller obviously cannot attain negative values

(recall that this would correspond to removing heparin from the system).6

The control horizon of the linear model-predictive controller was set at m = 40, which

at a 0.5 Hz sampling rate translates into 80 seconds. The state-penalty matrix Q was set

to 10Iq (a (q × q)-dimensional identity matrix), zeroing the cross-time-step penalties; and

the input-penalty scalar α fixed at unity. The maximum allowable output violation (δ) was

capped at 2 nM.

Now compare Figures 6.3, 6.5, and 6.7, of the proportional, FL, and model-predictive

controllers, respectively. Proportional control is evidently completely inadequate. The FL

controller, too, exhibits errors of up to 100%: here the degradation we saw in the previous

chapter (Figure 5.6) is exacerbated by the enforcement of stricter constraints. The linear

model-predictive controller, on the other hand, reduces this error by some 50%. Some6On the other hand, it certainly is possible to manipulate the system simultaneously with both a pro-

and anti-coagulant—and the theory exists for multi-input feedback linearization. This complicates themathematics, but more significantly would seriously complicate the drug delivery.

125

Page 144: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

attempt was made to optimize the performance of this controller, but there is additionally

reason to believe that the tracking could be improved even more by, e.g., increasing the

sampling rate and extending the control horizon. Unfortunately, both of these remedies do

exact a price in terms of computation time.

Finally, we get some idea of how the LMPCer achieved its superior results by comparing

the inputs of Figures 6.6 and 6.8. In particular, the model-predictive controller starts

increasing heparin input around 45 seconds in response to the predicted thrombin excess

some 50 seconds later that it “knows” it will not be able to suppress then, given the

constraints. The FL controller obviously issues no such anticipatory correction.

0 50 100 150 200 250 3000

10

20

30

40

50

60

70

80

90naive constraints

time (s)

conc

entra

tion

(nM

)

[IIa] Desired[IIa] Uncontrolled[IIa] Controlled

Figure 6.3: Simulated proportional control of thrombin during a clotting event in a patient

with factor-V Leiden, again at 2 Hz, but with input constrained to the range 0-5 nM/s.

6.4 Discussion and Conclusions

The goal of this chapter was to set the control techniques of the previous chapter on

a firmer foundation. Toward that end, we set out with the hope of providing a stronger

126

Page 145: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

0 50 100 150 200 250 3000

0.5

1

1.5

2

2.5

3

3.5

4

4.5

5Input

time (s)

conc

entra

tion

rate

(nM

/s)

Figure 6.4: The proportional controller’s discrete input for Figure 6.3, constrained to lie

between 0 and 5 nM/s

theoretical justification for our application of feedback linearization, viz. to demonstrate

that the control law Eq. 6.16, in conjunction with the synthetic input of Eq. 5.16, would

force our system to track the desired trajectory in a bounded way, i.e. with errors asymptot-

ically approaching zero. Attention was restricted throughout the chapter to our workhorse

example: the model of the pathology factor-V Leiden, treated by heparin, and attempting

to force the thrombin concentrations of a clotting event to track the thrombin trajectory

during healthy clotting.

Our attempt to prove the boundedness of tracking was guided by a sufficient condition

in the form of the stability of the zero dynamics of the system. To give these dynamics even

the chance of being stable, we had to remove the “redundant” equations from the system,

i.e. the ones corresponding to algebraic constraints on the state. These turned out to

be precisely the conservation-of-mass equations (or some linear combination thereof) that

inhere in mass-action formulations of the kinetics. We provided a mechanical procedure for

recovering these equations (in their most perspicuous and useful form) from the formulation

127

Page 146: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

0 50 100 150 200 250 3000

10

20

30

40

50

60

70

80

90naive constraints

time (s)

conc

entra

tion

(nM

)

[IIa] Desired[IIa] Uncontrolled[IIa] Controlled

Figure 6.5: Factor-V Leiden, treated by heparin via a feedback-linearization controller

operating at 2 Hz, and within the range 0-5 nM/s.

Eq. 6.10 of the system dynamics and a list of the uncomplexed proteins; and then used

these equations to reduce the system to a structurally identical (i.e. both are “bi-affine”)

one that evolves on a lower-dimensional manifold.

Now, the equilibrium point of interest in this new system is very difficult to solve for,

so a numerical procedure was employed instead, evaluating the zero dynamics along the

controlled trajectory. As the state approaches its equilibrium along this trajectory, the

rank of the Jacobian of the zero dynamics approaches zero (meaning some of its eigenvalues

are approaching the imaginary axis), so we conclude (with some surprise) that the zero

dynamics of this system are not exponentially stable. That means that, at least with the

theorem at hand, we cannot prove the boundedness of tracking (the apparently bounded

results of the previous chapter notwithstanding).

A second deficiency of the control techniques demonstrated in Chapter 5 was their state-

observation requirement; in particular, because current sensor technologies limit our ability

128

Page 147: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

0 50 100 150 200 250 3000

0.5

1

1.5

2

2.5

3

3.5

4

4.5

5Input

time (s)

conc

entra

tion

rate

(nM

/s)

Figure 6.6: The discrete input generated by the feedback-linearization scheme, clamped

between 0 and 5 nM/s, that produced Figure 6.5.

to measure protein concentrations in vivo in real time, we should prefer to rely on as few

state measurements as possible. A heuristic showed that the controller could get away with

observing as few as 29 (out of about 90) of these concentrations—but in fact observation of

the single protein thrombin suffices for proportional control (obviously), and proportional

control was in turn demonstrated to produce excellent tracking—at least as good as its FL

counterpart—under reasonable assumptions on the input bounds (the same assumptions

that were employed in Chapter 5).

This result was also quite surprising (to the author), so it merits a little discussion.

Examining again the equations (5.10) and (5.16) for the control, and considering the present

case of q = 2 for perspicuity, we see that they might be combined as:

u =[Kp

a(x)Kd

a(x)1

a(x)

]yd − ξ1

yd − ξ2

yd − b(x)

(6.22)

with a(x) and b(x) defined as above. Thus we can view control via feedback linearization

129

Page 148: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

0 50 100 150 200 250 3000

10

20

30

40

50

60

70

80

90LMPC

time (s)

conc

entra

tion

(nM

)

[IIa] Desired[IIa] Uncontrolled[IIa] Controlled

Figure 6.7: Control of thrombin using the LMPC scheme in conjunction with feedback

linearization. Again, the uncontrolled trajectory is a result of factor-V Leiden, and the

input rate of heparin is confined to the range 0-5 nM/s. Note the improvement over Figure

6.5.

with the correction terms of Eq. 5.16 as simply a (sophisticated) variant on the generic

proportional-derivative (PD) control. In particular, the controller penalizes deviations in

the output error and its derivatives, but the gains are state-dependent (via a(x)). To see

this more clearly, we present an alternative derivation of this control law.

Suppose we want to design a PDD (proportional, derivative, and second-derivative)

controller for the system. The proportional input is obviously just up := K ′pe, where e =

yd− y is the output error and for some gain K ′p. Now, since derivatives require information

“from the future,” these controllers must in general estimate the error derivatives. However,

suppose instead we derive explicit functions for the derivatives in terms of the state. Then

we can write the derivative feedback as ud := K ′de = K ′d(yd − ξ2), where ξ2 is the name of

the explicit function of the state for the derivative y. Now, if we impose a penalty on the

second derivative of the error along the same lines, we find that yd is a function not just of

130

Page 149: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

0 50 100 150 200 250 3000

0.5

1

1.5

2

2.5

3

3.5

4

4.5

5Input

time (s)

conc

entra

tion

rate

(nM

/s)

Figure 6.8: The discrete input (constrained to 0-5 nN/s) generated by the LMPCer and

producing Figure 6.7. Sampling rate is 0.5 Hz

the state but of the input, as well, so that our equation for the second-derivative feedback

udd looks like:

udd = Kdd[yd − b(x)− a(x)(udd + ud + up)], (6.23)

with the familiar a(x) and b(x) functions from feedback linearization, and where we have

broken the input into its three components. Rearranging terms, we write:

udd =Kdd

1 + aKdd(yd − b)−

aKdd

1 + aKddud −

aKdd

1 + aKddup,

where the dependence of a and b on x has been suppressed for brevity. Solving for u =

udd+ud+up and recalling the definitions for the proportional and derivative feedback yields:

u =Kdd

1 + aKdd(yd − b) +

K ′d1 + aKdd

e+K ′p

1 + aKdde. (6.24)

We might call this the “true control law” for our PDD controller.

Now consider the case where a(x)Kdd 1. Then the true control law reduces to:

u ≈ 1a

(yd − b) +1a

K ′dKdd

e+1a

K ′pKdd

e,

131

Page 150: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

and if we pick K ′d = KddKd and K ′p = KddKp, then this is just the (adjusted) feedback-

linearization control law, Eq. 6.22!

But what if, on the contrary, a(x)Kdd is on the order of 1? This corresponds, once

again, to the regions where the strict relative degree increases, i.e. where a(x) is close to

zero. Take the extreme case, in which a(x)Kdd ≈ 0. Then the true control law 6.24 reduces

instead to:

u ≈ Kdd(yd − b) +K ′de+K ′pe,

that is, a constant gain PD controller, with an additional penalty term. In fine, application

of the true control law in the regime near the singularity eliminates the state-dependence of

the gains. Starting with Eq. 6.24, then, and proceeding with just a modicum of boldness,

we might dare to drop the derivative penalties and retain a proportional controller:

u = K ′pe.

Oddly enough, then, we have inverted our picture: Instead of treating feedback lineariza-

tion as the theoretical impetus for our controller, and subsequently adding proportional and

derivative terms to account for errors, or plant-model mismatch, or whatever; we take the

PDD as primitive and derive its control law, of which the feedback-linearization control law

is a special case, viz. the case when a(x)Kdd 1.

This little excursion suggests that we employ the so-called true control law, Eq. 6.24,

throughout. It also indicates that proportional control will work best when a(x) is quite

small. Returning to the present study, we can now say something more specific. If the

input is heparin and the output thrombin, then a(x) = −kon[IIa], where kon is the rate at

which heparin and thrombin bind. Since this term depends only upon thrombin, then the

concentrations of other proteins cannot affect the proximity to the singularity (where the

relative degree increases), and hence cannot affect the applicability of the approximations

of the true control law that we have lately discussed. That in turn suggests that, no matter

what the disease, if thrombin is to be guided by heparin, proportional control should work

well. Whether this is also true of other drugs turns on their chemical equations, and in

particular how these ramify in the a(x) term.

132

Page 151: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

We proceed to qualify that last assertion: proportional control works well in this setting

provided the input constraints are sufficiently loose. As we have seen, increasing these

constraints degrades both the proportional and simple FL controller; but we have also now

seen that a (linear) model-predictive controller prevents a great deal of this degradation.

It is not, however, perfect (see again Figure 6.7), and here we explore (briefly) one of the

reasons why.

The constraints umax, umin, on the input u map into time-varying constraints vmax(x(t)),

vmin(x(t)) on the synthetic input v, so to compute the latter requires knowledge of the future.

Pretending that vmax(x(t)) (ditto for the minimum) is constant over the control horizon

hopelessly degrades the controller, so alternatively we (following (Kurtz and Henson, 1997))

use the final (m−1) elements of the vector of inputs (call it vi−1) generated at the previous

time step to simulate x ahead over the whole control horizon and then use this hypothetical

state to calculate vmax(x(t)). Only the final (m − 1) (where m is the number of steps in

the control horizon) are used in this simulation because the first element was actually used

at that previous step—whereas, in point of fact, none of these other elements of vi−1 will

actually be used. Moreover, vi−1 was the solution to the QP when a different (previous) set

of constraints was in place, which has the following curious consequence: The vi−1 used in

simulating over the control horizon may translate (via Eq. 6.16) into a (true) input u which

violates the constraints umax, umin—even though at the previous step these same synthetic

inputs translated into an acceptable u. This is possible because the translation has changed,

i.e. our estimates of future a(x) and b(x) have changed, as a consequence of our input at

the previous time step. (Kurtz and Henson (1997) do not seem to have noticed this fact.)

The infeasibility of the last m − 1 elements of vi−1 obviously vitiates the optimality of its

first element, i.e. the synthetic control used at the previous time step. Thus, the more

quickly a(x) and b(x) change, the more the accuracy of the LMPC scheme will be reduced.

Once again, in the present case we can say something more: Since a(x) = −kon[IIa], we

should expect the LMPCer to have difficulty wherever the slope of thrombin is steep—and,

yes, that is exactly where it does go wrong (once more consult Figure 6.7). Notice that the

proportional and FL controllers, on the other hand, do not exhibit tracking error where the

133

Page 152: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

slope increases, but rather at the moment the input reaches its upper bound—and this is

just what we should expect.

134

Page 153: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Chapter 7

Validation/Learning

7.1 Introduction

In the preceding chapters, we have so far demonstrated a methodology for modeling

blood clotting (hybrid Petri nets); a series of control techniques (and related analyses)

for manipulating it; and an analysis of the sensitivity of the model to its parameters.

The control techniques are robust to parameter variations—that is, their applicability is

not predicated on specific values of those parameters—except insofar as those parameters

(including the constraints on the input) degrade the feedback linearization, and even then

model-predictive control was shown to reduce (if not eliminate) this degradation. And

although some of our conclusions about control in Chapters 5 and 6 were contingent on

certain structural facts about the model (e.g., the relative degree of the system, or the

form of certain Lie derivatives), nevertheless the general conclusions about these techniques

depend on facts about the form of the ODEs which will be true in any correct model.

The choice of hybrid systems, too, and specifically hybrid Petri nets, was also justified on

grounds independent of specific parameter values. Only the conclusions on the sensitivity of

the system were tied to getting the rate constants largely right; and even here this analysis

was meant in part precisely to identify how badly errors in these constants, relative to each

other, would impugn the accuracy of the model.

135

Page 154: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

This independence of many of the foregoing analyses from the model parameters and

even from some structural aspects is a two-edged sword, however: The extent to which we

do exploit actual parameter values determines the extent to which we can predict specific

outcomes in, e.g., a clinical setting. The problem is that we don’t fully trust those parame-

ters! On the other hand, the model has been demonstrated to agree, at least qualitatively,

with some clinical results; and furthermore the parameters have some good claim to truth,

being the results of painstaking chemical experiments. The structure we have even more

confidence in. So the question is: how can we take advantage of the model to the extent

that it is indeed accurate, and lean on “external” gold-standard data where they exist and

contradict the model’s predictions?

The method proposed in this chapter, and explored briefly, is to take a machine-learning

approach. The model is to be thought of as a map from the space of initial conditions and

parameters into an output space, in which space different system states (e.g. hypercoag-

ulatory vs. hypocoagulatory vs. normal) are (hopefully) cleanly separated, or clustered.

The goal is to modify the clustering function so as to maximize the probability that two

sets of initial conditions giving rise to the same disease state get mapped into the same

cluster, and that two sets of ICs representing different disease states get mapped into dif-

ferent clusters—all subject to some constraints on the clustering procedure (e.g. the shapes

and number of clusters). Of course, which ICs correspond to diseases is only known for a

few cases; for the others we rely on the model. What we end up with is a clusterer whose

assignments of points to clusters is influenced both by the blood-clotting model and by our

prior knowledge of IC-disease pairs. How little the clusterer leans on the prior information

gives some indication of how useful the model (and the choice of output space) is, at least

for the task at hand: predicting hyper- and hypo- coagulatory states.

It would be nice to dispense with model simulations altogether, if possible. How might

we go straight from initial conditions to disease states, using the clusterer just described?

This would amount to a kind of generalization. It would be even nicer if the generalization

gave us the “distance” between one initial condition and the closest IC in a different output

class. Then we could give (e.g.) the minimal change in ICs to get from an unhealthy class

136

Page 155: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

into a healthy class. All of these considerations together suggest a decision tree (the classic

description is in (Breiman et al., 1984)). At each stage in a decision tree, the clusters in the

current branch are partitioned according to the most informative (in the sense of entropy

reduction) predictor variable. This process is continued either until the partitioning is

finished (each data vector has been assigned to a leaf in the tree) or until the entropy

reduction falls below some threshold.

7.2 Methods

Once again we focus only on the ODE portion of the HS model, but only for the

convenience of doing all the work within Matlab. There is no other obstacle to using the

full model. (As a matter of fact, availing ourselves of the entire model would allow us to

incorporate clotting time and cross-linking percentages into the output space, which should

be quite useful in separating healthy and diseased vectors.)

Only 15 of the 80 or so proteins have nonzero initial conditions, and we do not explore

the possibility of variations in the others. (As far as I know, this never occurs in vivo.)

Even so, the “curse of dimensionality” prevents dense sampling of the input space: m

samples of each of the 15 dimensions requires over a billion samples (i.e. a billion model

runs, each taking around 5 seconds) even for a relatively sparse m = 4. We rely on some

prior knowledge then, to decrease this number: Assuming that each non-standard initial

concentration of some zymogen or other is the result of some independent and rare event,

the chances of more than (say) two of these events happening at once (i.e. in the same

individual) are very small. Therefore, we sample only the pairwise variations in ICs, 5

samples per protein. These are distributed evenly between 2% of normal levels and 100%.

Thus, for every pair of proteins, there are 25 different initial conditions corresponding to

all the combinations of deficiencies of 2%, 26.5%, 51%, 75.5%, and 100%.

We desire an output space that will cleanly separate hyper-, hypo-, and normal

coagulation—where by “cleanly” we mean that near neighbors to each exemplar should

themselves be classified with that exemplar. Once again we turn to thrombin, in particular

137

Page 156: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

to its concentration profiles over the first 300 seconds of a clotting event. To reduce the com-

putational load on the clustering algorithm, we use only average thrombin concentrations

over 30 second intervals; thus the data live in a ten-dimensional space. We then reduce the

dimension further by rotating the data with a singular-value decomposition and throwing

away the dimensions corresponding to small singular values. This amounts to retaining just

two dimensions (with the felicitous consequence of letting us visualize the clusters).

Clustering was performed with the standard k-means algorithm (using Matlab’s im-

plementation), allowing the starting points (“seeds”) to be selected randomly. Now, I said

above that we want to rely on prior “gold standard” data where they exist. This was im-

plemented as follows. Three “observed” vectors were selected: one for normal coagulation

(all proteins at 100% of normal levels), one for severe hæmophilia A (factor VIII at 2% of

normal pre-injury levels, all others at 100%), and one vector for antithrombin deficiency

(AT at 26.5% of normal; all others unchanged). The k-means algorithm was then initialized

with k = 3, i.e. one cluster for each of the three states (too little, too much, and just the

right amount of clotting), and then k increased and the algorithm run again until the three

“known” vectors ended up in different clusters. Obviously, if this had required a k higher

than about 6, this method would have produced uninterpretable results (what do these clus-

ters correspond to?), and we should have been forced to take a different tack. In particular,

we are ready to adjust the output space, if need be: perhaps to consider concentrations of

some other protein, or to use a nonlinear transform on the thrombin concentrations. This

sounds rather ad hoc; a more principled version of this approach is described in Section

7.4.3 below. Nevertheless, in the present case, as we shall see presently, k = 4 sufficed to

separate the three known vectors.

The decision tree was also implemented using native Matlab functions. The probability

distribution over concentrations of the fifteen nonzero proteins was assumed to be uniform

for the purposes of the entropy calculation (we explore alternatives to this below). The full

(unpruned) tree was constructed, with splitting based on Gini’s diversity index (see e.g.

(Breiman et al., 1984)).

138

Page 157: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

7.3 Results

These results, it must be stressed, are preliminary, and are meant primarily to demon-

strate the uses of the techniques just described.

Using k = 3 clusters in the k-means algorithm pushed the hæmophiliac and normal

vectors into the same cluster, which is obviously not acceptable. However, four clusters

suffice to class all three priorly known vectors separately. These data appear in Figure 7.1.

−3500 −3000 −2500 −2000 −1500 −1000 −500 0−500

−400

−300

−200

−100

0

100

200

300

400

500

Figure 7.1: An output space in which hypo- (black x’s), hyper- (light gray +’s), and normal

(gray circles) coagulation are cleanly separated by k-means, with k = 4 The white hexagram,

black square, and white pentagram (resp.) are the known exemplars of the three classes.

See text for an interpretation of the fourth class (gray diamonds).

What are we to make of the fourth class? In all of them thrombomodulin is at or below

139

Page 158: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

25% —and in 90% of these cases, it is at 2%. We conclude that lack of thrombomodulin

induces extreme hypercoagulation, in (literally) a class of its own.

The decision tree shown in Figure 7.2 is unpruned: branches were created until each

leaf corresponded to single class. Left branches correspond to the labels at nodes, i.e. to

concentrations being less than some amount. Notice that, since the input space was sampled

only at pairwise deficiencies of initial conditions, only two left branches are ever required

to reach a leaf. Now, we can see (among other things) that, without knowing anything

about the values of other protein concentrations, the initial concentration of protein C is

most informative; so ceteris paribus if e.g. we can only measure one variable, we ought to

measure protein C. In fact (not shown in figure), having less than 63% of protein C (breaking

at the first node in the tree) makes hypercoagulation 75% likely. (Recall, however, that this

assumes uniform priors over the sampled concentrations.) Even more interestingly, we

can see (at a glance) the implications of certain deficiencies. For example, a patient with

antithrombin deficiency no worse than 38% of normal levels (1317 nM) cannot have healthy

clotting unless native concentrations of factor XI are also below about 87% (26 nM). Finally,

the tree makes evident certain minimal adjustments to move from one class to another, i.e

the fewest number of protein concentrations to change. The hypercoagulatory consequences

of protein-C deficiency, for instance, represented at the bottom of the leftmost main trunk,

can be rectified by decreasing native tissue-factor levels below 40% (0.002 nM) (as long as

protein C concentrations are at least 38%, that is).

7.4 Discussion

7.4.1 Clustering

The results of the clustering algorithm in some sense validate the model: a principled

choice of output space puts our exemplar initial conditions into different clusters, as they

ought to be. Of course, this conclusion depends (weakly) on whether k-means is the appro-

priate clustering algorithm; and this in turn depends on whether variance is the appropriate

140

Page 159: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

statistic to determine cluster scatter. But this is also in some sense the minimal assumption,

i.e. what we should assume in the absence of any other information on cluster scatter.

The validation is also fairly weak, in that it depends on classifying only three exemplars

properly. A more robust validation would include additional IC vectors for each of the

three classes. Knowing other initial-condition vectors that correspond to normal clotting

would be particularly helpful. A stronger validation would also check the boundaries of the

clusters.

Now, the idea was not just to classify the exemplars properly, but to lean on them

where the model failed—which we did, by using them to determine (indirectly) the number

of clusters. A principled approach (outlined below) will measure the trade-off between

these two—obviously, if the cluster choices were dictated entirely by the exemplars, the

model would not have been in any way validated—but the best we can say presently is that

both the exemplars and the model were separately necessary but insufficient to produce the

clusters. So the exemplars contribute to the clusterer—but how much extra information

does the model provide? One kind of answer is that without the model, clustering is not

even possible: the points in input space were selected uniformly, so obviously any set of

clusters would be as good as any other. The three exemplars do give some information,

but there is no reason to believe that vectors corresponding to (e.g.) factor IX deficiency

(hæmophilia B) would be placed in the same cluster as the hæmophilia A exemplar. In

the very least, much more prior information would have to be known (i.e. we should need

many more exemplars). The true test of the information provided by the model would be to

take (non-exemplar) vectors classified as belonging to class A, and then test these vectors

against, say, “hold-out” data.

7.4.2 Decision tree

The decision tree generalizes over the mappings from initial conditions to clusters in

providing summaries of these mappings in the most efficient manner. This is useful for

determining, for example, whether a deficiency of a certain clotting factor is alone enough

141

Page 160: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

to condemn the clotting to an unhealthy cluster. The tree also tells us what factors are

most informative about disease states (assuming that the clustering is more or less correct).

So, for instance, if we had to predict clotting outcome based on only three proteins, we

ought to measure protein C, antithrombin, and thrombomodulin. It is interesting to note

that all three are inhibitors.

I have called the generalizations provided by the tree “summaries” because they do not

contain more information than that in the initial conditions with their corresponding labels,

though they do provide that information in a much more useful form. We could additionally,

however, have used non-uniform priors for the initial concentrations, corresponding to the

observed clinical occurrences of the various deficiencies, in which case the tree would reflect

a compromise between the information given by the model plus the clusterer, on the one

hand, and those priors, on the other. Predictions from this tree might be quite useful in a

clinical setting.

7.4.3 A mathematical take

Some of the methods of this chapter were rather ad hoc. What follows is a more

principled (and technical) exposition follows. The map from disease states d into parameters

θ is taken to be a probabilistic function, P (θ|d), which in general we do not know (but about

which I shall say more shortly). The function P (s|θ), although it is written as a probability,

is a deterministic map given by the model, where s is a vector in some suitably chosen output

space—perhaps, as above, the average thrombin concentrations over 10-second intervals

during the first 300 seconds of a clotting event. The suitability of our choice will turn on

how well points in this space lend themselves to clustering by coagulatory (disease) state.

Finally, a clustering function Pφ(c|s) assigns points in the output space to clusters, based

on some parameter φ—setting, e.g., cluster size, or the cluster means, or etc.

The joint probability is then:

Pφ(c, s, θ, d) = P (d)P (θ|d)P (s|θ)Pφ(c|s), (7.1)

where P (D) is the prior probabilities of diseases. We are concerned, however, with Pφ(c|d),

142

Page 161: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

the probability of assigning a disease to a certain cluster. In fact, we shall choose the

number of clusters equal to the number of disease states, so that the goal will be to pick φ

so as to maximize the probability that, for each disease, all occurrences of that disease (i.e.,

for all corresponding parameters θ) will be mapped to the same cluster. Notice, then, that

from Eq. 7.1 we can write:

Pφ(c|d) =∫θP (θ|d)

∫sP (s|θ)Pφ(c|s)dsdθ, (7.2)

so that the objective is to find:

φ∗ := argmaxφ

∑i

Pφ(C = ci|D = di)

= argmaxφ

∑i

∫θP (θ|di)

∫sP (s|θ)Pφ(ci|s)dsdθ

.

Now, using Bayes’ rule, P (θ|d) = P (d|θ)P (θ)/P (d), and we shall assume P (θ) and P (d)

are uniformly distributed (uninformative priors). P (di|θ), on the other hand, is for certain

“observed” parameters θj (the ones which we know a priori to correspond to certain diseases)

just an indicator function of d. That is, if θj is a known parameter set for disease m, then:

P (di|θj) = 1 i = m

0 i 6= m.(7.3)

For all other θj , P (di|θj) = 1/p, where p is the number of clusters; that is, we assume nothing

about the relationship between those parameters and the disease state. Furthermore, P (s|θ)

is 1 only for the s (call it sθ) corresponding to θ, and zero elsewhere. Thus we can write

the objective function more simply as:

φ∗ := argmaxφ

∑i

∫θP (di|θ)Pφ(ci|sθ)dθ

,

with P (di|θj) as above. So φ∗ is the solution to:

∑i

∫θP (di|θ)

d

dφPφ(ci|sθ)dθ = 0. (7.4)

Solving this maximization problem for φ∗, then, gives “best” classification function

Pφ∗(ci|sθ), i.e. the one that best fits the exemplar data. Of course we shall also have

built into this function whatever assumptions we want to make about the relationship be-

tween the classes and the output space: unimodality, almost certainly, but perhaps even

143

Page 162: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

stronger assumptions. This allows us to be quite flexible, however; for example, we may

pick a distribution with nonzero skewness so that hypocoagulation can capture even the

extreme cases that had to be classed into an outlier class in the analysis above. The most

important constraint on this function is that the derivative in Eq. 7.4 be soluble.

7.5 Conclusions

The main point of this chapter was to present some techniques for using the model of

this dissertation, even in spite of any lack of fidelity, for clinical purposes. The idea was

to use the model in a “shallow” way, that is only to cluster initial conditions into three

(coarse) bins of coagulatory risk. The clustering required some, though very small, post

hoc adjustment in order to correctly classify three known exemplars of these bins: the

number of clusters had to be extended from three to four. The final class was explained as

“ultracoagulatory,” corresponding only to cases of severe (< 30%) deficiency in the inhibitor

thrombomodulin. A decision tree for summarizing the findings was also generated, from

which could be read off the minimal proteins to change in order to move a patient from an

unhealthy class to a healthy one. Such a tree could in practice also incorporate knowledge

of the a priori probabilities of certain protein deficiencies, in which case the nodes of the

tree could be used to determine the most predictive proteins in a clinical setting.

144

Page 163: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Figure 7.2: The decision tree. The class labels are M = medium, L = low, H = high, and

O = outliers for the four classes of coagulation levels.

145

Page 164: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Chapter 8

Conclusions & Future Work

8.1 Introduction

Having come this far, we survey the landscape we have traversed; discuss some terrain I

have skirted but would revisit for a fuller exploration; and (attempt to) descry the direction

of the main paths off toward the horizon.

The goals of this dissertation were:

• to build a complete model of the coagulation cascade, i.e. one embracing all the

relevant details but also open to (facile) incorporation of more accurate parameters

or even structural elements, if necessary;

• to use this model and some tools from control theory to devise methods for control-

ling coagulation, especially in pathological cases, by clever administration of anti- or

procoagulants;

• to say something about the mathematical properties of the system (e.g. sensitivity,

stability, controllability) and their clinical ramifications;

• and to exploit in a principled way our knowledge of coagulation as embodied in the

model, even in spite of its imperfections, to make useful predictions for the clinical

setting.

146

Page 165: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Now, since the “relevant details” include qualitative data and discrete events, the first

goal entailed modeling the cascade as “hybrid,” i.e. consisting of interacting discrete and

continuous state spaces. Simulations of thrombin profiles and clotting times under healthy

and pathological conditions were used to verify qualitatively the fidelity of the model. We

also explored the sensitivity of the model to its parameters.

Although approaching coagulation as a hybrid system allows (more) complete simula-

tions, it limits our analysis and control techniques. However, bracketing out portions of the

model leaves us with a purely continuous system, the control of which suffices (I claim: see

Section 8.3 below for the justification of this maneuver) to control clotting. We saw four

different control techniques applied to the simulation of various diseases in this model, each

appropriate to different conditions; and along the way produced a more compact represen-

tation of the dynamics—which turned out to be at best asymptotically stable. Lastly, with

respect to control, I made some general observations about the controllability of chemical

systems, and formulated a different (more general) derivation of the main control techniques.

Finally, I proposed using the model in conjunction with some machine-learning

techniques—clustering and a decision tree—in order to generate reliable predictions about

more shallow outcomes, viz. whether a set of initial conditions produces hypercoagulatory,

hypocoagulatory, or normal clotting. The main idea was to bias the predictions of the model

using what limited clinical data exist.

Below we discuss the merits and demerits of the results of this thesis in detail.

8.2 The Hybrid-System Approach

I have said repeatedly that the simulations of Chapter 4 demonstrate the viability of

a hybrid-systems model of blood clotting, inasmuch as it produces thrombin profiles and

clotting times congruent with the clinical literature. However, it must be stressed that

congruence is a weak criterion, in particular because of the underdetermination of the

model by clinical data. That means that although I have demonstrated the sufficiency

of such a model insofar as those data are concerned, I have not shown its necessity. Of

147

Page 166: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

course, a hybrid-systems approach can indeed capture information that cannot be forced

into ODEs, but I have not made any (falsifiable) predictions exploiting this aspect of the

model which I can also show cannot be made with an ODE model. The trouble is in large

part the dearth of clinical data. On the other hand, the model is built precisely so as to

easily incorporate clinical hypotheses, so it can be part of the solution to this problem (used

in conjunction with clinical tests). Hybrid-Petri nets are particularly useful in this regard,

being perspicuous representations.

Mention of the software platform the HPNs were built in has thus far been relegated

to a footnote, but a word is due now. The chemical equations of this model give rise

to differential equations that are “stiff,” which roughly corresponds to operating at very

different (many orders of magnitude) time scales; or, alternatively, to large eigenvalues in

the Jacobian of the ODE. Special numerical algorithms are required to avoid solving these

ODEs at the smallest time scale, and again our old friend Matlab has built-in functions

for this purpose (I used ode15s). Visual Object Net++, the HPN platform, does not.

Indeed, the algorithms used to solve stiff ODEs often react very poorly to discontinuities

(ode15s, e.g., is not recommended for use in hybrid models), so the solution to this difficulty

is not a simple matter of building a better numerical solver—nor is it within the scope of

this thesis, but it must be said that it obliges VON++ to run about a hundred times more

slowly than the equivalent simulation in Matlab.

Finally, we briefly explored the sensitivity of the system to its rate constants. These

sensitivities were then used to change the rate constants in a principled manner while

attempting to match the timing results of the PT test. It turned out that matching these

results could be achieved by changing a few (a tenth of the total number of parameters) by

60% or less; but (on the other hand) at the price of intolerably degrading the fidelity of the

original simulations.

148

Page 167: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

8.3 Control

8.3.1 Modeling assumptions

On what grounds can we restrict our control schemes to the continuous subsystem de-

scribed in Chapters 5 and 6 and still claim to be able to control clotting satisfactorily?

This amounts to asking how we can ignore the HPN which models contact-pathway initial-

ization, and the HPN which models the final stages of clotting. As for the former: under

normal circumstances, this portion of the cascade is thought to have limited clinical signif-

icance(Adcock et al., 2002). The latter is certainly significant, so in the control schemes of

this document I have assumed a worst-case requirement of reproducing the exact thrombin

profile of the healthy clotting. However, we should prefer to relax this requirement, as well

as to provide bounds for acceptable behavior of the initializing HPN (how diseased can this

portion be without significantly affecting the downstream portion of the cascade?); we defer

this to the section below on future work.

The control schemes of this thesis assumed the following simplifications. First, the model

captures the dynamics of the cascade in plasma, i.e. ignoring the effects spatial arrangements

on reactions. The ODEs are only an approximation, then, of the true dynamics. Likewise,

I have assumed no inflow or outflow of proteins—which, interestingly, can increase the

controllability of the system; see Chapter 5—which is to some extent justified by the limited

number of lipid binding sites available at the site of injury: even as new proteins flow in, the

number of reactions dwindle as the available binding sites are exhausted. This assumption

is nevertheless not strictly correct.

Finally, I made some assumptions about the anti-coagulant heparin: It was assumed

that the major action of heparin is on thrombin, whereas in reality it acts to some extent

on factor Xa, too. Some of the rate constants for heparin were also estimated, since in the

literature only the ratio of on- to off-rates has been reported.

149

Page 168: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

8.3.2 Conclusions

The motivation behind the control techniques was to improve upon the current state

of the art in clinical treatments, which make use of only the most basic data—roughly,

empirical correlations between administration of certain quantities of anti- and procoagu-

lants, on the one hand, and a reduction in thrombophilia and hæmophilia (resp.), on the

other. These methods demands daily visits to a clinic for blood drawing and testing, as

well as maintenance of a permanent level of some drug in the bloodstream. Both of these

requirements are obviously undesirable.

Four different control techniques were proposed and demonstrated, although two of them

turned out to be instances of a more general control law. In the absence of constraints on the

input, a learned, step-input controller could enforce reproduction of the healthy thrombin

profile in the cases of hæmophilia A and of factor-V Leiden with total errors in the region

of 5%. These controllers require no observation of the state, but their clinical application

is predicated on the (perfect) fidelity of the model.

On the other hand, and again in the absence of constraints, a feedback controller based

on feedback linearization as well as a simple proportional controller achieved the same

performance without any “training” requirement, and without being tied to the specific

rate constants of the model. In fact, the FL controller requires only that the model be

control-affine (and in fact even this requirement might be dropped (Henson and Seborg,

1996)), and the proportional controller’s requirements are even weaker.

Interestingly, we can view these two controllers as implementations of the same control

law, Eq. 6.24, operating at different regimes, dictated by proximity to the singularity where

the relative degree of the system changes. I have given a more general formulation of this

control law, deriving it via design of a PD controller. This derivation also makes it possible

to see, inter alia, feedback linearization as implementing a kind of PD controller with

state-dependent gains. Moreover, the success of the proportional controller can be seen, in

light of this control law, as a consequence of the proximity of the state to the singularity

throughout the control interval. This proximity would be the same in our model for any

150

Page 169: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

control of thrombin with heparin, suggesting that proportional control should be effective

for any diseases employing such treatment.

Is the bounded reproduction of the thrombin trajectory guaranteed? In real life,

plant/model mismatch and sensor noise make exact tracking impossible, so a guarantee

of bounded tracking is highly desirable. However, the answer is: not by the usual theorem,

which requires exponential stability of the so-called zero dynamics (the evolution of the

nonlinear subsystem with the feedback-linearized system held at zero). These dynamics are

at best asymptotically stable, at least near the equilibrium point of interest. In proving this,

I also provided a mechanical procedure for recovering the equations for mass conservation,

inherent in mass-action kinetics, from the differential equations for the system (provided

they are written in the form Eq. 5.12), and along with it a method for reducing the system

to a more compact form.

Finally, on this head, we considered how realistic constraints on the rate of heparin

injection might degrade the previous controllers. For factor-V Leiden, tight enough con-

straints do indeed degrade the FL and proportional controllers—but the addition of (linear)

model-predictive control overcame much of this degradation. It is not perfect, however, and

I have pointed out that the errors arise (at least in part) when the controller’s prediction

of future constraints on the “synthetic” input (the input to the linearized subsystem) goes

bad—and this happens whenever the Lie derivatives used to translate from real inputs to

synthetic inputs change fast.

A final word on control: How controllable is the system in general, i.e. using any input

(rather than just heparin)? I have shown that the dimension of the locally accessible mani-

fold is upper-bounded by the number of reactions in a chemical system, where bi-directional

reactions are counted just once. This allows one to state limits on the controllability of the

system without even constructing the system of ODEs; or, alternatively, to make demands

on the number of reactions and chemicals a therapeutic intervention must introduce if

complete control of the system is to be possible.

151

Page 170: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

8.4 Learning/Classification

A certain tension has characterized my claims about the clinical relevance of the model

of this dissertation. On the one hand, I have stressed that the current state of the art

in disease prediction and treatment rely on only the crudest knowledge—usually corre-

lations between pre-injury blood-factor concentrations (what I have been calling “initial

conditions”) and disease states or health risks. On the other hand, I have said that the

predictions made by my model (as well as others) are underdetermined by clinical data. So

the solution to the primitiveness of the medical professions’s predictions would appear to

be more refined predictions based on the use of models; but the more refined predictions

offered by a model cannot be verified, on account of the exiguousness of the clinical data!

This problem can only be solved by performing in silico experiments in conjunction with

their in vivo and in vitro counterparts—and this will take time, as well perhaps as the

development of better sensor technologies. In order to use the model as it stands, then, 1

certain compromises must be made. In particular, I suggested using the model to make the

“shallow” prediction of whether a certain initial state would correspond to a disease, but

also biasing the classification with the clinical data (“exemplar vectors” and their labels) I

have just animadverted upon.

There is evidently no hope of clustering (e.g.) hæmophilia in the input space, at least

not without many more exemplar vectors, since deficiencies of (to stick with the example)

factor VIII (hæmophilia A) cannot be generalized to deficiencies of factor IX (hæmophilia

B) by any clustering algorithm. It is the model that gives us that information—as well,

presumably, as mappings that are not known to clinicians, in contrast to the one just given.

The fact that clustering in this output space, then, separates the exemplar, is some kind of

vindication of the model with respect to this shallow criterion; even more so is its ability to

do so with very little ad hoc gerrymandering. Correct classification of three exemplar initial

condition vectors (one for each class) required only that the number of clusters be increased

from three to four. A more complete validation is proposed in the following section.1I have to stress again that the applicability of the control techniques, and the related analyses, do not

depend on the model being perfect; they require only that it not be radically wrong.

152

Page 171: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

The importance of the decision tree as it stands is its ability to summarize efficiently our

observations about the predictiveness of certain alterations in initial conditions. Its ultimate

usefulness, however, lies in its ability to incorporate knowledge about the prior probability

of protein deficiencies into these observations. So the original HPN incorporated as much

qualitative and quantitative knowledge as we have about the mechanisms of blood clotting

into a single model; the clustering analysis allowed for incorporation of prior knowledge

about correlations between initial blood-factor deficiencies and disease states; and the deci-

sion tree enabled us to integrate knowledge about the likelihood of certain deficiencies. This

last model is only intended for what I have called “shallow” predictions, but these are nev-

ertheless useful. So, for example, if a patient has a protein-C deficiency, having only about

30 to 65% of normal pre-injury concentrations, the decision tree tells us that a reduction in

the concentrations of any of factors II, V, VIII, IX, X, XI, VIIa, or of tissue factor (consult

Figure 7.2 for the precise values of these reductions) will suffice to bring clotting into the

normal regime. Perhaps this is cheaper or easier to implement than increasing the amount

of protein C. The tree also shows the interesting result that the initial concentrations of the

three major inhibitors—protein C, thrombomodulin, and antithrombin—are the best risk

predictors—at least assuming uniform priors on all the deficiencies.

8.5 Future Work

The model is not perfect but is easily modifiable and tested. Refinement of the model

in conjunction with clinicians is therefore an obvious next step. The discrepancy between

these simulations (and those of other models!) with the PT test must also be rectified.

Thirdly, given the difficulties with stiff equations just described, the model should also be

migrated over to another platform—like PIPE (Bonet et al., 2007)—where a different ODE

solver can be implemented.

In order to relax the constraints on the output, we might prove what range of thrombin

trajectories result in healthy clotting. This amounts to a reachability analysis on the HPN

for the final portion of the system. Likewise, determination of the range of initial conditions

153

Page 172: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

which generate a sufficiently small amount of factor XI to prevent perturbation of the con-

tinuous subsystem amounts to a reachability analysis on the initializing HPN. Reachability

on hybrid systems is computationally intensive, and in fact prohibitively so for dimensions

greater than five. Thus we might alternatively discretize these HPNs and perform a discrete

reachability analysis on them. Reachability is a solved problem for a large class of discrete

Petri nets (Bause and Kritzinger, 1996); and in fact one of the major requirements is that

the capacity of the places in the network be bounded, which conservation of mass ensures

in the present case. Relatedly, we performed our sensitivity analysis only on the ODEs,

but a similar analysis could be performed on these HPNs and hence the whole system.

(A technique for sensitivity analysis of hybrid systems with a differential-algebraic-discrete

structure is described in (Hiskens and Pai, 2000).)

In Chapter 3 we discussed a control scheme for the ODEs (Belta et al., 2002) that would

benefit from such a reachability analysis on the final (Fibrin) HPN; we return to it now in

some detail. The idea, recall, was to steer the trajectory of interest by specifying conditions

on the system of ODEs, x = f(x) + g(x)u, only at the vertices of some polytope. If these

ODEs are multi-affine in the state—and the ODEs of our model are—then specifications at

the vertices are sufficient to ensure that the trajectory exit through the desired facet of the

polytope.

The way to visualize this control scheme is to imagine a box in three-dimensional space

(imagining 80-dimensional space being out of the question), at every corner of which is

attached the base of an arrow that points in the direction of the flow of the vector fields

f(x) + g(x)u and whose length is proportional to the magnitude of the flow. We want to

set these arrows—using the only means we have, i.e. adjusting u—so that they point away

from all the faces of the box except one (for concreteness imagine it to be the “roof” of the

box). In order to achieve this we need only find controls that satisfy these requirements

on the vector fields at the corners of the box (that is the interesting result of (Belta et al.,

2002)). In the present case, the idea would be to steer the state trajectory through that

facet (hyperplane) perpendicular to the thrombin axis at which thrombin achieves some

minimum required concentration—the location of which would come from the reachability

154

Page 173: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

analysis. This would guarantee (assuming no deficiencies of blood factors in the Fibrin

HPN) that a blood clot form—eventually (see below). Similarly, after crossing the relevant

thresholds, thrombin levels could be driven back down through the same facet using the

same control scheme (though on a different, adjacent, polytope).

At first blush it may seem quixotic to effect this change in the vector fields via manip-

ulation of u—since u, being a single input, can change only a single vector field (namely

that of heparin, or whatever pharmaceutical we’re using). However, our control problem

is unconstrained in other aspects, and hence (perhaps) more easily satisfied. In particular,

most of the supporting hyperplanes of our polytope will be more or less arbitrary: we don’t

care about the concentrations of any factors but thrombin. So we can make our box as big

(or small) as need be till the vector fields at its corners satisfy the requirements. Of course,

this isn’t quite true since the box must always live in the positive orthant—blood factors

cannot take on negative values. But this is a help rather than a hindrance, since (by the

same token) vector fields lying along the face dividing two orthants cannot point toward

the negative orthant; and so at the corners of any polytope contiguous with these faces, the

vector fields must point in the correct direction with respect to that face, viz. inside the

box. (This is of course no guarantee that the other components of the vector field point in

the correct direction; the idea is just that, though we can’t extend the box any further in

that direction, we shall never need to.)

One significant advantage of this technique over the feedback controllers of this disser-

tation is that it does not require (the currently infeasible) real-time access to blood-factor

concentrations, except for that of thrombin—and this only in the case that we need to switch

control schemes in the middle of the clotting process, e.g. if we’re trying to prevent both

hyper- and hypocoagulation (a somewhat unlikely prospect). In general, the constraints

can be computed beforehand, and we need observe nothing.

One very serious disadvantage—though perhaps one that can be remedied—is that

this scheme affords no control over temporal aspects. In fact, that is why we have such

flexibility constructing our polytope: we can afford to be prodigal in our box size because

155

Page 174: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

we are ignoring how long the state trajectory meanders around in the state space. It is for

this reason that this scheme has not been employed in the present document.

A second very serious problem is computational load. Since the vector fields are in

general functions of the input u, the inequalities derived for each corner of the polytope will

impose constraints on that input. In particular, the constraints take the form of n systems

of 2n linear inequalities, each in m unknowns, where m is the number of inputs. At n = 80,

we have the remarkable number of (approximately) 1024 constraints, for each of 80 systems.

Luckily, for each system we can throw away a constraint as soon as we have considered it,

i.e. we keep only the strongest constraint. That means we shall never have to store more

than 76 constraint equations. Nevertheless, although this eliminates the issue of memory, it

is clear that 1022 is a show-stopping number for computation time. Some cleverness will be

required to avoid carrying out all of these computations. It may, e.g., be possible to assume

some of the constraints a priori, by (again e.g.) exploiting the mass-conserving properties

of the system.

I have only provided a preliminary investigation of the use of clustering and classification

with the model, so there is much to be done in this regard. The most useful contribution

would be cross-validation of the clustering. That is, as a true test of the clustering algorithm

and the model’s ability to classify according to the three clotting states listed above, a larger

set of exemplars should be assembled (this could be done)—perhaps one exemplar for each

of the known hyper- and hypocoagulatory diseases (assuming the disease is a matter of

initial conditions; cf. factor-V Leiden). This set should then be split into validation and

training data, the former used to form the clusters—using perhaps a constrained-clustering

algorithm: see e.g. (Davidson and Ravi, 2005)—and the latter used to determine how well

the clusters perform on unseen data.

The easiest and most obvious next step with the decision tree is to assign priors to the

distributions over factor deficiencies. In fact, these priors would also improve sampling of

the model, which could then be done in true Monte Carlo fashion, rather than the uniform,

deterministic sampling of Chapter 7.

156

Page 175: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Finally, an attempt should be made to control the system using the more general control

law derived in Chapter 6. This law might also be incorporated into a model-predictive

controller.

157

Page 176: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

Bibliography

K. Crookston, “Pathology outlines: Coagulation,” PathologyOutlines.com. [Online].Available: http://pathologyoutlines.com/coagulation.html

M. W. King, “The medical biochemistry page,” Indiana University School of Medicine.[Online]. Available: http://themedicalbiochemistrypage.org/blood-coagulation.html

T. Halkier, Mechanisms in blood coagulation, Fibrinolysis, and the Complement System.Cambridge, England: Cambridge University Press, 1991, translation from Danish intoEnglish by Paul Woolley.

S. Bungay, P. Gentry, and R. Gentry, “A mathematical model of lipid mediated thrombingeneration,” Mathematical Medicine and Biology, vol. 20, pp. 105–129, 2003.

D. Goltzman, “Approach to hypercalcemia, hhm/pthrp, and treatment,” in Diseases ofBone and Calcium Metabolism, A. Arnold, Ed. MDText.com, Inc, 2005, ch. 4. [Online].Available: http://www.endotext.org/parathyroid/parathyroid4/parathyroidframe4.htm

R. Røjkjær and A. Schmaier, “Activation of the plasma kallikrein/kinin system on endothe-lial cell membranes,” Immunopharmacology, vol. 43, pp. 109–114, 1999.

Z. Shariat-Madar, F. Mahdi, and A. Schmaier, “Assembly and activation of the plasmakallikrein system: a new interpretation,” International Immunopharmacology, vol. 2, pp.1841–1849, 2002.

I. H. Viera Stvrtinova, J. Jakubovsky, Inflammation and Fever. Academic ElectronicPress, 1995. [Online]. Available: http://nic.sav.sk/logos/books/scientific/

M. A. Panteleev, M. V. Ovanesov, D. A. Kireev, A. M. Shibeko, E. I. Sinauridze, N. M.Ananyeva, A. A. Butylin, E. L. Saenko, and F. I. Ataullakhanov, “Spatial propagationand localization of blood coagulation are regulated by intrinsic and protein c pathways,respectively,” Biophysical Journal, vol. 90, pp. 1489–1500, March 2006.

D. Luan, M. Zai, and J. D. Varner, “Computationally derived points of fragility of a humancascade are consistent with current therapeutic strategies,” PLoS Computational Biology,vol. 3, pp. 1347–1359, July 2007.

F. Rosner, Medicine in the Bible and the Talmud: Selections from Classical Jewish Sources.KTAV Publishing House, Inc., 1995.

S. Butenas, T. Orfeo, M. Gissel, K. Brummel, and K. Mann, “The significance of circulatingfactor ixa in blood,” The Journal of Biological Chemistry, vol. 279, no. 22, pp. 22 875–22 882, 2004.

158

Page 177: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

M. Panteleev, V. Zarnitsina, and F. Ataullakhanov, “Tissue factor pathway inhibitor - apossible mechanism of action,” European Journal of Biochemistry, vol. 269, pp. 2016–2031, 2002.

T. Adams, S. Everse, and K. Mann, “Predicting the pharmacology of thrombin inhibitors,”Journal of Thrombosis and Hæmostatis, vol. 1, pp. 1024–1027, 2002.

Y. Qiao, C. Xu, Y. Zeng, X. Xu, H. Zhao, and H. Xu, “The kinetic model and simulation ofblood coagulation–the kinetic influence of activated protein c,” Medical Engineering andPhysics, vol. 26, pp. 341–347, 2004.

C. Xu, Y. Zeng, and H. Gregersen, “Dynamics model of the role of platelets in the bloodcoagulation system,” Medical Engineering and Physics, vol. 24, pp. 587–593, 2002.

R. Leipold, T. Bozarth, A. Racanelli, and I. Dicker, “Mathematical model of serine proteaseinhibition in the tissue factor pathway to thrombin,” The Journal ob Biological Chemistry,vol. 270, no. 43, pp. 25 383–25 387, 1995.

A. Kogan, D. Kardakov, and M. Khanin, “Analysis of the activated partial thromboplastintime test using mathematical modeling,” Thrombosis Research, vol. 101, pp. 299–310,2001.

A. Mohan, “Modeling the growth and dissolution of clots in flowing blood,” Ph.D. disser-tation, Texas A & M University, 2005.

W. Mounts and M. Leibman, “Qualitative modeling of normal blood coagulation and itspathological states using stochastic activity networks,” International Journal of BiologicalMacromolecules, vol. 20, pp. 265–281, 1997.

J. Signorini and P. Gruessay, “Object-oriented specification of complex bio-computing pro-cesses: a case study of a network of proteolytic enzymes,” in Biologically Inspired Ap-proaches to Advanced Information Technology, ser. Lecture Notes in Computer Science,A. Ijspeert, M. Murata, and N. Wakamiya, Eds., no. XIV. Lausanne, Switzerland:Springer-Verlag, Jan. 2004, pp. 1–12.

J. Lygeros, S. Sastry, and C. Tomlin, “The art of hybrid systems,” July 2001, textbookon hybrid systems, forthcoming. [Online]. Available: http://robotics.eecs.berkeley.edu/∼sastry/ee291e/book.pdf

J. Lygeros, “Lecture notes on hybrid systems,” Feb. 2004, for the Departmentof Electrical and Computer Engineering, University of Patras. [Online]. Available:http://robotics.eecs.berkeley.edu/∼sastry/ee291e/lygeros.pdf

S. Engell, S. Kowalewski, C. Schulz, and O. Stursberg, “Continuous-discrete interactionsin chemical processing plants,” Proceedings of the IEEE, vol. 88, no. 7, pp. 1050–1068,2000.

O. Czogalla, R. Hoyer, and U. Jumar, “Modelling and simulation of controlled road traffic,”in Modelling, Analysis, and Design of Hybrid Systems, ser. Lecture Notes in Control andInformation Sciences, S. Engell, G. Frehse, and E. Schnieder, Eds. Springer-Verlag, 2002,vol. 279, pp. 419–435.

159

Page 178: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

R. Horowitz and P. Varaiya, “Control of an automated highway system,” Proceedings of theIEEE, vol. 88, no. 7, pp. 913–925, 2000.

M. Oishi, I. Mitchell, A. Bayen, C. Tomlin, and A. Degani, “Hybrid verification of an inter-face for an automatic landing,” in Proceedings of the 41st IEEE Conference on Decisionand Control, Las Vegas, NV, Dec. 2002.

R. Fierro, A. Das, V. Kumar, and J. Ostrowski, “Hybrid control of formations of robots,”in Proceedings of the 2001 IEEE Conference on Robotics and Automation, vol. 1, Seoul,Korea, May 2001, pp. 157–162.

T. Schlegl, M. Buss, and G. Schmidt, “Hybrid control of multi-fingered dexterous robotichands,” in Modelling, Analysis, and Design of Hybrid Systems, ser. Lecture Notes inControl and Information Sciences, S. Engell, G. Frehse, and E. Schnieder, Eds. Springer-Verlag, 2002, vol. 279, pp. 437–465.

S. Pettersson and B. Lennarton, “Stability analysis of hybrid systems a gear-box applica-tion,” in Nonlinear and Hybrid Systems in Automotive Control, S. Engell, G. Frehse, andE. Schnieder, Eds. Springer-Verlag, 2003, ch. 17, pp. 373–389.

P. Antsaklis and X. Koutsoukos, “Hybrid systems: review and recent progress,” in Software-Enable Control: Information Technologies for Dynamic Systems, T. Samad and G. Balas,Eds. IEEE Press, 2003, pp. 273–298.

S. Neuendorffer, “Modelling real-world control systems: beyond hybrid systems,” in Pro-ceedings of the 2004 Winter Simulation Conference, R. Ingalls, M. Rossetti, J. Smith,and B. Peters, Eds. Washington DC, USA: IEEE Press, Dec. 2004.

R. Ghosh and C. Tomlin, “Lateral inhibition through delta-notch signaling: a piece-wise affine hybrid model,” in Proceedings of the 4th International Workshop on HybridSystems: Computation and Control, ser. Lecture Notes in Computer Science, M. DiBenedetto and A. Sangliovanni-Vincentelli, Eds., vol. 2034, no. XIV. Rome, Italy:Springer-Verlag, Mar. 2001, pp. 232–246.

M. Chen and R. Hofestadt, “Quantitative Petri net model of gene regulated metabolicnetworks in the cell,” In Silico Biology, vol. 3, no. 3, pp. 347–365, 2003.

H. Matsuno, A. Doi, N. M, and S. Miyano, “Hybrid petri net representation of gene regu-latory network,” in Proceedings of the Pacific Symposium on Biocomputing (PSB), ser.Lecture Notes in Computer Science, R. Altman, K. Dunker, L. Hunter, T. Klein, andK. Lauderdale, Eds., vol. 5. Honolulu, Hawaii: World Scientific Press, Jan. 2000, pp.338–349.

H. Matsuno, Y. Tanaka, H. Aoshima, A. Doi, M. Matsui, and S. Miyano, “Biopathwaysrepresentation and simulation on hybrid functional petri net,” In Silico Biology, vol. 3,no. 3, pp. 389–404, 2003.

H. Jong, J. Gouze, C. Hernandez, M. Page, T. Sari, and J. Geiselmann, “Hybrid modelingand simulation of genetic regulatory networks: a qualitative approach,” in Proceedingsof the 6th International Workshop on Hybrid Systems: Computation and Control, ser.Lecture Notes in Computer Science, A. Pnueli and O. Maler, Eds., vol. 2623. Prague,Czech Republic: Springer-Verlag, Apr. 2003, pp. 267–282.

160

Page 179: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

S. Kowalewski, “Introduction to the analysis of hybrid systems,” in Modelling, Analysis,and Design of Hybrid Systems, ser. Lecture Notes in Control and Information Sciences,S. Engell, G. Frehse, and E. Schnieder, Eds. Springer-Verlag, 2002, vol. 279, pp. 153–171.

T. Henzinger, P. Kopke, A. Puri, and P. Varaiya, “What’s decidable about hybrid au-tomata?” Journal of Computer and Systems Sciences, vol. 57, no. 1, pp. 94–124, 1998.

R. Alur et al., “Charon,” Department of Computer and Information Science, Universityof Pennsylvania. [Online]. Available: http://www.cis.upenn.edu/mobies/charon/

A. Chutinan et al., “Checkmate,” Department of Electrical and Computer Engineering,Carnegie Mellon University. [Online]. Available: Homepage:http://www.ece.cmu.edu/∼webk/checkmate

T. Dang and O. Maler, “d/dt - reachability analysis of continuous and hybrid systems,”VERIMAG. [Online]. Available: http://www-verimag.imag.fr/∼tdang/ddt.html

T. Henzinger, P. Ho, and H. Wong-Toi, “Hytech,” Department of EECS, Universityof California, Berkeley. [Online]. Available: http://www-cad.eecs.berkeley.edu/∼tah/HyTech/

S. Miyano et al., “Genomic object net,” Human Genome Center, Miyano Laboratory.[Online]. Available: http://www.genomicobject.net/

E. Lee et al., “Hyvisual,” Department of EECS, University of California, Berkeley.[Online]. Available: http://ptolemy.eecs.berkeley.edu/hyvisual/

“Matlab,” The MathWorks, Inc. [Online]. Available: http://www.mathworks.com

C. Daws et al., “Kronos,” VERIMAG. [Online]. Available: http://www-verimag.imag.fr/TEMPORISE/kronos/

L. de Moura et al., “Symbolic analysis laboratory (sal),” SRI International. [Online].Available: http://sal.csl.sri.com/

A. Deshpande et al., “Shift,” California PATH. [Online]. Available: http://www-path.eecs.berkeley.edu/shift/

R. Drath, “Description of hybrid systems by modified petri nets,” in Modelling, Analysis,and Design of Hybrid Systems, ser. Lecture Notes in Control and Information Sciences(LNCIS). Springer-Verlag, July 2002, vol. 279, pp. 15–36.

K. Larsen et al., “Uppaal,” Uppsala University, Sweden and Aalborg University, Denmark.[Online]. Available: http://www.uppaal.com/

C. Belta, L. Habets, and V. Kumar, “Control of multi-affine systems on rectangles with ap-plications to hybrid biomolecular networks,” in Proceedings of the 41st IEEE Conferenceon Decision and Control 2002, vol. 1. Las Vegas, Nevada: IEEE, December 2002, pp.534–539.

C. Belta, P. Finin, L. C. Habets, Adam M. Halasz, M. Imielinski, R. V. Kumar, andH. Rubin, “Understanding the bacterial stringent response using reachability analysis ofhybrid systems,” in Hybrid Systems: Computation and Control, Proceedings of the 7thInternational Workshop. Springer-Verlag, 2004, pp. 111–125.

161

Page 180: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

J. Hu, W.-C. Wu, and S. Sastry, “Modeling subtilin production in bacillus subtilis usingstochastic hybrid systems,” in Hybrid Systems: Computation and Control, 7th Interna-tional Workshop, University of Pennsylvania, Philadelphia, March 25-27, 2004, Proceed-ings, ser. Lecture Notes in Computer Science, R. Alur and G. Pappas, Eds., vol. 2993.Springer-Verlag, 2004, pp. 417–431.

K. R. Joshi, N. Neogi, and W. H. S, “Dynamic partitioning of large discrete event biologicalsystems for hybrid simulation and analysis,” in in Proceedings of the 7th InternationalWorkshop on Hybrid Systems: Computation and Control (HSCC 2004), 2004, pp. 463–476.

B. Pohl, C. Beringer, M. Bomhard, and F. Keller, “The quick machine-a mathematicalmodel for the extrinsic activation of coagulation,” Hæmostasis, vol. 24, pp. 325–337,1994.

J. Makin, “A computational model of human blood clotting,” International ComputerScience Institute, 1947 Center Street; Berkeley, CA; 94703, Tech. Rep., Dec. 2008.[Online]. Available: http://www.icsi.berkeley.edu/pubs/techreports/tr-08-009.pdf

L. Muszbek, V. Yee, and Z. Hevessy, “Blood coagulation factor xiii: structure and function,”Thrombosis Research, vol. 94, pp. 271–305, 1999.

C. Greenberg, C. Miraglia, F. Rickels, and M. Shuman, “Cleavage of blood coagulationfactor xiii and fibrinogen by thrombin during in vitro clotting,” The Journal of ClinicalInvestigation, vol. 75, pp. 1463–1470, May 1985.

K. Standeven, R. Ariens, and P. Grant, “The molecular physiology and pathology of fibrinstructure/function,” Blood Reviews, 2005.

K. Lewis, D. Teller, J. Fry, G. Lasser, and P. Bishop, “Crosslinking kinetics of the humantransglutaminase, factor xiii[a2], acting on fibrin gels and γ-chain peptides,” Biochem-istry, vol. 36, pp. 995–1002, 1997.

K. Brummel, S. Paradis, S. Butenas, and K. Mann, “Thrombin functions during tissuefactor-induced blood coagulation,” Blood, vol. 100, no. 1, pp. 148–152, 2002.

M. Mosesson, “Fibrinogen functions and fibrin assembly,” Fibrinolysis & Proteolysis, vol. 14,no. 2, pp. 182–186, 2000.

A. Undas, K. Brummel, J. Musial, K. Mann, and A. Szczeklik, “PlA2 polymorphism of β3

integrins is associated with enchanced thrombin generation and impaired antithromboticaction of aspirin at the site of microvascular injury,” Circulation, vol. 104, pp. 2666–2672,2001.

C. van’t Veer, M. Kalafatis, R. Bertina, P. Simioni, and K. Mann, “Increased tissue factor-intiated prothrombin activitation as a result of the Arg506 → Gln mutation in factorVLEIDEN ,” The Journal of Biological Chemistry, vol. 272, no. 33, pp. 20 721–20 729,1997.

M. A. Khanin, D. V. Rakov1, and A. E. Kogan, “Mathematical model for the blood coag-ulation prothrombin time test,” Thrombosis Research, vol. 89, pp. 227–232, 1998.

D. Adcock, R. Jensen, and C. Johns, Coagulation Handbook. Esoterix Coagulation, 2002.

162

Page 181: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

F. Bause and P. Kritzinger, Stochastic Petri nets: an introduction to the theory. Vieweg-Verlag, 1996.

I. A. Hiskens and M. Pai, “Trajectory sensitivity analysis of hybrid systems,” IEEE Trans-actions on Circuits and Systems, vol. 47, no. 2, pp. 204–220, 2000.

K. Mann, K. Brummel, and S. Butenas, “What is all that thrombin for?” Journal ofThrombosis and Hæmostasis, vol. 1, pp. 1504–1514, 2003.

W. B. Blair, “Series solution to the general linear time varying system,” IEEE Transactionson Automatic Control, vol. 16, pp. 210–211, April 1971.

R. Wagenvoord, P. Hemker, and H. Hemker, “The limits of simulation of the clottingsystem,” Journal of Thrombosis and Hæmostatis, vol. 4, pp. 1331–1338, 2006.

I. Mitchell, A. M. Bayen, and C. J. Tomlin, “Validating a Hamilton-Jacobi approximation tohybrid system reachable sets,” in Hybrid Systems: Computation and Control, 4th Interna-tional Workshop, HSCC 2001, Rome, Italy, March 28-30, 2001, Proceedings, ser. LectureNotes in Computer Science, M. D. D. Benedetto and A. L. Sangiovanni-Vincentelli, Eds.,vol. 2034. Springer, 2001, pp. 418–432.

M. F. Hockin, K. M. Cawthern, M. Kalafatis, and K. G. Mann, “A model describingthe inactivation of factor va by apc: Bond cleavage, fragment dissociation, and productinhibition,” Biochemistry, vol. 38, pp. 6918–6934, 1999.

S. S. Sastry, Nonlinear Systems: Analysis, Stability, and Control. New York: Springer-Verlag, 1999.

G. Bastin and J. Levine, “On state accessibility in reaction systems,” IEEE Transactionson Automatic Control, vol. 38, no. 5, pp. 733–742, May 1993.

M. A. Henson and D. E. Seborg, “Feedback linearizing control,” in Nonlinear Process Con-trol, D. E. Seborg and M. A. Henson, Eds. Upper Saddle River, NJ: Prentice Hall, 1996,pp. 149–231.

S. T. Olson, “Transient kinetics of heparin-catalyzed protease inactivation by antithrombinIII,” The Journal of Biological Chemistry, vol. 263, no. 4, pp. 1698–1708, 1988.

J. O. Egan, M. Kalafatis, and K. G. Mann, “The effect of Arg306 → Ala and Arg506 → Glnsubstitutions in the inactivation of recombinant human factor va by activated protein cand protein s,” Protein Science, vol. 6, pp. 2016–2027, 1997.

S. Butenas, C. van’t Veer, and K. G. Mann, “‘Normal’ thrombin generation,” Blood, vol. 94,pp. 2169–2178, 1999.

A. Fersht, Structure and Mechanism in Protein Science. New York: W.H. Freeman andCompany, 1999.

C. J. Tomlin and S. S. Sastry, “Switching through singularities,” in Proc. 36th Conf. onDecision and Control, 1997, pp. 1–6.

J. Hauser, S. Sastry, and P. Kokotovic, “Nonlinear control via approximate input-outputlinearization: the ball and beam example,” in Proceedings of the 28th conference ondecision and control, Tampa, Florida, December 1989.

163

Page 182: A Computational Model of Human Blood Clotting: · PDF fileA Computational Model of Human Blood Clotting: Simulation, Analysis, Control, and Validation Joseph Gerard Makin Electrical

H. Hemker, P.Giesen, R. AlDieri1, V. Regnault, E. de Smed, R. Wagenvoord1, T. Lecompte,and S. Beguin, “The calibrated automated thrombogram(cat): a universal routine test forhyper- and hypocoagulability,” Pathophysiology of Haemostasis and Thrombosis, vol. 32,pp. 249–253, 2002.

M. J. Kurtz and M. A. Henson, “Input-output linearizing control of constrained nonlinearprocesses,” Journal of Process Control, vol. 7, pp. 3–17, February 1997.

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

P. Bonet, C. M. Llado, and R. Puigjaner, “Pipe v2.5: a petri net tool for performancemodeling,” in Proceedings of the 23rd Latin American Conference on Informatics, SanJose, Costa Rica, October 2007.

I. Davidson and S. S. Ravi, “Clustering with constraints: Feasibility issues and the k-meansalgorithm,” in Proc. 5th SIAM Data Mining Conference, 2005. [Online]. Available:http://www.cs.albany.edu/∼davidson/

164