12.07.2015 Views

A Clockwork Kidney - University of Melbourne

A Clockwork Kidney - University of Melbourne

A Clockwork Kidney - University of Melbourne

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

A C L O C KW O R K K I D N E YRO B E RT M O S SUsing hierarchical dynamical networks to modelemergent dynamics in the kidneySeptember 2008Department <strong>of</strong> Computer Science and S<strong>of</strong>tware EngineeringSchool <strong>of</strong> Engineering, The <strong>University</strong> <strong>of</strong> <strong>Melbourne</strong>Submitted in total fulfilment <strong>of</strong> the requirements<strong>of</strong> the degree <strong>of</strong> Doctor <strong>of</strong> Philosophy


Robert Moss. A <strong>Clockwork</strong> <strong>Kidney</strong>: Using hierarchical dynamicalnetworks to model emergent dynamics in the kidney© September 2008S U P E RV I S O R S:Edmund KazmierczakMichael KirleyPeter HarrisL O C AT I O N:<strong>Melbourne</strong>T I M E F R A M E:September 2008


The pure and simple truth is rarely pureand never simple.— Oscar WildeSee everything; overlook a great deal;correct a little.— Pope John XXIIIAll models are wrong. Some models are useful.— George E.P. BoxDedicated to my parents, for so many reasons.


simulation <strong>of</strong> whole-kidney function from the dynamics <strong>of</strong> individualnephrons is tractable. Furthermore, the work provides a basis forpredicting emergent effects <strong>of</strong> localised renal disease. With the continueddevelopment <strong>of</strong> this model, we hope that significant insightwill be gained into the onset, progression and treatment <strong>of</strong> renaldiseases.vi


D E C L A R AT I O NThis is to certify that(i) the thesis comprises only my original work towards the PhDexcept where indicated in the Preface;(ii) due acknowledgement has been made in the text to all othermaterial used;(iii) the thesis is less than 100,000 words in length, exclusive <strong>of</strong>tables, maps, bibliographies and appendices.<strong>Melbourne</strong>, September 2008Robert Moss


It is true that we live in a complex worldand strive to solve inherently complex problems,which <strong>of</strong>ten do require complex mechanisms.However, this should not diminish our desirefor elegant solutions, which convinceby their clarity and effectiveness.— Niklaus Wirth, 1984 Turing Award LectureA C K N O W L E D G E M E N T SI thank my supervisors Ed Kazmierczak and Mike Kirley, for theirboundless enthusiasm and support, their ideas and suggestions, andfor being so generous with their time. Making the difficult seem easyand the impossible merely difficult, I learnt everything I know <strong>of</strong> theworld <strong>of</strong> research under their tutelage.I owe a massive debt to Peter Harris, for his exceptional generosityand his readiness to share his expertise in a field to which I was anewcomer. My naive questions were always answered quickly andthoroughly, and his guidance was essential in ensuring that my workwas kept relevant to the world <strong>of</strong> renal modelling.Don Marsh and Randy Thomas have been very supportive <strong>of</strong> myefforts, and have provided both great assistance and invaluable feedbackon a frequent basis. I sincerely thank them, especially for all <strong>of</strong>their time that was spent patiently correcting my many beginners’mistakes.I discussed many <strong>of</strong> my ideas and problems with Gavin Baker andTariq Mahmood, receiving countless gems <strong>of</strong> advice and many helpfulsuggestions. Paul Fraser was always on hand to commiserate andempathise about the rigours <strong>of</strong> being a postgraduate student, and tocongratulate the hard-won successes.While conducting my research, I was financially supported by anAustralian Postgraduate Award (APA), funded by the Department <strong>of</strong>Education, Science and Training. I was also awarded the StawellScholarship by the School <strong>of</strong> Engineering to prepare a paper forpublication after the submission <strong>of</strong> this thesis.Finally, I thank my parents, who have been supportive not justthroughout my candidature, but who have also encouraged me froman early age to develop an interest in research. They are, almostcertainly, the only people who will read the published thesis in itsentirety.ix


C O N T E N T SI I N T RO D U C T I O N 11 I N T RO D U C T I O N 31.1 Aim 41.2 Overview 5II BAC K G RO U N D 72 T H E K I D N E Y 92.1 Physiology <strong>of</strong> the kidney 92.2 Renal modelling 122.3 Disease in the kidney 142.3.1 Symptoms <strong>of</strong> renal disease 142.3.2 The economic impact <strong>of</strong> renal disease 162.4 Modelling renal disease 162.5 Modelling the renal physiology as a network 173 N E T W O R K M O D E L S 193.1 Graph automata models 193.2 Complex network models 213.3 Hierarchy in real-world networks 233.4 Disturbed networks and disease 25III M O D E L S A N D E X P E R I M E N T S 274 M O D E L D E S I G N 294.1 Modelling approach 294.1.1 The effects <strong>of</strong> a discrete time model 304.1.2 Analysis <strong>of</strong> discrete network automata 324.2 Model structure 334.2.1 The single-nephron model 334.2.2 The multi-nephron model 344.2.3 The multi-column model 364.3 Network properties 384.4 State space equations and update rules 404.4.1 Choosing an appropriate time-step size 404.4.2 Discrete fluid flow and solute transport 414.4.3 Model input 434.4.4 State variables 434.4.5 State update rules 444.5 Summary 535 M O D E L DY NA M I C S 555.1 The single-nephron model 565.2 Two-nephron systems 625.3 An inhomogeneous 8-nephron system 675.4 A 72-nephron multi-column model 735.5 Summary <strong>of</strong> experiments 746 E M E R G E N T E FF E C T S O F L O C A L I S E D I M PA I R M E N T 796.1 Incorporating impairments into the model 796.1.1 Stochastic impairments 806.1.2 Stochastic spread <strong>of</strong> impairments 806.1.3 Deterministic impairments 81xi


xiiC O N T E N T S6.2 Choice <strong>of</strong> impairment 816.2.1 Implementation <strong>of</strong> the impairment 826.3 Measuring the effects <strong>of</strong> impairments 826.4 Impairments in an 8-nephron system 836.4.1 A single impaired nephron 846.4.2 Two impaired nephrons 886.4.3 Summary <strong>of</strong> experiments 926.5 Impairments on a larger scale 936.5.1 Impaired nephron pairs 946.5.2 Impaired columns 996.5.3 Summary <strong>of</strong> experiments 1066.6 Conclusions about system stability 1077 L A R G E-SCALE S I M U L AT I O N 1097.1 Introduction 1097.2 Performance <strong>of</strong> the implementation 1107.2.1 The single-threaded algorithm 1107.2.2 Performance analysis 1117.3 Improving performance with concurrency 1147.3.1 The concurrent algorithm 1147.3.2 Communication in the concurrent algorithm 1177.3.3 Performance <strong>of</strong> the implementation 1197.4 Automatic generation <strong>of</strong> large models 1227.4.1 Stochastic generation <strong>of</strong> large models 1227.4.2 Model growth as a rewriting system 123IV S U M M A RY O F R E S E A R C H 1278 D I S C U S S I O N A N D C O N C L U S I O N 1298.1 Contributions 1298.2 Limitations and further work 1308.2.1 Improved understanding <strong>of</strong> whole-kidney function 1308.2.2 Modelling <strong>of</strong> renal disease 1318.2.3 Pluggable virtual kidney 1318.2.4 Modelling kidney failure with reliability factors 1328.3 Conclusion 132B I B L I O G R A P H Y 133


L I S T O F F I G U R E SFigure 1 Physiology <strong>of</strong> the kidney. 9Figure 2 The functional segments <strong>of</strong> the nephron. 10Figure 3 The inverse-sigmoid signal generated by the TGFmechanism. 11Figure 4 An example <strong>of</strong> a 1D binary cellular automaton. 19Figure 5 The beehive rule. 20Figure 6 A logic gate implemented with the beehive rule. 21Figure 7 The iterative construction <strong>of</strong> a hierarchical network.24Figure 8 Choosing an appropriate sampling rate. 31Figure 9 The single-nephron model. 34Figure 10 The single-nephron tubule as a 1D network. 34Figure 11 Solute transport in the loop <strong>of</strong> Henle. 35Figure 12 The two-nephron network. 35Figure 13 A single column <strong>of</strong> interstitial fluid. 36Figure 14 The arterial tree for a multi-nephron model. 37Figure 15 The arterial tree for multiple columns <strong>of</strong> interstitialfluid. 37Figure 16 Diffusion between columns <strong>of</strong> interstitial fluid. 38Figure 17 Evolution <strong>of</strong> small-world network properties I. 39Figure 18 Evolution <strong>of</strong> small-world network properties II. 40Figure 19 Discrete fluid flow along tubule segments. 41Figure 20 Fluid flow into a sub-divided tubule segment. 42Figure 21 Solute transfer in the loop <strong>of</strong> Henle. 48Figure 22 Oscillations in medullary sodium concentration. 50Figure 23 Regular oscillations in single-nephron glomerularfiltration rate (SNGFR). 57Figure 24 Construction and maintenance <strong>of</strong> a medullarysodium gradient. 58Figure 25 The single-nephron dynamics in response to decreasedblood sodium concentration (k = 0.3 andP A = 110). 60Figure 26 The change in sodium concentration in the top(outer medulla) and bottom (inner medulla) layers<strong>of</strong> medullary fluid, in response to decreasedblood sodium concentration. 60Figure 27 The single-nephron dynamics in response to increasedblood sodium concentration (k = 0.3 andP A = 110). 61Figure 28 The change in sodium concentration in the top(outer medulla) and bottom (inner medulla) layers<strong>of</strong> medullary fluid, in response to increasedblood sodium concentration. 61Figure 29 Mean difference in filtration rates in a two-nephronsystem, over the parameter space (P A , k). 62Figure 30 Lyapunov stability over the parameter space (P A ,k). 63xiii


xivList <strong>of</strong> FiguresFigure 31 The change in SNGFR in response to changes inP A . 64Figure 32 The effect <strong>of</strong> varying the vascular coupling betweentwo identical nephrons I. 65Figure 33 The effect <strong>of</strong> varying the vascular coupling betweentwo identical nephrons II. 65Figure 34 The effect <strong>of</strong> varying the vascular coupling betweentwo inhomogeneous nephrons. 66Figure 35 The medullary sodium gradient in a two-nephronsystem when blood sodium concentration remainsat 135–145 mmol/L. 68Figure 36 The medullary sodium gradient in a two-nephronsystem when blood sodium concentration progressivelydrops to 110–120 mmol/L. 68Figure 37 The medullary sodium gradient in a two-nephronsystem when blood sodium concentration progressivelyrises to 160–170 mmol/L. 68Figure 38 Structure <strong>of</strong> the 8-nephron model. 69Figure 39 The medullary salt gradient in an eight-nephronsystem when blood sodium concentration is altered.70Figure 40 The salt concentration in the inner and outermedulla <strong>of</strong> an eight-nephron system when bloodsodium concentration is altered. 70Figure 41 The whole-system filtration rate (WSFR) <strong>of</strong> aneight-nephron system when blood sodium concentrationis altered. 70Figure 42 The WSFR <strong>of</strong> eight-nephron systems with differentlevels <strong>of</strong> vascular signalling. 71Figure 43 The WSFR power spectral density <strong>of</strong> eight-nephronsystems with different levels <strong>of</strong> vascular signalling. 72Figure 44 The high-level structure <strong>of</strong> the 72-nephron model. 73Figure 45 The WSFR <strong>of</strong> the 72-nephron system when γ =0. 75Figure 46 The individual whole-column filtration rates (WCFRs)in the 72-nephron system when γ = 0. 75Figure 47 Power spectral densities <strong>of</strong> the WSFR <strong>of</strong> the 72-nephron system for different values <strong>of</strong> γ. 76Figure 48 Power spectral densities <strong>of</strong> the WCFRs <strong>of</strong> the 72-nephron system for different values <strong>of</strong> γ. 77Figure 49 Average WCFRs and WSFRs <strong>of</strong> the 72-nephron systemfor different values <strong>of</strong> γ. 77Figure 50 The loops <strong>of</strong> Henle in the eight-nephron system.Each nephron is identified by a number from 1to 8, and the length <strong>of</strong> the loops are rated on ascale <strong>of</strong> 0 to 3. 83Figure 51 Variations in the WSFR <strong>of</strong> an eight-nephron system,in response to single impaired nephrons <strong>of</strong>different loop lengths. 85


List <strong>of</strong> FiguresxvFigure 52 Comparisons <strong>of</strong> WSFR between the original (unimpaired)eight-nephron system and the eight-nephronsystem where nephron 7 is impaired and D P =0.5. 86Figure 53 The power spectral densities <strong>of</strong> SNGFR in theeight-nephron system with a single impaired nephron. 87Figure 54 The difference in power spectral densities <strong>of</strong>SNGFR between the original eight-nephron systemand the system with one impaired nephron. 87Figure 55 Variations in the WSFR <strong>of</strong> an eight-nephron system,in response to pairs <strong>of</strong> impaired nephrons<strong>of</strong> different loop lengths. 89Figure 56 Comparisons <strong>of</strong> WSFR between the original eightnephronsystem and the eight-nephron systemwhere both longest-looped nephrons are impairedand D P = 0.5. 90Figure 57 The power spectral densities <strong>of</strong> SNGFR in theeight-nephron system where both longest-loopnephrons are impaired. 91Figure 58 The difference in power spectral densities <strong>of</strong>SNGFR between the original eight-nephron systemand the system with two impaired nephrons. 91Figure 59 The high-level structure <strong>of</strong> the 72-nephron model. 93Figure 60 Variations in the WSFR in the 72-nephron systemwhere all nephrons <strong>of</strong> a specific loop length areimpaired. 95Figure 61 Comparison <strong>of</strong> the WSFR in the original 72-nephronsystem to that <strong>of</strong> the 72-nephron system whereall nephrons <strong>of</strong> a specific loop length are impaired.96Figure 62 A comparison <strong>of</strong> the power spectral densities <strong>of</strong>WCFR between the original 72-nephron systemand the system where all nephrons <strong>of</strong> a specificloop length are impaired. 97Figure 63 The difference in power spectral densities <strong>of</strong>WCFR between the original 72-nephron systemand the system where all nephrons <strong>of</strong> a specificloop length are impaired. 98Figure 64 Variations in WCFR in the 72-nephron systemwhere all nephrons <strong>of</strong> loop length 3 are impaired. 99Figure 65 Variations in the WSFR in the 72-nephron simulationswhere all <strong>of</strong> the nephrons in a particularcolumn (x, y) are impaired. 100Figure 66 Comparison <strong>of</strong> the WSFR in the original 72-nephronsystem and the system where all <strong>of</strong> the nephronsin column 9 are impaired. 101Figure 67 A comparison <strong>of</strong> the spectral densities <strong>of</strong> WCFRbetween the original 72-nephron system and thesystem where all <strong>of</strong> the nephrons in column 9are impaired. 102


Figure 68 The difference in power spectral densities <strong>of</strong>WCFR between the original 72-nephron systemand the system where all <strong>of</strong> the nephrons in column9 are impaired. 103Figure 69 The difference in the power spectral density <strong>of</strong>column 9 in the original 72-nephron system andthe system where all <strong>of</strong> the nephrons in column9 are impaired. 104Figure 70 Variations in WCFR in the system where all <strong>of</strong> thenephrons in column 9 are impaired. 105Figure 71 Comparing 1-column, 2-column and 4-columnsimulations. 112Figure 72 Execution time is O(n 3 ) for n nephrons per column.113Figure 73 The master/slave relationship between columns. 115Figure 74 Communication between the controller and columns. 116Figure 75 Communication between columns when performingthe inter-column diffusion. 117Figure 76 Reordering the communication in Fig. 74, in orderto reduce the delay between time-steps. 117Figure 77 Amdahl’s law. 121Figure 78 The improvement in performance as the simulationis distributed over multiple computers. 121Figure 79 A Lindenmayer system that resembles the arterialtree <strong>of</strong> the kidney. 125Figure 80 A simple model <strong>of</strong> compliant vessels. 131L I S T O F TA B L E STable 1 Some <strong>of</strong> the solutes regulated by the nephron. 10Table 2 Properties <strong>of</strong> the n-nephron single-column network.39Table 3 Properties <strong>of</strong> the n-nephron c 2 -column network. 40Table 4 The solutes support by the model. 43Table 5 State variables in the nephron model. 45Table 6 Model parameters. 45Table 7 The behaviour <strong>of</strong> each tubule segment in thenephron model. 46Table 8 Notation for model equations. 47Table 9 Distal tubule responses to hormone concentrationsin the filtrate. 52Table 10 Steady-state concentration levels. 56Table 11 Systematic decrease blood sodium concentration.60Table 12 Systematic increase in blood sodium concentration.61Table 13 The difference in the sodium gradient at standardand altered blood sodium concentrations. 70xvi


Table 14Table 15Table 16Table 17Increases in spectral density peaks in the SNGFR<strong>of</strong> nephrons in the system with one impairednephron. 86Increases in spectral density peaks in the SNGFR<strong>of</strong> nephrons in the system with two impairednephrons. 92Performance comparison <strong>of</strong> two models <strong>of</strong> identicalsize. 114Message content in inter-process communication.118A C R O N Y M SAARADHARFBOINCCKDCPUCRFDCTECFEPOESRFGFRGNIgARASRRTSISSNGFRTGFTHPOWCFRWSFRafferent arteriole resistanceantidiuretic hormoneacute renal failureBerkeley Open Infrastructure for Network Computingchronic kidney diseasecentral processing unitchronic renal failuredistal convoluted tubuleextracellular fluiderythropoietinend-stage renal failureglomerular filtration rateglomerular nephritisImmunoglobulin Arenin-angiotensin systemrenal replacement therapysusceptible-infected-susceptiblesingle-nephron glomerular filtration ratetubuloglomerular feedbackthrombopoietinwhole-column filtration ratewhole-system filtration ratexvii


Part II N T R O D U C T I O N


I N T R O D U C T I O N1A complex system is a system comprised <strong>of</strong> many interconnectedparts, whose behaviours are both variable and dependent on thebehaviour <strong>of</strong> the other parts [161, 146]. Such systems exhibit emergentbehaviour—behaviour that may be deterministic, but which hasproperties that can only be examined at a higher level than the individualparts [3]. The difficulty in understanding such systems isthat the behaviour <strong>of</strong> the whole is highly dependent on the individualparts, and each part is highly dependent on other parts [17]. Moresimply, a complex system is holistic—it is more than just the sum <strong>of</strong>its parts—and we must understand how the interactions between theparts affect the whole.The kidney is one <strong>of</strong> the major organs responsible for homeostasis <strong>of</strong>the body. It regulates acid-base balance, extracellular fluid, electrolyteconcentrations, blood volume and blood pressure, as well as excretingmetabolic waste products. The kidney is also a complex system [133,173, 164], consisting <strong>of</strong> approximately 800,000 to 1,000,000 nephronsin the human [149]. The nephron is the basic filtration unit <strong>of</strong> thekidney, and is a convoluted tubule that adjusts the solute levels <strong>of</strong> thebody’s blood plasma. Nephrons are surrounded by interstitial fluid,into which solutes are secreted and reabsorbed, and by a complicatednetwork <strong>of</strong> capillaries that also exchange solutes with the surroundinginterstitial fluid. While the behaviour <strong>of</strong> individual nephrons canfluctuate widely and can even behave in a chaotic manner [112],the overall behaviour <strong>of</strong> the kidney remains stable. In particular,interactions occur between neighbouring nephrons that give rise tosynchronisations in pressure and flow oscillations—indicating thatgroups <strong>of</strong> nephrons are capable <strong>of</strong> producing emergent behaviour[166, 81, 112, 145, 167].In fact, the kidney is able to maintain its homeostatic functions overan astonishing range <strong>of</strong> bodily conditions, such as wide variations inblood pressure, extracellular fluid volume and solute concentrations.The kidney is also resilient to the effects <strong>of</strong> renal diseases and theloss <strong>of</strong> renal tissue (e.g. surgical removal)—a human can survive withonly one half <strong>of</strong> a single kidney. Although some <strong>of</strong> the regulatorymechanisms in the kidney are well understood, much less is knownabout how the functional stability <strong>of</strong> the entire kidney arises from theindividual components and their interactions.To the best <strong>of</strong> our knowledge, there has been relatively little researchinto how the many regulatory mechanisms affect higher-levelbehaviour in the kidney. Researchers have focused on understandingthe chemical processes and fluid mechanics at the cellular level, andmodelling the kidney’s vascular system. Many low-level properties<strong>of</strong> nephron tubule segments have been modelled [189, 177] and anumber <strong>of</strong> multi-nephron models have been proposed (mostly twonephronmodels [78, 81, 145], although a 22-nephron model alsoexists [112]). However, the standard methods and techniques in thisfield are not amenable to modelling and analysing the kidney as a3


4 I N T RO D U C T I O Ncomplex system [17, 201], due to the coupling between equations, thecomplexity <strong>of</strong> calculating solutions, and the inability to decouple thestructure and behaviour <strong>of</strong> the model. This thesis proposes a differentapproach that minimises the coupling between equations, reducesthe computational cost <strong>of</strong> solving the model equations, and decouplesthe model’s structure from its behaviour.1.1 A I MA knowledge gap exists between the understanding <strong>of</strong> the cellularprocesses and transporters that give rise to nephron function, andthe understanding <strong>of</strong> whole-kidney function. There has been littleresearch into how the dynamics <strong>of</strong> individual nephrons—and thecouplings, interactions and regulator mechanisms that arise betweennephrons—give rise to the dynamics observed at the whole-kidneylevel.The aim <strong>of</strong> this thesis is to address this gap by modelling the kidneyas a complex system and analysing it accordingly. The ultimate intent<strong>of</strong> this research is to provide an approach to renal modelling that iscapable <strong>of</strong> predicting whole-kidney function from the dynamics <strong>of</strong>individual nephrons, and can therefore be <strong>of</strong> practical use to clinicians.Specifically, this thesis makes the following contributions:• A modelling approach—hierarchical dynamical networks—whichcombines complex networks and graph automata into a singlemodel. This approach explicitly captures the structure and interactionsin multi-nephron systems, and decouples the structureand behaviour <strong>of</strong> the model. This approach allows emergentdynamics to be easily explored and analysed.• The development <strong>of</strong> a multi-nephron model that produces validbehaviour and renders the simulation <strong>of</strong> whole-kidney functioncomputationally tractable. Using this model, the emergent effects<strong>of</strong> the couplings and interactions between nephrons canbe investigated.• An investigation into the dynamics <strong>of</strong> multi-nephron systemsthat focuses on whole-system and hierarchical properties ratherthan the dynamics <strong>of</strong> individual nephrons. As part <strong>of</strong> this investigation,the dynamics <strong>of</strong> a 72-nephron system are analysed—asystem significantly larger than existing multi-nephron models.• A study <strong>of</strong> whole-system stability in response to localised impairmentsin nephron function. This is the first study <strong>of</strong> theemergent dynamics <strong>of</strong> impaired nephron function, and serves asan illustration <strong>of</strong> how the emergent dynamics produced by renaldiseases may be predicted and analysed. The impaired multinephronsystems are shown to exhibit very stable behaviour,which we contend is a feature <strong>of</strong> both the model and the kidneyproper.• The computational cost <strong>of</strong> the model is shown to be low enoughthat the simulation <strong>of</strong> whole-kidney function is feasible for thefirst time. It is also demonstrated that simulations can be easily


1.2 OV E RV I E W 5distributed across multiple computers, resulting in a significantgain in performance. An implementation <strong>of</strong> the model that supportsparallel and distributed execution is presented, based onthe Join Calculus [59, 60].• In order to predict whole-kidney function, a whole-kidney modelmust be constructed. This thesis proposes two approaches forautomatically generating such models.1.2 OV E RV I E WRenal physiology is introduced in Chapter 2, followed by a survey<strong>of</strong> existing tubule, whole-nephron and multi-nephron models. Thesemodels, which are able to capture precise low-level dynamics <strong>of</strong>the nephron tubule, are shown to be unsuitable for modelling andanalysing the kidney as a complex system. This chapter also presentsa brief discussion <strong>of</strong> renal disease and a review <strong>of</strong> renal diseasestudies.Graph automata and complex networks are introduced in Chapter3, and are shown to be better suited to modelling the kidneyas a complex system than the renal models presented in Chapter 2.Existing graph automata and complex network models are reviewed,followed by a discussion <strong>of</strong> how hierarchical properties <strong>of</strong> complexsystems can be captured with such models. The chapter concludes byreviewing the application <strong>of</strong> these models to studying the effects <strong>of</strong>disease.The modelling approach—hierarchical dynamical networks—is presentedin Chapter 4 and contrasted to continuous non-hierarchicalmodels. The structure <strong>of</strong> the multi-nephron model is presented andcompared to the structure <strong>of</strong> existing networks, and the update rules<strong>of</strong> the model are then given.Chapter 5 presents the experimental design used to validate themodel and to explore the emergent dynamics produced by the model.This is followed by the results <strong>of</strong> each experiment in turn: the validation<strong>of</strong> the single-nephron dynamics; the validation <strong>of</strong> the two-nephrondynamics; a further exploration <strong>of</strong> the interactions between competingcoupling mechanisms in the two-nephron system; and finally ananalysis <strong>of</strong> the interactions between competing coupling mechanismsin 8-nephron and 72-nephron systems.A study <strong>of</strong> whole-system stability in response to localised impairmentsin nephron function is presented in Chapter 6. The chapterbegins by discussing how localised impairments can be incorporatedinto the model, and how the emergent effects <strong>of</strong> such impairmentscan be studied. A simple impairment is introduced into 8-nephronand 72-nephron systems and the stability <strong>of</strong> these systems is analysed.The impaired multi-nephron systems are shown to exhibit stablebehaviour, which we contend is a feature <strong>of</strong> both the model and thekidney proper.The implementation <strong>of</strong> the model simulation is presented in Chapter7 and the performance <strong>of</strong> this implementation is analysed, demonstratingthat the simulation <strong>of</strong> whole-kidney function from the dynamics<strong>of</strong> individual nephrons is computationally tractable. It is alsodemonstrated that the simulations can easily be distributed across


6 I N T RO D U C T I O Nmultiple computers—a distributed implementation based on the JoinCalculus [59, 60] is presented and a performance analysis revealsthat the result in a significant gain in simulation performance. awhole-kidney model is required to predict whole-kidney function, andthe chapter concludes by proposing two methods for automaticallygenerating such models.The contributions <strong>of</strong> this thesis are summarised in Chapter 8. Thelimitations <strong>of</strong> this work are then discussed, and potential avenues<strong>of</strong> further research to build on the work are proposed. Finally, theresearch outcomes are evaluated against the research aim.


Part IIB A C K G R O U N D


T H E K I D N E Y2This chapter begins by introducing the reader to the physiology <strong>of</strong> thekidney—the structure and function <strong>of</strong> the kidney is briefly presented—and a review <strong>of</strong> existing nephron tubule models is presented. This isfollowed by an overview <strong>of</strong> the major renal diseases, their significance,and their effects on the kidney, and a review <strong>of</strong> models <strong>of</strong> kidneydisease. The chapter is concluded with a short description <strong>of</strong> themodelling approach that will be used to develop the nephron modelpresented in this thesis.2.1 P H Y S I O L O G Y O F T H E K I D N E YThe kidney is one <strong>of</strong> the major organs involved in whole-body homeostasis.It is responsible for regulating solute concentrations in theblood, the volume, composition and avid-base balance <strong>of</strong> extra-cellularfluid, and blood pressure in the body. The kidney also plays a partin the body’s endocrine system, secreting hormones such as reninas part <strong>of</strong> the renin-angiotensin system (RAS), erythropoietin (EPO)to regulate red blood cell production, and thrombopoietin (THPO) toregulate platelet production.The functional unit <strong>of</strong> the kidney is the nephron, a long, segmentedtubule into which blood plasma is filtered (see Fig. 1a). As the filtrateflows along the nephron tubule, water is reabsorbed into thebloodstream and solutes are both reabsorbed into the bloodstreamand secreted into the tubule (see Table 1). The filtrate flows from thenephron into the collecting ducts (labelled “A” in Fig. 1a), which drainthe fluid into the bladder via the urethra. The human kidney containsaround one million nephrons [149], which are arranged in renal lobes(see Fig. 1b).(a) A single nephron andthe peritubular capillaries.(b) Nephrons (shown as glomeruli connected to arcuatearteries) are grouped in renal lobes.Figure 1: Physiology <strong>of</strong> the kidney.9


10 T H E K I D N E YSolute Formula Reabsorbed SecretedSodium Na + Yes YesPotassium K + Yes YesProton H + No YesChloride Cl − Yes YesCalcium Ca 2+ Yes NoBicarbonate HCO − 3 Yes YesPhosphate PO 3−4 Yes NoUrea (NH 2 ) 2 CO Yes YesGlucose C 6 H 12 O 6 Yes NoAmino acids various Yes NoTable 1: Some <strong>of</strong> the solutes regulated by the nephron.1Cortex29738Outer Medulla46Inner Medulla5Figure 2: A diagram showing the functional segments <strong>of</strong> the nephron.1: the afferent arteriole, which provides the nephron with blood; 2:the glomerulus, a network <strong>of</strong> capillaries that filters a fraction <strong>of</strong> theblood plasma into the nephron tubule (the remaining blood flows intothe peritubular capillaries that surround the tubule—not shown here,but shown in Fig. 1a); 3: the proximal tubule; 4: the descending limb<strong>of</strong> Henle; 5: the thin ascending limb <strong>of</strong> Henle; 6: the thick ascendinglimb <strong>of</strong> Henle; 7: the distal tubule; 8: the collecting ducts, whichdrains the filtrate into the bladder; 9: the TGF control mechanism (aninteraction between the macula densa and the filtration process).The structure <strong>of</strong> a single nephron is shown in Fig. 2. The nephronis surrounded by connective tissue, which holds the nephron tubulesand renal blood vessels, and the interstitial fluid, which can be dividedinto compartments within the cortex, the outer medulla andthe inner medulla. Blood flows through the afferent arteriole (1) andinto the glomerulus (2), where a fraction <strong>of</strong> the blood plasma is filteredinto the nephron tubule, and the remaining blood flows intothe peritubular capillaries (not shown). The filtrate—the plasma thathas been filtered into the tubule—flows through the proximal tubule(3), the descending and ascending limbs <strong>of</strong> the loop <strong>of</strong> Henle (4–6),


2.1 P H Y S I O L O G Y O F T H E K I D N E Y 11the distal tubule (7) and finally the collecting ducts (8). The loop<strong>of</strong> Henle acts as a counter-current exchange mechanism for NaCl,increasing the NaCl concentration in the interstitial fluid as the loopdescends. In addition to this counter-current mechanism, the production<strong>of</strong> lactate in the tubule may contribute significantly to theaccumulation <strong>of</strong> sodium [76], and solutes such as urea also play arole in the concentrating mechanism [97, 98, 176]. Existing modelshave incorporated gradients—measured between the top and bottom<strong>of</strong> the loop <strong>of</strong> Henle—such as 150mmol/L [78], although the gradientcan be as large as 1200mmol/L in human kidneys [169] and over2000mmol/L in other species [158, 76].A major regulation mechanism (9) in the kidney is tubuloglomerularfeedback (TGF), a negative-feedback mechanism that regulates thefiltration rate <strong>of</strong> an individual nephron. A signal is sent to the afferentarteriole that controls its diameter (and thus the filtration rate),based on the delivery <strong>of</strong> solutes, mainly sodium chloride, to a regioncalled the macula densa (the end <strong>of</strong> the ascending limb <strong>of</strong> Henle,6–7 in Fig. 2). The filtration rate affects distal sodium delivery—highsodium levels indicate that the filtration rate is too high for sodium tobe reabsorbed effectively, while low sodium levels indicate that thefiltration rate is so low that too much sodium is being reabsorbed—and the TGF mechanism keeps the filtration rate within appropriatebounds (see Fig. 3). There is evidence that the TGF mechanism mayalso be mediated through the resistance <strong>of</strong> the efferent arteriole[151, 68, 23]—the arteriole that carries blood out <strong>of</strong> the glomerulus—and the permeability <strong>of</strong> the glomerular capillaries [130, 28, 181].Solute DeliveryFiltration RateFigure 3: The inverse-sigmoid signal generated by the TGF mechanism.Changes in the solute delivery to the macula densa result in aninverse response in the filtration rate. The system typically oscillatesin the highlighted region, about the equilibrium point indicated onthe graph.Specialised cells in the tubule wall generate a signal that controlsthe diameter <strong>of</strong> the afferent arteriole (which supplies blood to theglomerulus). The signal is not instantaneous, but has a delay <strong>of</strong> around


12 T H E K I D N E Y3–5 seconds [166, 145, 79]), and there is a further delay for thefiltrate to flow from the glomerulus to the macula densa (around10–15 seconds [166, 78, 79]). This negative-feedback mechanism hasbeen the focus <strong>of</strong> many multi-nephron models, and is known to havean inverse-sigmoid relationship [160, 166] with sodium delivery [159]at the macula densa (see Fig. 3).Two different types <strong>of</strong> interactions are known to occur betweennephrons in multi-nephron systems [166, 81, 112]: vascular signaling,where the TGF signal from one nephron also reaches nearbynephrons; and hemodynamic coupling, where the resistance <strong>of</strong> anafferent arteriole affects the hydrostatic pressure at nearby nephrons.Vascular signaling favours in-phase synchronisation [112], since anincrease or decrease in the TGF signal <strong>of</strong> one nephron is propagated toneighbouring nephrons. Conversely, hemodynamic coupling favoursanti-phase synchronisation [112], since the contraction <strong>of</strong> an afferentarteriole reduces the filtration rate <strong>of</strong> that nephron and increasesthe flow to—and therefore the filtration rate <strong>of</strong>—the neighbouringnephrons. One model has predicted that vascular signalling overwhelmsthe hemodynamic coupling, leading to in-phase synchronisationsin adjacent nephrons [112].The kidney has a hierarchical physical composition—the nephronconsists <strong>of</strong> a number <strong>of</strong> distinct segments; nephrons can be groupedinto interacting neighbourhoods (such as all nephrons joined to asingle cortical radial artery [112]); large numbers <strong>of</strong> these neighbourhoodsform renal pyramids; and the kidney consists <strong>of</strong> multiplesuch pyramids (e. g., from 8–18 in humans [55] to a single pyramidin smaller animals such as rodents and monotremes [211]). Thefunctional structure <strong>of</strong> the kidney is closely related to its physicalstructure—witness how the arrangement <strong>of</strong> the nephron tubule reflectsthe functional structure <strong>of</strong> the TGF mechanism—and thereforeexhibits a similar hierarchy. Hierarchy is a property <strong>of</strong> many complexsystems [164, 150, 47] and it would be desirable to capture this hierarchyexplicitly in our model, as opposed to existing nephron models,which only capture hierarchy implicitly (if at all). Hierarchy in modelsis discussed further in Sect. 3.3.2.2 R E NA L M O D E L L I N GThe study <strong>of</strong> renal physiology takes three forms: clinical studies thatfocus on diagnosing and treating kidney-related diseases; experimentalstudies that examine the behaviour and function <strong>of</strong> the kidney(in vivo or in vitro); and theoretical studies that propose models tocapture the behaviours and functions that have been observed experimentally.These fields encompass the study <strong>of</strong> more than just thebehaviour <strong>of</strong> the nephron tubule segments, and can be divided intothree general areas: cellular, tubular and vascular physiology.For the purpose <strong>of</strong> this thesis it is sufficient to concentrate ontubular physiology, since the models presented herein focus on thebehaviour <strong>of</strong> and interactions between nephrons. To this end, therange <strong>of</strong> theoretical models that have been proposed are now surveyed,as they not only provide a starting point for the design <strong>of</strong>new models, but also compare the models to the relevant portion


2.2 R E NA L M O D E L L I N G 13<strong>of</strong> the available experimental data. Several clinical studies are alsoreviewed in Sect. 2.4.The majority <strong>of</strong> nephron modelling has focused on replicating lowlevelbehaviour in individual tubule segments (e.g., the proximaltubule [190, 193], distal convoluted tubule [192, 192, 38, 37] andcollecting ducts [191, 191]). Experimental data shows that in thenephron, the flow rate and tubule pressure both oscillate, and modelshave shown that such oscillations can arise due to the TGF mechanism[78, 80]. Existing models are able to reproduce a wide range <strong>of</strong>other experimental data, but uncertainty still remains about the roles<strong>of</strong> transporters along the nephron tubule, due to the limitations <strong>of</strong>existing experimental techniques and the inability <strong>of</strong> current modelsto capture the essence <strong>of</strong> all observed phenomena [189].There are some published models that study the behaviour <strong>of</strong> multiplenephrons, such as their ability to regulate concentration gradientsin the medullary fluid [76, 108], or synchronisation phenomena thatcan arise due to pressure oscillations in blood flow [81, 112]. Thesemodels do not include the entire nephron, although at least onewhole-nephron model exists [170].These tubule models are generally formulated as systems <strong>of</strong> differentialequations that model the mechanics <strong>of</strong> fluid flow along tubulesegments and the transport <strong>of</strong> solutes and water across the tubulewalls. This approach has yielded models that are able to replicatespecific, detailed behaviours.For example, Weinstein [192] models the rat early distal convolutedtubule (DCT) as a tubule lined by DCT cells and intercellular space(both <strong>of</strong> which admit solute transport). Solute transport is governedby electrochemical potential gradients (via cotransporters) or coupledto metabolic energy (via an ATPase). This model updates previouswork, in drawing on new experimental data to refine the formulation(and inclusion) <strong>of</strong> several cotransporters, as well as including a morerealistic model <strong>of</strong> the tubule lining than previous models, allowingcell volume to vary and not requiring the intercellular and peritubularfluids to be at equilibrium. This model provides new observationsregarding cell volume homeostasis and new predictions concerningthe effects <strong>of</strong> alkalosis on the DCT.In short, this paper presents a model <strong>of</strong> the rat early DCT that isa significant extension <strong>of</strong> existing DCT models, by virtue <strong>of</strong> havingfewer underlying assumptions and more detailed cotransporter formulations.Despite these features, this model is a development <strong>of</strong>the published tubule models, and does not break new ground withrespect to the approach taken.Unlike discrete equations, which explicitly calculate the currentstate <strong>of</strong> a system based on the previous state <strong>of</strong> the system, systems<strong>of</strong> continuous equations only place constraints upon the system. Solvingsuch constraints can be computationally very expensive (<strong>of</strong>tenrequiring numerical solving methods which discretise the equations).A unique solution may not exist, in which case linear combination <strong>of</strong>the solutions must be considered.The accuracy <strong>of</strong> these kinds <strong>of</strong> models is not without drawbacks, asthe models can be computationally very expensive. Due to the couplednature <strong>of</strong> the equations, the addition <strong>of</strong> another variable or a change


14 T H E K I D N E Yto the equation governing an existing variable can have significanteffects on the solutions <strong>of</strong> other equations in the model. Another drawbackis that the tubule models cannot necessarily be linked togetherto form models <strong>of</strong> an entire nephron tubule, as they may have differentunderlying assumptions or incompatible formulations (althoughsome tubule models have been successfully combined). Our approachovercomes these drawbacks by using difference equations, whichdecouple the continuous equations and provide explicit solutions foreach variable.2.3 D I S E A S E I N T H E K I D N E YRenal diseases can prevent the kidney from functioning correctly, resultingin renal failure. Renal failure is divided into acute and chronicforms. acute renal failure (ARF) is a rapid loss <strong>of</strong> renal function dueto kidney damage, which may be reversible if it is treated promptly.chronic kidney disease (CKD) is a progressive loss <strong>of</strong> renal functionover a period <strong>of</strong> months and years, which is divided into five stages,based on the decrease in the kidney filtration rate. The final stageis referred to as end-stage renal failure (ESRF), and requires someform <strong>of</strong> renal replacement therapy (RRT)—either a kidney transplantor dialysis. ARF may also result in ESRF. Regardless <strong>of</strong> the cause, ESRFmanifests as an abnormally low glomerular filtration rate (GFR) suchthat the kidney is unable to perform its function.This review focuses on diseases that cause CKD. The most commoncauses <strong>of</strong> CKD and ESRF in western countries are:• Diabetic nephropathy—thickening <strong>of</strong> the glomeruli.• Hypertension—chronically elevated blood pressure.• Glomerulonephritis—inflammation <strong>of</strong> the glomeruli.Together, these three conditions account for around 75% <strong>of</strong> all adultcases <strong>of</strong> chronic renal failure (CRF) [41]. In particular, diabetic nephropathyis the most common cause <strong>of</strong> ESRF in the USA [178, 53].2.3.1 Symptoms <strong>of</strong> renal diseaseWith the aim <strong>of</strong> studying how the onset <strong>of</strong> renal disease in individualnephrons can affect the whole-system dynamics <strong>of</strong> multi-nephronsystems, it is the symptoms <strong>of</strong> the renal diseases that are <strong>of</strong> greatestinterest, as the symptoms indicate the changes in dynamics causedby the diseases. The most common symptoms <strong>of</strong> renal disease arenow introduced.ProteinuriaProteinuria is the presence <strong>of</strong> abnormally large amounts <strong>of</strong> proteinin the urine—which can cause the urine to become foamy—and isa sign <strong>of</strong> potential renal damage. It was understood that a chargebarrier in the glomerulus prevents protein from entering the nephrontubule [180, 185, 39, 135] and that the presence <strong>of</strong> protein in theurine is due to deterioration in this charge barrier, allowing proteins


2.3 D I S E A S E I N T H E K I D N E Y 15such as albumin to enter the nephron tubule and ultimately the urine[69, 25]. However, the influence <strong>of</strong> this electrostatic interaction hasbeen questioned [45, 138, 73, 155] and there is evidence that proteinenters the nephron tubule under normal conditions, and that thepresence <strong>of</strong> protein in the urine is due to the failure <strong>of</strong> a reabsorptionmechanism in the proximal tubule cells [116, 32, 21, 44]. Proteinuriawas first described by William Charles Wells [63, 33], who suggestedthat the serum in the urine was derived from the blood [194] and gavethe first definition <strong>of</strong> the limits <strong>of</strong> normal albuminuria [195] (similarfigures are still used today [33]).HematuriaHematuria is the presence <strong>of</strong> red blood cells in the urine. Aside fromrenal diseases, hematuria can indicate the presence <strong>of</strong> kidney stonesor a tumour. Hematuria is a symptom <strong>of</strong> glomerulonephritic diseasessuch as Immunoglobulin A (IgA) disease (also known as Berger’snephropathy), which is the most common type <strong>of</strong> glomerulonephritisin adults world-wide [65, 20, 128].HypoalbuminemiaHypoalbuminemia is a condition where the level <strong>of</strong> albumin in theblood is abnormally low. It can be caused as a result <strong>of</strong> proteinuria,and may also indicate liver failure, cirrhosis or chronic hepatitis. Asalbumin is the major protein in the human body (contributing around60% <strong>of</strong> the total plasma protein by mass), a significant reduction inalbumin levels significantly lowers the oncotic pressure <strong>of</strong> the blood.Such a reduction leads to excessive amounts <strong>of</strong> water diffusing fromthe blood into the extra-cellular fluid, causing general swelling <strong>of</strong> thebody (edema).HyperlipidemiaHyperlipidemia is the presence <strong>of</strong> abnormally large amounts <strong>of</strong> lipidsin the blood, which can arise when the liver responds to albumin lossby secreting more protein into the bloodstream.HypertensionHypertension is a condition where the blood pressure is chronicallyelevated, and is one <strong>of</strong> the risk factors for cardiovascular diseases, aswell as a major cause <strong>of</strong> CRF. One <strong>of</strong> the major causes <strong>of</strong> hypertensionis a diet that is high in sodium [156, 92].Nephritic SyndromeNephritic syndrome is a collection <strong>of</strong> signs associated with glomerulardisorders, characterised by proteinuria, hematuria and hypertension.It is not a specific diagnosis and so the underlying causes can vary, butthe common feature is an inflammation <strong>of</strong> the glomeruli, leading tosalt and water retention and ultimately a reduction in kidney function.


16 T H E K I D N E YNephrotic SyndromeNephrotic syndrome is a disorder where large amounts <strong>of</strong> proteinleak from the blood into the urine. It is characterised by proteinuria,hypoalbuminemia, hyperlipidemia and edema (and occasionallyhypertension). Nephrotic syndrome has many specific causes (suchas diabetes and glomerulonephritis), but the common mechanismis inflammation causing damage to the glomeruli, allowing proteinssuch as albumin to pass through the kidneys and into the urine.2.3.2 The economic impact <strong>of</strong> renal diseaseIncidence <strong>of</strong> renal disease is becoming more prevalent in both westernand developing nations, placing a rapidly growing burden onglobal health care [36, 13, 183, 184]. Long-term kidney disease alsoincreases the risk <strong>of</strong> developing further complications such as cardiovasculardiseases, respiratory infections, bone and muscle problems,and anaemia [13]. Once ESRF occurs, the only option is to pursuepalliative treatments known as RRT. Dialysis is the most common form<strong>of</strong> treatment—approximately 77% <strong>of</strong> global RRT patients undergo dialysis,with the remaining 23% having functional transplants [125]—butit is also the most costly form <strong>of</strong> treatment [62] and is not withoutdrawbacks [154].The burden <strong>of</strong> premature death from diabetes is similar to that <strong>of</strong>HIV/AIDS, with one in 20 deaths attributable to diabetes, and thenumber <strong>of</strong> people with diabetes is expected to more than double by2030 [136].These predictions <strong>of</strong> increased incidence <strong>of</strong> renal disease and theassociated increase in health-care costs highlight the importance<strong>of</strong> improving our understanding <strong>of</strong> how renal diseases affect renalfunction. Studying the effects <strong>of</strong> renal diseases through the use <strong>of</strong>models may lead to improvements in the diagnosis and treatment <strong>of</strong>renal diseases, helping to reduce the financial and social impacts <strong>of</strong>these diseases.2.4 M O D E L L I N G R E NA L D I S E A S EThe causes and mechanisms <strong>of</strong> renal disease have been investigated atthe cellular level [21, 32] and many clinical experiments [10, 86, 114,58, 91], laboratory experiments [95, 24, 84, 144] and case studies[172, 14, 148] have been performed. Rashidi and Khodarahmi presenta simple model <strong>of</strong> nephron loss in response to CRF [149], but there arefew other models that attempt to capture the effects <strong>of</strong> renal diseaseon the kidney or on multi-nephron systems. In particular, it does notappear that any existing tubule or nephron models have attempted tomodel the impact <strong>of</strong> renal disease on the nephron in the context <strong>of</strong>multi-nephron systems, precluding the study <strong>of</strong> emergent dynamicsproduced in the presence <strong>of</strong>, or in response to, renal disease.Thomson et al. present “the tubular hypothesis <strong>of</strong> glomerular filtration”[178], which uses a control theory model to explain severalnuances <strong>of</strong> kidney function in early diabetes. Early diabetes is knownto have strange effects on GFR, including diabetic hyperfiltration, a


2.5 M O D E L L I N G T H E R E NA L P H Y S I O L O G Y A S A N E T W O R K 17paradoxical effect <strong>of</strong> dietary sodium, and the renal response to dietaryprotein and amino acid infusion. The tubular hypothesis presumesthat these effects stem from primary effects on the proximal tubuleand loop <strong>of</strong> Henle, which impact the single-nephron glomerular filtrationrate (SNGFR) by feedback through the macula densa and theTGF mechanism. Such primary effects are categorised as “tubular”in comparison to all other possible effectors, which are classified as“vascular”. The difference between tubular and vascular effectorsis that vascular effectors have the same effect on GFR and sodiumdelivery to the macula densa—if the GFR is increased, so is the sodiumdelivery—while tubular effectors have contrary effects—if the GFR isincreased, the sodium delivery is decreased.The published data on the content <strong>of</strong> the early distal tubule indiabetes consistently lists sodium values substantially below normal,and experimental data shows that diabetic nephrons respond to thiswith hyperfiltration [72]. However, there is also a change in the TGFresponse which indicates that a vascular effect may have almost asmuch control over GFR.The diabetic kidney also exhibits a “salt paradox” where GFR variesinversely with dietary sodium intake, resulting in a lower GFR whensodium intake is increased, rather than the normal response <strong>of</strong> anelevated GFR. This cannot be explained by any vascular effectors, butcan be explained by the tubular hypothesis as long as the proximaltubule is the dominant mechanism linking GFR to extracellular fluid(ECF).Glycine infusion normally causes GFR to increase without alteringsodium levels at the macula densa, indicating that both vascular andtubular effects contribute to the increase in GFR, while in the diabetickidney the vascular response is minimal. Protein feeding alsoincreases GFR, by increasing tubular reabsorption to levels that decreasethe sodium delivery to the macula densa, despite the elevatedGFR. This tubular effect remains present in the diabetic kidney.In summary, Thomson et al. demonstrate that this tubular hypothesisand the associated control theory model can explain severalfactors that influence the behaviour <strong>of</strong> the kidney in early diabetes.Their paper does not focus on specific cellular mechanisms, preferringto treat the kidney as an integrated system <strong>of</strong> parts—an approachthat is reminiscent <strong>of</strong> our own. The major difference is that this modelstarts from experimental observations <strong>of</strong> the whole kidney to suggestthe effects that early diabetes has on the nephron tubule, while thefocus <strong>of</strong> our disease modelling is to start from known effects in thenephron tubule and to use our models to determine the resultingbehaviour at a higher level, such as the whole kidney.2.5 M O D E L L I N G T H E R E NA L P H Y S I O L O G Y A S A N E T W O R KAs will be discussed in the following chapter (Chapter 3), networkmodels are well suited to capturing both implicit and explicit hierarchiespresent in the system being modelled, and are also well suitedto studying the changes in system dynamics that arise in responseto alterations in the model structure and low-level behaviours. Inparticular, network models provide an advantage when exploring the


18 T H E K I D N E Yeffects <strong>of</strong> different couplings and interactions between parts <strong>of</strong> themodel, as these mechanisms are explicitly captured by the modelstructure.Furthermore, due to the use <strong>of</strong> discrete time equations, such modelsare able to scale to large systems without incurring the increase incomputational cost that is typical <strong>of</strong> continuous systems. This point isdiscussed further in Sect. 4.1.1 and the model presented in this thesisis demonstrated to scale to extremely large models in Chapter 7.


N E T W O R K M O D E L S3As mentioned in the introduction, there are two common classes <strong>of</strong>network model—networks with regular, fixed topologies and nodeswhose states evolve due to homogeneous update rules [201]; andnetworks with irregular topologies, where nodes have constant stateand dynamics are realised as changes in the network topology [139].However, the kidney has both an irregular static functional topology(closely tied to its physical structure) and localised states that evolveinhomogeneously over time. The model presented in this thesis willtherefore have features in common with both networks, and so bothclasses <strong>of</strong> network models are examined in this chapter. Firstly, definitions<strong>of</strong> cellular and graph automata are presented, followed by areview <strong>of</strong> existing automata models. This is followed by an introductionto common network properties and a review <strong>of</strong> existing complexnetworks models.3.1 G R A P H AU T O M ATA M O D E L SA cellular automata A = {S, C, N, f} consists <strong>of</strong> a finite state spaceS, a configuration <strong>of</strong> the cells C ∈ S Z for a finite <strong>of</strong> infinite discreteregular grid <strong>of</strong> cells Z, a neighbourhood function N : Z → S n andan update rule f : S n → S. Time is discrete and at any time t, theconfiguration <strong>of</strong> an automaton is C(t) ∈ S Z . For each cell c ∈ Z, thestate at time t + 1 is calculated from the state <strong>of</strong> the neighbourhoodN(c) at time t, as shown in Eq. 3.1.c(t + 1) = f( N(c) ) (3.1)(a) An update rule—cells change statewhen their neighbours’ states differ.(b) The update rule appliedto an automaton.Figure 4: An example <strong>of</strong> a 1D binary cellular automaton that has aring topology—the cells at both ends are connected together.An example <strong>of</strong> a 1D automaton with a binary state space is shownin Fig. 4. The update rule f for this automaton (Fig. 4a) changes thestate <strong>of</strong> a cell when the states <strong>of</strong> the two neighbouring cells differ. Anapplication <strong>of</strong> this rule to a configuration is shown in Fig. 4b.Graph (or network) automata are an extension <strong>of</strong> cellular automata,where the topology <strong>of</strong> the cells Z is generalised to a network, ratherthan being restricted to a regular grid.19


20 N E T W O R K M O D E L SA graph (or network) G = {V, E} consists <strong>of</strong> a set <strong>of</strong> nodes (V),and edges that connect pairs <strong>of</strong> nodes (E ⊆ V × V). A graph is bydefault undirected, meaning that the edges (v 1 , v 2 ) and (v 2 , v 1 ) areequivalent, and for any (v 1 , v 2 ) ∈ E it is also true that (v 2 , v 1 ) ∈ E.Graphs may also be directed, in which case the edges (v 1 , v 2 ) and(v 2 , v 1 ) are distinct, and an edge (v 1 , v 2 ) is said to be directed fromv 1 to v 2 . The neighbourhood <strong>of</strong> a cell (node) c in a graph automata isthe set <strong>of</strong> cells (nodes) that are directly connected to c in the graph.Graph automata have been applied to the study <strong>of</strong> biological phenomena,such as computational operations in reaction-diffusion systems[2] and genetic regulatory networks [203].Adamatzky et al. [2] show how basic computational operationscan be implemented in a three-state hexagonal automaton. This automatonmodels a reaction-diffusion system using the “beehive rule”[206], which is presented in Fig. 5. The beehive rule exhibits gliderdynamics—that is, starting from a random configuration, it producesregular patterns that move across the cells whilst retaining theirshape (i. e., gliders). Adamatzky et al. show that by interpreting the directionin which the gliders travel as truth values, the beehive rule iscapable <strong>of</strong> producing logic gates (as depicted in Fig. 6), and performingactions such as multiplication. The result is that the automatonis logically universal—any sequential machine can be constructed inthe system.S2 Neighbours6543210S0S2 S0S0 S0 S2S0 S2 S2 S0S0 S0 S2 S2 S0S0 S2 S2 S2 S1 S1S0 S1 S2 S1 S2 S0 S00 1 2 3 4 5 6S1 Neighbours(a) The beehive update rule in matrix formS0 S0S1 S1 S2S1 S2Beehive ruleS2(b) An example <strong>of</strong> the beehive rule, for a cell that hastwo S1 neighbours and two S2 neighbours.Figure 5: The beehive rule operates on a hexagonal grid whose cellshave a state-space S = {S0, S1, S2} [2].While this is a significant achievement—it could potentially be a theoreticalprototype <strong>of</strong> physically-realised reaction-diffusion processors—the analysis and interpretation <strong>of</strong> the automaton’s behaviour is toodeeply rooted in visual observations <strong>of</strong> patterns in the cells. While


3.2 C O M P L E X N E T W O R K M O D E L S 21(a) t = 1 (b) t = 2 (c) t = 3 (d) t = 4¬x∧yxx∧¬yy(e) The logic gateFigure 6: A logic gate implemented with the beehive rule [2].such an analysis might be automated once the behaviour and theirinterpretations are known, it can not be readily applied to findingand analysing patterns <strong>of</strong> behaviour in larger, more complex models.Indeed, the state-space <strong>of</strong> a single nephron is so complicated that isnot obvious how to visualise more than a few rudimentary properties(such as changes in concentration <strong>of</strong> a single solute), which precludesthe possibility <strong>of</strong> discerning behavioural patterns visually.This failure <strong>of</strong> analysis techniques to scale to large models is notisolated to this single paper, rather it is typical <strong>of</strong> the graph automatamodels. Indeed, Wolfram acknowledges both the necessity <strong>of</strong> visualanalysis and the difficulties it entails [201, 3§12 pp. 111–3], remarkingthat as a result he ignored the possibility <strong>of</strong> cellular automataproducing complex behaviour for several years.One other shortcoming <strong>of</strong> standard graph automata is that all nodesare homogeneous, sharing the same state-space and update rules. Asdiscussed in Sect. 2.1, nephrons and nephron tubule segments arenot homogeneous, and so a graph automata model <strong>of</strong> a nephron mustallow for nodes with different state-spaces and update rules.Although nephrons are inhomogeneous, they do share a commonbasic structure and set <strong>of</strong> transport mechanisms. Each tubule segmentis typically lined by the same types <strong>of</strong> cells regardless <strong>of</strong> whichnephron it belongs to—although some tubules may not contain allpossible cell types—and so all instances <strong>of</strong> each segment that sharethe same cell types can be modelled homogeneously. By taking thisapproach, one does not obtain a graph automata with homogeneousnodes and update rules, but rather an automata with classes <strong>of</strong> nodesand update rules.However, when attention is turned to modelling large groups <strong>of</strong>nephrons, it becomes clear that the analysis will necessarily be largelystatistical in nature, and that the models must have more complex,less regular topologies than the typical graph automata. Such networksare known as Complex Networks.3.2 C O M P L E X N E T W O R K M O D E L SLarge networks with complex topologies have been studied in thecontext <strong>of</strong> many biological phenomena, such as neural networks


22 N E T W O R K M O D E L S[54], protein interactions [90], metabolic reactions [4], and geneticregulation [132]. Some basic network properties are:D E G R E E The number <strong>of</strong> edges that connect to a node. For directedgraphs this can be extended to discern between edges thatlink to a node (INDEGREE) and edges that link from a node(OUTDEGREE).D E G R E E D I S T R I B U T I O N The relationship between the degree k andthe probability P(k) <strong>of</strong> a given node having that degree.D I A M E T E R The maximum length <strong>of</strong> the shortest paths between eachpair <strong>of</strong> nodes.C O N N E C T I O N P RO BA B I L I T Y The probability that two nodes selectedat random are connected by an edge.C L U S T E R I N G C O E FFI C I E N T For a specific node n, the fraction <strong>of</strong> thepossible edges between all neighbours <strong>of</strong> n that are present; fora network N, the average <strong>of</strong> the clustering coefficients for eachnode in the network.Based on these properties, complex networks can be classified intoa number <strong>of</strong> categories:R E G U L A R Each node has the same degree k and the network is saidto be K-REGULAR.R A N D O M Each pair <strong>of</strong> nodes is connected by an edge with uniformprobability p. The degree distribution <strong>of</strong> such networks is approximatelyGaussian.S M A L L W O R L D The shortest path between two nodes is likely tobe small as the shortest path length scales logarithmically orslower with network size n [188]. In contrast, the shortest pathlength scales linearly with n in a regular network.S C A L E-FREE The network has a power-law degree distribution. Many,perhaps even most, <strong>of</strong> the real-world networks that have beenstudied are scale-free [124]. The mechanism proposed as theexplanation for this observation is “growth by preferential attachment”[18]—when a node is added to the network, it isconnected to an existing node i with a probability proportionalto the degree <strong>of</strong> i.Wang and Chen [186], Newman [132] and Albert and Barabási [4]all review the progress and significant results in the current studies<strong>of</strong> various complex networks, including the World Wide Web, humanlanguage, and paper authorship. The dynamics <strong>of</strong> these models arederived from their topology, either by adding or removing nodes fromthe network, or by percolating node states across the network (e. g.,to model the spread <strong>of</strong> disease). One <strong>of</strong> the main areas <strong>of</strong> researchusing these models is the analysis <strong>of</strong> network robustness 1 in responseto changes in the network (e. g.the removal <strong>of</strong> vertices or a change inpropagation parameters).1 Also called resilience or stability.


3.3 H I E R A R C H Y I N R E A L-WORLD N E T W O R K S 23The analysis methods used to determine network resilience typicallyfocus on how statistical properties such as reachability areaffected by changes to the network. For example, percolation theoryis concerned with the propagation <strong>of</strong> state across a network (e. g.,states such as “working” and “failed”). The parameters governing thestate propagation are altered and changes in vertex patterns are thenanalysed for properties such as reachability and connected sub-sets.This approach is widely used for modelling phenomena such as thespread <strong>of</strong> diseases in societies and the spread <strong>of</strong> computer virusesover computer networks.While this approach might initially seem well-suited to our aim <strong>of</strong>measuring nephron resilience to the onset <strong>of</strong> renal disease, this is notthe case. Renal diseases affect factors such as filtration rates and thestrength <strong>of</strong> transport mechanisms, and it is not at all clear how theseeffects can be understood or analysed in terms <strong>of</strong> the usual kinds <strong>of</strong>statistical network properties.Also, the manner in which these network models capture the dynamics<strong>of</strong> a system is not ideally suited to modelling groups <strong>of</strong> nephronsfor the following reasons:• We require classes <strong>of</strong> nodes and update rules, so we cannotmodel state updates as a simple propagation or percolationacross the network.• The topology <strong>of</strong> a network <strong>of</strong> nephrons remains fixed (ignoringmajor events such as surgical removal), so we cannot modelstate updates as changes to the network topology.However, multiple graph automata can be connected together by acomplex network, which can be used to capture high-level structureand hierarchy in a system. This hybrid approach allows us to separatethe modelling into two distinct concerns: modelling nephrons as graphautomata; and modelling groups <strong>of</strong> nephrons as complex networks <strong>of</strong>these automata.3.3 H I E R A R C H Y I N R E A L-WORLD N E T W O R K SHierarchy is important in complex systems, and in complex networksthe hierarchy emerges from the network model. Explicit hierarchyis typically used when designing systems [74, 89, 202], while thehierarchy in real-world systems is not <strong>of</strong>ten evident and must beextracted from the model.Most real-world complex networks exhibit four main hierarchicalproperties: a scale-free degree distribution [132]; small-worldbehaviour [186, 27]; a hierarchical structure in clustering features[150, 132]; and degree-degree correlations [131].To extract physical knowledge from a complex system, one mustfocus on the right level <strong>of</strong> description for that system [66]. As a result,there is great interest in the hierarchical properties <strong>of</strong> real-worldnetworks [115, 157] and how networks can be generated so that theyexhibit these hierarchical properties [174, 26, 67]. An example <strong>of</strong>generating a network with hierarchical properties is illustrated inFig. 7.


24 N E T W O R K M O D E L S(a)n=0,N=5(b) n=1,N=25(c) n=2,N=125Figure 7: The iterative construction <strong>of</strong> a hierarchical network from afully-connected cluster <strong>of</strong> five nodes, as demonstrated by Ravasz andBarabási in [150].Existing studies <strong>of</strong> hierarchy in networks have focused on identifyinghighly clustered neighbourhoods in networks [150] and treatingthese neighbourhoods as networks in themselves [9]. In other words,the original system is divided into a number <strong>of</strong> modules. The definition<strong>of</strong> a module in a complex dynamical system, and the importance<strong>of</strong> inter-module dependencies, is not always unanimous nor wellquantified[187].Song et al. demonstrate that real-world networks consist <strong>of</strong> selfrepeatingpatterns on all scales [165]. This is unexpected, as it iswidely believed that such networks are not self-similar, due to thesmall-world property, which implies that the relationship between thenumber <strong>of</strong> nodes and the diameter <strong>of</strong> the network is exponential, andnot the power-law relation that is expected <strong>of</strong> self-similar networks.The self-similarity <strong>of</strong> these networks is demonstrated by calculatingtheir fractal dimension (using a “box-counting” method). They state:The box-counting method provides a powerful tool forfurther investigations <strong>of</strong> network properties because itenables a renormalization procedure, revealing that theself-similar properties and the scale-free degree distributionpersist irrespectively <strong>of</strong> the amount <strong>of</strong> coarse-graining<strong>of</strong> the network.These properties help explain why real-world networks have ascale-free distribution, and the presence <strong>of</strong> self-similarity in differentnetworks—the World Wide Web, protein-protein interaction networks,cellular networks and social networks—suggests that this diverserange <strong>of</strong> networks may share common self-organisation dynamics.Graph automata can also demonstrate hierarchy in behaviour. Indeed,Wolfram classifies such automata into four classes, dependingon the hierarchy present in the time-state diagram [201, 6§2 p231],but little else has been done to investigate their hierarchical dynamics.Hierarchical automata do exist [74, 75, 113, 119], but are typicallyused to design modular systems [89] rather than to study the dynamics<strong>of</strong> existing systems. In these applications, the automata are used


3.4 D I S T U R B E D N E T W O R K S A N D D I S E A S E 25to capture the intended dynamics <strong>of</strong> the system being designed, andare not used to study any emergent dynamics that were not explicitlydesigned into the system. In conclusion, although hierarchicalautomata have been used to design modular systems, there has beenlittle research into the emergent hierarchical properties <strong>of</strong> theseautomata, which is in stark contrast to the studies <strong>of</strong> hierarchy incomplex networks.3.4 D I S T U R B E D N E T W O R K S A N D D I S E A S EMany aspects <strong>of</strong> diseases have been studied through the use <strong>of</strong> networkmodels. For example, the development <strong>of</strong> breast cancer hasbeen modelled as an implicit graph automata [56] and the spread <strong>of</strong>sexually-transmitted diseases has been studied using social networkmodels [104].One interesting study concerning the spread <strong>of</strong> epidemics across biological,social and communication networks is that <strong>of</strong> Pastor-Satorrasand Vespignani [142, 143], which examines the susceptible-infectedsusceptible(SIS) model [15, 127] on complex networks. Each node <strong>of</strong>a network represents an individual system and each edge is a connectionalong which the infection can spread to other systems. Nodeshave two available states—“healthy” (i. e., susceptible to infection)and “infected”—and at each time-step healthy nodes are infected withrate ν if they are connected to one or more infected nodes. At thesame time, infected nodes are cured and again become susceptiblewith rate δ, which defines an effective spreading rate <strong>of</strong> λ = ν/δ.In networks such as random graphs, the most significant epidemiologicalresult is the existence <strong>of</strong> a non-zero epidemic threshold λ C[15]. In the case λ < λ C , the infection dies out exponentially fast;in the case λ > λ C , the infection spreads and becomes persistent.However, over a wide range <strong>of</strong> scale-free networks Pastor-Satorrasand Vespignani observed an absence <strong>of</strong> the epidemic threshold, whichimplies that scale-free networks are susceptible to persistent infectionregardless <strong>of</strong> the spreading rate—an observation that fits wellwith real-world data concerning the spread <strong>of</strong> computer viruses overthe Internet. Given that scale-free networks appear in a wide range <strong>of</strong>areas [4], the results obtained here may help improve our understanding<strong>of</strong> computer virus epidemics and could well have applications inother social and biological systems.However, regardless <strong>of</strong> the scale <strong>of</strong> the models (e.g., cellular, social),the focus is on how diseases or infections spread. Our aim is to usedisease modelling to investigate how the onset <strong>of</strong> renal disease inindividual nephrons can affect the stability <strong>of</strong> an entire system <strong>of</strong>nephrons. To achieve this aim, the ability <strong>of</strong> diseased nephrons toaffect neighbouring nephrons and the whole-system dynamics will bestudied, rather than studying how the disease is propagated. Withthis focus on system stability, not the spread <strong>of</strong> disease, the majority<strong>of</strong> the existing disease modelling literature is not directly relevant.However, at least one study [12] has investigated the system-wideeffects <strong>of</strong> localised failure, in the context <strong>of</strong> power grids. Ash andNewth start with a simple model for representing cascading failureson complex networks (as given in e. g., [126, 50]). Each edge (i, j)


26 N E T W O R K M O D E L Sis assigned a weight g ij —the efficiency <strong>of</strong> moving between i and jon this edge—in the range [0,1]. Initially, g ij = 1 if the edge (i, j) ispresent in the network, and g ij = 0 if it is not.Flow between two nodes takes the most efficient (shortest) pathand the load L i (t) on node i at time t is equal to the number <strong>of</strong> suchpaths that pass through i at time t. Each node has a finite capacityψ i = L i (0) × α, where α is a measure <strong>of</strong> the stress on the systemat time t = 0. The network is operating at maximum capacity whenα = 1, and the load on the network is light when α ≫ 1. Once thecapacity <strong>of</strong> a node i is exceeded, its performance is diminished asper Eq. 3.2. Initially, the capacity <strong>of</strong> some node i is reduced, and ateach time-step the load is redistributed over other nodes accordingto Eq. 3.2.⎧⎨g ij (t) ×g ij (t + 1) =⎩g ij (0)ψ iL i (t)if L i (t) > ψ iotherwise(3.2)The efficiency <strong>of</strong> an Erdös–Rényi random graph [57] is comparedto that <strong>of</strong> a network that has been optimised using an evolutionaryalgorithm. For α < 1.35 the optimised network considerably outperformsthe random graph, while for α 1.35 the networks displaysimilar behaviour. This clearly demonstrates that the optimisation algorithmhas worked—network efficiency is greatly improved when thenetwork is operating near maximum capacity and a node is impaired—and the study goes on to analyse optimised networks for topologicalregularities that might explain the source <strong>of</strong> the increase in resilience.This study, unlike other network models, began with a functioningsystem in which the function <strong>of</strong> a single node was impaired and theresulting effect on the entire system was measured. While it mayappear that this approach is identical to the study <strong>of</strong> system stabilityoutlined in Sect. 1.1, this is not the case. This study incorporates achange in state—the reduction in capacity ψ i <strong>of</strong> some node i—andthis change is then propagated across the network, due to the loadredistribution captured in Eq. 3.2. Accordingly, the efficiency <strong>of</strong> thenetwork is simply a weighted measure <strong>of</strong> the number <strong>of</strong> nodes whosecapacities have been reduced. As such, this study is an application<strong>of</strong> percolation theory, where nodes are initially in one state (i. e.,full capacity) and the percolation <strong>of</strong> another state (i. e., reducedcapacity) is measured. In contrast, the approach outlined in Sect. 1.1introduces a localised impairment to a system and observes the effecton the system function, where the impairment is contained and doesnot spread across the system. The contribution <strong>of</strong> an evolutionaryalgorithm to optimise the efficiency <strong>of</strong> a network is also unrelated tothis approach, as it relies on the ability to change the topology <strong>of</strong> thesystem, which is not feasible in the context <strong>of</strong> renal disease.


Part IIIM O D E L S A N D E X P E R I M E N T S


M O D E L D E S I G N4This chapter presents the design <strong>of</strong> the model. To begin, Sect. 4.1 introducesthe modelling approach—hierarchical dynamical networks—and contrasts it with other approaches. The benefits <strong>of</strong> using a discretetime model are described and a brief overview <strong>of</strong> the analysistechniques for this model is presented. The structure <strong>of</strong> the model isthen described in Sect. 4.2, followed by an analysis <strong>of</strong> the structuralproperties <strong>of</strong> the model in Sect. 4.3. Finally, the underlying equationsand update rules that capture the model dynamics are presented inSect. 4.4.4.1 M O D E L L I N G A P P ROAC HAs we have chosen to model the kidney as a complex system, thecouplings between the individual parts (i. e., the tubule segments) are<strong>of</strong> more interest than the dynamics <strong>of</strong> the individual parts in isolation.To elaborate, it is <strong>of</strong> greater importance to model the major couplingand interaction mechanisms that arise in multi-nephron systems thanto develop highly accurate formulations for the individual tubulesegments. The kidney is known to be a remarkably resilient organ [71],capable <strong>of</strong> maintaining its function across a wide range <strong>of</strong> conditions;operating on the assumption that the kidney is indeed a complexsystem, our intuition is that this resilience is due to the regulatorymechanisms and other interactions that arise between the lower-levelcomponents <strong>of</strong> the kidney, rather than being primarily due to thebehaviour <strong>of</strong> the individual nephron tubule segments in isolation.This approach, with less emphasis on the behaviour <strong>of</strong> individualparts than on their interactions, is similar to the approaches usedin fluid dynamics and gas models, where the individual particles areessentially ignored, in favour <strong>of</strong> higher-level emergent propertiessuch as mean flow rate and pressure. In the words <strong>of</strong> Nobel laureateRichard Feynman [77]:We have noticed in nature that the behavior <strong>of</strong> a fluiddepends very little on the nature <strong>of</strong> the individual particlesin that fluid. For example, the flow <strong>of</strong> sand is very similarto the flow <strong>of</strong> water or the flow <strong>of</strong> a pile <strong>of</strong> ball bearings.We have therefore taken advantage <strong>of</strong> this fact to invent atype <strong>of</strong> imaginary particle that is especially simple for usto simulate. This particle is a perfect ball bearing that canmove at a single speed in one <strong>of</strong> six directions. The flow <strong>of</strong>these particles on a large enough scale is very similar tothe flow <strong>of</strong> natural fluids.We treat the nephron tubule segments analogously—the tubulesegments are modelled as highly idealised “imaginary particles” andare the building blocks <strong>of</strong> our model—and as such, they will be modelledas individual nodes in the network automata (see Sect. 3.1).29


30 M O D E L D E S I G NExisting network automata have typically used 2-state [205, 204],3-state [206, 2] and even n-state [61, 118, 129] state spaces. Ourstate spaces are significantly more complex than [61, 118, 129] andthe visual analysis methods used by existing models (and reviewed inSect. 3.1) do not scale to the study <strong>of</strong> such complex models.Complex systems exhibit hierarchy in their organisation [150, 187,47, 11, 164] and the kidney is no different. Systems <strong>of</strong> nephrons exhibithierarchy in their organisation and interact with each other ina variety <strong>of</strong> ways. Multi-nephron systems are modelled as multiplesingle-nephron network automata, connected to each other via an arterialtree. This tree <strong>of</strong> nephron models forms a complex network (seeSect. 3.2) but unlike the majority <strong>of</strong> existing complex network models,the hierarchy <strong>of</strong> this model will be represented explicitly in the modelstructure. We define the name hierarchical dynamical networks torefer to this approach <strong>of</strong> modelling a system as a hierarchical network<strong>of</strong> graph automata.4.1.1 The effects <strong>of</strong> a discrete time modelContinuous models (in the form <strong>of</strong> systems <strong>of</strong> continuous equations)are the standard approach used in many fields, such as physics, chemistry,various engineering disciplines and commerce. The equationsin these models impose constraints upon the variables in the model,which are not always directly solvable. Various numerical methodsfor solving such systems exist, <strong>of</strong>ten by converting the equations intoa discrete form (and thus arriving at a discrete model). The systembeing modelled may in fact be discrete and the equations reflectinga more abstract view <strong>of</strong> the system—for example, although fluid dynamicsare modelled using continuous equations, the fluids beingmodelled consist <strong>of</strong> finite numbers <strong>of</strong> molecules.Unlike existing nephron models, the model presented here usesa discrete time model and the model dynamics are captured by systems<strong>of</strong> difference equations. When continuous systems are solvednumerically—as is the case for a number <strong>of</strong> nephron models (e. g.,[78])—the system is translated into discrete equations. Using thisapproach, solutions are calculated iteratively until a fixed point isfound, which differs from the discrete time approach used here. Thischoice <strong>of</strong> a discrete time model has several advantages for modellinglarge systems, such as the kidney:• The equations explicitly calculate the state evolution <strong>of</strong> themodel, rather than capturing constraints on the state <strong>of</strong> themodel.• Coupling between equations does not affect the computationalcost <strong>of</strong> solving the equations—the computational cost <strong>of</strong> themodel is the sum <strong>of</strong> the computational costs <strong>of</strong> each equation.• The system dynamics are analysed by simulating the state evolution<strong>of</strong> the model, not by obtaining a steady-state solution. Modelparameters and couplings can be changed during a simulation,and the dynamic response <strong>of</strong> the system can be observed.


4.1 M O D E L L I N G A P P ROAC H 31These advantages <strong>of</strong> the discrete-time approach are importantin order to meet our aims. Recall that our focus is on the nature<strong>of</strong> the interactions and couplings in multi-nephron systems. Basedon the assumption that the kidney is a complex system, it is theseinteractions that regulate the system dynamics. By taking a networkapproach to our model, we can explicitly represent these interactionsas edges in the network, and the use <strong>of</strong> discrete-time equations meansthat the equation coupling is equivalent to the explicit coupling in thenetwork. To illustrate this point, consider three state variables x(t),y(t) and z(t), such that:x(t + ∆) = F(x(t), y(t)) (4.1)y(t + ∆) = G(y(t), z(t)) (4.2)z(t + ∆) = H(z(t)) (4.3)Although x(t) is indirectly coupled to z(t) through y(t), the value<strong>of</strong> z(t) and the behaviour <strong>of</strong> H(z(t)) do not affect the calculation <strong>of</strong>x(t). At any given time t i in a model simulation, the value <strong>of</strong> all statevariables x(t i ), y(t i ) and z(t i ) are known. Calculating x(t i + ∆) issimple a matter <strong>of</strong> calculating F(x(t i ), y(t i )), where all the necessaryvalues are already known. The indirect coupling between x(t) andz(t) is realised, as the value <strong>of</strong> z(t i − ∆) affects y(t i ), which in turnaffects x(t i + ∆), but the computational cost <strong>of</strong> the whole system issimply the sum <strong>of</strong> the computational costs <strong>of</strong> the individual equations.The choice <strong>of</strong> the time-step size ∆ is important, as it determineswhich signals the model will be capable <strong>of</strong> reproducing [162] (illustratedin Fig. 8). It also affects the model equations, as the time-stepsize determines what activities occur in a single time-step. Finally,the time-step size may affect the number <strong>of</strong> time-steps that need tobe simulated for the system to reach its steady-state behaviour. Thechoice <strong>of</strong> an appropriate value for ∆ is discussed in Sect. 4.4.1.2SignalSamples (∆=2)1Value0-11 2 3 4 5Time (s)Figure 8: An inappropriate sampling rate can render a model incapable<strong>of</strong> reconstructing signals produced by the original system.


32 M O D E L D E S I G NWhen multiple parts <strong>of</strong> the model must reach an equilibrium in asingle time-step, systems <strong>of</strong> coupled equations can be used to capturethis explicit coupling (such an approach is used in Sect. 4.4.5). Forexample:x(t + ∆), y(t + ∆), z(t + ∆) = F(x(t), y(t), z(t)) (4.4)Simultaneous equations are generally more complex to solve thanindividual equations, but if the equations can be reduced to linearequations the system can be solved by Gaussian elimination. Theclassic algorithm uses O(n 3 ) arithmetic operations to solve for nequations in n variables. Improved algorithms have been developed,reaching O(n 2.807 ) [171] and O(n 2.376 ) [46], but these algorithmstend to have a larger constant and are only better suited to solvingsufficiently large systems <strong>of</strong> equations. So while the use <strong>of</strong> simultaneousequations can increase the computational complexity <strong>of</strong> themodel, the increase in complexity is bounded and can be determinedeasily. For small systems <strong>of</strong> simultaneous equations, the increasein complexity will have negligible impact on the complexity <strong>of</strong> themodel, which is not a guarantee that can be made for small systems<strong>of</strong> coupled differential equations.4.1.2 Analysis <strong>of</strong> discrete network automataPerhaps the single biggest drawback <strong>of</strong> using discrete-time and differenceequations is that the model does not have a “solution” <strong>of</strong>the kind generated by continuous models—instead, the dynamics <strong>of</strong>the model is determined by analysing how the model state evolvesover time. In order to understand the dynamics <strong>of</strong> such a model, itmust be determined when the model has reached its steady-state(which is likely to be a cycle <strong>of</strong> states rather than a single state). Ingeneral, there is no algorithm to determine how quickly a discretemodel will reach a steady-state behaviour (a generalised algorithmis not possible, as it would amount to solving the Halting Problem[182]).The standard approach is to simulate the model from an initial stateand to search for patterns in the model dynamics. Typical analysistechniques for graph automata are inherently visual in nature anddo not scale to large systems (see Sect. 3.1). For complex modelssuch as the nephron model proposed here, alternative analysis techniqueshave to be applied, which are typically statistical in nature.Since many <strong>of</strong> the phenomena exhibited by multi-nephron systemsare oscillatory in nature (e. g., filtration rates and the tubuloglomerularfeedback (TGF) mechanism), signals analysis techniques such asFourier analysis and power spectral densities will be used. Lyapunovstability measures will also be used, to determine how sensitive themodel dynamics are to changes in the system. This notion <strong>of</strong> stabilityor sensitivity to conditions will prove particularly useful for determininghow stable the model’s dynamics are when we incorporate theeffect <strong>of</strong> diseases into the model (Chapter 6).It is possible that the choice <strong>of</strong> initial state will affect the steadystatebehaviour <strong>of</strong> the model (much like the choice <strong>of</strong> initial conditions


4.2 M O D E L S T RU C T U R E 33for systems <strong>of</strong> continuous equations) and the relationship betweenthe steady-state dynamics and the choice <strong>of</strong> initial conditions mustbe considered in order to come to any conclusions concerning thedynamics <strong>of</strong> the model. Similarly, the stability <strong>of</strong> the model in responseto changes in parameters may be harder to determine thanfor continuous models, as different parameter values can be usedin simulations, but there is no general approach to determine thehow dynamics <strong>of</strong> the model vary as a function <strong>of</strong> a parameter value(for exceedingly simple models, such an approach may be possible).Again, statistical methods will be used to measure such properties.4.2 M O D E L S T RU C T U R E4.2.1 The single-nephron modelThe nephron tubule consists <strong>of</strong> a number <strong>of</strong> functionally distinctsegments—typical tubule models focus on the behaviour <strong>of</strong> a singlesuch segment—and the single nephron model reflects this structureby treating the nephron tubule as a sequence <strong>of</strong> nodes, one for eachsuch segment.The nephron does not exist in isolation—the tubule sits in interstitialfluid with which it exchanges solutes (shown in Fig. 2); it isalso surrounded by a network <strong>of</strong> peritubular capillaries that alsoexchange solutes with the interstitial fluid (shown in Fig. 1a). Thesingle-nephron model must therefore take into account both <strong>of</strong> thesestructures. Note that while both the tubule and peritubular capillariesexchange solutes with the interstitial fluid, the tubule usesmany active and secondary transport mechanisms to do so, whilethe peritubular capillaries only undergo diffusion with the interstitialfluid.The interstitial fluid is divided into horizontal layers (as shown inFig. 2) whose compositions differ, and so the single-nephron modelreflects this layered structure. The peritubular capillaries surroundthe tubule, and our model divides these capillaries into segmentsthat correspond to the tubule segments. Rather than using additionalnodes to model these capillaries, each tubule segment node alsomodels the corresponding peritubular capillaries. That is, the state<strong>of</strong> each tubule segment node encapsulates the state <strong>of</strong> the tubulesegment and the state <strong>of</strong> the surrounding peritubular capillaries.The model structure—showing fluid flow along the tubule and theTGF mechanism—is depicted in Fig. 9 and the complete network—including interstitial fluid nodes and edges for solute transport andTGF—is shown in Fig. 10.This network captures the functional topology <strong>of</strong> a nephron, butit is not detailed enough to permit behaviours such as the countercurrentexchange mechanisms in the loop <strong>of</strong> Henle. The model isrefined by dividing the limbs <strong>of</strong> Henle and the medullary fluid into anumber <strong>of</strong> layers (see Fig. 11) in order to model the counter-currentexchange mechanism—a good example <strong>of</strong> hierarchy in a complexsystem. Without dividing the medullary fluid into layers a salt gradientcould not exist, and without dividing the limbs into layers there wouldbe no counter-current exchange mechanism.


34 M O D E L D E S I G NCortexGlomerulusDistalTubuleProximalTubuleTGFOuter MedullaDescendingLimbAscendingThick LimbCollectingDuctsInner MedullaAscendingThin LimbFigure 9: The single-nephron model, showing tubule flow (solid arrows)and the TGF mechanism (dashed arrow). Solute transportsbetween the tubule and the surrounding fluid are not shown.GlomerulusProximalTubuleDescendingLimbAscendingThin LimbAscendingThick LimbDistalTubuleCollectingDuctsCortexOuterMedullaInnerMedullaFluid flow TG feedback Solute transportFigure 10: The single-nephron tubule as a 1D network automata,showing the edges that capture fluid flow, solute transport and TGF.Rather than dividing the nodes into sub-nodes, it would have beenpossible to model the layers as a number <strong>of</strong> state variables in theoriginal nodes. However, this approach would have hidden the hierarchicalstructure <strong>of</strong> the model, preventing the model from explicitlycapturing the layers and transports as nodes and edges. The number<strong>of</strong> layers into which the loop <strong>of</strong> Henle and medullary fluid aredivided is a parameter <strong>of</strong> the model, allowing for control over the balancebetween approximating a continuous counter-current exchangemechanism and having a simple model structure.4.2.2 The multi-nephron modelExisting models <strong>of</strong> two-nephron systems have focused on the interactionsthat arise due to the oscillations in filtration rate and tubulepressure [112, 166, 19, 81, 99, 19], taking into account the regulatoryrole played by the TGF mechanism. These models have reproduced arange <strong>of</strong> experimentally observed phenomena, such as in-phase synchronisation<strong>of</strong> both single-nephron glomerular filtration rate (SNGFR)and pressure oscillations in the proximal tubule. However, thesemodels fail to take into account any interactions that may arise viaother channels, such as changes in the interstitial fluid due to so-


4.2 M O D E L S T RU C T U R E 35Outer MedullaDescending LimbH 2OH 2OH 2OH 2OH 2OH 2ONaNaNaNaNaNaThick AscendingLimbThin AscendingLimb150 mmol/LNa300 mmol/LInner MedullaFigure 11: Solute transport in the loop <strong>of</strong> Henle, demonstrating howthe limbs and the interstitial fluid are divided into layers. Diffusion<strong>of</strong> H 2 O occurs between the descending limb and the interstitial fluid,while Na is actively transported from the ascending limbs into theinterstitial fluid. The division into layers is necessary for the loop <strong>of</strong>Henle to act as a counter-current exchange mechanism. An example<strong>of</strong> typical figures for the sodium gradient in the interstitial fluid [78]are shown on the right.lute secretion or reabsorption by a segment <strong>of</strong> the nephron tubule.The extension <strong>of</strong> the single-nephron network to a two-nephron networkis designed to allow for these kinds <strong>of</strong> interactions to arise andbe analysed—the network automata allows for parameters (such asinteractions and couplings) to be changed during a simulation.A two-nephron model was obtained by connecting two singlenephronmodels together, as shown in Fig. 12. The afferent arteriolesare treated as variable resistors placed in parallel across a fixed potential(the difference between arterial and venous pressure). Thus,by changing the resistance <strong>of</strong> its afferent arteriole, a nephron caninfluence the SNGFR <strong>of</strong> neighbouring nephrons—this interaction isreferred to as hemodynamic coupling. The afferent arterioles areassumed to be non-compliant, to simplify the arterial model.GlomerulusProximalTubuleDescendingLimbAscendingThin LimbAscendingThick LimbDistalTubuleCollectingDuctsCortexOuterMedullaInnerMedullaGlomerulusProximalTubuleDescendingLimbAscendingThin LimbAscendingThick LimbDistalTubuleCollectingDuctsFluid flow TG feedback Solute transportFigure 12: The two-nephron network, showing the edges that capturefluid flow, solute transport and TGF.The multi-nephron models are a straight-forward generalisation <strong>of</strong>the two-nephron model shown in Fig. 12, where multiple nephron


36 M O D E L D E S I G Nautomata share the same column <strong>of</strong> interstitial fluid (such a columnis shown in Fig. 13). Again, the afferent arterioles are assumed to benon-compliant and are treated as variable resistors.CortexOuter MedullaInner MedullaFigure 13: A single column <strong>of</strong> interstitial fluid, showing the division<strong>of</strong> the fluid into layers. Multiple nephron automata can be placed inthis column, sharing the interstitial fluid.For a small number <strong>of</strong> nephrons, it is sufficient to model thenephrons as parallel resistors placed across a potential equal tothe difference between the arterial and vascular pressures. However,larger multi-nephron models must take into account the structure <strong>of</strong>the arterial tree that supplies the nephrons with blood [112].The arterial tree for a column <strong>of</strong> interstitial fluid is modelled asa number <strong>of</strong> branches connected to a main vessel (see Fig. 14).Nephrons connected to the same branch are treated as analogous toresistors connected in parallel (with resistance equal to the resistance<strong>of</strong> the afferent arterioles) over a potential (the difference betweenarterial and venous blood pressure). Nephron resistance is modelledsolely as the afferent arteriole resistance because the tubule and capillariesare non-compliant and therefore have a constant resistance,which we ignore for simplicity. Between each branch is a resistancewhich accounts for the extra distance the blood must travel to reachthe lower branches.4.2.3 The multi-column modelFor sufficiently large multi-nephron models it is not reasonable for all<strong>of</strong> the nephrons to share the same interstitial fluid, as this would allowall <strong>of</strong> the nephrons in the model to directly interact with each other—something that does not occur in sufficiently large systems (suchas the kidney as a whole)—and it enforces a globally-homogeneousinterstitial fluid, which is not physiologically reasonable.To separate nephrons into local neighbourhoods, and to allow theinterstitial fluid to evolve inhomogeneously, the multi-column modeldivides the fluid into a number <strong>of</strong> columns, as shown in Fig. 15.Each column is supplied with blood from an arterial tree that hasan identical structure to the arterial tree inside each column. Thecolumns are arranged on a 2D grid, and may thus neighbour othercolumns.


4.2 M O D E L S T RU C T U R E 37P AP VR BR BNNNNFigure 14: The arterial tree for a multi-nephron model, shown assolid lines. Each node “N” represents a single nephron network automatain the column <strong>of</strong> fluid, whose resistance is equal to that <strong>of</strong>the nephron’s afferent arteriole R A . P A and P V are the arterial andvenous pressures, respectively, and the resistance <strong>of</strong> the arterial vesselsis modelled as additional resistances R B . The dotted lines showthe venous tree, which is not included in the model.P AP VR CR CFigure 15: The arterial tree for multiple columns <strong>of</strong> interstitial fluid.Each column contains an arterial tree (as shown in Fig. 14). P Aand P V are the arterial and venous pressures, respectively, and theresistance <strong>of</strong> the arterial vessels is modelled as additional resistancesR C . The dotted lines show the venous tree, which is not included inthe model.To simulate a single time-step for this model, we simulate the timestepfor each column <strong>of</strong> interstitial fluid, and then permit diffusionbetween neighbouring columns. We make two assumptions concern-


38 M O D E L D E S I G Ning the nature <strong>of</strong> this diffusion. Firstly, we only permit horizontaldiffusion (i.e., there is no diffusion between different layers <strong>of</strong> theinterstitial fluid). Secondly, we only permit the diffusion to act on onequarter <strong>of</strong> the fluid in each column (see Fig. 16)—this prevents anyportion <strong>of</strong> the interstitial fluid from undergoing inter-column diffusionwith more than one column, as each quarter <strong>of</strong> the fluid diffuses withthe neighbour on the nearest side (if any).diffusionFigure 16: At each time-step, diffusion is performed between the fluidlayers in neighbouring columns, where one quarter <strong>of</strong> the fluid ineach layer is allowed to diffuse with a neighbouring layer.Note that the arterial tree and the 2D column arrangement areentirely separate and can be structured independently <strong>of</strong> each other—e. g., columns attached to the same branch <strong>of</strong> the arterial tree neednot be located near each other. As hemodynamic coupling is capturedby the arterial tree and the inter-column diffusion is captured by the2D arrangement, these two mechanisms can be controlled independently,allowing for the interactions between these mechanisms to beinvestigated.4.3 N E T W O R K P RO P E RT I E SNow that the structure <strong>of</strong> the model has been defined, it is possible tocalculate several standard network properties (as defined in Sect. 3.2)that will allow the connectivity properties <strong>of</strong> models <strong>of</strong> differentsizes to be compared, as well as providing a reference point forcomparing the model structure to other network models and realworldnetworks. Note that these properties only take into accountdirect connections between pairs <strong>of</strong> nodes, and can not be usedto determine the likelihood <strong>of</strong> interactions between nodes that arenot directly connected. All nodes are connected via paths across thenetwork, but emergent behaviour cannot be inferred from the networktopology—it must be analysed using simulations. Similarly, althoughnephrons can be coupled (e. g., hemodynamic coupling occurs overarterial tree), these network properties cannot infer any properties<strong>of</strong> the resulting dynamics.The network properties <strong>of</strong> the single-column model are shownin Table 2. Not much can be inferred from the values for a singlenephronnetwork, due to the small size <strong>of</strong> the network (14 nodes).Only when the model is extended to larger numbers <strong>of</strong> nephrons willit be possible to draw some conclusions about the network properties<strong>of</strong> the model. If we consider a two-nephron network (Fig. 12), wesee that in comparison to the single-nephron network, the averagedegree has increased, the connection probability and the clusteringcoefficient have dropped, and the network diameter is unchanged.


4.3 N E T W O R K P RO P E RT I E S 39Note that the clustering coefficient decreases for n < 4 and increasesfor n 4. For n > 5, we have C > p (as shown in Fig. 17), which isan indication that the network is a small-world network.Property n = 1 n = 2 value lim n→∞Mean Degree 3.83 4.8446n7n+56.57Diameter 6 6 6 6Clustering (C) 0.097 0.065n2 + 9n−3 1 4n−21Connection Pr (p) 0.174 0.1357n+50.07123n49n 2 +63n+20Table 2: Properties <strong>of</strong> the n-nephron single-column network.00.18Cp0.120.0601 3 5 7 9nFigure 17: As n increases, the clustering coefficient (C) becomeslarger than the connection probability (p).The same network properties for multi-column models—with c 2columns <strong>of</strong> interstitial fluid arranged in a c × c grid, each containingn nephrons—are shown in Table 3. Note that although the networkdiameter tends to ∞, it grows linearly with the number <strong>of</strong> columnsin the model. There is essentially no difference in the mean degree,clustering coefficient or connection probability between the singlecolumnand c 2 -column networks. In both cases the mean degree tendsto 46/7 ≈ 6.57; the clustering coefficient tends to 1/14 ≈ 0.071; and theconnection probability tends to 0.Similar to the single-column network, we find that as the networksize grows, we have C > p, which again indicates that the model isa small-world network. In fact, for c = 1 we find that C > p whenn > 5 (as per the single-column network) and that for c > 1 we find∀n : C > p. Both values quickly approach their asymptotic values, sothat even for small values <strong>of</strong> n and c the difference between C and pis close to maximal (as shown in Fig. 18).


40 M O D E L D E S I G NProperty Value lim c→∞Mean Degree46nc 2 +12c(c−1)46n+12(7n+5)c 2 7n+5Diameter 6 + (c − 1) 2 ∞Clustering (C)Connection Pr (p)nc 22 + n3n(3n−1)+4−4/c + n2n(2n−1)+4−4/c(7n+5)c 223nc 2 +6c(c−1)(7n+5)c 2 ((7n+5)c 2 −1)n14n+10Table 3: Properties <strong>of</strong> the n-nephron c 2 -column network.00.10.050-0.05-0.113n57 1 23c4Figure 18: As either n or c increases, the difference between theclustering coefficient (C) and the connection probability (p) quicklyapproaches the asymptotic value <strong>of</strong> 1/14 ≈ 0.071, as shown in this plot<strong>of</strong> C − p.4.4 S TAT E S PAC E E Q UAT I O N S A N D U P DAT E RU L E S4.4.1 Choosing an appropriate time-step sizeAs mentioned previously (Sect. 4.1.1), the model uses discrete time.Each time-step maps the current state <strong>of</strong> the model, S(t) to the state<strong>of</strong> the model at the next time-step S(t + ∆) (the time-step size ∆is aparameter <strong>of</strong> the simulation). A simulation <strong>of</strong> the model is achievedby repeatedly applying this mapping, beginning with the initial modelstate. This mapping is realised as a number <strong>of</strong> equations, whichsimulate the fluid flow and solute transports in the model, and updatethe state variables in the model.More importantly, the kidney exhibits multi-rate dynamics—nephronresponses can range from a matter <strong>of</strong> seconds (e. g., the TGF mechanism)to the order <strong>of</strong> hours (e. g., the secretion <strong>of</strong> hormones). Specifically,the SNGFR tends to oscillate with a period <strong>of</strong> 30–40s [78, 166,111] and the TGF signal has a delay <strong>of</strong> 3–5s [166, 145]. The modelaims to recreate these signals, and therefore needs a sampling rate


4.4 S TAT E S PAC E E Q UAT I O N S A N D U P DAT E RU L E S 41capable <strong>of</strong> reproducing these signals [162]. For example, a time-stepsize <strong>of</strong> no greater than 1.5s is required to reconstruct a 3s delay inthe TGF signal. A time-step size <strong>of</strong> one second was ultimately chosen,as this is sufficient to reconstruct signals with frequencies less than1/2Hz [162] and therefore meets our needs.4.4.2 Discrete fluid flow and solute transportDue to the use <strong>of</strong> discrete time, bodies <strong>of</strong> fluid are modelled asdiscrete volumes, and flow along a vessel is treated as a sequence<strong>of</strong> these discrete volumes. A vessel is able to hold a fixed number<strong>of</strong> discrete volumes, and at each time-step an entering volume <strong>of</strong>fluid forces another volume <strong>of</strong> fluid to exit the vessel (depicted inFig. 19)—each vessel acts as a queue. In this model, all vessels (i. e.,the arterial tree, the nephron tubule and the peritubular capillaries)are non-compliant, meaning their radius is fixed regardless <strong>of</strong> thehydraulic pressure <strong>of</strong> the fluid in vessel.v 0v 1v 2v 3v 0v 1v 2v 3Figure 19: Discrete fluid flow along a segment <strong>of</strong> tubule that containsthree discrete volumes <strong>of</strong> fluid (v 1 , v 2 and v 3 ). In one time-step avolume <strong>of</strong> fluid v 0 enters the tubule, causing the volume v 3 to exit thetubule.Existing models have modelled the tubule as compliant [111], modelledtubule segments as having constant resistance [19], and modelledthe tubule and afferent arteriole as non-compliant [99]. Treatingthe tubule and capillaries as non-compliant allows the model to ignorethe effects <strong>of</strong> hydrostatic pressure on resistance to fluid flow.Tubular compliance affects the speed <strong>of</strong> pulse propagation in thetubule, and the TGF time-delay can be increased to compensate fornon-compliance [99].The model does not take into hydrostatic pressure anywhere inthe tubule (but hydrostatic pressure is used to determine blood flowthrough the arterial tree and to calculate SNGFRs). As the tubule isassumed to have no compliance, the pressure will vary in proportionto the flow rate along the tubule. The filtrate enters the tubule indiscrete volumes, each <strong>of</strong> which travels along the tubule at the samerate, rather than mixing with the neighbouring fluid volumes.The tubule segment nodes control solute transport—each nodecontains update rules that govern the transports between the tubulefiltrate, the interstitial fluid and the peritubular capillaries. Tubulesegments that have been divided into sub-nodes (i. e., the loops <strong>of</strong>Henle and the collecting ducts) perform the solute transport andfluid flow in several stages, in order to capture the dynamics <strong>of</strong> thecounter-current exchange mechanism (shown in detail in Fig. 20).The stages progress as follows:


42 M O D E L D E S I G N1. The fluid entering the tubule segment in the time-step is dividedinto sub-volumes, (the specific number is a parameter <strong>of</strong> themodel).2. A sub-volume enters the tubule segment.3. Solute transports are simulated for each sub-node.4. Steps 2 and 3 are repeated until all <strong>of</strong> the sub-volumes haveentered the tubule segment.V aV 0V 1V 1V aV b1V bV aVv bVv ba1Vv abV 2V 2V 1V 11V b1V b2V 3V 3V 2V 21V 11V 12V 3V 3V 3V 21V 21V 3Figure 20: A volume <strong>of</strong> fluid V 0 enters the tubule in the time-step.According to step 1, it is divided into two volumes, V a and V b . V b entersthe tubule, forcing V 3 to flow out <strong>of</strong> the tubule. Solute transportsare then performed for each sub-node (represented by the superscript“1”). V a then enters the tubule, forcing V2 1 to flow out <strong>of</strong> thetubule, and solute transports are performed again (represented bya superscript “2” for volumes that have undergone solute transporttwice, and by a superscript “1” for V a , which has undergone solutetransport once).For the sake <strong>of</strong> simplicity, the model ignores a number <strong>of</strong> factorsthat can affect solute transport and the composition <strong>of</strong> the tubulefiltrate:• Physiological factors such as conductivity, reflection and Michaelis-Menten kinetics.• Chemical processes in the tubule, such as glucose consumption,lactate production and hormone secretion.• Low-level structural features such as the various pathways betweenthe filtrate and the interstitial fluid.• Permeability is only taken into account for solute transports thatare dependant on a hormone concentration (e. g., sodium andwater reabsorption in the distal tubule and collecting ducts).


4.4 S TAT E S PAC E E Q UAT I O N S A N D U P DAT E RU L E S 434.4.3 Model inputThe input for the model consists <strong>of</strong> the blood entering the arterial treeat each time-step. The input is characterised by the concentrations<strong>of</strong> a number <strong>of</strong> solutes—the solutes used in the model are listed inTable 4. Except for water and red blood cells, all solutes are assumedto have a molarity <strong>of</strong> 0—they are not considered to contribute t<strong>of</strong>luid volume. The solute AvgProtein represents a typical protein inthe bloodstream and takes its properties from albumin, the mostabundant plasma protein in humans and other mammals.The input solute concentrations are specified as ranges, having aminimum and maximum value. At each time-step, the input concentrationsolute is chosen at random from the range, where each value inthe range has an equal likelihood <strong>of</strong> being chosen. Model parameterscan also be changed at any time-step in a simulation, making themodel interactive.Solute Molar Mass Has volumeWater 18.01 YesSodium 58.44 NoChlorine 35.453 NoPotassium 39.0983 Noantidiuretic hormone (ADH) 1.5e5 NoAldosterone 360.49 NoRBC 54.05e12 YesAvgProtein 6.5e4 NoTable 4: The solutes support by the model.4.4.4 State variablesThe glomerulus and ascending thick limb nodes have internal statevariables, which are used by the update rules to implement the dynamics<strong>of</strong> these tubule segments. These state variables are listed inTable 5: R A (t), C A (t), C E (t), P G (t), Π G (t), ξ(t) and φ(t) are the statevariables for the glomerulus, and Na M (t) is the state variable for thethick ascending limb. The variables C A (t) and C E (t) are calculatedfrom the composition <strong>of</strong> the plasma at the afferent and efferent arterioles,respectively. Likewise, Na M (t) is the history <strong>of</strong> sodium deliveryto the macula densa and is calculated from the composition <strong>of</strong> thefiltrate leaving the thick ascending limb. The update rules for theremaining state variables are:


44 M O D E L D E S I G NR A (t) = ξ(t) × exp(ΛH A ) (4.5)P G (t) = P V + P A × R E /R A (t) × C A (t)/C E (t)(4.6)(1 + R E /R A (t)) × C A (t)/C E (t)ψφ(t) = ξ max −(4.7)1 + exp(k[Na M (t − D T GF ) − Na 1/2 ])( 2ζφ(t) + ξ(t − ∆)(ω∆ + 2 )ξ(t − 2∆)(ω∆)ξ(t) =2 −(ω∆) 21 + 2ζω∆ + 1(4.8)(ω∆)( )2CA + CΠ G = ΠE, Π(C) = aC + bC 2 (4.9)2The glomerulus state variables, and their associated update rules,are discrete forms <strong>of</strong> the equations for the same variables in [78],while the formula for calculating oncotic pressure Π(C) (and the associatedconstants a, and b) come from Deen et al. [52]. Higher-orderequations for ξ(t) have been used to simulate higher-frequency phenomenon[112, 166], but we have chosen to use the simpler secondorderequation. The model parameters used in these update rules arelisted in Table 6. The value <strong>of</strong> D T GF controls the time delay in theTGF mechanism, and need not be a multiple <strong>of</strong> ∆—the value can becalculated by linear interpolation between time-steps.Hemodynamic coupling arises naturally from the arterial tree network.Vascular signalling, however, does not arise naturally from thearterial tree or from the single-nephron model. To incorporate thiscoupling mechanism into the model, the update rule for the statevariable R A (t) must be adjusted. The strength <strong>of</strong> the TGF signal isproportional to ξ i (t) for each adjacent nephron i on the same branch<strong>of</strong> the arterial tree, and each such signal is assumed to have decayedto a fraction γ <strong>of</strong> the original signal. Thus, the update rule is now:R A (t) =(ξ(t) + ∑ i∈Nγξ i (t))× exp(ΛH A ) (4.10)Vascular signalling has been modelled in existing models by anidentical approach [112, 166].4.4.5 State update rulesThe update rules for each tubule segment node control the flow <strong>of</strong>fluid and the transport <strong>of</strong> solutes along the tubule. A summary <strong>of</strong> thefunction <strong>of</strong> each node is given in Table 7. Solutes are transportedbetween the tubule filtrate and the interstitial fluid, and diffusionoccurs between the interstitial fluid and the peritubular capillaries.The model does not take into account hydrostatic pressure in theinterstitial fluid, so water is reabsorbed directly into the peritubularcapillaries to preserve the interstitial fluid volume.The state space for each node in the model covers the composition<strong>of</strong> the fluid in the node—R N for N solutes, including water—and thenode’s state variables—R S V for S V state variables—resulting in a


4.4 S TAT E S PAC E E Q UAT I O N S A N D U P DAT E RU L E S 45Variable Definition UnitR A (t) Afferent arteriole resistance. mmHg . min / nLC A (t)C E (t)Afferent arteriole protein concentration.Efferent arteriole protein concentration.P G (t) Hydrostatic pressure in theglomerulus.Π G (t) Oncotic pressure in theglomerulus.ξ(t)Afferent arteriole response toTGF.g/Lg/LmmHgmmHgmmHg . min / nLφ(t) The TGF signal. mmHg . min / nLNa M (t) Macula densa sodium delivery. mol/sTable 5: State variables in the nephron model.Parameter Definition UnitP A Arterial pressure. mmHgP V Venous pressure. mmHgP BCK fNa 1/2kD TGFωζH APressure at the start <strong>of</strong> thetubule.Permeability constant for theglomerulus.Sodium delivery for which theTGF response is half-maximal.Range parameter for the TGFsignal.Time delay in sending the TGFsignal.Natural frequency for the TGFmechanism.Damping coefficient for the TGFmechanism.mmHgnonemol/sL/mmolsrad/snoneΛ Viscous resistance coefficient. noneThe average afferent hematocrit.noneTable 6: Model parameters.


46 M O D E L D E S I G NSegmentBehaviourGlomerulusFilters a percentage <strong>of</strong> the plasma into thetubule.Proximal Tubule Reabsorbs 2/3 <strong>of</strong> the tubular Na and H 2 O.Descending LimbAscending LimbDistal TubuleCollecting DuctsDiffuses H 2 O to balance the Na concentrationsin the tubule and the medullary fluid.Actively transports sodium into the medullaryfluid to maintain a gradient between themedullary and tubular Na concentrations.Reabsorbs Na proportional to aldosterone concentrationand reabsorbs H 2 O proportional toADH concentration.Reabsorbs Na proportional to aldosterone concentrationand reabsorbs H 2 O proportional toADH concentration.Table 7: The behaviour <strong>of</strong> each tubule segment in the nephron model.state space <strong>of</strong> R N+S V . This state space is far more complex than thetypical n-state spaces <strong>of</strong> the cellular automata reviewed in Sect. 3.1.For simplicity, the update rules in the model describe only thetranslocation <strong>of</strong> sodium and water as essential components <strong>of</strong> thefiltrate. Equations for other solutes (such as chloride and potassium)can easily be incorporated into future, more complex models. Suchmodels will also need to account for the effects <strong>of</strong> changes in filtrationand reabsorption at the level <strong>of</strong> whole kidney function andthe consequent modulation <strong>of</strong> extracellular fluid volume and bloodpressure. In addition, the update rules take into account the effects <strong>of</strong>the hormones ADH and aldosterone, but neither hormone is secreted,reabsorbed or metabolised in the model at present.The notation used to present the update rules is described in Table8. The update rules are now presented.GlomerulusThe role <strong>of</strong> the glomerulus node is to filter a portion <strong>of</strong> the plasmafrom the blood entering the nephron via the afferent arteriole into thenephron tubule. As such, this node must calculate the SNGFR for thetime-step and divide the incoming blood accordingly. The equationused to calculate SNGFR is a standard equation [71, p. 317]:SNGFR = K f × (P G − P BC − Π G ) (4.11)Proximal tubuleThe proximal tubule reabsorbs 67% <strong>of</strong> the sodium and water in thetubule fluid. The water is transported directly to the capillaries (topreserve interstitial fluid volume), while the sodium is transported


4.4 S TAT E S PAC E E Q UAT I O N S A N D U P DAT E RU L E S 47SymbolNa(t)H 2 O(t)X(t) IX(t) IFX(t) I+IFX(t) RX(t) DX(t) AX(t) MX(t) D+MX(t) M+ANa GDefinitionQuantity <strong>of</strong> sodium at time t (mol).Volume <strong>of</strong> water at time t (L).Quantity <strong>of</strong> X entering the tubule segment.Quantity <strong>of</strong> X in the interstitial fluid that surroundsthe tubule segment.Quantity <strong>of</strong> X entering the tubule segmentand in the interstitial fluid that surrounds thetubule.Quantity <strong>of</strong> X reabsorbed by the tubule segment.Quantity <strong>of</strong> X in the descending limb.Quantity <strong>of</strong> X in the ascending limb.Quantity <strong>of</strong> X in the medullary fluid.Quantity <strong>of</strong> X in the medullary fluid and descendinglimb.Quantity <strong>of</strong> X in the medullary fluid and ascendinglimb.The sodium gradient maintained between theascending limb and medullary fluid (mol/L).Table 8: Notation for model equations.


48 M O D E L D E S I G Ninto the interstitial fluid and diffused into the capillaries. The updaterules are:Na(t) R = 0.67 × Na(t) I (4.12)H 2 O(t) R = 0.67 × H 2 O(t) I (4.13)This constant behaviour is clearly far simpler than existing proximaltubule models (such as [193] and [190]), but it is able to capture some<strong>of</strong> the essential dynamics <strong>of</strong> the tubule segment:• The proximal reabsorption is essentially iso-osmotic [177], sosodium and water are reabsorbed in equal proportion.• The proximal tubule normally reabsorbs about 2/3 <strong>of</strong> the filteredload <strong>of</strong> sodium and water [30, 31, 71, p. 333].Loop <strong>of</strong> HenleThe major difference between the model presented here and existingmodels <strong>of</strong> the loop <strong>of</strong> Henle [76, 175, 108] is that in this model,the reabsorption processes are assumed to maintain an equilibriumbetween the filtrate and the interstitial fluid. The descending limb onlypermits the diffusion <strong>of</strong> water and the ascending limb only permitsthe active transport <strong>of</strong> sodium, as shown in Fig. 21. There are a largenumber <strong>of</strong> other transport mechanisms that are known to occur inthe limbs (e. g., salt secretion in the descending limb [108], lactateproduction [76] and the recycling <strong>of</strong> urea [176]), which are currentlynot modelled.DiMAiNat DiH 2Ot DiH 2ONaNat M H 2 Ot MNat AiH 2Ot AiDescending LimbInterstitial FluidAscending LimbFigure 21: Solute transfer in the loop <strong>of</strong> Henle, showing only thequantities <strong>of</strong> sodium and water present in the nodes.Consider a horizontal slice <strong>of</strong> the loop <strong>of</strong> Henle at time t, as shownin Fig. 21. At each time-step, the model simulates the diffusion <strong>of</strong>water from the descending limb and the active transport <strong>of</strong> sodiumfrom the ascending limb. The model assumes that both the diffusionand active transport reach equilibrium during the time-step. Thismeans that:• Enough water is diffused to balance the ratio <strong>of</strong> sodium to waterin the interstitial fluid and the descending limb.• Enough sodium is actively transported to maintain a gradient<strong>of</strong> Na G mol/L between the interstitial fluid and the ascendinglimb.Given these assumptions and the state <strong>of</strong> slice at time t, we areable to calculate the state at time t + ∆ (before the fluid flow along


4.4 S TAT E S PAC E E Q UAT I O N S A N D U P DAT E RU L E S 49the tubule is simulated). Firstly, the model conserves the total waterand sodium in the system:H 2 O(t) D+M = H 2 O(t + ∆t) D+M (4.14)Na(t) M+A = Na(t + ∆t) M+A (4.15)The descending limb reaches an equilibrium with the medullaryfluid, maintaining a sodium concentration that is identical to thesodium concentration in the medullary fluid:Na(t) M Na(t)=DH 2 O(t + ∆t) M H 2 O(t + ∆t) D(4.16)H 2 O(t + ∆t) D =Na(t) D× H 2 O(t) D+MNa(t) D+M(4.17)H 2 O(t + ∆t) M =Na(t) MNa(t) D+M× (H 2 O(t) D+M ) (4.18)The ascending limb reaches an equilibrium with the medullary fluid,maintaining a sodium concentration that is Na G less than the sodiumconcentration in the medullary fluid:Na(t + ∆t) MH 2 O(t) MNa(t + ∆t) A == Na(t + ∆t) A+ Na G (4.19)H 2 O(t) A H 2O(t) A×H 2 O(t) M+A (4.20)(Na(t) M+A − Na G H 2 O(t) M )Na(t + ∆t) M =H 2O(t) MH 2 O(t) M+A×(Na(t) M+A − Na G H 2 O(t) A )(4.21)Simulations <strong>of</strong> the two-nephron model revealed a limitation <strong>of</strong> theabove equations—this system <strong>of</strong> equations assume that the loop <strong>of</strong>Henle reaches an equilibrium with the medullary fluid in isolation–theequations ignore the possibility that other loops <strong>of</strong> Henle are alsointeracting with the medullary fluid. When two nephrons share thesame interstitial fluid, the above equations cause large oscillationsin the medullary sodium concentration (see Fig. 22), as the loops <strong>of</strong>Henle both compensate for any imbalance in the equilibrium.The solution was for the equations to ensure that each layer <strong>of</strong>interstitial fluid reached equilibrium with all <strong>of</strong> the descending andascending limbs <strong>of</strong> Henle present in the fluid—a change that is onlya generalisation <strong>of</strong> the original equations. To achieve this new equilibrium,it was necessary for the transport equations to be solvedsimultaneously.Consider again a horizontal slice <strong>of</strong> the loop <strong>of</strong> Henle at time t, asshown in Fig. 21. At each time-step, the model simulates the diffusion<strong>of</strong> water from the descending limb and the active transport <strong>of</strong> sodiumfrom the ascending limb. The model assumes that both the diffusionand active transport reach equilibrium. This now means that:• Enough water is diffused to balance the ratio <strong>of</strong> sodium to waterin the interstitial fluid and all <strong>of</strong> the descending limbs.


50 M O D E L D E S I G N320310Sodium (mmol/L)300290280270260250100 110 120 130 140 150TimestepFigure 22: Oscillations in medullary sodium concentration.• Enough sodium is actively transported to maintain a gradient <strong>of</strong>X mol/L between the interstitial fluid and all <strong>of</strong> the ascendinglimbs.For a system <strong>of</strong> n nephrons in a single column <strong>of</strong> interstitial fluid,the equations for the loop <strong>of</strong> Henle are based on the assumption thatthe n nephrons cooperatively maintain the sodium gradient. Eachnephron in the system reabsorbs some volume w i <strong>of</strong> water from thedescending limb Di and some amount s i <strong>of</strong> salt from the ascendinglimb Ai for each i ɛ {1..n}. The state <strong>of</strong> a single layer <strong>of</strong> a loop <strong>of</strong> Henleat time t is shown in Fig. 21. At equilibrium, we obtain three equationsfor the water transport in the descending limbs: the conservationequations (4.22) and (4.23), and the equilibrium equation (4.24).H 2 O(t + ∆) Di = H 2 O(t) Di − w i (4.22)H 2 O(t + ∆) M = H 2 O(t) M + ∑ iw i (4.23)Na(t) Di Na(t)=M(4.24)H 2 O(t + ∆) Di H 2 O(t + ∆) MLikewise, we obtain three equations for the sodium transport in theascending limbs: the conservation equations (4.25) and (4.26), andthe equilibrium equation (4.27).Na(t + ∆) Ai = Na(t) Ai − s i (4.25)Na(t + ∆) M = Na(t) M + ∑ is i (4.26)Na(t + ∆) MH 2 O(t) M= Na(t + ∆) AiH 2 O(t) Ai+ Na G (4.27)Both systems <strong>of</strong> equations (water transport and sodium transport)can be transformed into the form Ax = b, where:


4.4 S TAT E S PAC E E Q UAT I O N S A N D U P DAT E RU L E S 51⎛⎞1 α 1 α 1 · · · α 1α 2 1 α 2 · · · α 2A =⎜.⎝ .... .. ⎟⎠α N α N α N · · · 1⎡ ⎤ ⎡ ⎤ ⎡ ⎤w 1 s 1β 1w 2s 2...β 2x =⎢ . or, b =⎣ .⎥ ⎢ ⎥ ⎢ . ⎦ ⎣ ⎦ ⎣ .⎥⎦w N s N β N(4.28)(4.29)The values <strong>of</strong> α i and β i depend on which system <strong>of</strong> equations weare solving. For the equations governing water reabsorption in thedescending limbs, we find:α i =Na(t) DiNa(t) M + Na(t) Di(4.30)β i = Na(t) MH 2 O(t) Di − Na(t) Di H 2 O(t) MNa(t) M + Na(t) Di(4.31)For the equations governing the active transport <strong>of</strong> sodium in theascending limbs, we find:α i =H 2 O(t) AiH 2 O(t) Ai + H 2 O(t) M(4.32)β i = Na(t) AiH 2 O(t) M − Na(t) M H 2 O(t) AiH 2 O(t) Ai + H 2 O(t) M+ Na (4.33)GH 2 O(t) Ai H 2 O(t) MH 2 O(t) Ai + H 2 O(t) MThe refined model solves both systems <strong>of</strong> linear equations usingGaussian elimination (with partial pivoting, to improve the stability<strong>of</strong> the solutions [179]), allowing equations 4.22, 4.23, 4.25 and 4.26to be solved.These equations enforce the sodium molarity at each macula densato be equal, assuming that the macula densas are located in thesame layer <strong>of</strong> interstitial fluid. This has the effect <strong>of</strong> making the TGFresponse <strong>of</strong> each nephron identical at every time-step, enforcingin-phase synchronisation. Experimental data, however, contradictsthis [167]. A more realistic response is obtained by altering the TGFmechanism to respond to sodium delivery (mmol/s) instead <strong>of</strong> sodiumconcentration, an approach used by Layton et al. [99]. As a result,the values <strong>of</strong> Na 1/2 used by the model are not directly comparable totypical values <strong>of</strong> 30–60 mmol/L [78].Distal tubuleThe first portion <strong>of</strong> the distal tubule—the early distal tubule—is similarin function to the thick ascending limb [71, p. 336], reabsorbingsodium but remaining impermeable to water. This portion <strong>of</strong> the


52 M O D E L D E S I G NADHAldosteroneMinimal 1 pM 0.01 nMHalf-maximal 10 pM 0.3 nMMaximal 100 pM 10 nMTable 9: Distal tubule responses to hormone concentrations in thefiltrate. The ADH response values are from Star et al. [168] and thealdosterone response values are from Christ et al. [40]. Note that theinhibitory effects <strong>of</strong> much higher concentrations [87] are ignored bythis model.tubule also forms part <strong>of</strong> the juxtaglomerular complex, generatingthe TGF signal that provides feedback control over the SNGFR. As illustratedin Fig. 9 and Fig. 10, the thick ascending limb node capturesthis behaviour.The distal tubule node focuses on capturing the dynamics <strong>of</strong> thelate distal tubule. This portion <strong>of</strong> the tubule has the following characteristics[71, p. 337]:• Sodium is reabsorbed at a rate that is regulated by hormones.In particular, the reabsorption rate is proportional to the aldosteroneconcentration in the filtrate.• The permeability <strong>of</strong> the tubule to water is controlled by the concentration<strong>of</strong> ADH. When ADH concentration is high, the tubuleis permeable to water, but in the absence <strong>of</strong> ADH the tubule isessentially impermeable to water.Accordingly, the update rules for the distal tubule reabsorb sodiumand water in amounts that are proportional to the filtrate concentrations<strong>of</strong> aldosterone and ADH, respectively. The minimal, half-maximaland maximal tubule responses are shown in Table 9.The update rules calculate the tubule response to ADH (Eq. 4.34)and aldosterone (Eq. 4.35)—the responses are kept within the range[0 . . . 1] (not shown in the equations)—and water and sodium arereabsorbed accordingly (Eq. 4.36 and Eq. 4.37). Note that theseequations do not account for any time delay between the hormoneconcentrations and the according tubule response [168, 40].R ADH (t) = log 10(ADH(t) I × 10 12 )2R Ald (t) = log 10(Aldosterone(t) I × 10 11 )(3)HH 2 O(t) R = R ADH (t) × H 2 O(t) I − Na(t) 2 O(t) I+IFINa(t) I+IF(4.34)(4.35)(4.36)Na(t) R = R Ald (t) × Na max × Na(t) I (4.37)The volume <strong>of</strong> reabsorbed water H 2 O R (t) is the volume <strong>of</strong> waterrequired to reach equilibrium with the interstitial fluid, multipliedby the tubule response R ADH (t). The quantity <strong>of</strong> reabsorbed sodium


4.5 S U M M A RY 53Na R (t) is the volume <strong>of</strong> filtrate sodium Na I (t) multiplied by the tubuleresponse R Ald (t) and a parameter Na max . This parameter representsthe maximal amount <strong>of</strong> sodium reabsorption that can be achieved bythe tubule, and was kept fixed at 0.9. The effects <strong>of</strong> this parameter onthe model dynamics were not investigated, as the model analysis didnot focus on the urine formation and concentration mechanisms.Collecting ductsThe cortical collecting ducts have functional characteristics similarto those <strong>of</strong> the late distal tubule [71, p. 336]. In particular, theyexhibit the same reabsorption characteristics for sodium and water.The medullary collecting ducts, while differing in function from thecortical ducts in several ways, also exhibit the same reabsorptioncharacteristics for sodium and water [71, pp. 337–8].Accordingly, the update rules for the collecting ducts are the sameas the update rules for the distal tubule—the collecting duct responsesto ADH and aldosterone are calculated (Eq. 4.38 and Eq. 4.39), andwater and sodium are reabsorbed accordingly (Eq. 4.40 and Eq. 4.41).R ADH (t) = log 10(ADH(t) I × 10 12 )2R Ald (t) = log 10(Aldosterone(t) I × 10 11 )(3)HH 2 O(t) R = R ADH (t) × H 2 O(t) I − Na(t) 2 O(t) I+IFINa(t) I+IF(4.38)(4.39)(4.40)Na(t) R = R Ald (t) × Na max × Na(t) I (4.41)4.5 S U M M A RYThis chapter has introduced the model design. A new modellingapproach—hierarchical dynamical networks—was introduced, followedby an evaluation <strong>of</strong> the benefits <strong>of</strong> this approach and anoverview <strong>of</strong> appropriate analysis techniques. The model structurewas then presented, and the structural properties <strong>of</strong> the model wereanalysed. Finally, the state variables and update rules <strong>of</strong> the modelwere documented. An experimental analysis <strong>of</strong> the model dynamics isnow presented in the following chapter.


M O D E L D Y N A M I C S5This chapter presents an analysis <strong>of</strong> the dynamics <strong>of</strong> the model describedin Chapter 4. The experimental design used to investigate thebehaviour <strong>of</strong> the model was as follows:• A single-nephron system was demonstrated to reproduce experimentallyobserved behaviour. Without performing this validation,the results <strong>of</strong> any experiments conducted on multi-nephronsystems would be invalid. The single-nephron system was shownto exhibit regular oscillations in single-nephron glomerular filtrationrate (SNGFR), due to the tubuloglomerular feedback (TGF)mechanism. It was also demonstrated that the system was capable<strong>of</strong> constructing and maintaining a stable sodium gradient inthe medullary fluid, over a wide range <strong>of</strong> sodium concentrations.• A two-nephron system was demonstrated to produce experimentallyobserved behaviour, consistent with existing investigations<strong>of</strong> two-nephron models. Specifically, the synchronisations thatarise between the filtration rates <strong>of</strong> the two nephrons wereanalysed, and the effects <strong>of</strong> hemodynamic coupling and vascularsignalling were shown to produce results similar to thoseobserved in existing models. The two-nephron system also exhibiteda larger, more stable sodium gradient in the medullaryfluid than the single-nephron system. This is consistent with theadditional concentrating power associated with multiple loops<strong>of</strong> Henle that is produced by existing models.• The strength <strong>of</strong> the hemodynamic coupling in the two-nephronsystem was explored for range <strong>of</strong> values for P A and k. By measuringthe Lyapunov stability <strong>of</strong> the two-nephron system fordifferent values <strong>of</strong> these parameters, it was confirmed that bothparameters had equivalent effects on the hemodynamic couplingin the system.• The analysis was then extended to an 8-nephron system, whichproduced a larger, more stable gradient than the two-nephronsystem, again consistent with results obtained from existingmodels.• The effects <strong>of</strong> hemodynamic coupling and vascular signallingon the 8-nephron system were then investigated. Rather thananalysing the dynamics <strong>of</strong> the individual nephrons, the analysisfocused on whole-system properties such as the whole-systemfiltration rate (WSFR). Power spectral densities were used todetermine the dominant frequencies present in the WSFR. Theresults showed that the degree <strong>of</strong> vascular signalling present inthe system not only affected the average WSFR, but also alteredthe dominant frequencies.• The analysis was finally extended to determine the effects <strong>of</strong>vascular signalling on a 72-nephron multi-column system. Again,55


56 M O D E L DY NA M I C Sthe analysis did not focus on single-nephron dynamics, butrather on whole-column filtration rate (WCFR) and WSFR. Thepower spectral densities <strong>of</strong> the WCFR and WSFR again demonstratedthat the degree <strong>of</strong> vascular signalling did not only affectthe average filtration rates, but also affected the dominant frequencies.For example, in the absence <strong>of</strong> vascular signalling theWSFR exhibited a dominant low frequency. However, when vascularsignalling was introduced, the WSFR exhibited significanthigher-frequency signals.This model is also available via the <strong>Kidney</strong>Grid portal, which letsresearchers access a number <strong>of</strong> kidney models from around the world[42, 43]. The purpose <strong>of</strong> the <strong>Kidney</strong>Grid is to support collaborativeresearch and shared access to knowledge, both within the renal sciencecommunity and for external audiences (e. g., for educationalpurposes), without depending on single laboratories or specific programmingenvironments [42]. Currently, <strong>Kidney</strong>Grid users are ableto run simulations over ranges <strong>of</strong> parameters for two-nephron andeight-nephron systems. Visualisation is provided in the form <strong>of</strong> graphs<strong>of</strong> time-series data such as SNGFR and WSFR.The experimental results obtained from the experiments outlinedabove are now presented.5.1 T H E S I N G L E-NEPHRO N M O D E LTo validate the dynamics <strong>of</strong> the single-nephron model, the analysisfocused on the regulation <strong>of</strong> SNGFR mediated by the TGF mechanism,and the ability <strong>of</strong> the loop <strong>of</strong> Henle to construct and maintain a sodiumgradient in the medullary fluid. Throughout this analysis, the soluteconcentrations in the blood are—unless stated otherwise—as shownin Table 10.Solute Min (mmol/L) Max (mmol/L) SourceSODIUM 135 145 [96]CHLORIDE 95.0 103.0 [96]POTASSIUM 3.5 5.0 [96]Molecule Min (g/L) Max (g/L) SourceADH 0.0 4.7 × 10 −9 [1]ALDOSTERONE 5.0 × 10 −8 2.5 × 10 −7 [1]PROTEIN 54.0 54.0 [19]Table 10: Steady-state concentration levels.Experimental studies <strong>of</strong> SNGFR regulation (e. g., [103, 102, 78]) haveobserved regular oscillations in SNGFR, and tubule models support thehypothesis that these oscillations are caused by the TGF mechanism([78, 19]).The single-nephron model produced regular oscillations in SNGFRover a wide range <strong>of</strong> parameter values. These oscillations are driven


5.1 T H E S I N G L E-NEPHRO N M O D E L 57by the TGF mechanism, so a decrease in the strength <strong>of</strong> the feedbackshould result in smaller oscillations [78]. The single-nephron modelproduced this behaviour—plots <strong>of</strong> SNGFR for different values <strong>of</strong> theparameter k (the range parameter for the TGF signal) are shown inFig. 23—indicating that the model has successfully captured thisfeature <strong>of</strong> the TGF mechanism. The dynamics <strong>of</strong> the SNGFR were notinvestigated further for the single-nephron model, as the purpose<strong>of</strong> the analysis was to validate the behaviour <strong>of</strong> the single-nephronmodel, not to exhaustively explore the range <strong>of</strong> dynamics that it couldproduce.28282727SNGFR (nL / min)262524232221SNGFR (nL / min)26252423222120201950 100 150 200 250 300 350 400 450 5001950 100 150 200 250 300 350 400 450 500TimestepTimestep(a) k = 0.1 and P A = 110(b) k = 0.2 and P A = 11028282727SNGFR (nL / min)262524232221SNGFR (nL / min)26252423222120201950 100 150 200 250 300 350 400 450 5001950 100 150 200 250 300 350 400 450 500TimestepTimestep(c) k = 0.3 and P A = 110(d) k = 0.4 and P A = 110Figure 23: Regular oscillations in SNGFR.The construction and maintenance <strong>of</strong> a solute gradient in themedullary fluid is <strong>of</strong> great importance to the regulation <strong>of</strong> waterand sodium levels in the body. For example, this gradient allows thekidney to reabsorb a maximal amount <strong>of</strong> water when the body isdehydrated, producing highly concentrated urine [71, pp. 353]. Thisexistence <strong>of</strong> this gradient is due to the ability <strong>of</strong> the loop <strong>of</strong> Henleto act as a counter-current exchange mechanism [71, pp. 352]. Accordingly,the ability <strong>of</strong> the single-nephron model to construct andmaintain a gradient in the medulla was investigated.Figure 24 shows the medullary solute gradients for the systemswhose SNGFRs are shown in Fig. 23. The larger oscillations in SNGFRresult in larger variations in the magnitude <strong>of</strong> the salt gradient, but allfour systems demonstrate an ability to construct and maintain a stablesalt gradient in the medullary fluid. The magnitude <strong>of</strong> the gradientis at least an order <strong>of</strong> magnitude smaller than the salt gradientsobserved experimentally—the gradient between the top and bottom<strong>of</strong> the loop can be as large as 1200mmol/L in human kidneys [169]and over 2000mmol/L in other species [158, 76]. This difference inmagnitude is caused by three factors:


58 M O D E L DY NA M I C S• It does not take into account other solutes such as chloride andurea (e. g., if chloride were reabsorbed in quantities similar tosodium, the magnitude <strong>of</strong> the gradient would be doubled);• It does not have the concentrating power <strong>of</strong> multiple loops <strong>of</strong>Henle [97, 98, 76, 108]—this increase in concentrating poweris demonstrated in Sect. 5.2; and• It does not take into account other processes such as the production<strong>of</strong> lactate in the tubule, which may contribute significantlyto accumulation <strong>of</strong> sodium [76].Given these limitations, the ability <strong>of</strong> the model to maintain sucha gradient demonstrates that the equations for the loop <strong>of</strong> Henle(Sect. 4.4) are able to capture the basic behaviour <strong>of</strong> the countercurrentexchange mechanism.7070Sodium Gradient (mmol/L)605040302010Sodium Gradient (mmol/L)605040302010050 100 150 200 250 300 350 400 450 500050 100 150 200 250 300 350 400 450 500TimestepTimestep(a) k = 0.1 and P A = 110(b) k = 0.2 and P A = 1107070Sodium Gradient (mmol/L)605040302010Sodium Gradient (mmol/L)605040302010050 100 150 200 250 300 350 400 450 500050 100 150 200 250 300 350 400 450 500TimestepTimestep(c) k = 0.3 and P A = 110(d) k = 0.4 and P A = 110Figure 24: Construction and maintenance <strong>of</strong> a medullary sodiumgradient. The graphs show the difference in sodium concentrationbetween the outermost and innermost layers <strong>of</strong> medullary fluid.The stability <strong>of</strong> the sodium gradient across the medullary fluid wasfurther examined by changing the sodium delivery to the maculadensa, to provoke minimal and maximal TGF responses. In orderto change the sodium delivery, blood sodium levels were decreasedto 110 mmol/L and increased to 170 mmol/L. These levels exceedphysiologically normal levels—the body rigorously regulates the bloodsodium concentration—and only serve to illustrate the range <strong>of</strong> theTGF response.In all <strong>of</strong> the experiments shown in Fig. 24, the blood sodium concentrationwas 135–145 mmol/L (as listed in Table 10), and so therange <strong>of</strong> sodium delivery to the macula densa did not explore thelimits <strong>of</strong> the TGF response in the model. The stability <strong>of</strong> the sodium


5.1 T H E S I N G L E-NEPHRO N M O D E L 59gradient across the medullary fluid was further examined by changingthe sodium delivery to the macula densa, to provoke minimal andmaximal TGF responses. In order to change the sodium delivery, twoexperiments were conducted. In the first experiment, blood sodiumconcentration was decreased to 110 mmol/L; in the second experiment,blood sodium concentration was increased to 170 mmol/L. Theresults <strong>of</strong> these two experiments are now presented.For the first experiment, the blood sodium concentration was decreasedby 5 mmol/L every 500 time-steps (the decreases in bloodsodium concentration over time are listed in Table 11). The resultingSNGFR and medullary salt gradient are shown in Fig. 25. As the bloodsodium concentration decreased, so too did the sodium delivery tothe macula densa, and the TGF mechanism responded by dilating theafferent arteriole and increasing the SNGFR. As the SNGFR approachedits maximal value, the size <strong>of</strong> the oscillations decreased.Despite the decrease in blood sodium concentration, the medullarysalt gradient increased by only 11% when the blood sodium concentrationreached the minimal range <strong>of</strong> 110–120 mmol/L. This gradientincrease arose because, although the sodium concentration in eachlayer <strong>of</strong> medullary fluid decreased, the sodium concentration in thetop layers (the outer medulla) decreased more than the sodium concentrationin the bottom layers (the inner medulla). The difference insodium concentrations for these layers <strong>of</strong> medullary fluid are shownin Fig. 26.In the second experiment, the blood sodium concentration wasincreased by 5 mmol/L every 500 time-steps—the blood sodium concentrationranges are listed in Table 12. The resulting SNGFR andmedullary salt gradient are shown in Fig. 27. As the blood sodiumconcentration increased, so too did the sodium delivery to the maculadensa, and the TGF mechanism responded by constricting the afferentarteriole and decreasing the SNGFR. As the SNGFR approached itsminimal value, the size <strong>of</strong> the oscillations decreased.Despite the increase in blood sodium concentration, the medullarysalt gradient decreased by 11% when the blood sodium concentrationreached the maximal range <strong>of</strong> 160–170 mmol/L. This gradientdecrease arose because, although the sodium concentration in eachlayer <strong>of</strong> medullary fluid increased, the sodium concentration in thetop layers (the outer medulla) increased less than the sodium concentrationin the bottom layers (the inner medulla). The difference insodium concentrations for these layers <strong>of</strong> medullary fluid are shownin Fig. 28.In conclusion, this analysis <strong>of</strong> the single-nephron model dynamicsdemonstrate that the model is able to reproduce several patterns<strong>of</strong> behaviour that have been experimentally established. The singlenephronmodel exhibited regular oscillations in SNGFR whose amplitudeand frequency were comparable to both experimental data andexisting models. The model was also able to create and maintain astable salt gradient in the medullary fluid, over a wider range <strong>of</strong> bloodsodium concentrations than is physiologically reasonable.


60 M O D E L DY NA M I C STime-step Min (mmol/L) Max (mmol/L)0. . . 499 135 145500. . . 999 130 1401000. . . 1499 125 1351500. . . 1999 120 1302000. . . 2499 115 1252500. . . 3000 110 120Table 11: Systematic decrease blood sodium concentration.2970SNGFR (nL / min)2827262524232221Sodium Gradient (mmol/L)605040302010201000 2000 3000Timestep00 1000 2000 3000Timestep(a) The SNGFR(b) The medullary salt gradientFigure 25: The single-nephron dynamics in response to decreasedblood sodium concentration (k = 0.3 and P A = 110).190180Outer medullaInner medullaSodium (mmol/L)170160150140130120110500 1000 1500 2000 2500 3000TimestepFigure 26: The change in sodium concentration in the top (outermedulla) and bottom (inner medulla) layers <strong>of</strong> medullary fluid, inresponse to decreased blood sodium concentration.


5.1 T H E S I N G L E-NEPHRO N M O D E L 61Time-step Min (mmol/L) Max (mmol/L)0. . . 499 135 145500. . . 999 140 1501000. . . 1499 145 1551500. . . 1999 150 1602000. . . 2499 155 1652500. . . 3000 160 170Table 12: Systematic increase in blood sodium concentration.2870SNGFR (nL / min)2726252423222120Sodium Gradient (mmol/L)605040302010191000 2000 3000Timestep00 1000 2000 3000Timestep(a) The SNGFR(b) The medullary salt gradientFigure 27: The single-nephron dynamics in response to increasedblood sodium concentration (k = 0.3 and P A = 110).210200Outer medullaInner medullaSodium (mmol/L)190180170160150140500 1000 1500 2000 2500 3000TimestepFigure 28: The change in sodium concentration in the top (outermedulla) and bottom (inner medulla) layers <strong>of</strong> medullary fluid, inresponse to increased blood sodium concentration.


62 M O D E L DY NA M I C S5.2 T W O-NEPHRO N S Y S T E M SWith the single-nephron model dynamics validated, the focus <strong>of</strong> theanalysis turns to multi-nephron systems, which allow the interactionsand couplings between nephrons to be investigated. The multinephronanalysis begins with the simplest multi-nephron system—-atwo-nephron model, where each nephron has an identical structureand location in the renal fluid layers, and vascular signalling is absent(γ = 0).The first experiment investigated the effects that changing thearterial pressure (P A ) and the steepness <strong>of</strong> the TGF response (k) haveon the SNGFRs <strong>of</strong> the two nephrons. The state-space <strong>of</strong> P A × k wasexplored for values <strong>of</strong> P A ranging from a normal level <strong>of</strong> 108.5 mmHg[78] to a maximum <strong>of</strong> 114 mmHg in 0.5 mmHg increments, and forvalues <strong>of</strong> k ranging from 0.25 to 0.4 [78] in increments <strong>of</strong> 0.01. AsP A or k was increased, the two nephrons were observed to changefrom in-phase synchronisation to anti-phase synchronisation, with aregion in the middle where the nephrons were neither in-phase oranti-phase. This trend is summarised in Fig. 29, which shows themean absolute difference in filtration rates <strong>of</strong> the two nephrons overthe parameter space. This demonstrates the increasing the arterialpressure (P A ) and increasing the steepness <strong>of</strong> the TGF response (k)have similar effects on the two-nephron system, as suggested in [19].∆ (nL/min)20151050108110112P A (mmHg)0.400.360.320.28114 0.24 k (L/mmol)Figure 29: Mean difference in filtration rates in a two-nephron system,over the parameter space (P A , k).To measure the stability <strong>of</strong> the model in response to these parameterchanges, Lyapunov exponents were calculated about severalpoints in the phase space. An exponent L y for point y around pointx indicates that the behaviour <strong>of</strong> the system at y approaches thebehaviour <strong>of</strong> the system at x no slower than exp L y. The negativeLyapunov exponents (i. e., those where the behaviours <strong>of</strong> the two systemsconverge) about each <strong>of</strong> the chosen points are shown in Fig. 30.These graphs show that the system exhibits stable behaviour alongthe lines P A − k = C, for some fixed C. This further reinforces the


5.2 T W O-NEPHRO N S Y S T E M S 63hypothesis that P A and k have equivalent effects on the system [19],as an increase <strong>of</strong> one can be countered by a decrease in the other,and vice versa.-0.04-0.02L y-0.000.40L y(a) Lyapunov exponents at P A = 110.0, k = 0.33108110112P A (mmHg)114 0.24 0.32k (L/mmol)-0.04-0.02L y-0.000.40L y(b) Lyapunov exponents at P A = 113.0, k = 0.35108110112P A (mmHg)114 0.24 0.32k (L/mmol)Figure 30: Lyapunov stability over the parameter space (P A , k).The behaviour <strong>of</strong> the model was also examined a wider range <strong>of</strong>arterial pressures P A (80–130 mmHg). At high pressures (115–130mmHg) the model exhibited anti-phase synchronisations, and dueto the non-compliant nature <strong>of</strong> the model, the difference betweenminimal and maximal filtration rates rose to 23 nL/min at P A = 130—shown in Fig. 31a. At low pressures (80–100 mmHg) the model exhibitedin-phase synchronisations, and the difference between minimaland maximal filtration rates decreased to 2.5 nL/min at P A = 80mmHg, again due to the non-compliant nature <strong>of</strong> the model—shownin Fig. 31b.


64 M O D E L DY NA M I C S552850SNGFR (nL / min)454035SNGFR (nL / min)272630500 600 700 800 900 100025500 600 700 800 900 1000TimestepTimestep(a) P A = 130 mmHg.(b) P A = 80 mmHg.Figure 31: The change in SNGFR in response to changes in P A .Anti-phase synchronisations have not been observed experimentally[167], as the vascular coupling between nephrons in the kidney overwhelmsthe hemodynamic coupling [112]. The two-nephron model hasbeen shown to produce anti-phase synchronisations in the absence<strong>of</strong> vascular coupling, and it will now be shown that incorporatingvascular coupling into the model is sufficient to prevent anti-phasesynchronisations from arising, which is consistent with the resultspresented by Marsh et al. [112] and further validates the model.As hemodynamic and vascular coupling reinforce opposing behaviours,the two-nephron model is a suitable candidate for analysinghow the combination <strong>of</strong> these opposing mechanisms affect nephronbehaviour. The analysis so far has been conducted with hemodynamiccoupling present in the model (via the vascular network) and withoutany vascular coupling (γ = 0). As expected, the hemodynamiccoupling enforced anti-phase synchronisation between the two filtrationrates when the coupling was sufficiently strong (see Fig. 29).Accordingly, the two-nephron system exhibited regular anti-phaseoscillations for a range <strong>of</strong> P A and k values.To investigate the ability <strong>of</strong> vascular coupling to prevent anti-phasesynchronisation, the values P A = 110.5 mmHg and k = 0.3 were chosen,and the SNGFRs <strong>of</strong> the two nephrons were analysed over a range<strong>of</strong> γ values. The results are shown in Fig. 32, which demonstratesthat as γ is increased, the anti-phase nature <strong>of</strong> the SNGFR oscillationsis decreased until the system exhibits in-phase synchronisation. Themagnitude <strong>of</strong> the oscillations also decrease as γ is increased, sincevascular signalling causes the TGF signals to each afferent arterioleto approach the same value (i. e., when γ = 1 the signals will beidentical).If the strength <strong>of</strong> the hemodynamic coupling is increased—that is,if one or both <strong>of</strong> P A and k are increased—in the presence <strong>of</strong> vascularsignalling, then the anti-phase synchronisation becomes more prevalent.This is illustrated in Fig. 33, where k has been increased from0.2 to 0.3. Under these conditions, the effect <strong>of</strong> γ = 0.010 is similar tothe effect <strong>of</strong> γ = 0.005 when k = 0.2 (compare Fig. 32a and Fig. 33a).Similarly, for the case <strong>of</strong> γ = 0.11 the dynamics are similar to theeffect <strong>of</strong> γ = 0.008 when k = 0.2 (compare Fig. 32b and Fig. 33b).With two identical nephrons, the coupling mechanisms competeto force in-phase and anti-phase synchronisations on the system. Ifsome asymmetry is introduced into the model (e. g., if the time delay


5.2 T W O-NEPHRO N S Y S T E M S 6534343232SNGFR (nL / min)302826SNGFR (nL / min)302826242422500 600 700 800 900 100022500 600 700 800 900 1000TimestepTimestep(a) γ = 0.005: regular anti-phase oscillationsare present.(b) γ = 0.008: anti-phase oscillationsare less regular than for γ = 0.005.34323432SNGFR (nL / min)302826SNGFR (nL / min)302826242422500 600 700 800 900 100022500 600 700 800 900 1000TimestepTimestep(c) γ = 0.010: little evidence <strong>of</strong> in-phaseor anti-phase oscillations.(d) γ = 0.012: in-phase oscillations arepresent.Figure 32: The effect <strong>of</strong> varying the strength <strong>of</strong> vascular couplingbetween two identical nephrons, where P A = 110.5 mmHg and k =0.2.34343232SNGFR (nL / min)302826SNGFR (nL / min)302826242422500 600 700 800 900 100022500 600 700 800 900 1000TimestepTimestep(a) γ = 0.010: anti-phase oscillationsare present, similar to the behaviourwhen k = 0.2 and 0.005 < γ < 0.008.(b) γ = 0.011: anti-phase oscillationsare present, similar to the behaviourwhen k = 0.2 and γ = 0.008.Figure 33: The effect <strong>of</strong> varying the strength <strong>of</strong> vascular couplingbetween two identical nephrons, where P A = 110.5 mmHg and k =0.3.in one <strong>of</strong> the TGF mechanisms is altered), then the TGF responses <strong>of</strong>the two nephrons will be driven to different frequencies. The effect<strong>of</strong> such an asymmetry was investigated and the results are shown inFig. 34, where an additional two second delay was added to the TGFmechanism <strong>of</strong> one <strong>of</strong> the nephrons.As was the case for two identical nephrons, without any vascularcoupling the hemodynamic coupling forces an anti-phase synchronisationbetween the two SNGFRs, despite the different time delays inthe two TGF mechanisms. As the vascular coupling is increased, themagnitude <strong>of</strong> the oscillations decrease and become less regular—as


66 M O D E L DY NA M I C SSNGFR (nL / min)34333231302928272625242322500 600 700 800 900 1000SNGFR (nL / min)34333231302928272625242322500 600 700 800 900 1000TimestepTimestep(a) γ = 0.005 and identical TGF delays.(b) γ = 0.005 and different TGF delays.31313030SNGFR (nL / min)292827SNGFR (nL / min)29282726500 600 700 800 900 100026500 600 700 800 900 1000TimestepTimestep(c) γ = 0.008 and identical TGF delays.(d) γ = 0.008 and different TGF delays.3030SNGFR (nL / min)2928SNGFR (nL / min)292827500 600 700 800 900 100027500 600 700 800 900 1000TimestepTimestep(e) γ = 0.010 and identical TGF delays.(f) γ = 0.010 and different TGF delays.3030SNGFR (nL / min)2928SNGFR (nL / min)292827500 600 700 800 900 100027500 600 700 800 900 1000TimestepTimestep(g) γ = 0.012 and identical TGF delays.(h) γ = 0.012 and different TGF delays.2929SNGFR (nL / min)28SNGFR (nL / min)2827500 600 700 800 900 100027500 600 700 800 900 1000TimestepTimestep(i) γ = 0.020 and identical TGF delays.(j) γ = 0.020 and different TGF delays.Figure 34: The effect <strong>of</strong> varying the strength <strong>of</strong> vascular couplingbetween two nephrons, whose TGF delays differ by two seconds.


5.3 A N I N H O M O G E N E O U S 8-NEPHRO N S Y S T E M 67shown in Fig. 34d for γ = 0.08—in the same manner as for the system<strong>of</strong> two identical nephrons. The two SNGFRs approach in-phase synchronisationas the vascular coupling is further increased. However, unlikethe system <strong>of</strong> two identical nephrons, some differences between thetwo SNGFRs remain even when γ = 0.20 (shown in Fig. 34j). Thus, thedifferent delays in the TGF mechanism prevent the two nephrons fromexhibiting identical SNGFRs.Now that the interactions between the hemodynamic and vascularcoupling mechanisms have been investigated, the analysis turns tothe medullary salt gradient produced by the two-nephron model.When the blood sodium concentration was within the normal range<strong>of</strong> 135–145 mmol/L, the two-nephron model was able to maintain astable salt gradient 46% larger than the gradient produced by thesingle-nephron model, as shown in Fig. 35. This increase in the saltgradient is due to the increased concentrating power <strong>of</strong> multipleloops <strong>of</strong> Henle.The two-nephron system also produced a more stable salt gradientthan the single-nephron system—Fig. 36 shows the salt gradientwhen blood sodium concentration was decreased as per Table 11, andFig. 37 shows the salt gradient when the blood sodium concentrationwas increased as per Table 12. In response to these changes inblood sodium concentration, the salt gradient in the two-nephronsystem exhibited smaller variations than the single-nephron system.These variations are summarised in Table 13. This result suggeststhat adding more nephrons to the model results in a larger, morestable gradient, which is consistent with existing medullary models[97, 98, 76, 108].5.3 A N I N H O M O G E N E O U S 8 - N E P H RO N S Y S T E MThe two-nephron system was used to analyse the interaction betweenthe competing coupling mechanisms <strong>of</strong> hemodynamic couplingand vascular signalling—and also to investigate the stability <strong>of</strong> themedullary salt gradient maintained by multiple loops <strong>of</strong> Henle—in thesimplest possible system. The analysis is now extended to an eightnephronsystem that consists <strong>of</strong> four nephron pairs, each pair havinglonger loops <strong>of</strong> Henle than the last. The structure <strong>of</strong> this system isshown in Fig. 38.The medullary salt gradient <strong>of</strong> the system was investigated over thesame changes in blood sodium concentration as the single-nephronand two-nephron systems (the changes are listed in Table 11 andTable 12). In comparison to these systems, the 8-nephron systemmaintained a larger, more stable medullary gradient, as illustratedin Fig. 39. The sodium concentrations at the top (outer medulla) andbottom (inner medulla) <strong>of</strong> the medullary fluid are shown in Fig. 40,and the oscillations in these concentrations are much smaller thanthose in the previous systems (e. g., compare to Fig. 36 and Fig. 37).This is further evidence that the combined dynamics <strong>of</strong> multipleloops <strong>of</strong> Henle result in a larger medullary gradient, which is againconsistent with existing medullary models [97, 98, 76, 108]. As shownin Fig. 41, in response to these changes in blood sodium concentration


68 M O D E L DY NA M I C SSodium Gradient (mmol/L)70605040302010Sodium (mmol/L)220210200190180170160150140130Outer medullaInner medulla00 1000 2000 3000120500 1000 1500 2000 2500 3000TimestepTimestep(a) The magnitude <strong>of</strong> the salt gradient.(b) The salt concentration in the innerand outer medulla.Figure 35: The medullary sodium gradient in a two-nephron systemwhen blood sodium concentration remains at 135–145 mmol/L.Sodium Gradient (mmol/L)70605040302010Sodium (mmol/L)220210200190180170160150140130Outer medullaInner medulla00 1000 2000 3000120500 1000 1500 2000 2500 3000TimestepTimestep(a) The magnitude <strong>of</strong> the salt gradient.(b) The salt concentration in the innerand outer medulla.Figure 36: The medullary sodium gradient in a two-nephron systemwhen blood sodium concentration progressively drops to 110–120mmol/L.70220Sodium Gradient (mmol/L)60504030201000 1000 2000 3000Sodium (mmol/L)210200190180170160150140130120Outer medullaInner medulla500 1000 1500 2000 2500 3000TimestepTimestep(a) The magnitude <strong>of</strong> the salt gradient.(b) The salt concentration in the innerand outer medulla.Figure 37: The medullary sodium gradient in a two-nephron systemwhen blood sodium concentration progressively rises to 160–170mmol/L.


5.3 A N I N H O M O G E N E O U S 8-NEPHRO N S Y S T E M 69CortexOuterMedullaInnerMedullaFigure 38: Structure <strong>of</strong> the 8-nephron model.the WSFR increases and decreases in the same manner as the SNGFRin the single-nephron and two-nephron systems.Note that while this 8-nephron system has exhibited a largermedullary salt gradient than the previous single-nephron and twonephronsystems, the increase is relatively small—21%—when comparedto the increase <strong>of</strong> 46% that was observed between the singlenephronand two-nephron systems. The likely causes for this are:the 8-nephron system may have approached the maximal medullarygradient that can be produced by the equations in Sect. 4.4.5; andthe value <strong>of</strong> Na 1/2 (see Table 6) is the same for each nephron, whichmay be unreasonable as the sodium delivery to the macula densa willdiffer depending on the length <strong>of</strong> the loop <strong>of</strong> Henle.While the two-nephron system was used to investigate the effects<strong>of</strong> vascular signalling on the synchronisations between individualnephrons, the 8-nephron system was used to investigate the effects<strong>of</strong> vascular signalling on the WSFR <strong>of</strong> a system <strong>of</strong> inhomogeneousnephrons. Figure 42 shows the WSFR <strong>of</strong> the 8-nephron system forseveral values <strong>of</strong> γ, the percentage <strong>of</strong> the TGF signal that leaks toneighbouring nephrons. These simulations demonstrated that thepresence <strong>of</strong> vascular signalling decreased the average WSFR and alsoled to smaller oscillations in WSFR.These trends are not surprising, since vascular signalling increasesthe TGF signal sent to each nephron—the nephron receives a fraction<strong>of</strong> neighbouring TGF signals in addition to its own TGF signal—whichleads to a decrease in the SNGFR <strong>of</strong> each nephron. And as shown inSect. 5.1, the size <strong>of</strong> the oscillations in SNGFR are decreased whenthe sodium delivery to the macula densa differs greatly from thehalf-maximal value Na 1/2 (see Fig. 3). Accordingly, increasing thedegree <strong>of</strong> vascular signalling in the 8-nephron system decreases theaverage SNGFR and the size <strong>of</strong> the oscillations in SNGFR, which resultsin similar effects on the WSFR.However, the power spectral densities <strong>of</strong> the different WSFRs (shownin Fig. 43) demonstrate that the degree <strong>of</strong> vascular signalling has amore complex effect on WSFR than simply reducing the average WSFRand the magnitude <strong>of</strong> WSFR oscillations. The different peaks exhibitedby the power spectral densities indicate that different dominant frequenciesare present in the WSFR, depending on the value <strong>of</strong> γ. Thismeans that the vascular signalling has affected the SNGFR frequencies<strong>of</strong> the individual nephrons in the system, due to vascular signalling


70 M O D E L DY NA M I C SSodium (mmol/L) Single nephron Two nephrons110–120 +9.8% +0.6%160–170 −13.0% −7.5%Table 13: The difference in the sodium gradient at standard (135–145mmol/L) and altered blood sodium concentrations, which demonstratesthat the two-nephron system maintains a more stable gradientthan the single-nephron system.7070Sodium Gradient (mmol/L)605040302010Sodium Gradient (mmol/L)60504030201000 1000 2000 300000 1000 2000 3000TimestepTimestep(a) Blood sodium concentration progressivelydropped to 110–120 mmol/L.(b) Blood sodium concentration progressivelyraised to 160–170 mmol/L.Figure 39: The medullary salt gradient in an eight-nephron systemwhen blood sodium concentration is altered.Sodium (mmol/L)230220210200190180170160150140130120110Outer medullaInner medulla500 1000 1500 2000 2500 3000Sodium (mmol/L)230220210200190180170160150140130120110Outer medullaInner medulla500 1000 1500 2000 2500 3000TimestepTimestep(a) Blood sodium concentration progressivelydropped to 110–120 mmol/L.(b) Blood sodium concentration progressivelyraised to 160–170 mmol/L.Figure 40: The salt concentration in the inner and outer medulla <strong>of</strong>an eight-nephron system when blood sodium concentration is altered.210210200200WSFR (nL/min)190180170WSFR (nL/min)190180170160160150500 1000 1500 2000 2500 3000150500 1000 1500 2000 2500 3000TimestepTimestep(a) Blood sodium concentration progressivelydropped to 110–120 mmol/L.(b) Blood sodium concentration progressivelyraised to 160–170 mmol/L.Figure 41: The WSFR <strong>of</strong> an eight-nephron system when blood sodiumconcentration is altered.


5.3 A N I N H O M O G E N E O U S 8-NEPHRO N S Y S T E M 71190185185180WSFR (nL/min)180175WSFR (nL/min)175170170165165500 1000 1500 2000 2500 3000160500 1000 1500 2000 2500 3000TimestepTimestep(a) γ = 0.000(b) γ = 0.001175164162170WSFR (nL/min)165160WSFR (nL/min)160158156154155500 1000 1500 2000 2500 3000152500 1000 1500 2000 2500 3000TimestepTimestep(c) γ = 0.002(d) γ = 0.003156144WSFR (nL/min)154152150148WSFR (nL/min)142140138146500 1000 1500 2000 2500 3000136500 1000 1500 2000 2500 3000TimestepTimestep(e) γ = 0.004(f) γ = 0.005Figure 42: The WSFR <strong>of</strong> eight-nephron systems with different levels <strong>of</strong>vascular signalling.


72 M O D E L DY NA M I C S2e-206e-201.8e-201.6e-205e-20Power1.4e-201.2e-201e-208e-216e-21Power4e-203e-202e-204e-212e-211e-2000 0.02 0.0400 0.02 0.04Normalised FrequencyNormalised Frequency(a) γ = 0.000(b) γ = 0.0012.5e-197e-202e-196e-205e-20Power1.5e-191e-19Power4e-203e-202e-205e-201e-2000 0.02 0.0400 0.02 0.04Normalised FrequencyNormalised Frequency(c) γ = 0.002(d) γ = 0.0031e-208e-21Power9e-218e-217e-216e-215e-214e-213e-212e-211e-21Power7e-216e-215e-214e-213e-212e-211e-2100 0.02 0.0400 0.02 0.04Normalised FrequencyNormalised Frequency(e) γ = 0.004(f) γ = 0.005Figure 43: The WSFR power spectral density <strong>of</strong> eight-nephron systemswith different levels <strong>of</strong> vascular signalling.


5.4 A 7 2-NEPHRO N M U LT I-COLUMN M O D E L 73between nephrons whose loops <strong>of</strong> Henle have different lengths. Moreinterestingly, the spectral densities for γ = 0.000 (Fig. 43a) andγ = 0.005 (Fig. 43f) are very similar, despite the spectral densities for0.000 < γ < 0.005 showing no such similarity. This suggests that theWSFRs shown in Fig. 42a and Fig. 42f have similar properties—despitethe obvious differences in average and oscillation size—which is asurprising observation.5.4 A 7 2 - N E P H RO N M U LT I-COLUMN M O D E LThe eight-nephron system was used to analyse the interaction betweenthe competing coupling mechanisms <strong>of</strong> hemodynamic couplingand vascular signalling—and also to investigate the stability <strong>of</strong> themedullary salt gradient maintained by multiple loops <strong>of</strong> Henle—in asystem <strong>of</strong> inhomogeneous nephrons. The analysis is now extended toa 72-nephron system consisting <strong>of</strong> nine columns arranged in a 3 × 3grid, with each column having a structure identical to that <strong>of</strong> theeight-nephron system presented in Sect. 5.3. The 2D arrangement<strong>of</strong> the columns and the structure <strong>of</strong> the arterial tree are shown inFig. 44.P AP VR Cy23 6 9R C1 2 312 5 8R C4 5 601 4 70 1 2x7 8 9(a) The 2D arrangement <strong>of</strong>interstitial fluid columns.(b) The arterial tree connecting thecolumns.Figure 44: The high-level structure <strong>of</strong> the 72-nephron model. Eachcolumn is assigned a number 1–9. The internal structure <strong>of</strong> eachcolumn is identical, and is shown in Fig. 38.In the absence <strong>of</strong> vascular signalling (i. e., γ = 0), the WSFR exhibitsoscillations <strong>of</strong> 100–200 nL/min (see Fig. 45a). The power spectraldensity shows that these oscillations predominantly occur at a singlefrequency—the frequency at which the longest-looped nephronsnaturally oscillate. This is different to the power spectral density <strong>of</strong>the 8-nephron system, where higher-frequency peaks are present(see Fig. 43a), and indicates that the majority <strong>of</strong> higher-frequencysignals (such as the oscillations <strong>of</strong> the shorter-looped nephrons) canceleach other out in this system. Figure 46 shows the individualWCFRs, and the power spectral densities clearly indicate the presence


74 M O D E L DY NA M I C S<strong>of</strong> higher-frequency oscillations in the WCFRs. Note that althoughthe structure <strong>of</strong> the individual columns are homogeneous, the WCFRplots and power spectral density plots differ, which indicates thatthe higher-level structure <strong>of</strong> the system (i. e., the 2D column arrangementand the arterial tree connecting the columns) is affecting thedynamics <strong>of</strong> each column.Once vascular signalling is introduced into the 72-nephron system,the patters in the WSFR become less clear. As shown in Fig. 47, theaddition <strong>of</strong> vascular signalling results in higher-frequency signalsbecoming more dominant than when gamma = 0, although this trendis not entirely consistent—witness Fig. 47c. Te individual WCFRs, however,show a consistent increase in the presence <strong>of</strong> higher-frequencysignals when compared to their behaviour when γ = 0—compare thepower spectral densities in Fig. 48 with those in Fig. 46.Finally, it was observed that as the degree <strong>of</strong> vascular signalling wasincreased, the average WSFR and WCFRs decreased—shown in Fig. 49—which is consistent with the results produced by the 8-nephron systemin Sect. 5.3.5.5 S U M M A RY O F E X P E R I M E N T SThe experiments in this chapter have demonstrated that the modelpresented in Chapter 4 produces valid dynamics across a range <strong>of</strong>multi-nephron systems.Firstly, a single-nephron system was shown to produce oscillationsin SNGFR mediated by the TGF mechanism, and was also shown tomaintain a stable salt gradient in the medullary fluid across thephysiological range <strong>of</strong> blood sodium concentrations.A two-nephron system was then used to investigate the interactionsthat arise between nephrons. Over the state space k × P A , the systemwas shown to transition between in-phase and anti-phase oscillationsin the absence <strong>of</strong> vascular signalling, and the parameters k and P Awere shown to have equivalent effects on these oscillations. Whenvascular signalling was introduced to the system, it was shown tooverwhelm the hemodynamic coupling present in the system and enforcein-phase oscillations, even when the nephrons featured differentdelays in the TGF mechanism. This system also maintained a larger,more stable medullary salt gradient than the single-nephron system,suggesting that the presence <strong>of</strong> multiple loops <strong>of</strong> Henle is responsiblefor the stability and magnitude <strong>of</strong> the gradient.An eight nephron system was then demonstrated to maintain amedullary salt gradient that was both larger and more stable than thetwo-nephron system, providing further evidence that additional loops<strong>of</strong> Henle lead to larger, more stable medullary salt gradients. Theeffects <strong>of</strong> vascular signalling on the WSFR were then investigated, andit was shown that increased levels <strong>of</strong> vascular signalling decreasedthe average WSFR. Moreover, the spectral densities <strong>of</strong> the WSFR demonstratedthat the degree <strong>of</strong> vascular signalling affected the dominantfrequencies in the WSFR.Finally, a 72-nephron system <strong>of</strong> 9 columns was analysed. Althoughthe columns were homogeneous, the WCFRs differed, depending onwhich branch <strong>of</strong> the arterial tree the column was connected to. An


5.5 S U M M A RY O F E X P E R I M E N T S 759e-178e-177e-176e-17Power5e-174e-173e-172e-171e-1700 0.02 0.04Normalised Frequency(a) The WSFR.(b) The WSFR power spectral density.Figure 45: The WSFR <strong>of</strong> the 72-nephron system when γ = 0.WCFR (nL/min)280260240220200180160Power4e-173.5e-173e-172.5e-172e-171.5e-171e-175e-18140500 1000 1500 2000 2500 3000Timestep00 0.02 0.04Normalised Frequency(a) The WCFR <strong>of</strong> columns 1–3.(b) The spectral density <strong>of</strong> columns 1–3.2602.5e-17WCFR (nL/min)240220200180160140Power2e-171.5e-171e-175e-18120500 1000 1500 2000 2500 3000Timestep00 0.02 0.04Normalised Frequency(c) The WCFR <strong>of</strong> columns 4–6.(d) The spectral density <strong>of</strong> columns 4–6.2402e-172201.8e-171.6e-17WCFR (nL/min)200180160Power1.4e-171.2e-171e-178e-186e-181404e-182e-18120500 1000 1500 2000 2500 300000 0.02 0.04TimestepNormalised Frequency(e) The WCFR <strong>of</strong> columns 7–9.(f) The spectral density <strong>of</strong> columns 7–9.Figure 46: The individual WCFRs in the 72-nephron system when γ = 0.


76 M O D E L DY NA M I C S9e-171.8e-178e-171.6e-177e-171.4e-176e-171.2e-17Power5e-174e-17Power1e-178e-183e-176e-182e-174e-181e-172e-1800 0.02 0.0400 0.02 0.04Normalised FrequencyNormalised Frequency(a) γ = 0.000.(b) γ = 0.001.3e-173e-172.5e-172.5e-172e-172e-17Power1.5e-17Power1.5e-171e-171e-175e-185e-1800 0.02 0.0400 0.02 0.04Normalised FrequencyNormalised Frequency(c) γ = 0.002.(d) γ = 0.003.2.5e-179e-188e-182e-177e-18Power1.5e-171e-17Power6e-185e-184e-183e-185e-182e-181e-1800 0.02 0.0400 0.02 0.04Normalised FrequencyNormalised Frequency(e) γ = 0.004.(f) γ = 0.005.Figure 47: Power spectral densities <strong>of</strong> the WSFR <strong>of</strong> the 72-nephronsystem for different values <strong>of</strong> γ.


5.5 S U M M A RY O F E X P E R I M E N T S 771e-174e-189e-183.5e-188e-187e-183e-18Power6e-185e-184e-18Power2.5e-182e-181.5e-183e-182e-181e-181e-185e-1900 0.02 0.0400 0.02 0.04Normalised FrequencyNormalised Frequency(a) Columns 1–3 when γ = 0.003.(b) Columns 1–3 when γ = 0.005.9e-183.5e-188e-183e-187e-186e-182.5e-18Power5e-184e-18Power2e-181.5e-183e-182e-181e-181e-185e-1900 0.02 0.0400 0.02 0.04Normalised FrequencyNormalised Frequency(c) Columns 4–6 when γ = 0.003.(d) Columns 4–6 when γ = 0.005.3.5e-181.8e-183e-181.6e-182.5e-181.4e-181.2e-18Power2e-181.5e-18Power1e-188e-191e-186e-194e-195e-192e-1900 0.02 0.0400 0.02 0.04Normalised FrequencyNormalised Frequency(e) Columns 7–9 when γ = 0.003.(f) Columns 7–9 when γ = 0.005.Figure 48: Power spectral densities <strong>of</strong> the WCFRs <strong>of</strong> the 72-nephronsystem for different values <strong>of</strong> γ.18001700WSFR (nL/min)160015001400WCFR (nL/min)2202001801601404-67-913000 0.001 0.002 0.003 0.004 0.0050.001γ0.0031-30.005Columnγ(a) The change in average WSFR.(b) The change in average WCFR.Figure 49: Average WCFRs and WSFRs <strong>of</strong> the 72-nephron system fordifferent values <strong>of</strong> γ.


78 M O D E L DY NA M I C Sanalysis <strong>of</strong> the power spectral densities showed the arterial tree alsoinfluenced the dominant frequencies in the WCFRs. Increasing thedegree <strong>of</strong> vascular signalling in this system was shown to result inthe presence <strong>of</strong> higher-frequency signals in the WCFRs and WSFR, andalso to reduce the average WCFRs and WSFR.


E M E R G E N T E F F E C T S O F L O C A L I S E D I M PA I R M E N T6The dynamics <strong>of</strong> several multi-nephron systems have been analysedand presented in Chapter 5, providing a benchmark against whichthe behaviour <strong>of</strong> similar systems can be compared. In particular, suchbenchmarks allow the system-wide effects <strong>of</strong> localised changes indynamics (hereafter referred to as impairments) to be determined.The magnitude <strong>of</strong> the effect <strong>of</strong> an impairment on the whole-systemdynamics forms a measure <strong>of</strong> the stability or resilience <strong>of</strong> the system.The most obvious application <strong>of</strong> this approach—particularly usefulwhen applied to a whole-kidney model—is to model renal diseases asimpairments and to study how the diseases affect the whole-systemdynamics. Such studies could investigate which symptoms <strong>of</strong> a renaldisease have the most significant impact on renal function, or determinewhat fraction <strong>of</strong> the nephron population must become impairedto reach a particular stage <strong>of</strong> chronic kidney disease (CKD). The mostcommon symptoms <strong>of</strong> renal disease have been introduced in Sect. 2.3,and the application <strong>of</strong> network models and automata to studyingdisease has been reviewed in Sect. 3.4.This chapter presents an example <strong>of</strong> how to incorporate an impairmentinto the model and how to determine the impact <strong>of</strong> theimpairment on the dynamics <strong>of</strong> multiple systems and at differentlevels <strong>of</strong> hierarchy. This example introduces a simple impairment thatattempts to capture a single effect <strong>of</strong> proteinuria—the presence <strong>of</strong>abnormally large amounts <strong>of</strong> protein in the urine (see Sect. 2.3 forfurther details). This impairment only serves to illustrate the generalapproach, and not to predict the effects that proteinuria might haveon kidney behaviour.To begin, the method by which impairments are incorporated intothe model is described, and the control that the modeller has over theincidence and spread <strong>of</strong> an impairment is discussed. Secondly, thechoice <strong>of</strong> the specific impairment is justified, and the design <strong>of</strong> theimpairment is presented. Thirdly, several techniques for measuringthe effects <strong>of</strong> the impairment on the system dynamics are outlined,focusing on the dynamics at several levels <strong>of</strong> hierarchy.The dynamics <strong>of</strong> several impaired systems are then presented andcompared to the dynamics <strong>of</strong> the original (unimpaired) systems. Twoexperiments investigate the effects that a single impaired nephronand pairs <strong>of</strong> impaired nephrons have on an 8-nephron system; tw<strong>of</strong>urther experiments investigate the effect <strong>of</strong> impairments in the context<strong>of</strong> a 72-nephron system. The chapter concludes by drawing someconclusions about the stability <strong>of</strong> the model in response to the chosenimpairment.6.1 I N C O R P O R AT I N G I M PA I R M E N T S I N T O T H E M O D E LFor an impairment to change the model dynamics, it must changeeither the model structure (Sect. 4.2) or the underlying equations79


80 E M E R G E N T E FF E C T S O F L O C A L I S E D I M PA I R M E N T(Sect. 4.4). The high-level structure <strong>of</strong> the model (Fig. 14 and Fig. 15)captures the physical structure <strong>of</strong> the kidney, which should not bealtered by an impairment. The lower-level structure (Fig. 12) capturesthe functional structure <strong>of</strong> the nephrons, which is closely tied to thephysical structure <strong>of</strong> the nephrons and should also not be alteredby an impairment. Instead, it is clear that impairments should onlyintroduce changes to the underlying model equations.To allow for both impaired and unimpaired nephrons to exist in amulti-nephron system, it is necessary for each nephron to includean impair state that indicates whether the nephron is impaired ornot. When the model equations are applied to a particular nephron,the impair state is examined and the appropriate modifications tothe model equations are applied. It is straightforward to extend thisapproach to support multiple impairments.Once the impairments have been defined, the evolution <strong>of</strong> thenephron impair states over time must be specified. All nephrons ina model are initially unimpaired and thus exhibit normal behaviour.The different ways in which a nephron may become impaired are nowdescribed.6.1.1 Stochastic impairmentsEach impairment is assigned a failure probability, which defines thelikelihood that a nephron will become impaired over a single timestep.This probability can be defined with respect to time (i. e., a fixedprobability for each time-step) or with respect to blood flow to thenephron at that time-step (i. e., a probability that can vary with eachtime-step).Similarly, each impaired is assigned a recovery probability, whichdefines the likelihood that an impaired nephron will become unimpaired.This probability can also be defined with respect to time or toblood flow to the nephron.Note that this approach models the transition between impair statesas a Markov Chain.6.1.2 Stochastic spread <strong>of</strong> impairmentsImpairments are also assigned a spreading probability that definesthe likelihood <strong>of</strong> that impairment spreading from an impaired nephronto a neighbouring nephron over a single time-step. For this to be welldefined,a definition <strong>of</strong> what constitutes a neighbouring nephron mustbe given.Nephrons in a single column <strong>of</strong> interstitial fluid are arranged in anumber <strong>of</strong> branches (see Sect. 4.2.2 and Fig. 14). If two nephronsin the same branch are next to each other, it is natural to considerthem as neighbours. The first nephron in a branch is also consideredto be a neighbour <strong>of</strong> the first nephron in the adjacent branchesabove and below it (again, see Fig. 14). More concisely, nephrons areneighbours if they are adjacent in either a depth-first or breadth-firsttraversal <strong>of</strong> the column’s nephron tree. Finally, the first nephron in anentire column is considered to be a neighbour <strong>of</strong> the first nephrons inadjacent columns (see Fig. 15 and Fig. 16).


6.2 C H O I C E O F I M PA I R M E N T 816.1.3 Deterministic impairmentsBy allowing impairments to arise, spread and recover in a stochasticmanner, it can be observed how they spread across the model, ina similar manner to the network models used in epidemiology (seeSect. 3.4). However, studying the effects that impaired nephrons haveon a system <strong>of</strong> nephrons—as opposed to studying the manner in whichthe impairments spread—requires explicit control over the impairstate <strong>of</strong> the nephrons.To this end the model also supports deterministic impairments,where any number <strong>of</strong> impairments can be introduced to particularnephrons in a system at specific time-steps. These impairments occurin parallel with the stochastic methods described previously, andfor absolute control over the impair states it is necessary to set thefailure, spreading and recovery probabilities <strong>of</strong> the impairment tozero.6.2 C H O I C E O F I M PA I R M E N TAs outlined in Sect. 2.3, there are many diseases that affect thekidney. For the purposes <strong>of</strong> disease modelling, however, only theactual symptoms <strong>of</strong> the disease are <strong>of</strong> interest—the specific causes(e.g., buildup <strong>of</strong> scar tissue, calcium deposits) are not directly relevant.As such, the modelling concern does not focus on a specific renaldisease, but rather on the symptoms <strong>of</strong> that disease. By designingimpairments for these symptoms, the effects <strong>of</strong> renal diseases can bemodelled as specific combinations <strong>of</strong> these impairments.As was stated in Sect. 2.3, diabetic nephropathy, hypertensionand glomerulonephritis account for around 75% <strong>of</strong> all adult cases <strong>of</strong>chronic renal failure (CRF) [41], with diabetic nephropathy the mostcommon cause <strong>of</strong> CRF and end-stage renal failure (ESRF) in the USA[53]. The focus <strong>of</strong> the impairment example to be presented here istherefore on the symptoms presented by these diseases:E D E M A Excess fluid in the body.H E M AT U R I A Blood present in the urine.H Y P E R L I P I D E M I A Raised levels <strong>of</strong> lipids in the blood.H Y P E RT E N S I O N Chronically elevated blood pressure.H Y P OA L B U M I N E M I A Abnormally low albumin levels in the blood.P ROT E I N U R I A Protein present in the urine.Some <strong>of</strong> these symptoms are caused by the presence <strong>of</strong> othersymptoms—e.g., proteinuria can result in hypoalbuminemia, whichreduces the oncotic pressure in the blood and can lead to edema—and are not relevant in isolation. Other symptoms can be caused bychanging model parameters—e.g., hypertension can be simulatedby increasing the arterial pressure parameter—having a systemwideeffect as opposed to the per-nephron effects <strong>of</strong> impairmentsin our model, and must be controlled separately. Of the symptoms


82 E M E R G E N T E FF E C T S O F L O C A L I S E D I M PA I R M E N Tdescribed here, the impairment that will now be presented only considersproteinuria—this impairment serves only as an example <strong>of</strong>this approach, and not as a complete, detailed model <strong>of</strong> one or moresymptoms.6.2.1 Implementation <strong>of</strong> the impairmentProteinuria is the presence <strong>of</strong> abnormally large amounts <strong>of</strong> protein inthe urine. As proteins are not secreted by the tubule, this conditionarises due to the presence <strong>of</strong> proteins in the tubule filtrate. Whilethis can have a number <strong>of</strong> effects on the function <strong>of</strong> the tubule, theonly effect that will be captured by the impairment is that on thereabsorption <strong>of</strong> water in the descending limb <strong>of</strong> Henle.The presence <strong>of</strong> protein in the filtrate increases the oncotic pressure<strong>of</strong> the filtrate and thus retards the diffusion <strong>of</strong> water from thefiltrate into the surrounding interstitial fluid. Accordingly, the impairmentwill capture this reduction in reabsorption by modifying theequations that govern the reabsorption <strong>of</strong> water in the limbs <strong>of</strong> Henle(as given in Sect. 4.4.5).Specifically, for each descending limb that is affected by the impairment,the amount <strong>of</strong> water reabsorbed into the interstitial fluidis some fraction D i <strong>of</strong> the amount that would be reabsorbed werethe tubule not affected by proteinuria. This is modelled by changingthe w i coefficients in Eq. 4.29 to take into account D i , resulting inEq. 6.1. For impaired nephrons, the value <strong>of</strong> D i is a model parameter,while for unimpaired nephrons D i = 1.⎡w 1′ w2′ x =⎢ . ⎣ .wN′⎤⎥⎦, w ′ i = w iD i(6.1)6.3 M E A S U R I N G T H E E FF E C T S O F I M PA I R M E N T SThe goal <strong>of</strong> this example impairment is to determine how systemdynamics are affected by impaired nephrons, and so the focus <strong>of</strong> theanalysis will be on whole-system dynamics rather than the dynamics<strong>of</strong> individual tubule segments.Existing multi-nephron models have focused on single-nephronglomerular filtration rate (SNGFR) [99] and tubular pressure [112, 81];due to the design <strong>of</strong> the model, pressure is not calculated anywherein the tubule. Pressure can be treated as varying in proportion t<strong>of</strong>low rate, if the tubule is assumed to be non-compliant or only slightlycompliant. This point is discussed further in Sect. 4.4.2.In the experiments performed previously (Chapter 5), the analysis<strong>of</strong> the system dynamics focused on SNGFR—in particular, the frequency<strong>of</strong> the oscillations and the phase (relative to other nephrons)—andso SNGFR will form the starting point <strong>of</strong> the analysis here. With theemphasis on whole-system dynamics, the analysis will also extend to


6.4 I M PA I R M E N T S I N A N 8-NEPHRO N S Y S T E M 83comparing the whole-system filtration rate (WSFR) <strong>of</strong> unimpaired andimpaired systems.Based on the assumption that the filtration rate at any level <strong>of</strong> hierarchy(e. g., single nephrons, multiple columns) can be characterisedby a small number <strong>of</strong> measurements (e. g., frequency, amplitude,phase and mean value), the low-level dynamics can be ignored. Instead,the analysis can determine the effects <strong>of</strong> the impairment—andthe degree to which the effects permeate across the system—bycomparing these characteristic measurements in the impaired andunimpaired systems at multiple levels <strong>of</strong> hierarchy.Once these effects have been analysed and quantified, it may transpirethat the impairment can be simplified—i. e., some <strong>of</strong> the changesto the model equations may be removed without affecting the systemdynamics—allowing the modeller to determine which aspects <strong>of</strong> theimpairment have the largest effects. To draw an analogy with flocks <strong>of</strong>birds [152, 153]—where three exceedingly simple rules for individualnavigation are sufficient to generate realistic aggregate motion foran entire flock—a thorough, detailed set <strong>of</strong> impairments may prove tobe more complex than necessary to model a chosen phenomena, suchas a particular renal disease.6.4 I M PA I R M E N T S I N A N 8 - N E P H RO N S Y S T E MThe first investigation was to analyse the effect that impaired nephronshave on the dynamics <strong>of</strong> the eight-nephron system presented inSect. 5.3. This system consists <strong>of</strong> four nephron pairs, each pair havinglonger loops <strong>of</strong> Henle than the last–this arrangement is shown inFig. 50.01 2 3 4 5 6 7 8Length123Figure 50: The loops <strong>of</strong> Henle in the eight-nephron system. Eachnephron is identified by a number from 1 to 8, and the length <strong>of</strong> theloops are rated on a scale <strong>of</strong> 0 to 3.Two separate experiments were conducted, to investigate the effects<strong>of</strong> a single impaired nephron and the effects <strong>of</strong> two impairednephrons. These experiments are now presented and compared, andthe effects <strong>of</strong> impaired nephrons on the eight-nephron system aresummarised.In the following discussion, nephrons are frequently referred to bythe length <strong>of</strong> their loop <strong>of</strong> Henle (0 . . . 3) and their nephron number(1 . . . 8). The SNGFR oscillation frequency <strong>of</strong> the nephron pair with loop<strong>of</strong> Henle length L is also referred to as the natural frequency <strong>of</strong> looplength L, or the loop length L frequency.


84 E M E R G E N T E FF E C T S O F L O C A L I S E D I M PA I R M E N T6.4.1 A single impaired nephronThe first experiment was to investigate how a single impaired nephroncould affect the eight-nephron system. For each nephron in the system,simulations were run where the nephron was deterministicallyinfected at t = 1, 200 and the simulation ran until t = 3, 000. A simulationwas run for each value <strong>of</strong> D P from 0.5 to 0.95—in steps <strong>of</strong>0.05—and the model parameters and conditions were otherwise identicalto the experiments performed upon the system in Sect. 5.3.A comparison <strong>of</strong> the WSFR for each <strong>of</strong> these simulations is presentedin Fig. 51—Fig. 51a shows the average WSFRs and Fig. 51b shows themagnitude <strong>of</strong> the oscillations in WSFRs. From Fig. 51a it is clear thatthe impairment decreased the average WSFR, and that the magnitude<strong>of</strong> this decrease is proportional to both D 1 and the length <strong>of</strong> thePimpaired nephron’s loop <strong>of</strong> Henle. However, Fig. 51b shows that theoscillations in WSFR were not affected by the impairment unless one<strong>of</strong> the longest-loop nephrons was impaired and D P was small ( 0.6).A more detailed comparison between the original eight-nephronsystem and one <strong>of</strong> the impaired systems will now be given. Fig. 51shows that the impaired system that differed the most from the originalsystem was when one <strong>of</strong> the longest-loop nephron was infectedand D P = 0.5, and so it is with this impaired system that the originalsystem will be compared. Specifically, in this impaired system theimpaired nephron is nephron 7.The WSFR <strong>of</strong> the two systems are shown in Fig. 52a, where thedecrease in average WSFR due to the impairment can clearly be seen.The power spectral densities <strong>of</strong> the two WSFRs are shown in Fig. 52b,and the impaired system can be seen to have a more pronounced peakat a low frequency. This suggests that the SNGFRs <strong>of</strong> the longer-loopednephrons are experiencing larger oscillations in the impaired systemthan in the original system.The SNGFRs <strong>of</strong> the individual nephrons in the impaired system willnow be examined. The power spectral densities <strong>of</strong> the SNGFRs arepresented in Fig. 53. This figure shows that while nephrons 3–8 areessentially oscillating at a single frequency, nephrons 1 and 2 (thenephrons with the shortest loops) are oscillating at a combination<strong>of</strong> three different frequencies. From the other power spectral densities,it is apparent that the spikes in the power spectral densities <strong>of</strong>nephrons 1 and 2 correspond to the natural frequencies <strong>of</strong> the looplengths 0, 1 and 3—there does not appear to be a spike at the naturalfrequency <strong>of</strong> loop length 2. The frequency <strong>of</strong> the oscillations in SNGFRis clearly proportional to the length <strong>of</strong> the loops <strong>of</strong> Henle, as expected(see Sect. 5.3). Note that the impaired nephron (nephron 7) has alarger spike than nephron 8, while the other nephron pairs exhibitpower spectral densities that are identical.To compare these spectral densities to those <strong>of</strong> the same nephronsin the original system, plots <strong>of</strong> the differences in spectral densitiesbetween the original and impaired systems are given in Fig. 54.It is evident that nephrons 7 and 8 have larger oscillations in theimpaired system than in the original system, and as previously notedthe impaired nephron 7 is exhibiting larger oscillations than nephron8. Conversely, nephrons 3–6 exhibit a reduction in the oscillation


6.4 I M PA I R M E N T S I N A N 8-NEPHRO N S Y S T E M 85Average WSFR (nL/min)178177.6177.2176.8176.401Loop Length 23 0.4 0.6ImpairedOriginal10.8D P(a) The average WSFR over the simulations.Range in WSFR (nL/min)4.443.63.201Loop Length 23 0.4 0.6ImpairedOriginal10.8D P(b) The size <strong>of</strong> oscillations in WSFR over the simulations.Figure 51: Variations in the WSFR <strong>of</strong> an eight-nephron system, inresponse to single impaired nephrons <strong>of</strong> different loop lengths.magnitude. Nephrons 1 and 2 have increased oscillations at thenatural frequencies <strong>of</strong> loop lengths 1 and 3, but have decreasedoscillations at their own natural frequency (loop length 0).From the spikes at the natural frequency <strong>of</strong> loop length 3 thatare present in the power spectral densities <strong>of</strong> other nephron pairs,it is evident that the infection <strong>of</strong> nephron 7 has affected the SNGFR<strong>of</strong> nephrons 1 and 2. Also, although nephrons 3–6 exhibit reducedoscillations in SNGFR, their spectral densities show no spikes at thenatural frequency <strong>of</strong> the impaired nephron, which would appear to indicatethat these nephrons are somehow less affected by the impairednephron. The changes in spectral density peaks are summarised inTable 14.


86 E M E R G E N T E FF E C T S O F L O C A L I S E D I M PA I R M E N T180WSFR (nL/min)178176174Impaired172Original1200 1800 2400 3000Timestep(a) The WSFR in both the original and impaired eight-nephron systems.3e-202.5e-20ImpairedOriginal2e-20Power1.5e-201e-205e-2100 0.02 0.04Normalised Frequency(b) The power spectral density <strong>of</strong> WSFR in both the original and impaired eightnephronsystems.Figure 52: Comparisons <strong>of</strong> WSFR between the original (unimpaired)eight-nephron system and the eight-nephron system where nephron7 is impaired and D P = 0.5.Nephron Peak Increase Nephron Peak Increase1 0.25% 2 0.25%3 -77% 4 -77%5 -4.7% 6 -4.7%7 183% 8 153%Table 14: Increases in spectral density peaks in the SNGFR <strong>of</strong> nephronsin the system with one impaired nephron.


6.4 I M PA I R M E N T S I N A N 8-NEPHRO N S Y S T E M 872.52Nephron 1Nephron 23025Nephron 3Nephron 4Power1.51Power2015100.5500 0.02 0.0400 0.02 0.04Normalised FrequencyNormalised Frequency(a) Nephrons 1 and 2 (loop length 0).(b) Nephrons 3 and 4 (loop length 1).80007000Nephron 5Nephron 690008000Nephron 7Nephron 860007000Power5000400030002000Power600050004000300020001000100000 0.02 0.04Normalised Frequency00 0.02 0.04Normalised Frequency(c) Nephrons 5 and 6 (loop length 2).(d) Nephrons 7 and 8 (loop length 3).Figure 53: The power spectral densities <strong>of</strong> SNGFR in the eight-nephronsystem with a single impaired nephron.Power0.40.30.20.10-0.1-0.2-0.3Nephron 1Nephron 20 0.02 0.04Normalised FrequencyPower100-10-20-30-40-50-60-70-80-90-100Nephron 3Nephron 40 0.02 0.04Normalised Frequency(a) Nephrons 1 and 2 (loop length 0).(b) Nephrons 3 and 4 (loop length 1).Power500-50-100-150-200-250-300-350-400Nephron 5Nephron 60 0.02 0.04Normalised FrequencyPower40003500300025002000150010005000Nephron 7Nephron 80 0.02 0.04Normalised Frequency(c) Nephron 5 and 6 (loop length 2).(d) Nephrons 7 and 8 (loop length 3).Figure 54: The difference in power spectral densities <strong>of</strong> SNGFR betweenthe original eight-nephron system and the system with oneimpaired nephron.Returning attention to Fig. 52, despite changes in SNGFR spectraldensities at several frequencies, the WSFR spectral density only exhibitsan increase at the natural frequency <strong>of</strong> the impaired nephron,and a decrease in the average WSFR <strong>of</strong> 0.7%. Despite the spectral


88 E M E R G E N T E FF E C T S O F L O C A L I S E D I M PA I R M E N Tdensities <strong>of</strong> individual nephrons varying from -77% to 183%, theaverage WSFR is not significantly affected, demonstrating that thiseight-nephron system can exhibit stable whole-system behaviourwhen the function <strong>of</strong> an individual nephron is impaired.6.4.2 Two impaired nephronsThe second experiment was to determine the effects <strong>of</strong> two impairednephrons on the same system. In this experiment, D P was kept fixedat 0.5 and for each pair <strong>of</strong> nephrons in the system, a simulationwas run where both nephrons were impaired at t = 1, 200 and thesimulation ran until t = 3, 000.A comparison <strong>of</strong> the WSFR for each <strong>of</strong> these simulations is presentedin Fig. 55—Fig. 55a shows the average WSFRs and Fig. 55b showsthe magnitude <strong>of</strong> the oscillations in WSFRs. From Fig. 55a it is clearthat the impairment decreased the average WSFR in proportion tothe lengths <strong>of</strong> the infected nephrons’ loops <strong>of</strong> Henle, as was the casefor a single impaired nephron. The magnitude <strong>of</strong> WSFR oscillations(Fig. 55b) also shows a similarity to the previous experiment, as it wasnot affected by the impairment unless at least one <strong>of</strong> the longest-loopnephrons was impaired.A comparison between the original system and the impaired systemwhere both longest-loop nephrons are impaired will now be given.This particular impaired system was chosen as it exhibits the lowestaverage WSFR <strong>of</strong> the systems under consideration (see Fig. 55a).The WSFR <strong>of</strong> the two systems are shown in Fig. 56a, where thedecrease in average WSFR due to the impairment is apparent. Thepower spectral densities <strong>of</strong> the two WSFRs are shown in Fig. 56b, andthe impaired system can be seen to have peak seven times larger atthe natural frequency <strong>of</strong> loop length 3. This peak is approximatelyfive times larger than the equivalent peak in the system with a singleimpaired nephron (Fig. 56b), indicating that the two impairednephrons have a much larger influence on the system dynamics thanthe single impaired nephron in Sect. 6.4.1.The analysis now turns to the SNGFRs <strong>of</strong> the individual nephrons.The power spectral densities <strong>of</strong> the SNGFRs are presented in Fig. 57.This figure shows that while nephrons 3–8 are essentially oscillating ata single frequency, nephrons 1 and 2 (the nephrons with the shortestloops) are oscillating at a combination <strong>of</strong> different frequencies. Thesespectral densities exhibit three main differences from the previoussystem that contained only a single impaired nephron:• Nephrons 1 and 2 feature a significantly larger peak at the naturalfrequency <strong>of</strong> loop length 3 (compare Fig. 53a and Fig. 57a).• Nephrons 3 and 4 exhibit smaller peaks, with differing magnitudes,which accentuate the small peaks at the natural frequencies<strong>of</strong> loop lengths 0 and 3 (compare Fig. 53b and Fig. 57b).These additional peaks are present in the system with one impairednephron (Fig. 53b), but were minuscule in comparisonto the dominant peak.


6.4 I M PA I R M E N T S I N A N 8-NEPHRO N S Y S T E M 89Average WSFR (nL/min)17817717617501Loop Length 2ImpairedOriginal321 Loop Length3 0(a) The average WSFR over the simulations.Range in WSFR (nL/min)7654301Loop Length 2ImpairedOriginal321 Loop Length3 0(b) The size <strong>of</strong> oscillations in WSFR over the simulations.Figure 55: Variations in the WSFR <strong>of</strong> an eight-nephron system, inresponse to pairs <strong>of</strong> impaired nephrons <strong>of</strong> different loop lengths.• Nephrons 7 and 8 again exhibit peaks <strong>of</strong> different magnitude(compare Fig. 53d and Fig. 57d), but in this system both nephronshave identical impair states.The difference in spectral densities between nephrons 3 and 4, andbetween nephrons 7 and 8 is unexpected—both pairs <strong>of</strong> nephronshave identical impair states—but one possible explanation is that thedifferent peak sizes may be due to the relationship between the phase<strong>of</strong> the individual nephron and the phases <strong>of</strong> the other nephrons in thesystem. This phenomena will not be investigated in any further detailhere—the changes in spectral densities will now be examined.Plots <strong>of</strong> the differences in spectral densities between the originaland impaired systems are given in Fig. 58. The plot for nephrons 1and 2 exhibit the large peak at the natural frequency <strong>of</strong> loop length 3


90 E M E R G E N T E FF E C T S O F L O C A L I S E D I M PA I R M E N T180178WSFR (nL/min)176174172Impaired170Original1200 1800 2400 3000Timestep(a) The WSFR in both the original and impaired eight-nephron systems.1.6e-191.4e-19ImpairedOriginal1.2e-19Power1e-198e-206e-204e-202e-2000 0.02 0.04Normalised Frequency(b) The power spectral density <strong>of</strong> WSFR in both the original and impaired eightnephronsystems.Figure 56: Comparisons <strong>of</strong> WSFR between the original eight-nephronsystem and the eight-nephron system where both longest-loopednephrons are impaired and D P = 0.5.(see Fig. 58a) and also show that the increase at their own naturalfrequency is approximately half <strong>of</strong> the increase shown when a singlenephron is impaired (compare Fig. 54a and Fig. 58a). Nephrons 3–6 again exhibit a decrease at their natural frequencies (Fig. 58b,Fig. 58c) and there is a difference in the magnitude <strong>of</strong> the decreasesin nephrons 3 and 4, but it is small enough to be almost invisible(Fig. 58b). The plot for nephrons 7 and 8 clearly shows the differencein peak amplitudes (Fig. 58d) that was observed in Fig. 57d.It is again evident that nephrons 7 and 8 have larger oscillations inthe impaired system than in the original system, and the impairment<strong>of</strong> nephrons 7 and 8 has affected the SNGFR <strong>of</strong> nephrons 1 and 2.Although nephrons 3 and 4 do not exhibit an increase at the natural


6.4 I M PA I R M E N T S I N A N 8-NEPHRO N S Y S T E M 912.5Nephron 1Nephron 298Nephron 3Nephron 427Power1.51Power65430.52100 0.02 0.0400 0.02 0.04Normalised FrequencyNormalised Frequency(a) Nephrons 1 and 2 (loop length 0).(b) Nephrons 3 and 4 (loop length 1).80007000Nephron 5Nephron 670006000Nephron 7Nephron 860005000Power500040003000Power40003000200020001000100000 0.02 0.04Normalised Frequency00 0.02 0.04Normalised Frequency(c) Nephrons 5 and 6 (loop length 2).(d) Nephrons 7 and 8 (loop length 3).Figure 57: The power spectral densities <strong>of</strong> SNGFR in the eight-nephronsystem where both longest-loop nephrons are impaired.frequency <strong>of</strong> loop length 3 (Fig. 58b), the large decrease in the peak attheir own natural frequency has accentuated the peak at the naturalfrequency <strong>of</strong> loop length 3 (Fig. 57b). The changes in spectral densitypeaks are summarised in Table 15.Power1.41.210.80.60.40.20-0.2Nephron 1Nephron 20 0.02 0.04Normalised FrequencyPower200-20-40-60-80-100-120Nephron 3Nephron 40 0.02 0.04Normalised Frequency(a) Nephrons 1 and 2 (loop length 0).(b) Nephrons 3 and 4 (loop length 1).Power1000-100-200-300-400-500-600-700-800-900Nephron 5Nephron 60 0.02 0.04Normalised FrequencyPower2000180016001400120010008006004002000-200Nephron 7Nephron 80 0.02 0.04Normalised Frequency(c) Nephrons 5 and 6 (loop length 2).(d) Nephrons 7 and 8 (loop length 3).Figure 58: The difference in power spectral densities <strong>of</strong> SNGFR betweenthe original eight-nephron system and the system with twoimpaired nephrons.


92 E M E R G E N T E FF E C T S O F L O C A L I S E D I M PA I R M E N TNephron Peak Increase Nephron Peak Increase1 -3% 2 -3%3 -95% 4 -95%5 -10% 6 -10%7 140% 8 111%Table 15: Increases in spectral density peaks in the SNGFR <strong>of</strong> nephronsin the system with two impaired nephrons.Returning attention to Fig. 56, despite changes in SNGFR spectraldensities at several frequencies, the WSFR spectral density only exhibitsan increase at the natural frequency <strong>of</strong> the impaired nephrons,and a decrease in the average WSFR <strong>of</strong> 1.4%—twice the decrease thatarose with a single impaired nephron, even though the WSFR spectraldensity showed a peak five times larger than in the system with asingle impaired nephron (compare Fig. 52b and Fig. 56b). Despite thespectral densities <strong>of</strong> individual nephrons varying from -95% to 140%,the average WSFR is not significantly affected, demonstrating that thiseight-nephron system exhibits stable whole-system behaviour whenthe function <strong>of</strong> the two longest-loop nephrons are impaired.6.4.3 Summary <strong>of</strong> experimentsThe results from both experiments indicate that the dynamics <strong>of</strong> thiseight-nephron system are more sensitive to changes in the longerloopednephrons than the shorter-looped nephrons. A possible explanationfor this observation is that the longer-looped nephrons playa more dominant role in the maintenance <strong>of</strong> the salt gradient in themedullary fluid, and that any change to their behaviour will thereforehave a larger impact on the system dynamics. The SNGFR <strong>of</strong> thelonger-looped nephrons also oscillates more slowly than the shorterloopednephrons—due to the longer time delay between the filtrateentering the tubule and reaching the macula densa—and accordinglyan impaired nephron might only affect neighbouring nephrons withshorter loops <strong>of</strong> Henle, as the shorter-looped nephrons can respondto imbalances more quickly than the longer-looped nephrons.Similarly, although the average WSFR only decreased by 0.7–1.4%,the amplitude <strong>of</strong> the oscillations in WSFR increased, which may indicatethat the system as a whole was producing a larger dynamic responseto the changed conditions. This might mean that the impairedsystems are less able to respond to further changes in conditions, asthey are oscillating closer to their maximal range. Such an effect <strong>of</strong>impairments on multi-nephron systems—that the system may be lessable to respond to changes in conditions, rather than just exhibitingchanged steady-state dynamics—will not be investigated further here.


6.5 I M PA I R M E N T S O N A L A R G E R S C A L E 936.5 I M PA I R M E N T S O N A L A R G E R S C A L EThe second investigation was to analyse the effect that impairednephrons have at a larger scale—the model used for this experimentwas a 72-nephron system, consisting <strong>of</strong> nine columns arranged in a 3 ×3 grid, the structure <strong>of</strong> each column being identical to that <strong>of</strong> the eightnephronsystem used in Sect. 6.4 (see Fig. 50). The 2D arrangement<strong>of</strong> the interstitial fluid columns is shown in Fig. 59a and the arterialtree structure is shown in Fig. 59b. In each simulation, nephronswere impaired at t = 1, 200 with D P = 0.5, and the simulations wererun until t = 3, 000.P AP VR CR C1 2 3y213 6 92 5 8R C4 5 601 4 70 1 2x7 8 9(a) The 2D arrangement <strong>of</strong>interstitial fluid columns.(b) The arterial tree connecting the columns.Figure 59: The high-level structure <strong>of</strong> the 72-nephron model. Eachcolumn is assigned a number 1–9.Two separate experiments were conducted. The first investigatedthe effects <strong>of</strong> impairing all nephrons in the model whose loops <strong>of</strong>Henle were the same length, to determine how significantly the length<strong>of</strong> the loop affects the system dynamics. The difference between thisexperiment and the one conducted on the eight-nephron system(Sect. 6.4.1) is that the 72-nephron model incorporates inter-columndiffusion and a more complex arterial tree, which may reduce theeffect <strong>of</strong> the impairment.The second experiment investigated the effects <strong>of</strong> impairing thenephrons in an entire column, to determine how the impairmentaffects neighbouring columns. This experiment allowed the effects<strong>of</strong> the 2D column arrangement and the arterial tree structure to becompared. These experiments are now presented and compared, andthe effects <strong>of</strong> impaired nephrons on the 72-nephron system are thensummarised.


94 E M E R G E N T E FF E C T S O F L O C A L I S E D I M PA I R M E N T6.5.1 Impaired nephron pairsIn this experiment, for each loop length L (see Fig. 50) all <strong>of</strong> thenephrons whose loops were length L were impaired at t = 1, 200. Thisexperiment is essentially a larger-scale equivalent <strong>of</strong> the experimentdetailed in Sect. 6.4.2.A comparison <strong>of</strong> the WSFR for each <strong>of</strong> these simulations is presentedin Fig. 60. The average WSFR decreased in proportion to the looplength <strong>of</strong> the impaired nephrons (Fig. 60a), as per the experimentsin Sect. 6.4. When the longest-loop nephrons in the model wereimpaired, the decrease in WSFR was 1.3%, which is almost identicalto the decrease observed in Sect. 6.4.2. As shown in Fig. 60b, thesize <strong>of</strong> the oscillations in WSFR varied with no discernible regularity,in contrast to the eight-nephron system (Fig. 55b).The remainder <strong>of</strong> this analysis contrasts the dynamics <strong>of</strong> the original72-nephron system with the dynamics <strong>of</strong> the system where allnephrons <strong>of</strong> loop length 3 are impaired, as this system demonstratedthe largest decrease in WSFR (see Fig. 60).The WSFR <strong>of</strong> both the original and impaired systems are presentedin Fig. 61, where the decrease in average WSFR is not immediatelyapparent (Fig. 61a). The power spectral densities <strong>of</strong> the two WSFRs(Fig. 61b) show that the impaired nephrons have caused larger peaksat the natural frequencies <strong>of</strong> loop lengths 1 and 3. There is also anadditional minor spike at a normalised frequency <strong>of</strong> approximately0.033, which is a higher frequency than even that <strong>of</strong> the shortestloopnephrons. The increases shown here are much smaller than theincreased spikes that arose when the two longest-loop nephrons wereimpaired in the eight-nephron model (compare Fig. 56b and Fig. 61b).The analysis now turns to the WCFRs <strong>of</strong> the individual columns inthe model, which is one level <strong>of</strong> hierarchy lower than the WSFR but isalso a level <strong>of</strong> hierarchy above the SNGFR. At this level <strong>of</strong> hierarchythe WCFR is directly comparable to the WSFR <strong>of</strong> the eight-nephronmodels that were investigated in Sect. 6.4, and thus any observationsmade in these previous experiments may be relevant here.The spectral densities <strong>of</strong> the WCFR for each column in the model areshown in Fig. 62. Columns 1–3 exhibited two dominant peaks, both <strong>of</strong>which have increased in size in the impaired system—the increase inthe peak at the natural frequency <strong>of</strong> the nephrons with loop length 3has increased more than the secondary peak at the natural frequency<strong>of</strong> nephrons with loop length 1 (Fig. 62a). Columns 4–9 all exhibitedthree dominant peaks—at the natural frequencies <strong>of</strong> nephrons withloop length 0, 1 and 2. There was a difference in relative sizes <strong>of</strong> thepeaks, with columns 4–6 having a larger peak at the frequency <strong>of</strong>loop length 1 (Fig. 62b), and columns 7–9 having a larger peak at thefrequency <strong>of</strong> loop length 0 (Fig. 62c).All <strong>of</strong> the columns exhibited a larger spike at the natural frequency<strong>of</strong> the longest-loop nephrons (Fig. 63), which is consistent with theexperiments conducted on the eight-nephron system. As was shown inSect. 6.4, this does not mean that nephrons with different natural frequenciesare unaffected. Apart from this trend, there is no observablecommon pattern in the changes in power spectral densities.


6.5 I M PA I R M E N T S O N A L A R G E R S C A L E 95Average WSFR (nL/min)1730172017101700ImpairedOriginal16900 1 2 3Loop Length(a) The average WSFR over the simulations.Range in WSFR (nL/min)250240230220210200ImpairedOriginal1900 1 2 3Loop Length(b) The size <strong>of</strong> oscillations in WSFR over the simulations.Figure 60: Variations in the WSFR in the 72-nephron system where allnephrons <strong>of</strong> a specific loop length are impaired.It is also important to note that the lowest-frequency peak in allthree graphs covers the natural frequencies <strong>of</strong> nephrons with looplengths 2 and 3. For columns 1–3, the loop length 3 frequency dominatedand the loop length 2 frequency formed a shoulder (Fig. 62a),while for columns 4–9 it was the loop length 2 frequency that dominatedand the loop length 3 frequency that formed a shoulder(Fig. 62b and Fig. 62c). This indicates that the impaired nephronsare possibly having a larger effect on columns 1–3, as these were theonly columns to show a clear increase in the peak size for the looplength 3 frequency.However, as was demonstrated in earlier experiments (Sect. 6.4),an increased peak at the loop length 3 frequency does not mean thatthe impaired nephrons are having a significant effect—in Fig. 56b


96 E M E R G E N T E FF E C T S O F L O C A L I S E D I M PA I R M E N T18501800ImpairedOriginalWSFR (nL/min)17501700165016001200 1800 2400 3000Timestep(a) The WSFR in both the original and impaired 72-nephron systems.1.2e-161e-16ImpairedOriginal8e-17Power6e-174e-172e-1700 0.02 0.04Normalised Frequency(b) The power spectral densities <strong>of</strong> WSFR in both the original and impaired72-nephron systems.Figure 61: Comparison <strong>of</strong> the WSFR in the original 72-nephron systemto that <strong>of</strong> the 72-nephron system where all nephrons <strong>of</strong> a specific looplength are impaired.this peak is greatly exaggerated in the impaired system, and yet theimpaired system exhibits a decrease in average WSFR <strong>of</strong> only 1.4%.In this experiment, columns 1–3 exhibited the smallest decrease inaverage WCFR and columns 7–9 exhibited the largest decrease, asdemonstrated in Fig. 64a, despite columns 1–3 also having the largestincrease in the loop length 3 peak (the changes in power spectraldensity <strong>of</strong> the columns are shown in Fig. 63).In summary, the WSFR and WCFR power spectral densities showedsmaller differences between the original and impaired 72-nephronsystems than those observed in the equivalent experiment on theeight-nephron system (Sect. 6.4.2), and the decrease in average WSFR


6.5 I M PA I R M E N T S O N A L A R G E R S C A L E 974.5e-174e-17ImpairedOriginal3.5e-173e-17Power2.5e-172e-171.5e-171e-175e-1800 0.02 0.04Normalised Frequency(a) Columns 1–3.3e-172.5e-17ImpairedOriginal2e-17Power1.5e-171e-175e-1800 0.02 0.04Normalised Frequency(b) Columns 4–6.Power2e-171.8e-171.6e-171.4e-171.2e-171e-178e-186e-184e-182e-18ImpairedOriginal00 0.02 0.04Normalised Frequency(c) Columns 7–9.Figure 62: A comparison <strong>of</strong> the power spectral densities <strong>of</strong> WCFRbetween the original 72-nephron system and the system where allnephrons <strong>of</strong> a specific loop length are impaired.


98 E M E R G E N T E FF E C T S O F L O C A L I S E D I M PA I R M E N TPower1.2e-171e-178e-186e-184e-182e-180-2e-18-4e-18-6e-18-8e-180 0.02 0.04Normalised Frequency(a) Columns 1–3.Power7e-186e-185e-184e-183e-182e-181e-180-1e-18-2e-18-3e-180 0.02 0.04Normalised Frequency(b) Columns 4–6.1.5e-181e-185e-19Power0-5e-19-1e-18-1.5e-18-2e-180 0.02 0.04Normalised Frequency(c) Columns 7–9.Figure 63: The difference in power spectral densities <strong>of</strong> WCFR betweenthe original 72-nephron system and the system where all nephrons <strong>of</strong>a specific loop length are impaired.


6.5 I M PA I R M E N T S O N A L A R G E R S C A L E 99Average WCFR (nL/min)2102001901801700x1ImpairedOriginal2 0 1y2(a) The average WCFR for each column in the original and impaired systems.Range in WCFR (nL/min)1301201101000x1ImpairedOriginal2 0 1y2(b) The range in WCFR for each column in the original and impaired systems.Figure 64: Variations in WCFR in the 72-nephron system where allnephrons <strong>of</strong> loop length 3 are impaired.<strong>of</strong> 1.3% is almost identical to the decrease <strong>of</strong> 1.4% observed in theeight-nephron experiment.6.5.2 Impaired columnsIn this experiment, for each column in the model a simulation wasperformed where all <strong>of</strong> the nephrons in that column were impairedat t = 1, 200. A comparison <strong>of</strong> the WSFR for each <strong>of</strong> these simulationsis presented in Fig. 65. Fig. 65a shows that the average WSFR decreasedthe most when the bottom-most columns in the arterial tree(columns 7–9) were impaired—a maximal decrease <strong>of</strong> 0.3%. All <strong>of</strong>the impaired simulations exhibited a smaller range in WSFR than theoriginal system (Fig. 65b), with the smallest range being produced


100 E M E R G E N T E FF E C T S O F L O C A L I S E D I M PA I R M E N Twhen one <strong>of</strong> columns 4–6 was impaired. A potential explanation forthis observation is that columns 4–6 undergo inter-column diffusionwith columns at both top and bottom branches <strong>of</strong> the arterial tree,possibly resulting in a more wide-spread influence over the dynamics<strong>of</strong> the other columns.Average WSFR (nL/min)1730172017100x1ImpairedOriginal2 0 1y2(a) The average WSFR over the simulations.Range in WSFR (nL/min)2302202102001900x1ImpairedOriginal2 0 1y2(b) The size <strong>of</strong> oscillations in WSFR over the simulations.Figure 65: Variations in the WSFR in the 72-nephron simulations whereall <strong>of</strong> the nephrons in a particular column (x, y) are impaired.The remainder <strong>of</strong> the analysis will contrast the dynamics <strong>of</strong> theoriginal 72-nephron system with the dynamics <strong>of</strong> the system whereall nephrons in column 9 are impaired, as this is one <strong>of</strong> the threesystems that demonstrates the largest decrease in WSFR (shown inFig. 65a).The WSFRs <strong>of</strong> the original and impaired systems are plotted inFig. 66a—the decrease in average WSFR is not discernible, as thedifference is 0.3%. The power spectral densities <strong>of</strong> the two WSFRs are


6.5 I M PA I R M E N T S O N A L A R G E R S C A L E 101shown in Fig. 66b, where the impaired system is again shown to havean enlarged peak at the natural frequency <strong>of</strong> loop length 3, and alsoa decreased peak at the natural frequency <strong>of</strong> loop length 1—a trendnot observed in the previous experiments.18501800ImpairedOriginalWSFR (nL/min)175017001650160015501200 1800 2400 3000Timestep(a) The WSFR in both the original and impaired 72-nephron systems.Power1e-169e-178e-177e-176e-175e-174e-173e-172e-171e-17ImpairedOriginal00 0.02 0.04Normalised Frequency(b) The power spectral densities <strong>of</strong> WSFR in both the original and impaired72-nephron systems.Figure 66: Comparison <strong>of</strong> the WSFR in the original 72-nephron systemand the system where all <strong>of</strong> the nephrons in column 9 are impaired.The analysis now turns to the WCFRs <strong>of</strong> the individual columns in themodel. The power spectral densities <strong>of</strong> the WCFRs—shown in Fig. 67—were very similar to those in the previous experiment (Fig. 62), exceptthat the mountain at the natural frequencies <strong>of</strong> loop lengths 2 and3 exhibited a small valley between the two peaks. Apart from thesechanges, there appears to be little difference between the spectraldensities <strong>of</strong> the WCFRs in the original and impaired systems.Fig. 68 shows the changes in WCFR spectral densities as a result <strong>of</strong>the impairment in column 9. In comparison to the changes in WCFR


102 E M E R G E N T E FF E C T S O F L O C A L I S E D I M PA I R M E N T4e-173.5e-17ImpairedOriginal3e-17Power2.5e-172e-171.5e-171e-175e-1800 0.02 0.04Normalised Frequency(a) Columns 1–3.2.5e-172e-17ImpairedOriginalPower1.5e-171e-175e-1800 0.02 0.04Normalised Frequency(b) Columns 4–6.Power2e-171.8e-171.6e-171.4e-171.2e-171e-178e-186e-184e-182e-18ImpairedOriginal00 0.02 0.04Normalised Frequency(c) Columns 7–8.Figure 67: A comparison <strong>of</strong> the spectral densities <strong>of</strong> WCFR between theoriginal 72-nephron system and the system where all <strong>of</strong> the nephronsin column 9 are impaired.


6.5 I M PA I R M E N T S O N A L A R G E R S C A L E 1031e-185e-19Power0-5e-19-1e-18-1.5e-180 0.02 0.04Normalised Frequency(a) Columns 1–3.3e-182.5e-182e-18Power1.5e-181e-185e-190-5e-19-1e-180 0.02 0.04Normalised Frequency(b) Columns 4–6.3e-182e-181e-18Power0-1e-18-2e-18-3e-180 0.02 0.04Normalised Frequency(c) Columns 7–8.Figure 68: The difference in power spectral densities <strong>of</strong> WCFR betweenthe original 72-nephron system and the system where all <strong>of</strong> thenephrons in column 9 are impaired.


104 E M E R G E N T E FF E C T S O F L O C A L I S E D I M PA I R M E N T3e-172.5e-17ImpairedOriginal2e-17Power1.5e-171e-175e-1800 0.02 0.04Normalised Frequency(a) A direct comparison <strong>of</strong> the spectral densities.Power9e-188e-187e-186e-185e-184e-183e-182e-181e-180-1e-18-2e-180 0.02 0.04Normalised Frequency(b) The difference in spectral densities.Figure 69: The difference in the power spectral density <strong>of</strong> column 9in the original 72-nephron system and the system where all <strong>of</strong> thenephrons in column 9 are impaired.spectral densities from the previous experiment (Fig. 63), severalobservations can be made.• Columns 1–3 exhibit similar trends in both experiments (compareFig. 68a and Fig. 63a), but the magnitude <strong>of</strong> the increasesand decreases in peak size are approximately 5–10 times smaller.• Columns 4–6 also exhibit similar trends in both experiments, butagain the magnitude <strong>of</strong> the changes is reduced—this time 2–3times smaller— and there is no peak at the natural frequency <strong>of</strong>loop length 0 in this experiment.• Columns 7–8 do not exhibit trends similar to those in the previousexperiment, and are also the only columns that exhibit


6.5 I M PA I R M E N T S O N A L A R G E R S C A L E 105Average WCFR (nL/min)2102001901801700x1ImpairedOriginal2 0 1y2(a) The average WCFR for each column in the impaired and original systems.Range in WCFR (nL/min)1301201100x1ImpairedOriginal2 0 1y2(b) The range in WCFR for each column in the impaired and original systems.Figure 70: Variations in WCFR in the system where all <strong>of</strong> the nephronsin column 9 are impaired.larger changes in these spectral densities than in the previousexperiment. This is most likely due to the impaired state <strong>of</strong> column9 exerting a larger influence on these two columns—via thearterial tree—than the effects <strong>of</strong> the impairment in the previousexperiment.The spectral density <strong>of</strong> the WCFR <strong>of</strong> column 9 is shown in Fig. 69a,and while it shares a similar shape with the WCFRs <strong>of</strong> the othercolumns (Fig. 67), it is the only column that exhibited such markedincreases in the natural frequencies <strong>of</strong> loop lengths 2 and 3. Thechanges in the spectral density are shown in Fig. 68, which is uniquein that all peaks except one are positive and that the peak size isinversely proportional to the normalised frequency.


106 E M E R G E N T E FF E C T S O F L O C A L I S E D I M PA I R M E N TA direct comparison <strong>of</strong> the individual WCFRs in the original andimpaired systems is given in Fig. 70. Only column 9 shows evidence<strong>of</strong> a change in WCFR (Fig. 70a), which would indicate that the othercolumns are unaffected by the impairment <strong>of</strong> column 9. However,examining the size <strong>of</strong> the oscillations in WCFR reveals that column9 has influenced other columns (Fig. 70b)—columns 1–3 show adecrease <strong>of</strong> 0.5%, columns 4–6 show a decrease <strong>of</strong> 2.6%, column 7shows an increase <strong>of</strong> 0.6%, column 8 an increase <strong>of</strong> 0.8% and column9 an increase <strong>of</strong> 12%.Ignoring column 9, the changes in WCFR oscillation size are essentiallyuniform for all <strong>of</strong> the columns sharing the same branch <strong>of</strong> thearterial tree—there is only a 0.03% difference in the oscillation sizes<strong>of</strong> columns 5 and 6, even though column 6 undergoes inter-columndiffusion with the impaired column and column 5 does not. This wouldappear to indicate that the inter-column diffusion mechanism is ableto dampen out local changes, and that the arterial tree has a moredominant role in propagating the effects <strong>of</strong> the impairment acrossthe system.6.5.3 Summary <strong>of</strong> experimentsThe first experiment conducted on the 72-nephron system—the longestloopednephrons in each column were impaired—demonstrated a1.3% decrease in average WSFR. The same experiment, when conductedon the eight-nephron system (Sect. 6.4.2), demonstrated a1.4% decrease in average WSFR. These values were expected to besimilar, as the individual columns in this model share a structureidentical to that <strong>of</strong> the eight-nephron model, and the impairments inboth experiments were identical. The inter-column diffusion couldnot have caused any difference in average WSFR, since the interstitialfluid in each column evolved identically over time, due to the identicalstructure and impairment shared by all <strong>of</strong> the columns. It is possiblethat the difference in average WSFR between these experiments isdue to the arterial tree structure connecting the columns, but thedifference is so small that a more detailed investigation is required todetermine this for certain.In contrast, when the impairment was localised to a single columnin the 72-nephron system, the impairment did not appreciably affectthe dynamics <strong>of</strong> the system—average WSFR was decreased by only0.3% and neighbouring columns were not significantly affected (seeFig. 70). This is due to the inter-column diffusion (depicted in Fig. 16),which acts as an averaging mechanism for solute concentration gradientsin the interstitial fluid. The arterial tree was able to affectthe magnitude <strong>of</strong> the oscillations in WSFR (shown in Fig. 65b), butthe inter-column diffusion was able to keep the average WSFR almostconstant (shown in Fig. 65a), regardless <strong>of</strong> the choice <strong>of</strong> impairedcolumn. As the choice <strong>of</strong> column affects the number <strong>of</strong> neighbouringcolumns with which inter-column diffusion occurs, this indicates thatthe inter-column diffusion is a relatively strong agent for overcomingthe localised effects <strong>of</strong> this impairment.


6.6 C O N C L U S I O N S A B O U T S Y S T E M S TA B I L I T Y 1076.6 C O N C L U S I O N S A B O U T S Y S T E M S TA B I L I T YAs mentioned at the start <strong>of</strong> the chapter, the impairment presentedhere and the accompanying experiments serve only to illustrate theapproach <strong>of</strong> modelling and analysing the effects <strong>of</strong> a localised impairmenton multi-nephron systems. However, the experiments conductedhere demonstrated that a system consisting <strong>of</strong> a single column <strong>of</strong> interstitialis strongly resistant to the effects <strong>of</strong> the impairment. Furthermore,for a system consisting <strong>of</strong> multiple columns, the inter-columndiffusion mechanism further reduced the effects <strong>of</strong> the impairment, tothe point that there were little discernible effects on the whole-systemdynamics.Without making any predictions or observations concerning thereal-world effects <strong>of</strong> renal diseases on the kidney, these experimentsdemonstrated that the multi-nephron systems maintain remarkablystable dynamics when the function <strong>of</strong> individual nephrons are impaired.Despite the simplicity <strong>of</strong> the impairment model that was used,we contend that this stability is a feature <strong>of</strong> the model—and <strong>of</strong> thekidney proper—rather than a deficiency in the impairment model.


L A R G E - S C A L E S I M U L AT I O N77.1 I N T RO D U C T I O NSince one <strong>of</strong> the motivations for this model is to make whole-kidneysimulations computationally tractable, our attention now turns to howthe performance <strong>of</strong> the simulations scale as the size <strong>of</strong> the modelincreases. The experiments conducted in Chapter 5 and Chapter 6have only examined the behaviour <strong>of</strong> the model in the context <strong>of</strong> multinephronsystems whose sizes pale in comparison to that <strong>of</strong> an entirekidney. To discuss the simulation performance, several distinctionsmust be made:M O D E L The network model, as presented in Chapter 4.A L G O R I T H M A computational method for calculating the dynamics<strong>of</strong> a given model.I M P L E M E N TAT I O N A computer program that executes an algorithm.P RO C E S S A sequence <strong>of</strong> computations, performed in serial.S I M U L AT I O N The act <strong>of</strong> using an implementation to calculate thedynamics <strong>of</strong> a given model.This chapter begins by presenting the algorithm that was used forthe experiments in Chapter 5 and Chapter 6, followed by a performanceanalysis <strong>of</strong> the resulting implementation. Despite the fact thatthis implementation is single-threaded, and is therefore unable tomake use <strong>of</strong> more than one central processing unit (CPU), it is shownthat using this implementation to simulate a whole-kidney model iscomputationally tractable.Secondly, the simulation performance can be greatly improved byutilising more than one CPU, and a new algorithm is presented thatperforms each time-step as a number <strong>of</strong> concurrent processes, each<strong>of</strong> which can be performed on a different CPU. This algorithm wasimplemented using the Join Calculus [60, 59], which allows a singlesimulation to be distributed over multiple computers. A performanceanalysis demonstrates that this results in a significant performanceincrease when more than one CPU is available.Finally, to calculate the dynamics <strong>of</strong> an entire kidney, it is notsufficient for the simulation to be computationally tractable—themodel must first be constructed. The structure <strong>of</strong> the models used inChapter 5 and Chapter 6 were manually specified—the spatial parameters<strong>of</strong> each tubule segment, each nephron and the arterial tree—anapproach which clearly will not scale to models <strong>of</strong> 10 6 nephrons. Althoughno solution has currently been implemented, two techniquesfor automatically generating such large models are proposed.109


110 L A R G E-SCALE S I M U L AT I O N7.2 P E R F O R M A N C E O F T H E I M P L E M E N TAT I O N7.2.1 The single-threaded algorithmA pseudocode description <strong>of</strong> the single-threaded algorithm is given inListing 7.1.1 # determine the resistance <strong>of</strong> the arterial tree2 for column in model.columns:3 for nephron in column.nephrons:4 # calculated by an update rule (Eq. 4.5)5 nephron.update_resistance()6 # calculated from the nephron resistances (Fig. 14)7 # Ohm’s law and Kirch<strong>of</strong>f’s laws are used8 column.update_resistance()9 # calculated from the column resistances (Fig. 15)10 # Ohm’s law and Kirch<strong>of</strong>f’s laws are used11 R T = model.update_resistance()1213 # update the model parameters (Sect. 4.4.3)14 model.update_params()1516 # generate the input for the next time-step17 # volume = P A−P VR T(Sect. 4.2.2)18 input = model.input_blood(R T )1920 # distribute the blood to the nephrons21 for column in model.columns:22 # depends on the tree <strong>of</strong> column resistances (Fig. 15)23 column_input = model.distribute(input, column)24 for nephron in column.nephrons:25 # depends on the tree <strong>of</strong> nephron resistances (Fig. 14)26 nephron_input = column.distribute(column_input, nephron)27 nephron.set_input(nephron_input)2829 # simulate the time-step for each tubule segment (Sect. 4.4.5)30 # also update the state variables (Sect. 4.4.4)31 for column in model.columns:32 column.simulate_timestep()3334 # perform diffusion between neighbouring columns (Fig. 16)35 for column1, column2 in model.column_neighbours:36 column1.diffuse_with(column2)3738 # update the nephron disease states (Sect. 6.1)39 for column in model.columns:40 for nephron in column.nephrons:41 nephron.update_disease_state()Listing 7.1: The single-threaded algorithm in pseudocode.✆


7.2 P E R F O R M A N C E O F T H E I M P L E M E N TAT I O N 1117.2.2 Performance analysisThe single-threaded algorithm was implemented in OCaml [101] andwas compiled with the high-performance native-code compiler OCAM-LOPT. The simulations were performed on an Intel Core 2 Duo E6600with 2Gb <strong>of</strong> RAM, running Ubuntu Linux. The parameter values donot affect the simulation performance, as they do not affect the calculationsused to solve the model equations.The memory usage <strong>of</strong> the simulations (shown in Fig. 71a) growslinearly with the number <strong>of</strong> nephrons in the model, and the number <strong>of</strong>columns into which the model is divided appears to have little effect.However, the execution times show significant differences, dependingon the number <strong>of</strong> columns in the model (Fig. 71b). The execution timegrows at O(n 3 ) for n nephrons per column, as demonstrated by thecurves fitted to the execution time plots (see Fig. 72). This is to beexpected, as solving the n × n matrix governing the loops <strong>of</strong> Henle(described in Sect. 4.4.5) is solved by Gaussian elimination using thestandard algorithm, which has a complexity <strong>of</strong> O(n 3 ).As demonstrated by Fig. 71c, when the number <strong>of</strong> nephrons ineach column is held constant, the execution time grows linearly withthe number <strong>of</strong> columns in the model. Thus, for a model containingc columns and n nephrons per column, the execution time for thesimulation is O(cn 3 ). This explains why the n 3 coefficients in Fig. 72decrease by a factor <strong>of</strong> four when c is doubled—n is halved to conservethe number <strong>of</strong> nephrons in the model, and so:c ′ = 2 × c (7.1)n ′ = n 2((O c ′ n ′3) ( n) ) ( 3 cn3)= O 2c × = O2 4(7.2)(7.3)For example, consider two models <strong>of</strong> 100,000 nephrons (Table 16).The first consists <strong>of</strong> 100 columns arranged in a 10 × 10 grid with1,000 nephrons per column; the second consists <strong>of</strong> 10,000 columnsarranged in a 100 × 100 grid with 10 nephrons per column. As theperformance <strong>of</strong> a model is O(cn 3 ), the performance <strong>of</strong> the first modelis O(10 11 ), compared to O(10 7 ) for the second model. This is not onlya huge improvement in performance, as the structure <strong>of</strong> the secondmodel more closely approximates a continuous body <strong>of</strong> interstitialfluid—the performance <strong>of</strong> the model improves the more the interstitialfluid structure approximates a continuous model.Using this implementation, a simulation <strong>of</strong> a 512-nephron model(8 nephrons per column) progressed at 444 nephron time-steps persecond and used 28Mb <strong>of</strong> RAM (56kb per nephron). Assuming thatthe linear relationships in Fig. 71a and Fig. 71c hold for a model <strong>of</strong>10 6 nephrons, this indicates that the simulation <strong>of</strong> a single time-stepfor a whole-kidney model would require 37.5 minutes and 53.4Gb<strong>of</strong> RAM. These orders <strong>of</strong> magnitude demonstrate that the implementationscales up to a whole-kidney model. We are not aware <strong>of</strong> anyexisting nephron models for which a whole-kidney simulation is computationallyfeasible.


112 L A R G E-SCALE S I M U L AT I O N981 column2 columns4 columns7RAM (MB)6543210 20 30 40 50 60 70Model size (nephrons)(a) Memory usage9008007001 column2 columns4 columns600Time (s)500400300200100010 20 30 40 50 60 70Model size (nephrons)(b) Execution time1401208 nephrons/column12 nephrons/column16 nephrons/column100Time (s)8060402000 10 20 30 40 50 60 70Number <strong>of</strong> nephrons(c) Execution timeFigure 71: Comparing 1-column, 2-column and 4-column simulations.


7.2 P E R F O R M A N C E O F T H E I M P L E M E N TAT I O N 113900800Execution TimeBest FitTime (s)7006005004003002001005.7×10 -2 n 2 + 2.5×10 -3 n 3010 20 30 40 50 60 70Number <strong>of</strong> nephrons (n)(a) Single-column simulations300250Execution TimeBest FitTime (s)2001501002.1 + 0.5n + 2.2×10 -2 n 2 + 6.1×10 -4 n 350010 20 30 40 50 60 70Number <strong>of</strong> nephrons (n)(b) Two-column simulationsTime (s)14012010080604020Execution TimeBest Fit1.2 + 0.7n + 8.4×10 -3 n 2 + 1.6×10 -4 n 3010 20 30 40 50 60 70Number <strong>of</strong> nephrons (n)(c) Four-column simulationsFigure 72: Execution time is O(n 3 ) for n nephrons per column.


114 L A R G E-SCALE S I M U L AT I O NModel 1 Model 2Nephrons 10 5 10 5c 10 2 10 4n 10 3 10 1Time O(10 11 ) O(10 7 )Table 16: Performance comparison <strong>of</strong> two models <strong>of</strong> identical size.7.3 I M P ROV I N G P E R F O R M A N C E W I T H C O N C U R R E N C YAlthough whole-kidney simulations are achievable with the monolithicalgorithm, the performance <strong>of</strong> such simulations can readily beimproved. While reductions in execution time may be achieved byoptimising the monolithic implementation, we show that much largerimprovements can be achieved by dividing the simulation into multipleprocesses that can be executed concurrently. If the processescan be distributed over multiple computers—employing significantlymore resources than available on a single computer—the resultingperformance gain <strong>of</strong> the distributed system can be remarkable [94].Typical distributed systems such as Berkeley Open Infrastructurefor Network Computing (BOINC) [35, 7], SETI@home [94, 8] and <strong>Kidney</strong>Grid[42, 43], and distributed algorithms such as MapReduce [51]operate by dividing the computational work into a number <strong>of</strong> distinctprocesses, executing each process independently, and performingsome analysis or calculation on the results.In contrast, the model presented in Chapter 4 requires the modelstate at time t to be calculated before the state at time t + ∆ canbegin to be calculated. As a result, if the simulation were broken intomultiple processes, the processes would need to communicate witheach other after every time-step. This is in contrast to the distributedsystems described previously, which perform multiple jobs concurrentlyand then combine the results in some manner. Accordingly, aconcurrent algorithm for the model simulation will necessarily differfrom such systems and cannot be implemented using an identicalapproach.7.3.1 The concurrent algorithmThe single-threaded algorithm Listing 7.1 can be divided into a number<strong>of</strong> separate processes, based on the hierarchical structure <strong>of</strong>the model. Once the input has been generated and distributed tothe columns, the time-step can be simulated for each column independently,with the exception <strong>of</strong> the inter-column diffusion. Theinter-column diffusions can also be performed independently: foreach pair <strong>of</strong> neighbouring columns, assign one the “master”, which isresponsible for simulating the diffusion with the “slave” (see Fig. 73).This algorithm consists <strong>of</strong> a process for each column in the model(referred to as column processes and abbreviated to “Col1”, “Col2”,etc., in diagrams), and a process for the generation and distribution


7.3 I M P ROV I N G P E R F O R M A N C E W I T H C O N C U R R E N C Y 115<strong>of</strong> the model input (referred to as the “controller” and abbreviatedto “Ctrl” in diagrams). The algorithm is presented in pseudocodeform in Listing 7.2 and Listing 7.3, and the communication betweenthe processes is shown in Fig. 74 and Fig. 75. To reduce the delaybetween time-steps, the column processes can be changed so thatonce they receive input, they calculate the resistance <strong>of</strong> the arterialtree and immediately send this value to the controller (demonstratedin Fig. 76).1 # receive the resistances <strong>of</strong> each column in the model2 for column in model.columns:3 resistance = receive_message(column, send_resistance)4 R T = update_resistance(column, resistance)56 # generate the input for the next time-step7 # volume = P A−P VR T(Sect. 4.2.2)8 input = model.input_blood(R T )910 # distribute the blood to the columns11 for column in model.columns:12 # depends on the tree <strong>of</strong> column resistances (Fig. 15)13 column_input = model.distribute(input, column)14 send_message(column, column_input)Listing 7.2: The controller process in pseudocode.✆MasterMaster(a)Slave(b)SlaveFigure 73: The master/slave relationship between columns, for simulatingthe inter-column diffusion (see Fig. 16). The arrow (a) indicatesthat the column labelled “Master” performs the diffusion with thehighlighted column; the arrow (b) indicates that the highlighted columnperforms the diffusion with the column labelled “Slave”.The presentation <strong>of</strong> the column algorithm (Listing 7.2) has ignoredthe complexities <strong>of</strong> performing the inter-column diffusion. Each columnmust not only perform the diffusion between itself and neighbouringslaves, it must also communicate with neighbouring masters(see Fig. 75). This more detailed algorithm is presented in Listing 7.4.A significant feature <strong>of</strong> this concurrent algorithm is the total absence<strong>of</strong> shared state between processes—the only shared data ispassed between processes using messages. This greatly simplifies


116 L A R G E-SCALE S I M U L AT I O N1 # determine the resistance <strong>of</strong> the arterial tree2 for nephron in column.nephrons:3 # calculated by an update rule (Eq. 4.5)4 nephron.update_resistance()5 # calculated from the nephron resistances (Fig. 14)6 R C = column.update_resistance()78 # send the column resistance to the controller9 send_message(controller, R C )1011 # receive the input blood for the next time-step12 receive_message(column_input)1314 # distribute the blood to the nephrons15 for nephron in column.nephrons:16 # depends on the tree <strong>of</strong> nephron resistances (Fig. 14)17 nephron_input = column.distribute(column_input, nephron)18 nephron.set_input(nephron_input)1920 # simulate the time-step for each tubule segment (Sect. 4.4.5)21 # also update the state variables (Sect. 4.4.4)22 column.simulate_timestep()2324 # perform diffusion between neighbouring columns (Fig. 16)25 for neighbour in column.slaves:26 column.diffuse_with(slave)2728 # update the nephron disease states (Sect. 6.1)29 for nephron in column.nephrons:30 nephron.update_disease_state()Listing 7.3: The column process in pseudocode.✆Ctrl Col1 Col2 ... ColNMessagesSend resistanceSend inputActionsGenerate inputSimulate time-stepFigure 74: Communication between the controller and columns. Ateach time-step, the controller sends input (blood) to each column; thecolumns then simulate the time-step; finally the columns send theirupdated arterial resistances to the controller.the implementation and excludes the possibility <strong>of</strong> problems suchas deadlock, livelock and race conditions, which are commonly as-


7.3 I M P ROV I N G P E R F O R M A N C E W I T H C O N C U R R E N C Y 117MasterSlaveMessagesRequest fluidSend fluidSend changesActionsCalculate changesUpdate fluidFigure 75: Communication between columns when performing theinter-column diffusion.Ctrl Col1 Col2 ... ColNMessagesSend resistanceSend inputActionsGenerate inputSimulate time-stepFigure 76: Reordering the communication in Fig. 74, in order toreduce the delay between time-steps.sociated with concurrent programs. Perhaps the biggest advantageis that the implementation can easily be distributed across multiplecomputers, since the lack <strong>of</strong> shared memory between computers hasno impact.7.3.2 Communication in the concurrent algorithmThe concurrent algorithm that has been described involves regularcommunication between the individual processes—for each time-step,two messages are sent between each column and the controller(Fig. 74) and three messages are sent between each pair <strong>of</strong> neighbouringcolumns (Fig. 75). If the columns are distributed across anetwork <strong>of</strong> computers, the limitations <strong>of</strong> the network hardware canimpact upon the performance <strong>of</strong> this communication, and thus reducethe simulation performance.


118 L A R G E-SCALE S I M U L AT I O N1 # save the current state <strong>of</strong> the interstitial fluid2 F C = copy(column.interstitial_fluid)34 # ask neighbouring slaves for their interstitial fluid5 for neighbour in column.slaves:6 send_message(neighbour, request_interstitial_fluid)78 # count how many slaves and masters have not completed the diffusion9 slaves = column.slave_count10 masters = column.master_count1112 while true:13 # receive messages from neighbouring columns14 message = receive_message()1516 # a master has requested the column’s interstitial fluid17 if message == request_interstitial_fluid:18 send_message(message.sender, F C )19 # a master has adjusted the interstitial fluid20 elif message == adjust_interstitial_fluid:21 column.adjust_interstitial_fluid(message.changes)22 master_count -= 123 # a slave has sent its interstitial fluid24 elif message == send_renal_fluid:25 # calculate the changes to the fluid in both columns26 local, remote = diffusion_changes(F C , message.fluid)27 # inform the slave to adjust its interstitial fluid28 send_message(message.sender, adjust_renal_fluid(remote))29 # adjust this column’s interstitial fluid30 column.adjust_interstitial_fluid(local)31 slave_count -= 132 # stop when all diffusions have been completed33 if slaves == 0 and masters = 0:34 breakListing 7.4: The algorithm for inter-column diffusion in pseudocode.✆Message Data Size (bytes)Send resistance double 8Send input double list (fluid) 8 × SRequest interstitial fluid none 0Send interstitial fluid fluid list 8 × S × LAdjust interstitial fluid fluid list 8 × S × LTable 17: Message content in inter-process communication. Themodel supports S solutes and each column has L layers <strong>of</strong> interstitialfluid.The content <strong>of</strong> each message sent between the individual processesis documented in Table 17. Based on this data, the maximum volume<strong>of</strong> data sent by a column in a single time-step is given by Eq. 7.4.32 × S × L + 8 × S + 8 (7.4)


7.3 I M P ROV I N G P E R F O R M A N C E W I T H C O N C U R R E N C Y 119The model currently supports eight solutes (Sect. 4.4.3) and themodels used in Chapter 5 and Chapter 6 used up to sixteen layers <strong>of</strong>interstitial fluid, resulting in a maximum volume <strong>of</strong> 4168 bytes (4.07kb).To examine the worst-case volume <strong>of</strong> communication, consider awhole-kidney model (i. e., 10 6 nephrons) that consists <strong>of</strong> 10 6 columns,each containing a single nephron. Using Eq. 7.4, the volume <strong>of</strong> datasent at each time-step would be 3.88Gb—such a large volume <strong>of</strong>communication would certainly reduce the simulation performance.This volume can be decreased by reducing the number <strong>of</strong> columnsin the model, but there are two further options that do not affectthe model structure: locating multiple columns on the same computer(i. e., sharing a single CPU or using multiple CPUs); and allowingcolumns to simulate the behaviour <strong>of</strong> multiple columns. Both <strong>of</strong> theseoptions were explored with the implementation.7.3.3 Performance <strong>of</strong> the implementationAlthough the concurrent algorithm differs significantly from the originalalgorithm (Listing 7.1), the concurrent implementation sharesmost <strong>of</strong> its code with the original implementation (Sect. 7.2.2). Theconcurrent implementation was written in JoCaml [110, 147, 100],an extension <strong>of</strong> OCaml that incorporates the Join Calculus [60, 59] toallow for concurrent and distributed programming.The Join Calculus—like other process calculi such as CCS [120],the π Calculus [121, 122], the Fusion Calculus [140, 141] and theAmbient Calculus [34]—is used to model concurrent systems. TheJoin Calculus is equivalent to the π Calculus [59], but is easier toimplement in a distributed setting [59].The Join Calculus uses a message-passing concurrency model verysimilar to the one used to present the concurrent algorithm (Listing7.2 and Listing 7.3), which greatly simplified the implementationprocess. To illustrate how well the JoCaml semantics match the concurrentalgorithm, the implementation <strong>of</strong> the inter-column diffusion(Listing 7.4) is presented in Listing 7.5. The only difference is thatthe loop and the variables slaves and masters are captured by themessage diff_state(column, F C , slaves, masters).Amdahl’s law [6] establishes the theoretical maximum speedup thatcan be gained by performing a computation with multiple processors,where a fraction p <strong>of</strong> the computation can be executed in paralleland the computation is performed on N identical processors (Eq. 7.5).According to this law, we expect to see a significant speedup asthe simulation is distributed over a greater number <strong>of</strong> processors,although the return diminishes and approaches a limit (see Eq. 7.6and Fig. 77).1S(N) =(1 − p) + p/Nlim S(N) = 1N→∞ 1 − p(7.5)(7.6)This law is based on the implicit assumption that p is independent<strong>of</strong> N, which is only the case if the size <strong>of</strong> the problem is fixed. If thisis not the case and the problem size can scale with the number <strong>of</strong>


120 L A R G E-SCALE S I M U L AT I O N101 # start the inter-column diffusion2 diff_start(column) =3 # save the current state <strong>of</strong> the interstitial fluid4 F C = copy(column.interstitial_fluid)5 # ask neighbouring slaves for their interstitial fluid6 for neighbour in column.slaves:7 spawn request_fluid(column, neighbour)8 # wait to receive messages from neighbouring columns9 diff_state(column, F C , column.slaves, column.masters)11 # no more masters or slaves to wait for12 diff_state(column, F C , 0, 0) =13 # the inter-column diffusion is complete14 diff_complete(column)1516 # a master requests the column’s interstitial fluid17 diff_state(column, F C , slaves, masters)18 & request_fluid(master, column) =19 # send the interstitial fluid to the master20 send_fluid(master, column, F C )21 # still have to wait for all <strong>of</strong> the masters and slaves22 diff_state(column, F C , slaves, masters)2324 # a master changes the column’s interstitial fluid25 diff_state(column, F C , slaves, masters)26 & adjust_fluid(master, column, changes) =27 # adjust the interstitial fluid28 column.adjust_fluid(changes)29 # there is now one less master to wait for30 diff_state(column, F C , slaves, masters - 1)3132 # a slave sends its interstitial fluid33 diff_state(column, F C , slaves, masters)34 & send_fluid(column, slave, fluid) =35 # calculate the changes to the fluid in both columns36 local, remote = diffusion_changes(F C , message.fluid)37 # inform the slave to adjust its interstitial fluid38 adjust_fluid(column, slave, remote)39 # adjust this column’s interstitial fluid40 column.adjust_interstitial_fluid(local)41 # there is now one less slave to wait for42 diff_state(column, F C , slaves - 1, masters)Listing 7.5: The JoCaml implementation <strong>of</strong> inter-column diffusion inpseudocode form.✆processors (e. g., the complexity <strong>of</strong> the model may be increased byrefining the equations, incorporating additional equations, or augmentingthe structure), this will usually increase the relative amount<strong>of</strong> the computation that can be executed in parallel. As a result, thespeedup gained by using additional processors can significantly exceedthe original limit proposed by Amdahl’s law (e. g., by an order <strong>of</strong>magnitude [70]).For the model presented here, the problem size can be expanded byincreasing the complexity <strong>of</strong> the model as described in the previous


7.3 I M P ROV I N G P E R F O R M A N C E W I T H C O N C U R R E N C Y 121Speedup2018161412108642Parallel %95%90%75%50%01 8 64 512 4096 32768N (processors)Figure 77: Amdahl’s law for the theoretical maximum speedup whenusing multiple processors.876Measurementsp = 95%p = 90%p = 85%Speedup543212 4 6 8 10 12N (processors)Figure 78: The improvement in performance as the simulation isdistributed over multiple computers.paragraph, and also by increasing the number <strong>of</strong> nephrons in themodel, although it is not sensible to have a model that containsmore nephrons than are present in a kidney. Since the portion <strong>of</strong>the implementation that must be executed in serial is very small,and the complexity <strong>of</strong> the model will only increase as it is improvedand extended, it is evident that the distributed implementation willsignificantly benefit from being executed over many processors.To determine the maximal speedup obtainable by distributing thesimulation over multiple processors, a system <strong>of</strong> 384 nephrons wassimulated for 1, 000 time-steps using 1, 2, 3, 4, 6 and 12 computers(not including the computer running the controller process). Thetiming measurements are plotted in Fig. 78, which clearly shows that


122 L A R G E-SCALE S I M U L AT I O Nthe distributed implementation corresponds to p ≈ 90%. Accordingly,the maximal speedup that can be obtained is given by Eq. 7.7. Thenetwork capacity may have reduced the performance somewhat, andit is possible that a larger model would have reduced the frequency<strong>of</strong> communication and resulted in a larger performance increase. Alarger model was not used to test this hypothesis, due to the limitations<strong>of</strong> model construction—for a description <strong>of</strong> this issue and aproposed solution, see Sect. 7.4.lim S(N) = 1N→∞ 1 − p ≈ 1= 10 (7.7)1 − 0.9Therefore, given sufficient hardware, the distributed implementationshould be capable <strong>of</strong> providing a 10-fold performance increaseover the original single-threaded implementation. Based on the performanceestimates for a whole-kidney simulation given in Sect. 7.2.2,this would result in the simulation <strong>of</strong> a single time-step for a wholekidneymodel in 3.75 minutes (i. e., 3:45).7.4 AU T O M AT I C G E N E R AT I O N O F L A R G E M O D E L STo investigate the dynamics <strong>of</strong> sufficiently large multi-nephron systems(e. g.whole-kidney models), it is necessary to first constructthe model. The structure <strong>of</strong> the multi-nephron systems presented inChapter 5 and Chapter 6 were specified manually, but this approachis clearly infeasible for larger systems, so there is a need for automatedmodel growth. The kidney has a regular, but inhomogeneous,structure that cannot be easily replicated using well-known networkgrowth techniques [4] (e. g., preferential attachment).The automatic growth <strong>of</strong> models has been explored in the context <strong>of</strong>bone growth [29], and the dynamics <strong>of</strong> complex networks in responseto the addition <strong>of</strong> nodes has been studied across many fields [4]. Suchapproaches are not directly applicable to the growth <strong>of</strong> multi-nephronmodels because the addition <strong>of</strong> a nephron to an existing model is adiscrete action that cannot be performed at random, and the resultingmodel must have a physically-realistic structure.Instead, two alternative approaches to generate large multi-nephronsystems are proposed: a stochastic approach that uses statistical constraintson the model structure to generate a model that satisfiesthe constraints, and a rewriting system approach that generates thearterial tree and attaches nephrons to the tree in a stochastic manner.These two approaches are now presented in turn.7.4.1 Stochastic generation <strong>of</strong> large modelsThe stochastic approach allows the modeller to specify statisticalconstraints on various aspects <strong>of</strong> the model structure, from which amodel that satisfies these constraints is generated. Constraints canbe applied at three different levels <strong>of</strong> hierarchy: nephrons, interstitialfluid columns, and the entire model. Nephrons and interstitial fluidcolumns can also be divided into classes, each <strong>of</strong> which has its ownset <strong>of</strong> constraints.


7.4 AU T O M AT I C G E N E R AT I O N O F L A R G E M O D E L S 123For example, nephrons are <strong>of</strong>ten classified as cortical or juxtamedullary—thisapproach allows the modeller to define the constraintson these two classes independently. Similarly, columns <strong>of</strong>interstitial fluid can be divided into classes (e. g., based on how farinto the renal pyramid they descend) and have independent sets <strong>of</strong>constraints placed on them.The constraints on each nephron class are:• The length and location <strong>of</strong> each tubule segment in the interstitialfluid.• The value <strong>of</strong> nephron-specific model parameters such as k (seeTable 6).The constraints on each column class are:• The number <strong>of</strong> nephrons in the column.• The number <strong>of</strong> branches in the column’s arterial tree.• The distribution <strong>of</strong> nephron classes in the column.The constraints on the entire model structure are:• The number <strong>of</strong> interstitial fluid columns in the model.• The number <strong>of</strong> branches in the model’s arterial tree.• The distribution <strong>of</strong> column classes in the model.The stochastic nature <strong>of</strong> this approach is provided by acceptingthree different types <strong>of</strong> value for the constraints detailed above:• A single value (i. e., a constant);• A range from which a value will be chosen at random (e. g.,0 . . . 1); and• A list <strong>of</strong> values that associates a likelihood with each givenvalue.Neighbouring columns on a branch <strong>of</strong> the arterial tree are alsodefined to be neighbours in the 2D arrangement <strong>of</strong> columns, andthe only other neighbouring columns are those from neighbouringbranches. These limits on the model structure may reduce the ability<strong>of</strong> the generated models to accurately mimic a real kidney, but theyare easily solved. By allowing more general constraints (such asassigning likelihoods to columns <strong>of</strong> the same class being neighbours),it can not be guaranteed that the constraints will be solvable.7.4.2 Model growth as a rewriting systemAn alternative, more complex, approach to growing such large modelsis to generate the arterial tree for the entire model and then attachnephrons to this tree. This requires an understanding <strong>of</strong> the renalarterial tree structure and a method for using this information togenerate an arterial tree for the model. Based on experimental studies<strong>of</strong> vascular systems and the application <strong>of</strong> modelling techniques to


124 L A R G E-SCALE S I M U L AT I O Ncapture their structural properties, this approach should yield anaccurate model <strong>of</strong> the renal arterial tree, to which individual nephronscan then be attached.Vascular systems have been demonstrated to exhibit self-similarity,allowing their structure to be described by fractal models [22, 197,163, 207]. Studies <strong>of</strong> these fractal models suggest that there is an evolutionaryadvantage to branching systems having a fractal dimensions[196]—the fractal models are significantly less sensitive to structuralvariations than the classical scaling models [163, 196, 197].Lindenmayer systems [105, 115, 106, 107]—hereafter referredto as L-systems—are a class <strong>of</strong> rewriting system that are able torepresent fractal objects obtained by use <strong>of</strong> recursive transformations[85] (see Fig. 79 for an example), such as the fractal models that havebeen applied to vascular systems. Indeed, such systems have alreadybeen used to generate biological models such as virtual animals[82, 83], cellular processes [64], swarms [88] and vascular systems[207] including the human bronchial tree [22] and the human retina[93].In particular, the branching properties <strong>of</strong> the renal arterial tree[210] and the cardiovascular system [208, 209] have been shown toexhibit variations in branching parameters, resulting in multiple fractaldimensions. Radial, length and connectivity data for the arterialand venous trees <strong>of</strong> the rat kidney are also available [134].The fractal dimension <strong>of</strong> the renal arterial tree has been estimatedto be in the range 1.61–1.64 [49], which is consistent with valuesfound in artificial fractal structures generated by diffusion-limitedaggregation processes [200, 199, 198, 117] and similar to values proposedfor the retinal vasculature [123, 109]. However, the processesresponsible for the morphogenesis <strong>of</strong> the renal vasculature are stillnot fully understood [48].A straightforward approach to building the renal arterial tree is togrammatically evolve a rewriting system so that it builds a tree withthe desired fractal dimension [137] without analysing any graphicalrepresentations <strong>of</strong> the system [5]. This approach fails to take intoaccount the multifracticality <strong>of</strong> the renal arterial tree, but it is notimmediately clear what an impact (if any) this would have on themodel. It also uses the fractal dimension <strong>of</strong> the resulting tree as themeasure <strong>of</strong> fitness, although this is not necessarily a suitable metric,as noted by Zamir in [207]:Two such structures may have the same fractal dimensionby virtue <strong>of</strong> their space-filling properties, but havewidely different structures in their fluid dynamic designand function. A way <strong>of</strong> discussing fractal properties thatis particularly suited to open tree structures . . . is that<strong>of</strong> considering L-system models by which the tree can begenerated.The alternative approach suggested by Zamir allows the renal arterialtree to exhibit multiple fractal dimensions, by using a parametricL-system [207] that allows the branching parameters to change asthe tree descends. To determine the changes in the parameters wouldrequire dividing the renal arterial tree into a finite number <strong>of</strong> layers—each <strong>of</strong> which would ideally demonstrate little variation in its branch


7.4 AU T O M AT I C G E N E R AT I O N O F L A R G E M O D E L S 125L × 0.7940ᴼL × 0.7940ᴼ40ᴼL × 0.79²LL40ᴼ(a) The rule takes a line <strong>of</strong> length Land attaches two lines to one end, withlength L × 0.79, at angles <strong>of</strong> ±40 ◦ tothe original line.(b) The third stage <strong>of</strong> the system.(c) The sixth stage <strong>of</strong> the system.(d) The tenth stage <strong>of</strong> the system.Figure 79: An example <strong>of</strong> a Lindenmayer system that superficiallyresembles the arterial tree <strong>of</strong> the kidney, as presented in [207].


126 L A R G E-SCALE S I M U L AT I O Nstructure—and derive independent parameter values for each layer.The data provided by Nordsletten et al. would be a useful referencein determining how to divide the renal arterial tree into appropriatelayers [134].Both <strong>of</strong> the approaches outlined above rely on measurements <strong>of</strong> thebranching characteristics <strong>of</strong> the renal arterial tree. For the purposes<strong>of</strong> growing a model <strong>of</strong> the renal arterial tree, it is enough to measurethe length, diameter and branching angle <strong>of</strong> each branch in the tree[207]. Notwithstanding the techniques used in [134], an automatedprocess built to capture these parameters would be extremely useful.Baker has demonstrated that Model-Image Registration can be usedto fit a parametric 3D model to the curvature <strong>of</strong> the human cochlea,based on medical images [16]. This suggests that a similar approachmay be able to determine the branching parameters based on medicalimages <strong>of</strong> the renal arterial tree.If such a rewriting system were used to generate a representation<strong>of</strong> the vascular network for the model, this representation wouldcontain sufficient information to not only generate the arterial tree<strong>of</strong> the model, but to also give each arterial branch a spatial location.Given a desired resolution for the interstitial fluid columns, their 2Darrangement would then be specified by this spatial information. Asa result, this hypothetical arterial network generated by a rewritingsystem would specify the entire structure <strong>of</strong> the model, except for that<strong>of</strong> the individual nephrons. To specify the structure <strong>of</strong> each nephronin the model, a stochastic approach similar to the one outlined inSect. 7.4.1 would be sufficient.


Part IVS U M M A RY O F R E S E A R C H


D I S C U S S I O N A N D C O N C LU S I O N8The ultimate intent <strong>of</strong> this research was to provide an approach torenal modelling that is capable <strong>of</strong> predicting whole-kidney functionfrom the dynamics <strong>of</strong> individual nephrons, and can therefore be<strong>of</strong> practical use to clinicians. In this chapter, the contributions <strong>of</strong>this thesis are summarised. The limitations <strong>of</strong> the work are thendiscussed and several areas <strong>of</strong> further research to build on this workare nominated. Finally, the research outcomes are evaluated againstthe research aim.8.1 C O N T R I B U T I O N SThe standard methods and techniques used to model nephron behaviourare not well suited to studying interactions, couplings andemergent dynamics in large multi-nephron systems. This is due to severallimitations, chiefly the complexity <strong>of</strong> solving systems <strong>of</strong> coupledequations and the inability to decouple the structure and behaviour<strong>of</strong> the model. This thesis has proposed a new approach for modellingmulti-nephron systems, which minimises the coupling betweenequations, reduces the computational cost <strong>of</strong> solving the model equations,and decouples the structure and behaviour <strong>of</strong> the model. Thecontributions <strong>of</strong> this thesis are:• A modelling approach—hierarchical dynamical networks—whichcombines complex networks and graph automata into a singlemodelling framework. This approach explicitly captures thestructure and interactions in multi-nephron systems, and decouplesthe structure and behaviour <strong>of</strong> the model. This approachallows emergent dynamics to be easily explored and analysed,as discussed in Chapter 4.• The development <strong>of</strong> a multi-nephron model that produces validbehaviour in single-nephron and multi-nephron systems—asdemonstrated in Chapter 5—and renders the simulation <strong>of</strong>whole-kidney function from the dynamics <strong>of</strong> individual nephronscomputationally tractable. Using this model, the emergent effects<strong>of</strong> the couplings and interactions between nephrons canbe investigated.• An investigation into the dynamics <strong>of</strong> multi-nephron systemsthat focused on whole-system and hierarchical properties ratherthan the dynamics <strong>of</strong> individual nephrons, conducted in Chapter5. As part <strong>of</strong> this investigation, the dynamics <strong>of</strong> a 72-nephronsystem were analysed—a system significantly larger than existingmulti-nephron models.• A study <strong>of</strong> whole-system stability in response to localised impairmentsin nephron function, conducted in Chapter 6. This wasthe first study <strong>of</strong> the emergent dynamics <strong>of</strong> impaired nephron129


130 D I S C U S S I O N A N D C O N C L U S I O Nfunction, and served as an illustration <strong>of</strong> how the emergentdynamics produced by renal diseases may be predicted andanalysed. The impaired multi-nephron systems exhibited verystable behaviour, which we contend is a feature <strong>of</strong> both themodel and the kidney proper.• In Chapter 7, the computational cost <strong>of</strong> the model was shown tobe low enough that the simulation <strong>of</strong> whole-kidney function isfeasible for the first time. It was also demonstrated that simulationscould be easily distributed across multiple computers, resultingin a significant gain in performance. An implementation<strong>of</strong> the model that supports parallel and distributed executionwas presented, based on the Join Calculus [59, 60].• In order to predict whole-kidney function, a whole-kidney modelmust be constructed. Two approaches for automatically generatingsuch models were proposed (but not implemented) inChapter 7.8.2 L I M I TAT I O N S A N D F U RT H E R W O R KThe model presented in this thesis is built on a number <strong>of</strong> simplifyingassumptions, described in Sect. 4.4.2. Most notably, all vessels arenon-compliant and the function <strong>of</strong> each tubule segment was modelledat a much higher level than in existing tubule models, which attemptto capture precise low-level dynamics. However, both <strong>of</strong> these limitationscan be removed by refining the update rules in the model.The update rules for each tubule node can be replaced by discreteversions <strong>of</strong> the continuous equations used in existing tubule modelswithout incurring unexpected increases in computational cost (seeSect. 4.1.1). Similarly, compliance can be modelled by changing howthe flow through a vessel is related to the pressure difference acrossthe vessel, as illustrated in Fig. 80.The other significant limitation on this work is the lack <strong>of</strong> a methodfor automatically generating models <strong>of</strong> large multi-nephron systems—without such a method, the structure <strong>of</strong> such models much be specifiedmanually, which is infeasible for sufficiently large systems. Anapproach to automatically generate multi-nephron models from statisticalconstraints on the model structure was presented in Sect. 7.4.1,but did not provide control over the structure <strong>of</strong> the arterial tree andthe 2D arrangement <strong>of</strong> the interstitial fluid columns. An improvedapproach was presented in Sect. 7.4.2, using a rewriting system togenerate the renal arterial tree and attaching nephrons to this tree.By analysing branching characteristics in medical images <strong>of</strong> renalarterial trees, this approach would generate models with realisticarterial trees and model structure.Several areas <strong>of</strong> further research that build on this thesis are nowproposed.8.2.1 Improved understanding <strong>of</strong> whole-kidney functionThe simulation <strong>of</strong> whole-kidney function combined with the interactivity<strong>of</strong> the model would allow for an improved understanding how the


8.2 L I M I TAT I O N S A N D F U RT H E R W O R K 131CompliantNon-compliantFlowPressureFigure 80: A simple model <strong>of</strong> compliant vessels, using a cubic equationto capture the relationship between pressure and flow, rather thana linear equation. The highlighted region shows the pressure rangeacross which the cubic equation approximates the dynamics <strong>of</strong> acompliant vessel.whole-kidney function adapts to changes in bodily conditions. Oneapplication for this knowledge would be to design dialysis machinesthat can better replicate whole-kidney function.8.2.2 Modelling <strong>of</strong> renal diseaseBased on the simple example presented in Chapter 6, detailed models<strong>of</strong> individual renal diseases can be developed and incorporatedinto whole-kidney models, in order to perform clinically-relevant experimentssuch as investigating how the presence <strong>of</strong> a particularrenal disease affects whole-kidney function. The model also allowsfor impairments to spread across the system, which would allowfor experiments into the progressive loss <strong>of</strong> kidney function to beconducted.8.2.3 Pluggable virtual kidneySince the modelling approach presented in Sect. 4.1 decouples theindividual equations, arbitrary differential equations can be convertedinto difference equations and incorporated into the model. Due to thelow computational cost <strong>of</strong> the model and the distributed simulationframework, a virtual kidney model could be produced that would allowresearchers to experiment with incorporating systems <strong>of</strong> equationsfrom different tubule models and to analyse the resulting dynamics atmultiple levels <strong>of</strong> hierarchy, from the individual tubule to the wholekidneyscale.


132 D I S C U S S I O N A N D C O N C L U S I O N8.2.4 Modelling kidney failure with reliability factorsA simple failure model could also be incorporated into a whole-kidneymodel, whereby each nephron n would be assigned a reliability factorR n = f(t, v) that is a function <strong>of</strong> the age <strong>of</strong> the nephron t and thevolume <strong>of</strong> fluid v that has flown through that nephron. Simulations <strong>of</strong>a particular model would be used to determine the specific R n valuesfor some t ′ (e. g., 60 years). These R n values could then be appliedat the start <strong>of</strong> further simulations, allowing the dynamics <strong>of</strong> an agedkidney to be analysed, since the simulation would effectively havestarted at time t = t ′ . The feasibility <strong>of</strong> this study relies on choosinga reasonable function f(t, v)—given any such f(t, v), analysis <strong>of</strong> themodel dynamics and disease modelling experiments should proveto be useful benchmarks to compare the resulting dynamics againstthose observed in patients <strong>of</strong> the particular age t ′ .8.3 C O N C L U S I O NThe ultimate intent <strong>of</strong> this research was to provide the basis foran approach to renal modelling that is capable <strong>of</strong> predicting wholekidneyfunction from the dynamics <strong>of</strong> individual nephrons, and cantherefore be <strong>of</strong> practical use to clinicians.A unique modelling approach was proposed in Chapter 4, and theresulting models were shown to produce valid behaviour in Chapter 5.Chapter 5 also demonstrated that the models could be used to studyemergent dynamics in multi-nephron systems, and the application<strong>of</strong> this analysis to studying the effects <strong>of</strong> localised impairments onwhole-system dynamics was illustrated in Chapter 6. Finally, the computationalcost <strong>of</strong> the model was investigated in Chapter 7, whichdemonstrated that the simulation <strong>of</strong> whole-kidney function was computationallytractable, and that the performance <strong>of</strong> the simulationscould be improved by utilising multiple computers.In conclusion, the modelling and analysis techniques presentedin this thesis allow for emergent dynamics to be studied in largemulti-nephron systems. This work demonstrates that, for the firsttime, simulation <strong>of</strong> whole-kidney function from the dynamics <strong>of</strong> individualnephrons is tractable. Furthermore, the work provides a basisfor predicting emergent effects <strong>of</strong> localised renal disease. With thecontinued development <strong>of</strong> this model, we hope that significant insightwill be gained into the onset, progression and treatment <strong>of</strong> renaldiseases.


B I B L I O G R A P H Y[1] ADAM Medical Encyclopedia. A.D.A.M. Medical Encyclopedia.A.D.A.M. Inc, Atlanta, 2005. URL http://www.nlm.nih.gov/medlineplus/encyclopedia.html. (Cited on page 56.)[2] A. Adamatzky, A. Wuensche, and B. De Lacy Costello. Gliderbasedcomputing in reaction-diffusion hexagonal cellular automata.Chaos, Solitons & Fractals, 27(2):287–295, 2006.(Cited on pages 20, 21, and 30.)[3] E. Ahmed, A. S. Elgazzar, and M. I. Shehata. On differencesbetween fractional and integer order differential equationsfor dynamical games. Z Naturfors Sect A-J Phys Sci, 63a(3/4):152–155, 2008. (Cited on page 3.)[4] R. Albert and A.-L. Barabási. Statistical mechanics <strong>of</strong>complex networks. Reviews <strong>of</strong> Modern Physics, 74:47,2002. URL http://www.citebase.org/abstract?id=oai:arXiv.org:cond-mat/0106096. (Cited on pages 22, 25, and 122.)[5] M. Alfonseca and A. Ortega. Determination <strong>of</strong> fractal dimensionsfrom equivalent l systems. IBM J Res & Dev, 45(6):797–805, 2001. (Cited on page 124.)[6] G. M. Amdahl. Validity <strong>of</strong> the single processor approach toachieving large scale computing capabilities. In AFIPS ConferenceProceedings, vol. 30 (Apr. 18–20, 1967), pages 483–485.AFIPS Press, 1967. (Cited on page 119.)[7] D. P. Anderson. BOINC: A System for Public-Resource Computingand Storage. In Grid 2004: Proceedings <strong>of</strong> the FifthIEEE/ACM International Workshop on Grid Computing, pages4–10, Washington, DC, 2004. IEEE Computer Society. (Citedon page 114.)[8] D. P. Anderson, J. Cobb, E. Korpela, M. Leb<strong>of</strong>sky, andD. Werthimer. SETI@home: an experiment in public-resourcecomputing. Commun ACM, 45(11):56–61, 2002. (Cited onpage 114.)[9] R. F. S. Andrade, J. G. V. Miranda, S. T. R. Pinho, and T. P.Lobão. Characterization <strong>of</strong> complex networks by higher orderneighborhood properties. Eur Phys J B, 61(2):43–52, 2008.(Cited on page 24.)[10] L. Antiga, M. Piccinelli, G. Fasolini, B. Ene-lordache, P. Ondei,S. Bruno, G. Remuzzi, and A. Remuzzi. Computed tomographyevaluation <strong>of</strong> autosomal dominant polycystic kidney diseaseprogression: A progress report. Clin J Am Soc Nephrol, 1(4):754–760, 2006. (Cited on page 16.)[11] A. Arenas, A. Díaz-Guilera, and CJ C. J. Pérez-Vicente. Synchronizationreveals topological scales in complex networks. PhysRev Lett, 96(11):114102, 2006. (Cited on page 30.)133


134 B I B L I O G R A P H Y[12] J. Ash and D. Newth. Optimizing complex networks for resilienceagainst cascading failure. Physica A, 380(1):673–683,2007. (Cited on page 25.)[13] Australian Institute <strong>of</strong> Health and Welfare. Chronic kidneydisease in Australia. AIHW, 2005. URL http://www.aihw.gov.au/. Cat. No. PHE68. (Cited on page 16.)[14] K. F. Badr. Filtration function in glomerulonephritis. <strong>Kidney</strong>Int, 68(4):1905–1919, 2005. (Cited on page 16.)[15] N. T. J. Bailey. The Mathematical Theory <strong>of</strong> Infectious Diseases.Griffin, London, second edition, 1975. (Cited on page 25.)[16] G. Baker. Shape Modelling and Analysis <strong>of</strong> Anatomical Objects.PhD thesis, 2008. (Cited on page 126.)[17] Y. Bar-Yam. Dynamics <strong>of</strong> Complex Systems. Westview Press,2003. ISBN 0-81-334121-3. (Cited on pages 3 and 4.)[18] A.-L. Barabási and R. Albert. Emergence <strong>of</strong> scaling in randomnetworks. Science, 286(5439):509–512, 1999. (Cited onpage 22.)[19] M. Barfred, E. Mosekilde, and N. H. Holstein-Rathlou. Bifurcationanalysis <strong>of</strong> nephron pressure and flow regulation. Chaos,6(3):280–287, 1996. (Cited on pages 34, 41, 56, 62, and 63.)[20] L. P. Bartosik, G. Lajoie, L. Sugar, and D. C. Cattran. Predictingprogression in IgA nephropathy. Am J <strong>Kidney</strong> Dis, 38(4):881–883, 2001. (Cited on page 15.)[21] H. Birn, J. C. Fyfe, C. Jacobsen, F. Mounier, P. J. Verroust,H. Ørskov, T. E. Willnow, S. K. Moestrup, and E. I. Christensen.Cubilin is an albumin binding protein important for renal tubularalbumin reabsorption. J Clin Invest, 105(10):1353–1361,2000. (Cited on pages 15 and 16.)[22] M. Blank and W. A. Kalender. Construction and use <strong>of</strong> a virtualphantom <strong>of</strong> the human bronchial tree. Radiologe, 40(3):240–245, 2000. (Cited on page 124.)[23] R. C. Blantz and V. Vallon. Tubuloglomerular feedback responses<strong>of</strong> the downstream efferent resistance: Unmasking arole for adenosine? <strong>Kidney</strong> Int, 71(9):837–839, 2007. (Cited onpage 11.)[24] R. C. Blantz, F. C. Rector, Jr, and D. W. Seldin. Effect <strong>of</strong> hyperoncoticalbumin expansion upon glomerular ultrafiltration inthe rat. <strong>Kidney</strong> Int, 6(4):209–221, 1974. (Cited on page 16.)[25] K. Blouch, W. M. Deen, J. P. Fauvel, J. Bialek, G. Derby, and B. D.Myers. Molecular configuration and glomerular size selectivityin healthy and nephrotic humans. AJP Renal Physiol, 273(3):430–437, 1997. (Cited on page 15.)[26] S. Boccaletti, D.-U. Hwang, and V. Latora. Growing hierarchicalscale-free networks by means <strong>of</strong> nonhierarchical processes. IntJ Bif Chaos, 17(7):2447–2452, 2007. (Cited on page 23.)


B I B L I O G R A P H Y 135[27] B. Bollobás and O. Riordan. The diameter <strong>of</strong> a scale-freerandom graph. Combinatorica, 24(1):5–34, 2004. (Cited onpage 23.)[28] B. M. Brenner, J. L. Troy, T. M. Daugharty, W. M. Deen, andC. R. Robertson. Dynamics <strong>of</strong> glomerular ultrafiltration in therat ii. plasma-flow dependence <strong>of</strong> GFR. Am J Physiol, 223(5):1184–1190, 1972. (Cited on page 11.)[29] M. Bro-Nielsen, C. Gramkow, and S. Kreiborg. Non-rigid imageregistration using bone growth model. In CVRMed-MRCAS,volume 1205 <strong>of</strong> Lecture Notes in Computer Science, pages3–12. Springer, Mar 1997. ISBN 3-540-62734-0. URL http://citeseer.ist.psu.edu/107563.html. (Cited on page 122.)[30] F. P. Brunner, F. C. Rector, Jr, and D. W. Seldin. Mechanism <strong>of</strong>glomerulotubular balance: I. effect <strong>of</strong> aortic constriction andelevated ureteropelvic pressure on glomerular filtration rate,fractional reabsorption, transit time, and tubular size in theproximal tubule <strong>of</strong> the rat. J Clin Invest, 45(4):590–602, 1966.(Cited on page 48.)[31] F. P. Brunner, F. C. Rector, Jr, and D. W. Seldin. Mechanism <strong>of</strong>glomerulotubular balance. ii. regulation <strong>of</strong> proximal tubularreabsorption by tubular volume, as studied by stopped-flowmicroperfusion. J Clin Invest, 45(4):603–611, 1966. (Cited onpage 48.)[32] N. J. Brunskill, J. Stuart, A. B. Tobin, J. Walls, and S. Nahorski.Receptor-mediated endocytosis <strong>of</strong> albumin by kidney proximaltubule cells is regulated by phosphatidylinositide 3-kinase. JClin Invest, 101(10):2140–2150, 1998. (Cited on pages 15and 16.)[33] J. S. Cameron. Milk or albumin? The history <strong>of</strong> proteinuriabefore Richard Bright. Nephrol Dial Transplant, 18(7):1281–1285, 2003. (Cited on page 15.)[34] L. Cardelli and A. D. Gordon. Mobile ambients. In Foundations<strong>of</strong> S<strong>of</strong>tware Science and Computation Structures: First InternationalConference, FOSSACS ’98. Springer-Verlag, 1998. URLhttp://citeseer.ist.psu.edu/cardelli98mobile.html. (Cited onpage 119.)[35] D. Carroll, C. Rahmlow, T. Psiaki, and G. Wojtaszczyk. Distributingscience, July 2005. URL http://boinc.berkeley.edu/distributing_science_final.pdf. (Cited on page 114.)[36] A. Cass, S. Chadban, J. Craig, K. Howards, S. McDonald,G. Salkeld, and S. White. The Economic Impact <strong>of</strong> End-Stage<strong>Kidney</strong> Disease in Australia. <strong>Kidney</strong> Health Australia, 2006.URL http://www.kidney.org.au/. (Cited on page 16.)[37] H. Chang and T. Fujita. A numerical model <strong>of</strong> acid-base transportin rat distal tubule. AJP Renal Physiol, 281(2):222–243,2001. (Cited on page 13.)


136 B I B L I O G R A P H Y[38] H. Chang and T. Fujita. A numerical model <strong>of</strong> the renal distaltubule. AJP Renal Physiol, 276(6):931–951, 1999. (Cited onpage 13.)[39] R. L. S. Chang, W. M. Deen, C. R. Robertson, and B. M. Brenner.Permselectivity <strong>of</strong> the glomerular capillary wall: III. Restrictedtransport <strong>of</strong> polyanions. <strong>Kidney</strong> Int, 8(4):212–218, 1975. (Citedon page 14.)[40] M. Christ, K. Douwes, C. Eisen, G. Bechtner, K. Theisen, andM. Wehling. Rapid effects <strong>of</strong> aldosterone on sodium transportin vascular smooth muscle cells. Hypertension, 25(1):117–123,1995. (Cited on page 52.)[41] Chronic kidney disease. Wikipedia, 15/11/2007. URL http://en.wikipedia.org/wiki/Chronic_kidney_disease. (Cited onpages 14 and 81.)[42] X. Chu, A. Lonie, P. Harris, S. R. Thomas, and R. Buyya. <strong>Kidney</strong>grid:a grid platform for integration <strong>of</strong> distributed kidneymodels and resources. In MCG ’06: Proceedings <strong>of</strong> the 4th internationalworkshop on Middleware for grid computing, pages5–11. ACM, 2006. ISBN 1-59593-581-9. (Cited on pages 56and 114.)[43] X. Chu, A. Lonie, P. Harris, S. R. Thomas, and R. Buyya. Aservice-oriented grid environment for integration <strong>of</strong> distributedkidney models and resources. Concurr Comput-Pract Exp, 20(9):1095–1111, 2008. (Cited on pages 56 and 114.)[44] W. L. Clapp, C. H. Park, K. M. Madsen, and C. C. Tisher. Axialheterogeneity in the handling <strong>of</strong> albumin by the rabbit proximaltubule. Lab Invest, 58(5):549–558, 1988. (Cited on page 15.)[45] W. D. Comper and E. F. Glasgow. Charge selectivity in kidneyultrafiltration. <strong>Kidney</strong> Int, 47(5):1242–1251, 1995. (Cited onpage 15.)[46] D. Coppersmith and S. Winograd. Matrix multiplication viaarithmetic progressions. J Symbolic Comput, 9(3):251–280,1990. (Cited on page 32.)[47] L. da F. Costa. The hierarchical backbone <strong>of</strong> complex networks.Phys Rev Lett, 93(9):098702, 2004. (Cited on pages 12 and 30.)[48] F. Costantini. Renal branching morphogenesis: concepts, questions,and recent advances. Differentiation, 74(7):402–421,2006. (Cited on page 124.)[49] S. S. Cross, R. D. Start, P. B. Silcocks, A. D. Bull, D. W. K. Cotton,and J. C. E. Underwood. Quantitation <strong>of</strong> the renal arterial treeby fractal analysis. J Pathol, 170(4):479–484, 2005. (Cited onpage 124.)[50] P. Crucitti, V. Latora, and M. Marchiori. Model for cascadingfailures in complex networks. Phys Rev E, 69(4):045104, 2004.(Cited on page 25.)


B I B L I O G R A P H Y 137[51] J. Dean and S. Ghemawat. MapReduce: Simplified Data Processingon Large Clusters. In OSDI ’04, pages 137–150, Berkeley,CA, 2004. USENIX Association. (Cited on page 114.)[52] W. M. Deen, C. R. Robertson, and B. M. Brenner. A model <strong>of</strong>glomerular ultrafiltration in the rat. AJP Renal Physiol, 223(5):1178–83, 1972. (Cited on page 44.)[53] Diabetic nephropathy. Wikipedia, 15/11/2007. URL http://en.wikipedia.org/wiki/Diabetic_nephropathy. (Cited on pages 14and 81.)[54] S. N. Dorogovtsev and J. F. F. Mendes. Evolution <strong>of</strong> Networks:From Biological Nets to the Internet and WWW. Oxford <strong>University</strong>Press, 2003. ISBN 0-198-51590-1. (Cited on page 22.)[55] R. El-Galley. Surgical management <strong>of</strong> renal tumors. Radiol ClinN Am, 41(5):1053–1065, 2003. (Cited on page 12.)[56] H. Enderling, M. A. J. Chaplain, A. R. A. Anderson, and J. S.Vaidya. A mathematical model <strong>of</strong> breast cancer development,local treatment and recurrence. J Theor Biol, 246(2):245–259,2007. URL http://www.sciencedirect.com/science/article/B6WMD-4MJJH3R-7/2/bf50edb2044aa6b7a7bc33b4b113696a. (Citedon page 25.)[57] P. Erdös and A. Rényi. On random graphs. Publ Math, 6:290–297, 1959. (Cited on page 26.)[58] A. K. Exadaktylos, S. Hauser, J. Luterbacher, U. Marti, H. Zimmermann,and U. E. Studer. The Moon and the Stones. Can theMoon’s Attractive Forces Cause Renal Colic? J Emerg Med, 22(3):303–305, 2002. (Cited on page 16.)[59] C. Fournet and G. Gonthier. The reflexive CHAM and the joincalculus.In POPL’96, pages 372–385. ACM, Jan 1996. (Citedon pages 5, 6, 109, 119, and 130.)[60] C. Fournet and C. Laneve. Bisimulations in the join-calculus.TCS, 266(1–2):569–603, 2001. (Cited on pages 5, 6, 109, 119,and 130.)[61] L. S. Furtado and M. Copelli. Response <strong>of</strong> electrically coupledspiking neurons: A cellular automaton approach. Phys Rev E,73(1):011907, 2006. (Cited on page 30.)[62] T. I. Garner and R. Dardis. Cost-effectiveness analysis <strong>of</strong> endstagerenal disease treatments. Medical Care, 25(1):25–34,1987. (Cited on page 16.)[63] C. R. P. George. William Charles Wells (1757–1815)—a nephrologist<strong>of</strong> the Scottish enlightenment. Nephrol Dial Transplant,11(12):2513–2517, 1996. (Cited on page 15.)[64] J. L. Giavitto and O. Michel. Modeling the topological organization<strong>of</strong> cellular processes. Biosystems, 70(2):149–163, 2003.(Cited on page 124.)


138 B I B L I O G R A P H Y[65] Glomerulonephritis. Wikipedia, 15/11/2007. URL http://en.wikipedia.org/wiki/Glomerulonephritis. (Cited on page 15.)[66] N. Goldenfield and L. P. Kadan<strong>of</strong>f. Simple lessons from complexity.Science, 284(5411):87–89, 1999. (Cited on page 23.)[67] A. Grabowski and R. A. Kosiński. Properties <strong>of</strong> an evolvingdirected network with local rules and intrinsic variables. Int JMod Phys C, 18(1):43–52, 2007. (Cited on page 23.)[68] Z. Guan, D. A. Osmond, and E. W. Inscho. Purinoceptors inthe kidney. Exp Biol Med, 232(6):715–726, 2007. (Cited onpage 11.)[69] A. Guasch, W. M. Deen, and B. D. Myers. Charge selectivity <strong>of</strong>the glomerular filtration barrier in healthy and nephrotic humans.J Clin Invest, 92(5):2274–2282, 1993. (Cited on page 15.)[70] J. L. Gustafson. Reevaluating Amdahl’s law. Commun ACM, 31(5):532–533, 1988. ISSN 0001-0782. doi: http://doi.acm.org.ezproxy.lib.unimelb.edu.au/10.1145/42411.42415. (Cited onpage 120.)[71] A. C. Guyton and J. E. Hall. Textbook <strong>of</strong> Medical Physiology.Elsevier Saunders, eleventh edition, 2006. ISBN 0-7216-0240-1.(Cited on pages 29, 46, 48, 51, 52, 53, and 57.)[72] T. P. Hannedouche, A. G. Delgado, D. A. Gnionsahe, C. Boitard,B. Lacour, and J.-P. Grünfeld. Renal hemodynamics and segmentaltubular reabsorption in early type 1 diabetes. <strong>Kidney</strong>Int, 37(4):1126–1133, 1990. (Cited on page 17.)[73] B. Haraldsson and J. Sörensson. Why do we not all have proteinuria?An update <strong>of</strong> our current understanding <strong>of</strong> the glomerularbarrier. News Physiol Sci, 19(1):7–10, 2004. (Cited on page 15.)[74] D. Harel. Statecharts: A visual formalism for complex systems.Science <strong>of</strong> Computer Programming, 8(3):231–274, June 1987.URL citeseer.csail.mit.edu/harel87statecharts.html. (Citedon pages 23 and 24.)[75] D. Harel and A. Naamad. The STATEMATE semantics <strong>of</strong> statecharts.ACM Transactions on S<strong>of</strong>tware Engineering andMethodology, 5(4):293–333, 1996. URL citeseer.csail.mit.edu/harel96statemate.html. (Cited on page 24.)[76] S. Hervy and S. R. Thomas. Inner medullary lactate productionand urine-concentrating mechanism: a flat medullary model.AJP Renal Physiol, 284(1):65–81, 2003. (Cited on pages 11, 13,48, 57, 58, and 67.)[77] W. D. Hillis. Richard Feynman and the Connection Machine.Physics Today, 42(2):78–83, 1989. (Cited on page 29.)[78] N. H. Holstein-Rathlou and D. J. Marsh. A dynamic model <strong>of</strong>the tubuloglomerular feedback mechanism. AJP Renal Physiol,258(5 Pt 2):1448–1459, 1990. (Cited on pages 3, 11, 12, 13,30, 35, 40, 44, 51, 56, 57, and 62.)


B I B L I O G R A P H Y 139[79] N. H. Holstein-Rathlou and D. J. Marsh. Oscillations <strong>of</strong> tubularpressure, flow, and distal chloride concentration in rats. AJPRenal Physiol, 256(6):1007–1114, 1989. (Cited on page 12.)[80] N. H. Holstein-Rathlou, A. J. Wagner, and D. J. Marsh. Tubuloglomerularfeedback dynamics and renal blood flow autoregulationin rats. AJP Renal Physiol, 260(1 Pt 2):53–68, 1991.(Cited on page 13.)[81] N. H. Holstein-Rathlou, K. P. Yip, O. V. Sosnovtseva, andE. Mosekilde. Synchronization phenomena in nephron-nephroninteraction. Chaos, 11(2):417–426, 2001. (Cited on pages 3,12, 13, 34, and 82.)[82] G. S. Hornby and J. B. Pollack. Evolving L-systems to generatevirtual creatures. Comp & Graph, 25(6):1041–1048, 2001.(Cited on page 124.)[83] G. S. Hornby and J. B. Pollack. Creating high-level componentswith a generative representation for body-brain evolution. ArtificialLife, 8(3):223–246, 2002. (Cited on page 124.)[84] T. H. Hostetter, J. L. Olson, H. G. Rennke, M. A. Venkatachalam,and B. M. Brenner. Hyperfiltration in remnant nephrons: a potentiallyadverse response to renal ablation. AJP Renal Physiol,241(1):85–93, 1981. (Cited on page 16.)[85] K. Culik II and S. Dube. L-systems and mutually recursivefunction systems. Acta Informat, 30(3):279–302, 1993. (Citedon page 124.)[86] C. Iodice, M. M. Balletta, R. Minutolo, P. Giannattasio, S. Tuccillo,V. Bellizzi, M. D’Amora, G. Rinaldi, G. Signoriello, G. Conte,and L. De Nicola. Maximal suppression <strong>of</strong> renin-angiotensinsystem in nonproliferative glomerulonephritis. <strong>Kidney</strong> Int, 63(6):2214–2221, 2003. (Cited on page 16.)[87] Y. Maeda J. S. Han and M. A. Knepper. Dual actions <strong>of</strong> vasopressinand oxytocin in regulation <strong>of</strong> water permeability interminal collecting duct. AJP Renal PHysiol, 265(1):26–34,1993. (Cited on page 52.)[88] C. Jacob and S. von Mammem. Swarm grammars: growingdynamic structures in 3D agent spaces. Digital Creativity, 18(1):54–64, 2007. (Cited on page 124.)[89] D. N. Jansen, H. Hermanns, and J. Katoen. A probabilisticextension <strong>of</strong> UML statecharts. In Proceedings <strong>of</strong> FTRTFT 2002,volume 2469 <strong>of</strong> Lecture Notes in Computer Science, pages355–374. Springer-Verlag, 2002. (Cited on pages 23 and 24.)[90] H. Jeong, S. P. Mason, A.-L. Barabási, and Z. N. Oltvai. Lethalityand centrality in protein networks. Nature, 411:41–42,2001. URL http://www.citebase.org/abstract?id=oai:arXiv.org:cond-mat/0105306. (Cited on page 22.)


140 B I B L I O G R A P H Y[91] L. I. Juncos. Intrarenal mechanisms <strong>of</strong> salt and water retentionin the nephritic syndrome. <strong>Kidney</strong> Int, 61(3):1182–1195, 2002.(Cited on page 16.)[92] F. M. Karlsen, P. P. Leyssac, and N.-H. Holstein-Rathlou. Tubuloglomerularfeedback in Dahl rats. AJP Regul Integr CompPhysiol, 274(6):1561–1569, 1998. (Cited on page 15.)[93] G. Kokai and Z. Toth. Modelling blood vessels <strong>of</strong> the eye withparametric L-systems using evolutionary algorithms. Artif IntellMed, 1620:433–442, 1999. (Cited on page 124.)[94] E. Korpela, D. Werthimer, D. P. Anderson, J.Cobb, and M. Leb<strong>of</strong>sky.SETI@home—massively distributed computing for SETI.Comput Sci Eng, 3(1):78–83, 2001. (Cited on page 114.)[95] S. Kuroda, H. S. Aynedjian, and N. Bank. A micropuncturestudy <strong>of</strong> renal sodium retention in nephrotic syndrome in rats:Evidence for increased resistance to tubular fluid flow. <strong>Kidney</strong>Int, 16(5):561–571, 1979. (Cited on page 16.)[96] Lab Tests Online. Lab tests online, 2007. URL http://labtestsonline.org/map/aindex.html. (Cited on page 56.)[97] A. T. Layton and H. E. Layton. A region-based mathematicalmodel <strong>of</strong> the urine concentrating mechanism in the rat outermedulla: I. Formulation and base-case results. AJP Renal Physiol,58(6):1346–1366, 2005. (Cited on pages 11, 58, and 67.)[98] A. T. Layton and H. E. Layton. A region-based mathematicalmodel <strong>of</strong> the urine concentrating mechanism in the rat outermedulla: II. Parameter sensitivity and tubular inhomogeneity.AJP Renal Physiol, 58(6):1367–1381, 2005. (Cited on pages 11,58, and 67.)[99] H. E. Layton, E. B. Pitman, and L. C. Moore. Bifurcation analysis<strong>of</strong> TGF-mediated oscillations in SNGFR. AJP Renal Physiol, 261(5 Pt 2):904–919, 1991. (Cited on pages 34, 41, 51, and 82.)[100] F. Le Fessant and L. Maranget. Compiling join-patterns. ElectronicNotes in Computer Science, 16(2), 1998. (Cited onpage 119.)[101] X. Leroy, D. Doligez, J. Garrigue, D. Rémy, and J. Vouillon. TheObject Caml system 3.10. INRIA, May 2007. URL http://caml.inria.fr/pub/docs/manual-ocaml/. (Cited on page 111.)[102] P. P. Leyssac and L Baumbach. An oscillating intratubularpressure response to alterations in henle loop flow in the ratkidney. Acta Physiol Scand, 117(3):415–419, 1983. (Cited onpage 56.)[103] P. P. Leyssac and N. H. Holstein-Rathlou. Effects <strong>of</strong> varioustransport inhibitors on oscillating tgf pressure responses in therat. Pflugers Arch, 407(3):285–291, 1986. (Cited on page 56.)[104] F. Liljeros, C. R. Edling, L. A. N. Amaral, H. E. Stanley, andY. Åberg. The web <strong>of</strong> human sexual contacts. Nature, 411(6840):907–908, 2001. (Cited on page 25.)


B I B L I O G R A P H Y 141[105] A. Lindenmayer. Developmental systems without cellular interactions,their languages and grammars. J Theor Biol, 30(3):455–484, 1971. (Cited on page 124.)[106] A. Lindenmayer. Mathematical models for cellular interactionsin development, part 1. J Theor Biol, 18(3):280–299, 1968.(Cited on page 124.)[107] A. Lindenmayer. Mathematical models for cellular interactionsin development, part 2. J Theor Biol, 18(3):300–315, 1968.(Cited on page 124.)[108] P. Lory. Effectiveness <strong>of</strong> a salt transport cascade in the renalmedulla: computer simulations. AJP Renal Physiol, 252(6):1095–1102, 1987. (Cited on pages 13, 48, 58, and 67.)[109] M. A. Mainster. The fractal properties <strong>of</strong> retinal vessels : embryologicaland clinical implications. Eye, 4(1):235–241, 1990.(Cited on page 124.)[110] L. Mandel and L. Maranget. The JoCaml language 3.10. IN-RIA, July 2007. URL http://jocaml.inria.fr/manual/. (Citedon page 119.)[111] D. J. Marsh, O. V. Sosnovtseva, K. H. Chon, and N. H. Holstein-Rathlou. Nonlinear interactions in renal blood flow regulation.AJP Regul Integr Comp Physiol, 288(5):1143–1159, 2005.(Cited on pages 40 and 41.)[112] D. J. Marsh, O. V. Sosnovtseva, E. Mosekilde, and N. H. Holstein-Rathlou. Vascular coupling induces synchronization, quasiperiodicity,and chaos in a nephron tree. Chaos, 17(1):015114,2007. (Cited on pages 3, 12, 13, 34, 36, 44, 64, and 82.)[113] B. Martin. A geometrical hierarchy on graphs via cellularautomata. Fund Inform, 52(1–3):157–181, 2002. (Cited onpage 24.)[114] M. May, K.-P. Braun, C. Helke, W. Richter, H. Vogler, B. Hoschke,and M. Siegsmund. Lunar phases and zodiac signs do notinfluence quality <strong>of</strong> radical cystectomy–a statistical analysis <strong>of</strong>452 patients with invasive bladder cancer. Int Urol Nephrol,39(4):1023–1030, December 2007. (Cited on page 16.)[115] S. McGregor and C. Fernando. Levels <strong>of</strong> description: a novelapproach to dynamical hierarchies. Artificial Life, 11(4):459–472, 2005. (Cited on pages 23 and 124.)[116] S. J. McTaggart. Albumin Endocytosis in Renal Proximal TubuleCells. PhD thesis, <strong>University</strong> <strong>of</strong> <strong>Melbourne</strong>, 2003. (Cited onpage 15.)[117] P. Meakin. A new model for biological pattern formation. JTheor Biol, 118(1):101–113, 1986. (Cited on page 124.)[118] I. Mezić. Fkg inequalities in cellular automata and coupledmap lattices. Physica D, 103(1–4):491–504, 1997. (Cited onpage 30.)


142 B I B L I O G R A P H Y[119] E. Mikk, Y. Lakhnech, and M. Siegel. Hierarchical automata asmodel for statecharts. In ASIAN ’97: Proceedings <strong>of</strong> the ThirdAsian Computing Science Conference on Advances in ComputingScience, pages 181–196, London, UK, 1997. Springer-Verlag. ISBN 3-540-63875-X. (Cited on page 24.)[120] R. Milner. A calculus <strong>of</strong> communicating systems. Lecture Notesin Computer Science, 92, 1980. (Cited on page 119.)[121] R. Milner, J. Parrow, and D. Walker. A calculus <strong>of</strong> mobile processes,part 1. Information and Computation, 100(1):1–40, Sep1992. (Cited on page 119.)[122] R. Milner, J. Parrow, and D. Walker. A calculus <strong>of</strong> mobile processes,part 2. Information and Computation, 100(1):41–77,Sep 1992. (Cited on page 119.)[123] G. P. Mission, G. Landini, and P. I. Murray. Fractals and ophthalmology.Lancet, 339(8797):872, 1992. (Cited on page 124.)[124] M. Mitchell. Complex systems: Network thinking. Artif Intell,170(18):1194–1212, 2006. (Cited on page 22.)[125] S. Moeller, S. Gioberge, and G. Brown. Esrd patients in 2001:global overview <strong>of</strong> patients, treatment modalities and developmenttrends. Nephrol Dial Transplant, 17(5):2071–2076, 2002.(Cited on page 16.)[126] A. E. Motter and Y.-C. Lai. Cascade-based attacks on complexnetworks. Phys Rev E, 66(6):065102, 2002. (Cited on page 25.)[127] J. D. Murray. Mathematical Biology. Springer-Verlag, Berlin,1993. (Cited on page 25.)[128] R. Nair and P. D. Walker. Is IgA nephropathy the commonest primaryglomerulopathy among young adults in the USA? <strong>Kidney</strong>Int, 69(8):1455–1458, 2006. (Cited on page 15.)[129] K. Nakamura. Asynchronous cellular automata and their computationalability. Systems, Computers, Controls, 5(5):58–66,1974. (Cited on page 30.)[130] L. G. Navar, G. Saccomani, and K. D. Mitchell. Synergisticintrarenal actions <strong>of</strong> angiotensin on tubular reabsorption andrenal hemodynamics. Am J Hypertens, 4(1 Pt 1):90–96, 1991.(Cited on page 11.)[131] M. E. J. Newman. Assortative mixing in networks. Phys RevLett, 89(20):208701, 2002. (Cited on page 23.)[132] M. E. J. Newman. The structure and function <strong>of</strong> complex networks.SIAM Review, 45:167, 2003. URL http://www.citebase.org/abstract?id=oai:arXiv.org:cond-mat/0303516. (Cited onpages 22 and 23.)[133] D. Noble. The rise <strong>of</strong> computational biology. Nat Rev Mol CellBiol, 3(6):459–463, 2002. (Cited on page 3.)


B I B L I O G R A P H Y 143[134] D. A. Nordsletten, S. Blackett, M. D. Bentley, E. L. Ritman,and N. P. Smith. Structural morphology <strong>of</strong> renal vasculature.AJP Heart Circ, 291(1):291–309, 2006. (Cited on pages 124and 126.)[135] M. Ohlson, J. Sörensson, and B. Haraldsson. Glomerular sizeand charge selectivity in the rat as revealed by FITC-Ficoll andalbumin. AJP Renal Physiol, 279(1):84–91, 2000. (Cited onpage 14.)[136] World Health Organization. Diabetes Action Now. WHO Press,2004. (Cited on page 16.)[137] A. Ortega, A. A. Dalhoum, and M. Alfonseca. Grammaticalevolution to design fractal curves with a given dimension. IBMJ Res & Dev, 47(4):483–493, 2003. (Cited on page 124.)[138] T. M. Osicka and W. D. Comper. Glomerular charge selectivityfor anionic and neutral horseradish peroxidase. <strong>Kidney</strong> Int, 47(6):1630–1637, 1995. (Cited on page 15.)[139] J. Park and M. E. J. Newman. The statistical mechanics<strong>of</strong> networks. Physical Review E, 70:066117,2004. URL http://www.citebase.org/abstract?id=oai:arXiv.org:cond-mat/0405566. (Cited on page 19.)[140] J. Parrow and B. Victor. The fusion calculus: Expressiveness andsymmetry in mobile processes. In V. Pratt, editor, Proceedings<strong>of</strong> LICS ’98, pages 176–185. IEEE Computer Society Press,1998. (Cited on page 119.)[141] J. Parrow and B. Victor. The tau-laws <strong>of</strong> fusion. In D. Sangiorgiand R. de Simone, editors, International Conference onConcurrency Theory, volume 1466 <strong>of</strong> Lecture Notes in ComputerScience, pages 99–114. Springer-Verlag, 1998. (Cited onpage 119.)[142] R. Pastor-Satorras and A. Vespignani. Epidemic dynamics andendemic states in complex networks. Phys Rev E, 63(6):066117,2001. (Cited on page 25.)[143] R. Pastor-Satorras and A. Vespignani. Epidemic spreading inscale-free networks. Phys Rev Lett, 86(14):3200–3203, 2001.(Cited on page 25.)[144] J. P. Pennell and J. J. Bourgoignie. Adaptive changes <strong>of</strong> juxtamedullaryglomerular filtration in the remnant kidney. PflügersArchiv, 389(2):131–135, 1981. (Cited on page 16.)[145] E. B. Pitman, R. M. Zaritski, K. J. Kesseler, L. C. Moore, andH. E. Layton. Feedback-mediated dynamics in two couplednephrons. Bull Math Biol, 66(6):1463–1492, 2004. (Cited onpages 3, 12, and 40.)[146] P. E. Plsek and T. Greenhalgh. The challenge <strong>of</strong> complexity inhealth care. Brit Med J, 323(7313):625–628, 2001. (Cited onpage 3.)


144 B I B L I O G R A P H Y[147] M. Qin and L. Maranget. Compiling pattern-matching in joinpatterns.In CONCUR 2004, volume 3170 <strong>of</strong> LNCS, pages417–431. Springer-Verlag, 2004. (Cited on page 119.)[148] A. Raes, R. Donckerwolcke, M. Craen, M. C. Hussein, andJ. Vande Walle. Renal hemodynamic changes and renal functionalreserve in children with type I diabetes mellitus. PediatrNephrol, 22(11):1903–1909, 2007. (Cited on page 16.)[149] A. Rashidi and I. Khodarahmi. Mathematical modelling <strong>of</strong> thecourse <strong>of</strong> chronic renal failure. Nephrol Dial Transplant, 21(11):3344–3345, November 2006. (Cited on pages 3, 9, and 16.)[150] E. Ravasz and A.-L. Barabási. Hierarchical organization incomplex networks. Phys Rev E, 67(2):026112, 2003. (Cited onpages 12, 23, 24, and 30.)[151] Y. Ren, J. L. Garvin, and O. A. Carretero. Efferent arterioletubuloglomerular feedback in the renal nephron. <strong>Kidney</strong> Int,59(1):222–229, 2001. (Cited on page 11.)[152] C. W. Reynolds. Flocks, herds, and schools: A distributed behavioralmodel. Computer Graphics, 21(4):25–34, 1987. (Citedon page 83.)[153] C. W. Reynolds. Interaction with groups <strong>of</strong> autonomous characters.In Game Developers Conference 2000, 2000. URL http://www.red3d.com/cwr/papers/2000/pip.html. (Cited on page 83.)[154] S. A. H. Rizvi, S. A. A. Naqvi, Z. Hussain, A. Hashmi, F. Akhtar,M. Hussain, E. Ahmed, M. N. Zafar, S. Hafiz, R. Muzaffar, andF. Jawad. Renal transplantation in developing countries. <strong>Kidney</strong>Int Suppl, 63(S83):96–100, 2003. (Cited on page 16.)[155] L. M. Russo, G. L. Bakris, and W. D. Comper. Renal handling<strong>of</strong> albumin: a critical review <strong>of</strong> basic concepts and perspective.Am J <strong>Kidney</strong> Dis, 39(5):899–919, 2002. (Cited on page 15.)[156] F. M. Sacks, L. P. Svetkey, W. M. Vollmer, L. J. Appel, G. A. Bray,D. Harsha, E. Obarzanek, P. R. Conlin, E. R. Miller, 3rd, D. G.Simons-Morton, N. Karanja, and P. H. Lin. Effects on bloodpressure <strong>of</strong> reduced dietary sodium and the Dietary Approachesto Stop Hypertension (DASH) diet. N Engl J Med, 344(1):3–10,2001. (Cited on page 15.)[157] M. Sales-Pardo, R. Guimerà, A. A. Moreira, and L. A. N. Amaral.Extracting the hierarchical organization <strong>of</strong> complex systems.P Natl Acad Sci USA, 104(39):15224–15229, 2007. (Cited onpage 23.)[158] B. Schmidt-Nielsen and R. O’Dell. Structure and concentratingmechanism in the mammalian kidney. Am J Physiol, 200(6):1119–1124, 1961. (Cited on pages 11 and 57.)[159] J. Schnermann, F. S. Wright, J. M. Davis, W. v. Stackelberg, andG. Grill. Regulation <strong>of</strong> superficial nephron filtration rate bytubulo-glomerular feedback. Pflügers Archiv, 318(2):147–175,June 1970. (Cited on page 12.)


B I B L I O G R A P H Y 145[160] J. Schnermann, A. E. G. Persson, and B. Ågerup. Tubuloglomerularfeedback: Nonlinear relation between glomerular hydrostaticpressure and loop <strong>of</strong> henle perfusion rate. J Clin Invest,52(4):862–869, April 1973. (Cited on page 12.)[161] C. R. Shalizi. Methods and techniques <strong>of</strong> complex systemsscience: An overview. Complex Systems Science in Biomedicine,pages 33–114, 2006. URL http://www.citebase.org/abstract?id=oai:arXiv.org:nlin/0307015. (Cited on page 3.)[162] C. E. Shannon. Communication in the presence <strong>of</strong> noise. Proc.IEEE, 86(2):447–457, 1998. (Cited on pages 31 and 41.)[163] M. F. Shlesinger and B. J. West. Complex fractal dimension<strong>of</strong> the bronchial tree. Phys Rev Lett, 67(15):2106–2108, 1991.(Cited on page 124.)[164] H. A. Simon. The architecture <strong>of</strong> complexity. Proc Am PhilosSoc, 106(6):467–482, 1962. (Cited on pages 3, 12, and 30.)[165] C. Song, S. Havlin, and H. A. Makse. Self-similarity <strong>of</strong> complexnetworks. Nature, 433(7024):392–395, 2005. (Cited onpage 24.)[166] O. V. Sosnovtseva, D. E. Postnova, E. Mosekilde, and N. H.Holstein-Rathlou. Synchronization <strong>of</strong> tubular pressure oscillationsin interacting nephrons. Chaos, Solitons and Fractals, 15(2):343–369, 2003. (Cited on pages 3, 12, 34, 40, and 44.)[167] O. V. Sosnovtseva, A. N. Pavlov, E. Mosekilde, K. P. Yip, N. H.Holstein-Rathlou, and D. J. Marsh. Synchronization amongmechanisms <strong>of</strong> renal autoregulation is reduced in hypertensiverats. AJP Renal Physiol, 293(5):1545–1555, 2007. (Cited onpages 3, 51, and 64.)[168] R. A. Star, H. Nonoguchi, R. Balaban, and M. A. Knepper. Calciumand cyclic adenosine monophosphate as second messengersfor vasopressin in the rat inner medullary collecting duct.J Clin Invest, 81(6):1879–1888, 1988. (Cited on page 52.)[169] J. L. Stephenson. Urinary concentration and dilution: models,chapter 30, pages 1349–1408. Oxford <strong>University</strong> Press, NewYork, second edition, 1992. (Cited on pages 11 and 57.)[170] J. L. Stephenson, R. Mejia, and R. P. Tewarson. Model <strong>of</strong> soluteand water movement in the kidney. P Natl Acad Sci USA, 73(1):252–256, 1976. (Cited on page 13.)[171] V. Strassen. Gaussian elimination is not optimal. NumerischeMathematik, 13(4):354–356, 1969. (Cited on page 32.)[172] S. Swaminathan, N. Leung, D. J. Lager, L. J. Melton, E. J.Bergstralh, A. Rohlinger, and F. C. Fervenza. Changing incidence<strong>of</strong> glomerular disease in Olmsted County, Minnesota:a 30-year renal biopsy study. Clin J Am Soc Nephrol, 1(3):483–487, 2006. (Cited on page 16.)


146 B I B L I O G R A P H Y[173] G. Székely. An approach to the complexity <strong>of</strong> the brain. BrainRes Bull, 55(1):11–28, 2001. (Cited on page 3.)[174] K. Takemoto and C. Oosawa. Modeling for evolving biologicalnetworks with scale-free connectivity, hierarchical modularity,and disassortativity. Math Biosci, 208(2):454–468, 2007. (Citedon page 23.)[175] S. R. Thomas. Inner medullary lactate production and accumulation:a vasa recta model. AJP Renal Physiol, 279(3):468–481,2000. (Cited on page 48.)[176] S. R. Thomas. Cycles and separations in a model <strong>of</strong> the renalmedulla. AJP Renal Physiol, 275(5):671–690, 1998. (Cited onpages 11 and 48.)[177] S. R. Thomas, A. T. Layton, H. E. Layton, and L. C. Moore.<strong>Kidney</strong> modeling: Status and perspectives. Proc IEEE, 94(4):740–752, 2006. (Cited on pages 3 and 48.)[178] S. C. Thomson, V. Vallon, and R. C. Blantz. <strong>Kidney</strong> function inearly diabetes: the tubular hypothesis <strong>of</strong> glomerular filtration.AJP Renal Physiol, 286(1):8–15, 2004. (Cited on pages 14, 16,and 17.)[179] L. N. Trefethen and R. S. Schreiber. Average-case stability <strong>of</strong>Gaussian elimination. SIAM J Matrix Anal Appl, 11(3):335–360,1990. (Cited on page 51.)[180] K. Tryggvason. Unraveling the mechanisms <strong>of</strong> glomerularultrafiltration: nephrin, a key component <strong>of</strong> the slit diaphragm.J Am Soc Nephrol, 10(11):2440–2445, 1999. (Cited on page 14.)[181] B. J. Tucker and R. C. Blantz. An analysis <strong>of</strong> the determinants<strong>of</strong> nephron filtration rate. AJP Renal Physiol, 232(6):477–483,1977. (Cited on page 11.)[182] A. M. Turing. On computable numbers, with an applicationto the entscheidungsproblem. Proc London Math Soc, 42(s2):230–265, 1937. (Cited on page 32.)[183] U.S. Renal Data System. USRDS 2002 Annual Data Report:Atlas <strong>of</strong> End-Stage Renal Disease in the United States.Bethesda, MD: National Institute <strong>of</strong> Diabetes and Digestiveand <strong>Kidney</strong> Diseases, National Institutes <strong>of</strong> Health, 2002. URLhttp://www.usrds.org/adr_2002.htm. (Cited on page 16.)[184] U.S. Renal Data System. USRDS 2007 Annual Data Report:Atlas <strong>of</strong> Chronic <strong>Kidney</strong> Disease and End-Stage Renal Diseasein the United States. Bethesda, MD: National Institute <strong>of</strong> Diabetesand Digestive and <strong>Kidney</strong> Diseases, National Institutes<strong>of</strong> Health, 2007. URL http://www.usrds.org/adr.htm. (Cited onpage 16.)[185] V. M. Vehaskari, E. R. Root, F. G. Germuth, Jr, and A. M. Robson.Glomerular charge and urinary protein excretion: Effects <strong>of</strong>systemic and intrarenal polycation infusion in the rat. <strong>Kidney</strong>Int, 22(2):127–135, 1982. (Cited on page 14.)


B I B L I O G R A P H Y 147[186] X. F. Wang and G. Chen. Complex networks: Small-world,scale-free and beyond. IEEE Circuits and Systems, 3(1):6–20, 2003. URL http://www.citebase.org/abstract?id=oai:arXiv.org:cond-mat/0303516. (Cited on pages 22 and 23.)[187] R. A. Watson and J. B. Pollack. Modular interdependency incomplex dynamical systems. Artificial Life, 11(4):445–457,2005. (Cited on pages 24 and 30.)[188] D. J. Watts and S. H. Strogatz. Collective dynamics <strong>of</strong> ‘smallworld’networks. Nature, 393(6684):440–442, 1998. (Cited onpage 22.)[189] A. M. Weinstein. Mathematical models <strong>of</strong> renal fluid and electrolytetransport: acknowledging our uncertainty. AJP RenalPhysiol, 284(5):871–884, 2003. (Cited on pages 3 and 13.)[190] A. M. Weinstein. A mathematical model <strong>of</strong> the rat proximaltubule. AJP Renal Physiol, 250(5 Pt 2):860–873, 1986. (Citedon pages 13 and 48.)[191] A. M. Weinstein. A mathematical model <strong>of</strong> rat collecting duct,part 1. AJP Renal Physiol, 283(6):1237–1251, 2002. (Cited onpage 13.)[192] A. M. Weinstein. A mathematical model <strong>of</strong> rat distal convolutedtubule, part 1. AJP Renal Physiol, 289(4):699–720, 2005. (Citedon page 13.)[193] A. M. Weinstein. Osmotic diuresis in a mathematical model <strong>of</strong>the rat proximal tubule. AJP Renal Physiol, 250(5):874–884,1986. (Cited on pages 13 and 48.)[194] W. C. Wells. Observations on the dropsy, which succeeds scarletfever. Trans Soc Improv Med Chir Knowl, 3:167–186, 1812.(Cited on page 15.)[195] W. C. Wells. On the presence <strong>of</strong> the red matter and serum <strong>of</strong>blood in the urine <strong>of</strong> dropsy, which has not originated fromscarlet fever. Trans Soc Improv Med Chir Knowl, 3:194–240,1812. (Cited on page 15.)[196] B. J. West. Fractal forms in physiology. Int J Mod Phys B, 4(10):1629–1669, 1990. (Cited on page 124.)[197] B. J. West, V. Bhargava, and A. L. Goldberger. Beyond theprinciple <strong>of</strong> similitude: renormalization in the bronchial tree. JAppl Physiol, 60(3):1089–1097, 1986. (Cited on page 124.)[198] T. A. Witten, Jr and P. Meakin. Diffusion-limited aggregationat multiple growth sites. Phys Rev B, 28(10):5632–5642, 1983.(Cited on page 124.)[199] T. A. Witten, Jr and L. M. Sander. Diffusion-limited aggregation.Phys Rev B, 27(9):5686–5697, 1983. (Cited on page 124.)[200] T. A. Witten, Jr and L. M. Sander. Diffusion-limited aggregation:a kinetic critical phenomenon. Phys Rev Lett, 47(19):1400–1403, 1981. (Cited on page 124.)


148 B I B L I O G R A P H Y[201] S. Wolfram. A New Kind <strong>of</strong> Science. Wolfram Media, 2002.ISBN 1-57955-008-8. URL http://www.wolframscience.com.(Cited on pages 4, 19, 21, and 24.)[202] B. Wong, A. Slivkins, and E. G. Sirer. Approximate matchingfor peer-to-peer overlays with cubit. Technical Reporthttp://hdl.handle.net/1813/10826, Computing and InformationScience, Cornell <strong>University</strong>, May 2008. (Cited on page 23.)[203] A. Wuensche. Basins <strong>of</strong> attraction in network dynamics: A conceptualframework for biomolecular networks. In G. Schlosserand G. P. Wagner, editors, Modularity in Development and Evolution,chapter 13, pages 288–311. Chicago <strong>University</strong> Press,2004. (Cited on page 20.)[204] A. Wuensche. Classifying cellular automata automatically: Findinggliders, filtering, and relating space-time patterns, attractorbasins, and the z parameter. Complexity, 4(3):47–66, 1999.(Cited on page 30.)[205] A. Wuensche. Complexity in one-d cellular automata: Gliders,basins <strong>of</strong> attraction and the z parameter. Technical Report94-04-025, Santa Fe Institute, 1994. (Cited on page 30.)[206] A. Wuensche. Self-reproduction by glider collisions: the beehiverule. In J. Pollack et al, editor, ALife9 Proceedings, pages 286–291. MIT Press, 2004. (Cited on pages 20 and 30.)[207] M. Zamir. Arterial branching within the confines <strong>of</strong> fractalL-system formalism. Comp & Graph, 118(3):267–276, 2001.(Cited on pages 124, 125, and 126.)[208] M. Zamir. Fractal dimensions and multifractility in vascularbranching. J Theor Biol, 212(2):183–190, 2001. (Cited onpage 124.)[209] M. Zamir. On fractal properties <strong>of</strong> arterial trees. J Theor Biol,197(4):517–526, 1999. (Cited on page 124.)[210] M. Zamir and S. Phipps. Morphometric analysis <strong>of</strong> the distributingvessels <strong>of</strong> the kidney. Can J Physiol Pharmacol, 65(16):2433–2440, 1987. (Cited on page 124.)[211] B. Zimmerhackl, C. R. Robertson, and R. L. Jamison. Themicrocirculation <strong>of</strong> the renal medulla. Circ Res, 57(5):657–667,1985. (Cited on page 12.)


C O L O P H O NThis thesis was typset with LAT E X 2ε using the Bera, Euler and Timestype faces. The typographic style is available for LAT E X via CTAN as“classicthesis”, except that this thesis uses Bera as the main fontand Times for smallcaps.Final Version as <strong>of</strong> 1st September 2008 at 21:00.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!