14
www.ietdl.org Published in IET Control Theory and Applications Received on 7th August 2011 Revised on 19th January 2012 doi: 10.1049/iet-cta.2011.0475 ISSN 1751-8644 Real-time control of human coagulation J.G. Makin 1 S. Narayanan 2 1 EECS Department, University of California, Berkeley, USA 2 International Computer Science Institute, 1947 Center Street, Berkeley, CA 94704, USA E-mail: [email protected] Abstract: In previous work, the authors showed that the dynamics of human blood clotting could be fruitfully modelled and simulated as a hybrid system (HS), that is, one with interacting continuous and discrete parts. Here, the authors show that, although a complete analysis of the HS is (computationally) infeasible, analysis and control techniques can indeed be applied to a large, critical subsystem a set of about 100 ordinary differential equations. The theory is outlined behind the control techniques and then demonstrate in a series of simulations their application to control of pathological blood clotting, both hypercoagulatory (factor-V Leiden) and hypocoagulatory (hæmophilia A). In particular, steering is simulated during a clotting event of the crucial blood–protein thrombin, via the controlled injection of (recombinant) factor VIII (for hæmophilia) or the anti-coagulant heparin (for FV Leiden). It remains to remedy the shortcomings of this control technique, and to extend it to the remainder of the HS of the previous work; methods for these are proposed, and addressed in a subsequent article. 1 Introduction The complexity of the coagulation process renders treatment of its disorders difficult. Indeed, current treatments make no attempt to manipulate coagulation in real time; rather, periodic (of the order of 1 day to 1 week) measurements and interventions are made with the aim of keeping cer- tain known risk factors within safe ranges. This limitation is a consequence of two facts: first, the state of current sensor technology, which precludes real-time measurements of blood proteins (and hence real-time feedback control); and second, the lack of theoretical techniques for such con- trol. This paper addresses the second of these limitations by applying techniques from computer science and from mathematical control theory and demonstrating their valid- ity – and limitations – both mathematically and in a series of simulations. By providing part of the solution to the theoretical problem of controlling blood clotting, we hope to provide an impetus for the development of the relevant sensor technologies. The simulations of our previous work [1] demonstrate the utility of the hybrid-system (HS) model described therein for certain purposes, namely prediction, theoretical inves- tigation and sensitivity analysis. However, ultimately we should like to use it as a basis for the real-time control of blood-clotting. Now, the standard technique for control of HSs is numerical solution of a set of partial differen- tial equations – the Hamilton–Jacobi equations. For systems of dimension greater than about five, however, solution of the equations is computationally infeasible (a consequence of the curse of dimensionality) [2]. Unfortunately, our sys- tem has upwards of 100 state variables; so this technique is patently unworkable. The approach taken here is to consider only a subsys- tem of the model, a set of non-linear ordinary differential equations (ODEs) that were originally drawn largely from [3, 4]. In fact, the model can be decomposed into this (purely continuous) subsystem and two HSs, the three inter- acting only through two state variables, as we shall see below (Section 3.1), so that control of clot formation in the complete model can be approached as a problem of continuous (non-linear) control. (A more elaborate justifi- cation for this decomposition can be found in Chapter 4 of [5].) Thus in the present study, we content ourselves with deriving and simulating the control techniques for the system of ODEs alone – which ODEs moreover are, in con- sequence of the mass-action kinetics they model, merely second-order polynomial functions of the state. Specifically, the control task is to steer the concentration of thrombin (activated blood factor II) along a desired trajectory dur- ing a clotting event, by controlled rate of injection of an exogenous pharmaceutical (e.g. heparin). We demonstrate two different techniques to effect this control: the first, following [6], using non-linear feedback and a change of variables to partially linearise the ODEs, and then control- ling the linearised, single-input/single-output (SISO) system using standard methods; and the second based on step- input control. In the simulations, the input is either the anti-coagulant heparin or (recombinant) factor VIII, and the output is thrombin, the most important enzyme product of the coagulation cascade (see below). We present the theory of the control techniques first, along with some preparatory results on alternative approaches, before presenting a series of simulations. We then simulate the clotting process in a patient with the pro-coagulatory disorder factor-V Leiden and in a patient with moderate 2630 IET Control Theory Appl., 2012, Vol. 6, Iss. 17, pp. 2630–2643 © The Institution of Engineering and Technology 2012 doi: 10.1049/iet-cta.2011.0475

ISSN 1751-8644 Real-time control of human coagulation

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: ISSN 1751-8644 Real-time control of human coagulation

www.ietdl.org

Published in IET Control Theory and ApplicationsReceived on 7th August 2011Revised on 19th January 2012doi: 10.1049/iet-cta.2011.0475

ISSN 1751-8644

Real-time control of human coagulationJ.G. Makin1 S. Narayanan2

1EECS Department, University of California, Berkeley, USA2International Computer Science Institute, 1947 Center Street, Berkeley, CA 94704, USAE-mail: [email protected]

Abstract: In previous work, the authors showed that the dynamics of human blood clotting could be fruitfully modelled andsimulated as a hybrid system (HS), that is, one with interacting continuous and discrete parts. Here, the authors show that,although a complete analysis of the HS is (computationally) infeasible, analysis and control techniques can indeed be appliedto a large, critical subsystem a set of about 100 ordinary differential equations. The theory is outlined behind the controltechniques and then demonstrate in a series of simulations their application to control of pathological blood clotting, bothhypercoagulatory (factor-V Leiden) and hypocoagulatory (hæmophilia A). In particular, steering is simulated during a clottingevent of the crucial blood–protein thrombin, via the controlled injection of (recombinant) factor VIII (for hæmophilia) or theanti-coagulant heparin (for FV Leiden). It remains to remedy the shortcomings of this control technique, and to extend it tothe remainder of the HS of the previous work; methods for these are proposed, and addressed in a subsequent article.

1 Introduction

The complexity of the coagulation process renders treatmentof its disorders difficult. Indeed, current treatments makeno attempt to manipulate coagulation in real time; rather,periodic (of the order of 1 day to 1 week) measurementsand interventions are made with the aim of keeping cer-tain known risk factors within safe ranges. This limitationis a consequence of two facts: first, the state of currentsensor technology, which precludes real-time measurementsof blood proteins (and hence real-time feedback control);and second, the lack of theoretical techniques for such con-trol. This paper addresses the second of these limitationsby applying techniques from computer science and frommathematical control theory and demonstrating their valid-ity – and limitations – both mathematically and in a seriesof simulations. By providing part of the solution to thetheoretical problem of controlling blood clotting, we hopeto provide an impetus for the development of the relevantsensor technologies.

The simulations of our previous work [1] demonstrate theutility of the hybrid-system (HS) model described thereinfor certain purposes, namely prediction, theoretical inves-tigation and sensitivity analysis. However, ultimately weshould like to use it as a basis for the real-time controlof blood-clotting. Now, the standard technique for controlof HSs is numerical solution of a set of partial differen-tial equations – the Hamilton–Jacobi equations. For systemsof dimension greater than about five, however, solution ofthe equations is computationally infeasible (a consequenceof the curse of dimensionality) [2]. Unfortunately, our sys-tem has upwards of 100 state variables; so this technique ispatently unworkable.

2630© The Institution of Engineering and Technology 2012

The approach taken here is to consider only a subsys-tem of the model, a set of non-linear ordinary differentialequations (ODEs) that were originally drawn largely from[3, 4]. In fact, the model can be decomposed into this(purely continuous) subsystem and two HSs, the three inter-acting only through two state variables, as we shall seebelow (Section 3.1), so that control of clot formation inthe complete model can be approached as a problem ofcontinuous (non-linear) control. (A more elaborate justifi-cation for this decomposition can be found in Chapter 4of [5].) Thus in the present study, we content ourselveswith deriving and simulating the control techniques for thesystem of ODEs alone – which ODEs moreover are, in con-sequence of the mass-action kinetics they model, merelysecond-order polynomial functions of the state. Specifically,the control task is to steer the concentration of thrombin(activated blood factor II) along a desired trajectory dur-ing a clotting event, by controlled rate of injection of anexogenous pharmaceutical (e.g. heparin). We demonstratetwo different techniques to effect this control: the first,following [6], using non-linear feedback and a change ofvariables to partially linearise the ODEs, and then control-ling the linearised, single-input/single-output (SISO) systemusing standard methods; and the second based on step-input control. In the simulations, the input is either theanti-coagulant heparin or (recombinant) factor VIII, and theoutput is thrombin, the most important enzyme product ofthe coagulation cascade (see below).

We present the theory of the control techniques first, alongwith some preparatory results on alternative approaches,before presenting a series of simulations. We then simulatethe clotting process in a patient with the pro-coagulatorydisorder factor-V Leiden and in a patient with moderate

IET Control Theory Appl., 2012, Vol. 6, Iss. 17, pp. 2630–2643doi: 10.1049/iet-cta.2011.0475

Page 2: ISSN 1751-8644 Real-time control of human coagulation

www.ietdl.org

hæmophilia A, and then repeat the simulation under appli-cation of the control techniques. Finally, we discuss boththe relevance of this control task to the overall task of con-trolling blood clotting, as well as implementation issues, andthen propose an alternative technique to remedy some of thedefects of the present approach.

2 Theory

The model to be controlled is a system of about 100 coupled,non-linear ODEs, of the most general form

x = f(x, u) (1)

where u is the (single) control variable (say, an anti- orpro-coagulant). The control task is to force one of the statevariables (thrombin) to track a desired trajectory. In fact,the ODEs can be expressed, as we shall see, in a less gen-eral form; and the approach described in this paper willbe to exploit some of the peculiarities of the system thatdistinguish it from the most general case.

Now, control of linear systems is comparatively easy, soa standard approach is to design the controller around a lin-ear approximation to the true system, found by consideringonly the first term of the Taylor expansion, the Jacobian[∂f/∂x](x0), near an equilibrium point x0. Unfortunately,the Jacobian in our model is singular, so (by the Hartman–Grobman theorem) the linearisation is not guaranteed toapproximate the true system.

Alternatively, the system may be exactly linearised (asopposed to merely approximated by a linear system) bychoosing the appropriate non-linear feedback u = ψ(h(x))and looking at the system through a change of variablesξ = �(x)

ξ = Aξ + bv

y = ξ1 (2)

where in fact (A, b) are in controllable canonical form, sothe system is completely controllable. Here, h(x) is an outputfunction that reflects our observations of the state variables,and v is a synthetic input related to the true input u by aknown function. (This discussion of feedback linearisationfollows [6], but see also [7].) From this perspective, we canask whether or not, given the dynamics in (1), there existsan output function rich enough to support the linearisation.The answer is particularly straightforward if, as in our case,the ODEs are affine in the control, that is can be written as

x = f(x) + g(x)u (3)

in which case necessary and sufficient conditions for thelinearisation can be given in the form of conditions on thevector fields f(x) and g(x) and iterated Lie brackets thereof.Specifically, the matrix of vector fields

[g(x), adfg(x), . . . , adn−2f g(x), adn−1

f g(x)] (4)

often referred to as the strong-accessibility distribution, musthave full rank (n) in the region of interest; and the set ofvector fields

g(x), adfg(x), . . . , adn−3f g(x), adn−2

f g(x) (5)

must be involutive in the region. (A set of vector fields isinvolutive if the Lie bracket of any two of those vector fields

IET Control Theory Appl., 2012, Vol. 6, Iss. 17, pp. 2630–2643doi: 10.1049/iet-cta.2011.0475

is within the span of the original set, where the Lie bracket[f, g] of two vector fields is defined as [∂g/∂x]f − [∂f/∂x]g.)Here, by recursive definition

adkf g := [f, adk−1

f g], k > 0

adkf g := g, k = 0

If these conditions are not met and full-state linearisation isnot possible, one may attempt to linearise the system par-tially. Here, the linearisation is carried out with respect tosome given output function, which will admit the formula-tion of some number q of new state variables with lineardynamics. Of course, if full-state linearisation is not possi-ble, then q, called the relative degree of the affine-controlsystem, is strictly less than n, the dimension of the statespace.

Consider again the differential equations of our system asthe affine-control system of (3), where this time h(x) = y,the variable we wish to control. Differentiating the outputwith respect to time yields

y = (∇xh)x

= (∇xh)f(x) + (∇xh)g(x)u

= Lfh(x) + Lgh(x)u (6)

where Lfh(x) := (∇xh)f(x), the Lie derivative of the func-tion h along the vector field given by f. Now if Lgh(x) isnon-zero for all x, then the system is said to have a strictrelative degree of one, and we can force it to track a desiredtrajectory yd by choosing

ud = 1

Lgh(x)(yd − Lfh(x)) (7)

and making sure that the initial conditions match (y(0) =yd(0), y(0) = yd(0)). Changing variables according to ξ =�(x) := h(x) and defining for the nonce yd =: v, a syn-thetic input, we see a one-dimensional (1D) linear systemξ = 0 · ξ + v and an (n − 1)-dimensional non-linear systemη = λ(ξ , η). (The variable v is called a synthetic input invirtue of its role as the input to this linear system.) Notethat the linear system is decoupled from the non-linear one,in the sense that ξ is not a function of η.

If, on the other hand, Lgh(x) is zero for all x, then wedifferentiate y a second time to obtain

y = L2f h(x) + LgLfh(x)u (8)

and, if for all x, LgLfh(x) �= 0 (i.e. the system has strictrelative degree two), choose

ud = 1

LgLfh(x)(yd − L2

f h(x)) (9)

this time making sure that y(0) = yd(0) as well. Choosingξ = �(x) := [h(x), Lf h(x)]T again yields a linear system,this time 2D, which is decoupled from the remaining (n −2)-dimensional non-linear system.

In general, for a system of strict relative degree q, thecontrol law is

ud = 1

LgLq−1f h(x)

(v − Lqf h(x)) (10)

where we choose the synthetic input v = y(q)

d and set all theinitial conditions yd(0), yd(0), . . . , y(q)

d (0) appropriately. Note

2631© The Institution of Engineering and Technology 2012

Page 3: ISSN 1751-8644 Real-time control of human coagulation

www.ietdl.org

Fig. 1 A conceptual rendering of feedback linearisationThe arrangement can be viewed either as a controller in a feed-back loop with the original control-affine system (solid lines), plusan output function; or as an equivalent output function along withcoupled linear and non-linear systems (dashed lines), with the for-mer evolving independently of the latter, and the output dependingsolely on the former

that, if q = n, the dimension of the state space, then

�(x) := [h(x), Lfh(x), . . . , Ln−1f h(x)]T (11)

is a valid change of coordinates (locally diffeomorphic),which in fact puts the system into controllable canonicalform, that is, (3) is fully linearisable by state feedback. Theprocedure is illustrated in Fig. 1.

Finally, we have so far ignored the case where LgLq−1f h(x)

is zero for some values of x, and non-zero for others. In thiscase, the relative degree of the system is not well defined,and complications ensue. (We shall see in the sequel thatthis is indeed the case in the present system.)

3 Application to the model

3.1 Decomposition

The set of ODEs in the present model leaves out two impor-tant sections of the coagulation cascade: (i) the intiation ofthe intrinsic pathway; and (ii) the final portion of the cas-cade, in which thrombin activates factors I and XIII and aclot forms. We claim, however, that control of the cascadethat is indeed represented in the ODEs of this paper suffices,under certain general conditions, to ensure proper clotting.

We address the beginning of the intrinsic (contact-activation) pathway first. Although it is theoretically possiblefor defects in the intrinsic pathway to affect clotting, this hasnot been observed clinically. Indeed, deficiencies of the clot-ting factors of this early portion of the intrinsic pathway donot have hemorrhagic consequences [8].

Now, this consideration might make us wonder what,in fact, the point of the contact-activation system is. Theanswer appears to be that it offers a method for initiatingclotting after blood-vessel trauma even without the action oftissue factor (viz., through the action of negatively chargedsurfaces on factor XII ) – a redundancy of sorts. (Onecan imagine circumstances under which TF would fail tobe released into the blood stream, or alternatively underwhich too little of it will have been been synthesised bythe endothelial cells.) The control techniques of this paper,then, assume the presence of sufficient tissue factor.

The situation with the final portion of the cascade isslightly more complicated. Now, obviously, a control mech-anism that does not include a model of the generation of

2632© The Institution of Engineering and Technology 2012

factors XIIIa and Ia cannot guarantee healthy clotting ifthe production of these factors or the initial values of theirunactivated forms are pathological; so use of the controllersderived in this paper assumes at least the the healthy func-tioning of this final portion of the cascade. However, thereis the additional wrinkle that thrombin complexes with bothfactors I and XIII, which binding is not modelled by the sys-tem of ODEs, and which will perforce decrease the amountof free thrombin available for the modelled interactions. Weassume, however, following [9], that very little of eithercomplex forms.

Armed with these considerations, we restrict our atten-tion from here on to a set of ODEs that describe the entirecoagulation cascade except the initiation of the intrinsicpathway and the final portion of the cascade (factors I andXIII). Now, to prove that control of these ODEs sufficesto induce healthy clotting, we should at least have to showthat the amount of factor XIa produced by the former isnegligible compared with the amount produced by recipro-cal activation by factors Xa and IIa. In fact, we might beinterested in showing for what range of initial conditionsin this ‘upstream’ subsystem this result obtains. Similarly, aproof would require us to show that the thrombin trajectorieswe control are sufficient to produce a clot, and within a rea-sonable amount of time. This question, too, might be posedin terms of ranges: what range of thrombin is sufficient toeffect this end?

Both of these questions are questions of reachability; theyare addressed in detail in [5]. Presently, however, we makedo without any proof. The idea, then, is to steer modelthrombin concentrations, with the hope that it can be proventhat these trajectories suffice to induce proper clotting. Lack-ing such a proof (and not wanting to rely on hope), we willthroughout aim at the tightest control possible: that is, wewill attempt to force thrombin concentrations in diseased-patient models to match the trajectory of thrombin during ahealthy clotting event. This may be overkill in the sense thatlooser criteria might be adduced, via the reachability anal-ysis lately outlined, to guarantee proper clotting, but it iscertainly sufficient (at least insofar as the model is correct).

Finally, if our controller is going to neglect these twoportions of the cascade, then certainly the input must notinteract with them. Fortunately, the major drugs for themajor clotting problems do not. In the following demonstra-tions, we focus on the two most common bleeding disorders,factor-V Leiden and hæmophilia A. The latter is treated with(recombinant) factor VIII, which lies in the subsystem mod-elled by the ODEs under consideration in this paper; and theformer is treated with heparin, which – as we shall see –increases the efficacy of antithrombin in inhibiting thrombinand factor Xa, all of which proteins are again modelled bythe ordinary differential equations appearing below.

3.2 Feedback linearisation

Full-state linearisation as lately described faces a signif-icant issue when applied to the present model. Althoughverifying the two conditions associated with (4) and (5) ismathematically straightforward, it is computationally inten-sive. Computing the vector fields requires the computationof two Jacobians for every Lie derivative, although one ofthem, ∂f/∂x, need only be computed once. However, eachof the approximately 100 vector fields of (4) needs to becomputed, and computation of each of the associated Jaco-bian matrices entails 1002 derivatives of polynomials. Evenneglecting the multiplication and addition operations, this

IET Control Theory Appl., 2012, Vol. 6, Iss. 17, pp. 2630–2643doi: 10.1049/iet-cta.2011.0475

Page 4: ISSN 1751-8644 Real-time control of human coagulation

www.ietdl.org

brings the total to a million computations – each of whichis a (symbolic) derivative of a polynomial, increasing thetotal number of computations yet more. And most unfortu-nately of all, each derivative (at least potentially) generatesmore variables, via the chain rule of differentiation, so thepolynomials have increasingly more terms in later vectorfields.

Apart from computational difficulties, it remains to con-struct the output function, h(x). No mechanical procedurefor this construction exists.

In fact, we can circumvent the computational obstacles.As lately noted, the rank of the strong-accessibility distribu-tion, (4), must be full for the system to be fully linearisable.However, as [10] have pointed out, in the case of reac-tion systems like the present one, this rank is bounded fromabove by the rank of a matrix D (the ‘accessibility matrix’)that is independent of x, and is computed very simply. Theexistence of such a matrix depends once again on being ableto express the governing ODEs in an even more specificform. In particular, (3) can be written

x = Cr(x) + bu (12)

where each element in the vector r(x) is a monomial cor-responding to one of the chemical reactions of the system(i.e. one of the arrows in the chemical formulae); and b isa constant vector, that is, does not depend on the state. Thepresent case is a special case even of the one discussed in[10], since our model includes neither inflow nor outflowof reactants. For this system, the accessibility matrix is theaugmented matrix D := [b, C], which we now show.

Consider a vector field s(x) in the span of D, that is

s(x) =∑

i

diφi(x) (13)

for arbitrary functions of x, φi, and where di are the columnsof D as defined above. Then the Lie bracket of the driftvector field f with s is

[f, s] = ∂s

∂xf(x) − ∂f

∂xs(x)

=∑

i

di(∇xφi)f(x) − C∂r

∂xs(x) (14)

Now, since (∇xφi)f(x) is just another arbitrary function ofx, the first term is within the span of D (compare this termwith (13)). Similarly, the second term lies within the spanof C and thus a fortiori lies within the span of D. Thus, wesee that the Lie bracket of f, the drift term of our ODE, withany vector field in the span of D, yields a vector field that isitself in the span of D. Since the control vector field g(x) ishere just the constant vector b, it too is in Sp{D}; and hencead1

f g = [f, g] is as well, by the fact just noted. Moreover,now since ad1

f g ∈ Sp{D}, then so is ad2f g = [f, ad1

f g] – andso on, extending the claim to all the vector fields in thestrong-accessibility distribution. Thus we can say

∀x, rk(D) ≥ rk([g(x), adfg(x), . . . , adn−2f g(x), adn−1

f g(x)])(15)

Crucially, the rank of D is easy to compute, being but aconstant matrix, and we avoid grinding through the iteratedLie derivatives of the right-hand side.

For a hæmophiliac patient, we use factor VIII as the sin-gle input, giving rk(D) = 83 < n = 98 – so the system is

IET Control Theory Appl., 2012, Vol. 6, Iss. 17, pp. 2630–2643doi: 10.1049/iet-cta.2011.0475

not fully linearisable by state feedback; that is, there is nosingle-variable output function h(x) that can be constructedto support such feedback. Now, factor-V Leiden implies theabsence of certain reactions (and corresponding state vari-ables), viz. those pertaining to the inactivation of factor Va.(The modelling of factor-V Leiden is explained at lengthbelow.) Treatment with heparin in turn requires the additionof several reactions and state variables. However, none ofthese changes, nor the different control vector field b – withits single non-zero entry in a different row – suffice to elim-inate the deficit between the rank of D and the dimension ofthe state: rk(D) = 73 < n = 89. The meaning of this deficitwe shall discuss later.

So the system is not completely controllable. However,there is reason to believe that clotting can be controlledjust by controlling the concentration of thrombin (activatedfactor II): clotting occurs downstream in the coagulation cas-cade of the ODEs of this study, and interacts with them onlyvia thrombin (as we saw in the previous section). Thus weattempt to partially linearise the system of (3) with throm-bin the output y = h(x). Indeed, using heparin as the controlvariable and augmenting the system with the appropriateterms from the heparin chemical reactions, the strict relativedegree of the system is two. (Lacking complete controlla-bility, is there something the system has gained? That is,is there a trade-off in the area? An affirmative answer canbe given if one considers the computational complexity: inorder to compute the feedback law, we need only calcu-late [two] twice-iterated Lie brackets; whereas if the systemwere completely controllable, the two vector fields wouldrequire the calculation of n-iterated Lie brackets. In a sys-tem as complicated as ours, this would be a serious problem,even given a simple output function. Additionally, we needonly specify two, rather than n, initial conditions. This ques-tion is even more relevant when an output function that willprovide strict relative degree n is available.) Alternatively,using factor VIII as the input (for hæmophilia) yields a strictrelative degree of three.

However, the relative degree for neither of these sys-tems is well-defined; in particular, the relevant Lie derivativeis zero at the initial condition, x0. We circumvent thisdifficulty by allowing the plant to run uncontrolled untilthe concentration of thrombin is some distance δ from theorigin, that is, until the state has drifted away from thesingularity at which the relative degree changes, beforeturning the controller on. The parameter δ is tuned man-ually for ideal tracking. (The non-regularity of the sys-tem and its consequences are discussed in more detail inSection 6.2 below.)

3.2.1 Error correction: Now, in practice, small numer-ical discrepancies arise – for example, from a mismatchbetween the plant and the model, or from numerical approx-imations – so that the trajectory produced by the control lawof (10) deviates from the desired trajectory. Compensationis made in the form of additional elements in the feedbackloop, usually some variant on the PID controller. A propor-tional and an integral term are generally used [11], but herewe content ourselves with proportional and derivative terms(as many as the strict relative degree of the system) since(i) the integral term did not seem to improve tracking in oursimulations; and (ii) it can be shown [6] that, if the systemof (3) is ‘globally exponentially minimum-phase’, then thiscontrol law guarantees bounded trajectory tracking, that is,the tracking error and its derivatives tend asymptotically to

2633© The Institution of Engineering and Technology 2012

Page 5: ISSN 1751-8644 Real-time control of human coagulation

www.ietdl.org

zero. That is, we replace the synthetic input v in (10) with

v(t) = y(q)

d (t) + [Kp K1d · · · Kq−1

d ][yd − ξ] (16)

where yd is the vector of the desired output and its deriva-tives:

yd = [yd yd · · · y(q−1)]T

and ξ is the linearised state, which in our system correspondsto the actual output y and its derivatives.

The question of whether the system is indeed globallyexponentially minimum-phase turns out to be complicated;it is discussed more fully in in a companion article, buthere we summarise those conclusions. The usual test is tomake sure that the ‘zero-dynamics’ – the evolution of theuncontrolled, non-linear subsystem η = λ(ξ , η) (see Fig. 1),constrained to the manifold where the linear-system vari-ables ξ are zero – is itself exponentially stable at the relevantequilibrium point. Unfortunately, the change of coordinates(ξ, η) = �(x) breaks down (loses rank) precisely here owingto the non-regularity of the system – due in particularto the zero concentration of thrombin. More sophisticatedapproaches [12, 13] require computing the relative degree ofthe system at the singularity; but being much higher than 2 inour model, this turns out to incur similar, prohibitive, com-putational penalties in iterated Lie brackets as the ones latelydiscussed in generating the full accessibility distribution.

Failure to apply the usual tests does not, of course, showthat the system is not exponentially minimum-phase, andmoreover there are two additional reasons for suspectingthe non-linear dynamics to be stable. First, the coagulationcascade as modelled here – even in the case of factor-VLeiden – ultimately eventuates in the binding of all activeproteases into inactive complexes. (Whether this happensexponentially rather than merely asymptotically is not asobvious.) The introduction of the anti-coagulatory heparininto the system is unlikely to upset this stability. Second,the control schemes do appear to achieve bounded trackingin continuous-time simulation below.

The gains were calculated by treating the system as alinear-quadratic regulator, with – unless otherwise noted –state-penalty matrix set to the identity matrix and the inputpenalty set to unity. Solution of the optimisation problemwith Matlab’s lqr function gives Kp = 1, Kd = 1.7321 forcontrol under factor-V Leiden. However, control of the non-linear system was observed to be fairly insensitive to thechoice of cost parameters. Increasing or decreasing the ratioof the state penalty to the input penalty by as much as anorder of magnitude yielded qualitatively similar thrombintrajectories (but see the discussion in Section 6.2 below).

3.2.2 Discretisation: Finally, practical application willalso require discretisation of the controller. In the simu-lations below, this was achieved simply by allowing it toupdate only every 0.5 s. The model was then simulated incontinuous time using a zero-order hold. Now, the samplingrate of the controller affects the controllability of the sys-tem: under a sufficiently low rate the theory lately outlinedwill fail to achieve exact output tracking. Two hertz wastherefore chosen simply by determining, via trial and error,the lowest rate that would not severely degrade the results.

3.3 Step-input control

A much simpler control technique was also applied; therationale for it is discussed below. This strategy is predi-cated on the assumption that a single-step input might suffice

2634© The Institution of Engineering and Technology 2012

to force the system to track the desired thrombin trajec-tory ‘reasonably well’. More precisely, the assumption isthat the step input that matches both the desired peak con-centration of thrombin and the occurrence of this peak willresult in a thrombin trajectory that deviates very little overits entirety from its desired counterpart. Therefore this tech-nique was implemented by performing a parameter searchover repeated trials for that step input that would minimisethe peak-concentration and peak-time discrepancies. Thusthe input u was modified on successive trials according to:

u ← u + α(

maxt

{y(t)} − maxt

{yd(t)})

+ β(argmaxt{y(t)} − argmaxt{yd(t)}

)(17)

where α and β-positive for anti-coagulant inputs and neg-ative for pro-coagulants – scale the relative contribution ofeach term, and were adjusted by hand. The search was per-formed as follows. First, the time and location of the desiredthrombin peak were computed from the desired throm-bin trajectory (for its computation, see Section 4 below).Then, for fixed α and β and an input u initialised at 0,the cascade was simulated, from which the time and loca-tion of the thrombin peak was calculated. The input wasthen modified according to (17) and the process repeateduntil convergence. Since the relative weighting of peak-concentration and peak-time discrepancies yielded better andworse matches of the overall trajectory, it was adjusted (viaα and β) by hand, another search performed, and so on;until visual inspection showed a good match. Useful valuesof α and β are reported in the results below (in general, agood ratio of the former to the latter was about five).

4 Methods

We draw our chemical reactions from three main sources,the first two of which collect their own data from a numberof primary sources (the last is itself a primary source): [3,4, 14]. A description of each follows. The entire system issummarised in Fig. 2.

Forty-eight chemical reactions for normal blood clotting,with no exogenous intervention, were taken directly from[3], from which also the numerical values of the rate con-stants were supplied. They are reproduced here in Table 1.The blood factors are referred to by their Roman-numeraldesignations, a lowercase ‘a’ denoting the activated form.Other abbreviations include: mIIa for meizothrombin, LBSfor the concentration of lipid binding sites, PS for proteinS, PC for protein C, APC for its activated form, TFPI fortissue-factor-pathway inhibitor, Tm for thrombomodulin, ATfor antithrombin and TF for tissue factor. A subscripted‘i’ indicates the inactivated form of an enzyme, and asubscripted ‘L’ indicates the lipid-bound form.

The simple mechanism of factor Va inactivation of the[3] model consisted of a single chemical equation. We sub-stituted for this equation the model of factor-Va inactivationof [4], which was designed specifically to match data forfactor-V Leiden. (For completeness: in fact, we used allof the chemical equations from [3] except two: the deac-tivation of FVaL by APC; and the unbinding of FVaiL, thelipid-bound, inactivated form of factor Va, from its phos-pholipid surface – since, after all, FVaiL never shows up inour model, being replaced by various cleaved forms of fac-tor Va.) These 34 chemical equations are listed in Table 2.Factor-V Leiden is a genetic disorder in which the amino

IET Control Theory Appl., 2012, Vol. 6, Iss. 17, pp. 2630–2643doi: 10.1049/iet-cta.2011.0475

Page 6: ISSN 1751-8644 Real-time control of human coagulation

www.ietdl.org

Fig. 2 Coagulation cascadeTriangle-headed arrows indicate transformations to activated forms;flat-headed lines connect the catalysts to the reaction they catalyse;line-arrowheaded arrows indicate binding reactions; and ball-endedlines denote inhibitory reactions. Reactions that are not treated inthis model (see Section 3.1) have been greyed out. The input fortreatment of factor-V Leiden, heparin, is in pale grey. The remain-ing arrows (true-arrowheaded) mark the initiation points of thecascade. Some reactions have been omitted to reduce clutter: thelipid bindings, those involving meizothrombin (mIIa; since its roleis identical to thrombin), the sub-reactions that compose factor-Vdeactivation (Table 2), and the sub-reactions of heparin-mediatedthrombin inactivation (Table 3)

acid Arg505 of the heavy chain of factor V is replaced byGln505, which effectively disables APC-mediated cleavageof factor V at this point. The disorder was discovered onlyas recently as 1994, but it is in fact quite common, affect-ing perhaps 5% of the white population, among whom it isthe most prevalent of the inherited hypercoagulatory disor-ders. Now, [4] were able to match data for inactivation ofFVaLEIDEN by assuming that the form of factor V cleavedat Arg505, Va5, has no effect on clotting, that is, it is ‘inac-tivated’. (This is so even despite its reported 60% efficacyin complex with factor Xa as prothrombinase. Since it hasno overall effect, and for the sake of simplicity, we treatedVa5L as inert.)

We made a further simplification in the incorporation ofthis model: [4] model the inactivation of factor Va by APC,rather than the more efficacious APC:PS complex, as in [3].Now in fact, uncomplexed APC competes with APC:PS forbinding with factor Va, but [3] leaves out this competition,that is, only APC:PS is modelled as binding with Va – pre-sumably because, under normal conditions, very little freeAPC circulates: the vast majority of it is immediately boundup by protein S. We follow this expedient in appropriatingthe [4] model, replacing APC with APC:PS in all the rele-vant chemical equations. Does not this substitution requirean alteration of the rate constants? After all, APC:PS issupposed to be more efficacious. Yes: and here we followthe results of [15], who found that the increased efficacyof the complex results from a factor of three increase in

IET Control Theory Appl., 2012, Vol. 6, Iss. 17, pp. 2630–2643doi: 10.1049/iet-cta.2011.0475

the cleavage rate of Va at Arg306. (Previously it had beenreported that the factor was twenty rather than three, but as[15] notes, this would abolish the thrombophilic effect offactor-V Leiden: in the presence of protein C, the cleavageat Arg306 would dominate the cleavage at Arg505, so thatthe disabling of the latter in FVL would have no overalleffect on clotting.) In the interest of perspicuity, the fac-tor of three and the original rate constant from [4] are listedexplicitly in Table 2. Of course, if we wanted to model (e.g.)protein-S deficiency, we should need to put the equations for(uncomplexed) APC inactivation of FVa into the model.

Does this complicated replacement for the single equationof [3] yield similar results under normal (healthy) condi-tions? In fact, there is a significant change: under the morecomplicated model, thrombin peaks during normal clottingaround 90 nM, whereas under [3]’s model, it peaked atalmost 40 nM (although, curiously, the time of the peakis the same in both models). Now [3] verifies their modelby comparing it to clinical observations of thrombin curvesunder various physiological conditions, in particular by com-parison with Fig. 5 of [16]. However, oddly enough, theseconditions assume no protein C or protein S, so [3]’s sim-ulations do not actually verify this part of their model. Inpoint of fact, [3]’s ‘verifications’ are anyway qualitative: thethrombin curves from [16], while peaking at similar con-centrations, are somewhat delayed compared with those of[3]. Resolution of this issue, too, we defer to the discus-sion section below. For now it should be said that the moredetailed model of [4] is probably more accurate than thesingle equation of [3], so we proceed with some confidence.

The heretofore unexplained abbreviations appearing in thetable are as follows. The numbers 3, 5, and 6 denote the siteor sites at which the species of factor Va has been cleaved,respectively, Arg306, Arg505 and Arg662. Thus, Va53 is factorVa cleaved at both Arg306 and Arg505; and VaA3 is factorV’s A2 peptide chain, cleaved at Arg306. LC denotes thelight peptide chain of factor Va and HC its heavy chain;and VaLC is the light chain and A1 peptide chain. Thesubscripted L again denotes the lipid-bound species.

A word on the lipid-binding: the reader may notice thatnot all of the proteins subscripted with an L have equa-tions for dissociating from their lipid substrates. In this,we are for the most part following [3], in which onlythe simple proteins, but not the compounds, are allowedto dissociate. For the remainder, we have followed [4] innot allowing the uncomplexed proteins LCL, Va5L, Va3L,Va53L, Va36L, Va56L, Va536L and VaLCL to dissociatefrom their substrates. Now, in fact, the experiments in [4]assume a saturating amount of lipids; whereas in our model,the phospholipid concentration is something less than satu-rating. Technically, then, these proteins should be allowedto dissociate. However, assuming on- and off-rates equalto that for (uncleaved) factor Va, these dissociations havean entirely negligible effect even on the concentrations ofthe FVa-variants themselves, let alone on thrombin levels –which is to be expected, since phospholipids never fall belowabout 1500 nM throughout the simulations, at which level-binding forces vastly predominate over unbinding forces (cf.the rate constants of chemical reaction 4). Note, also, thatthe heavy-chain variants of factor Va are not lipid bound: thebinding site for lipids lies in the light chain of factor V [4].

Finally, on this head, a correction was required: [3] claimthat thrombin is released from the lipid surface when itis activated. However, conservation of mass requires theselipid-binding sites to be returned to the pool of uncom-plexed LBS; thus in chemical reaction 33 we have added

2635© The Institution of Engineering and Technology 2012

Page 7: ISSN 1751-8644 Real-time control of human coagulation

www.ietdl.org

2636© The Institutio

Table 1 Chemical reactions and rate constants drawn from [3]; they govern all the reactionsexcept factor Va inactivation and the interaction of heparin with the system

No. Reaction kon, nM−1 s−1 koff, s−1 kcat, s−1

1 II + LBS � IIL 0.0043 1 –2 mIIa + LBS � mIIaL 0.05 0.475 –3 V + LBS � VL 0.05 0.145 –4 Va + LBS � VaL 0.057 0.17 –5 VII + LBS � VIIL 0.05 0.66 –6 VIIa + LBS � VIIaL 0.05 0.227 –7 VIII + LBS � VIIIL 0.05 0.1 –8 VIIIa + LBS � VIIIaL 0.05 0.335 –9 IX + LBS � IXL 0.05 0.115 –10 IXa + LBS � IXaL 0.05 0.115 –11 X + LBS � XL 0.01 1.9 –12 Xa + LBS � XaL 0.029 3.3 –13 APC + LBS � APCL 0.05 3.5 –14 PS + LBS � PSL 0.05 0.2 –15 VIIIai + LBS � VIIIai,L 0.05 0.335 –16 PC + LBS � PCL 0.05 11.5 –17 TFL + VIIaL � TF:VIIaL 0.5 0.005 –18 TFL + VIIL � TF:VIIL 0.005 0.005 –19 TF:VIIaL + IXL � TF:VIIa:IXL

→ TF:VIIaL + IXaL 0.01 2.09 0.3420 TF:VIIaL + XL � TF:VIIa:XL → TF:VIIa:XaL 0.1 32.5 1.521 TF:VIIa:XaL → TF:VIIaL + XaL – 1 –22 TF:VIIL + XaL � TF:VII:XaL → TF:VIIaL + XaL 0.05 44.8 15.223 IXaL + VIIIaL � IXa:VIIIaL 0.1 0.2 –24 XaL + VaL � Xa:VaL 1 1 –25 IXa:VIIIaL + XL � IXa:VIIIa:XL

→ IXa:VIIIaL + XaL 0.1 10.7 8.326 VL + XaL � V:XaL → VaL + XaL 0.1 1 0.04327 VIIIL + XaL � VIII:XaL → VIIIaL + XaL 0.1 2.1 0.02328 VL + IIa � V:IIaL → VaL + IIa 0.1 6.94 0.2329 VIIIL + IIa � VIII:IIaL → VIIIaL + IIa 0.1 13.8 0.930 Xa:VaL + IIL � Xa:Va:IIL 0.1 100 –31 Xa:VaL + mIIaL � Xa:Va:mIIaL 0.1 66 –32 Xa:Va:IIL → Xa:Va:mIIaL – – 1333 Xa:Va:mIIaL → Xa:VaL + IIa + LBS – – 1534 VIIL + XaL � VII:XaL → VIIaL + XaL 0.05 44.8 15.235 XI + IIa � XI:IIa → XIa + IIa 0.1 10 1.4336 APC:PSL + VIIIaL � APC:PS:VIIIaL

→ APC:PSL + VIIIai,L 0.1 1.6 0.437 TFPI + Xa � TFPI:Xa 0.016 0.000333 –38 TFPI:Xa + TF:VIIaL � TFPI:Xa:TF:VIIaL 0.01 0.0011 –39 IXa + AT → IXa:AT 4.9 ×10−7 – –40 Xa + AT → Xa:AT 2.3 ×10−6 – –41 IIa + AT → IIa:AT 6.83 ×10−5 – –42 VL + mIIaL � V:mIIaL → VaL + mIIaL 0.1 6.94 1.03543 VIIIL + mIIaL � VIII:mIIaL → VIIIaL + mIIaL 0.1 13.8 0.944 IIa + TmL � IIa:TmL 1 0.5 –45 IIa:TmL + PCL � IIa:Tm:PCL → IIa:TmL + APCL 0.1 6.4 3.646 mIIa + AT → mIIa:AT 6.83 ×10−6 – –47 APCL + PSL � APC:PSL 0.1 0.5 –48 XIa + IXL � XIa:IXL → XIa + IXaL 0.01 1.417 0.183

an LBS term that was missing from [3]. Of course, as wehave just finished saying, the concentration of binding sitesis in excess over the other proteins, so this addition makesno discernible difference to the simulations.

The form of the heparin equations comes from [14] andappear in Table 3. The abbreviated proteins are heparin,thrombin (activated factor II), and antithrombin; (IIa-AT)

n of Engineering and Technology 2012

is a stable complex, which can no longer dissociate intoinhibitor and protease.

Such are the forms of the heparin reactions; however, toour knowledge, the exact values of the on- and off-ratesunder physiological conditions have never been determined.In [14], however, are found estimates of the ratios of off-to on-rates; the following expedient was therefore adopted:

IET Control Theory Appl., 2012, Vol. 6, Iss. 17, pp. 2630–2643doi: 10.1049/iet-cta.2011.0475

Page 8: ISSN 1751-8644 Real-time control of human coagulation

www.ietdl.org

IET Control Theordoi: 10.1049/iet-ct

Table 2 Chemical reactions and rate constants of the [4] model of factor-Va deactivation;factor-V Leiden is modelled as knocking out the reactions that cleave factor V at the Arg505

site; and proteins cleaved at this site are marked by a 5 (see the text for abbreviations)

No. Reaction kon, nM/s koff, s−1 kcat, s−1

49 LCL + HC � VaL 2.63 × 10−6 1.72 × 10−5 –50 LCL + HC5 � Va5L 2.63 × 10−6 1.72 × 10−5 –51 LCL + HC3 � Va3L 2.63 × 10−6 1.72 × 10−5 –52 LCL + HC53 � Va53L 2.63 × 10−6 1.72 × 10−5 –53 LCL + HC36 � Va36L 2.63 × 10−6 1.72 × 10−5 –54 LCL + HC56� Va56L 2.63 × 10−6 1.72 × 10−5 –55 LCL + HC536 � Va536L 2.63 × 10−6 1.72 × 10−5 –56 LCL + APC:PS L � LC:APC:PSL 2.63 × 10−6 1.72 × 10−5 –57 VaL + APC:PSL � Va:APC:PSL 0.1 0.7 –58 Va5L + APC:PSL � Va5:APC:PSL 0.1 0.7 –59 Va3L + APC:PSL � Va3:APC:PSL 0.1 0.7 –60 Va53L + APC:PSL � Va53:APC:PSL 0.1 0.7 –61 Va36L + APC:PSL � Va36:APC:PSL 0.1 0.7 –62 Va56L + APC:PSL � Va56:APC:PSL 0.1 0.7 –63 Va536L + APC:PSL � Va536:APC:PSL 0.1 0.7 –64 VaLCL + APC:PSL � VaLC:APC:PSL 0.1 0.7 –65 VaLCL + VaA3 � Va3L 2.57 × 10−6 0.028 –66 VaLCL +VaA53 � Va53L 2.57 × 10−6 0.028 –67 VaLCL +VaA36 � Va36L 2.57 × 10−6 0.028 –68 VaLCL +VaA356 � Va536L 2.57 × 10−6 0.028 –69 VaLC:APC:PSL + VaA3 � Va3:APC:PSL 2.57 × 10−6 0.028 –70 VaLC:APC:PSL + VaA53 � Va53:APC:PSL 2.57 × 10−6 0.028 –71 VaLC:APC:PSL + VaA36 � Va36:APC:PSL 2.57 × 10−6 0.028 –72 VaLC:APC:PSL + VaA356 � Va536:APC:PSL 2.57 × 10−6 0.028 –73 Va:APC:PSL ⇀ Va3:APC:PSL – – 3*0.06474 Va5:APC:PSL ⇀ Va53:APC:PSL – – 3*0.06475 Va5:APC:PSL ⇀ Va56:APC:PSL – – 5.2 × 10−4

76 Va3:APC:PSL ⇀ Va36:APC:PSL – – 5.2 × 10−4

77 Va53:APC:PSL ⇀ Va536:APC:PSL – – 5.2 × 10−4

78 Va56:APC:PSL ⇀ Va536:APC:PSL – – 3*0.06479 Va:APC:PSL ⇀ Va5:APC:PSL – – 180 Va3:APC:PSL ⇀ Va53:APC:PSL – – 181 Va36:APC:PSL ⇀ Va536:APC:PSL – – 1

Table 3 Chemical equations of the heparin reactions; the equations weredrawn from [14], but the rate constants were estimated as in the text

No. Reaction kon, nM/s koff, s−1 kcleavage, s−1

82 Hep + IIa � Hep:IIa 0.1 0.1 ∗ 3 × 104 –83 Hep + AT � Hep:AT 0.1 0.1 ∗ 2 × 102 –84 Hep:IIa + AT � Hep:AT:IIa 0.1 0.1 ∗ 20 –85 Hep:AT + IIa � Hep:AT:IIa 0.1 0.1 ∗ 3 × 103 –86 IIa:AT + AT � Hep:AT:IIa 0.1 0.1 ∗ 0.6 –87 Hep:AT:IIa → IIa-AT – – 5

On-rates were set to an intermediate value within the biolog-ically normal range for enzyme-substrate reactions (see forexample Table 4.4 in [17], but also [3]), viz. 0.1 s/nM. Thecorresponding off-rates were then computed by multiplyingthe aforementioned ratios by the on-rates, that is, 0.1. Again,this is represented explicitly in the table. This approximationdoes not vitiate the theoretical apparatus, but it does havenumerical consequences; we take up this question again inthe discussion below.

The final rate constant is a catalytic rate, and as suchis taken directly from [14]. However, we also deviate inone other particular from that paper: whereas [3] model ATbinding to IIa as irreversible, with a very slow on-rate, [14]considers the reaction to be reversible, with a very high

y Appl., 2012, Vol. 6, Iss. 17, pp. 2630–2643a.2011.0475

off-to-on ratio. These mechanisms are certainly not equiv-alent, and we use the more recent mechanism from [3].However, in any case, this reaction will be dominated bythe much higher affinity reaction between AT and heparin,the primary path to the Hep:AT:IIa complex, so the choicehas little practical consequence.

Finally: heparin is also thought to facilitate the inacti-vation of factor Xa, but once again for simplicity theseinteractions have been neglected.

Table 4 lists all the non-zero initial concentrations. Theconcentration of lipid binding sites and the initial activat-ing amount of tissue factor were chosen to match [16], asin [3]. All other concentrations are normal physiologicalvalues.

2637© The Institution of Engineering and Technology 2012

Page 9: ISSN 1751-8644 Real-time control of human coagulation

www.ietdl.org

2638© The Institution of

Table 4 Non-zero initial conditions; all other proteins are initialised at zero

Species Conc., nM Species Conc., nM Species Conc., nM

tissue factor 0.005 factor VIII 0.7 TFPI 2.5factor II 1400 factor IX 90 antithrombin 3400factor V 20 factor XI 30 protein C 60factor VII 10 factor X 170 protein S 300factor VIIa 0.1 thrombomodulin 1 LBS 3396

The chemical reactions of Tables 1–3 are equivalent toa set of coupled, non-linear, ODEs. For completeness weillustrate the transformation for the first chemical equation,II + LBS � IIL. This is equivalent to

d[II]dt

= −kon[II][LBS] + koff[IIL]d[LBS]

dt= −kon[II][LBS] + koff[IIL]

d[IIL]dt

= kon[II][LBS] − koff[IIL]

where square brackets denote (time-varying) concentration.More succinctly, as in (12), we may write

d

dt

[ [II][LBS][IIL]

]=

[−1 1−1 11 −1

] [kon[II][LBS]

koff[IIL]]

(18)

Incorporating the other chemical reactions amounts toadding monomials to each ODE as necessary, and of courseadding an ODE for each new chemical introduced; or, alter-natively, to adding new columns to the numerical matrix,C, for each new rate constant, and adding a new row foreach new chemical. This simple structure is a consequenceof the mass-action kinetics of the coagulation reactions, andin turn makes possible the algebraic proof exploited above(Section 3.2).

All simulations were performed in MATLAB [18] usingthe stiff ODE solver ode15s. Computation of the strict rel-ative degree of the system was also performed in MATLAB,using the symbolic toolbox. For all simulations, the ‘desiredtrajectory’ (time course) of thrombin concentration was pre-computed by first simulating healthy coagulation, that is,numerically solving the unmodified system of ODEs. Sim-ulation of the control techniques then took place in modelsthat had been modified appropriately to account for thedisease of interest, as above.

5 Results

5.1 Factor-V Leiden: unconstrained, continuous,feedback-linearising control

We first simulate coagulation in a patient with the hyper-coagulatory disorder factor-V Leiden, with and withoutintervention by the anti-coagulant heparin, as well as normal(non-pathological) clotting. Initiation of the clotting event isassumed to take place via the intrinsic pathway and is there-fore modelled by initialising tissue factor at a concentrationof 5 pM (following [3]).

Fig. 3 shows the thrombin profile during normal clot-ting (dark line), clotting in a ‘patient’ with factor-V Leiden(light line), and that clotting in the same patient but with

Engineering and Technology 2012

the feedback-linearising controller (medium line). Exact out-put tracking has been achieved; however, several defects areimmediately obvious. First, the controller operates continu-ously, that is, updating as often as the numerical simulationof the ODEs does, whereas any practical controller must bedigital. The input, appearing in at right in Fig. 3, is obviouslynot feasible for any imaginable drug-delivery mechanism.In fact, to avoid numerical errors, the input was clamped at100 nM/s. (We explain this issue in the discussion sectionbelow.) Finally, and relatedly, a realistic controller will havea maximum and minimum input rate; certainly, the ratecannot be negative, since this implies withdrawal of hep-arin from the site of the injury. The present controller wasallowed us to do just that.

5.2 Factor-V Leiden: constrained, continuous,feedback-linearising control

We address the final defect first. Note, however, that thetheory outlined above does not guarantee the success of anythese remedies. As lately noted, the input cannot be a neg-ative number, nor can it be much higher than a single-digitnanomolarity per second. (Continuous intravenous infusionsover a 24 h period total between 20 000 and 40 000 Inter-national units, which comes to about 10 pM/s. Since ourtreatment lasts only over the course of a few minutes, we canpresumably use higher doses over this short interval with-out overmuch hæmophilic risk downstream.) We impose anupper bound (somewhat arbitrarily) at 20 nM/s and a lowerbound of zero by ‘squashing’ inputs outside this range, thatis, setting any u calculated from (10) that exceeds one ofthese bounds to the value of the bound. The result appearsin Fig. 4.

There is indeed some degradation from the unconstrainedcontroller; in particular, the controller overshoots the peaktrajectory somewhat, and is unable to compensate for thefinal undershoot. These defects are respectively the conse-quences of the maximum and minimum constraints; but thelatter at least might be remedied by some form of anti-cipation; that is, if the controller could predict the finalundershoot, it might ease off the input earlier. This pos-sibility is explored in the discussion below. As for theovershoot, it can be eliminated by raising the maximuminput to 50 nM/s (graph not shown). However, in the absenceof an established maximum, it might be safer to try toaddress this issue with an anticipatory control as well.

5.3 Factor-V Leiden: constrained, discrete,feedback-linearising control

The corresponding input at right in Fig. 4 still exhibitssome chatter, naturally, since it is allowed to vary con-tinuously. That shortcoming is rectified by discretising thecontroller, which results in the output and input of Fig. 5(solid lines). Here the output again overshoots the desired

IET Control Theory Appl., 2012, Vol. 6, Iss. 17, pp. 2630–2643doi: 10.1049/iet-cta.2011.0475

Page 10: ISSN 1751-8644 Real-time control of human coagulation

www.ietdl.org

0 50 100 150 200 250 3000

10

20

30

40

50

60

70

80

90

time (s)

conc

entra

tion

(nM

)

[IIa] Desired[IIa] Uncontrolled[IIa] Controlled

0 50 100 150 200 250 300−80

−60

−40

−20

0

20

40

60

80

100

time (s)

conc

entra

tion

rate

(nM

/s)

Fig. 3 Graphs showing thrombin profile during normal clottingTop: simulated control of thrombin concentration during a clottingevent in a patient with factor-V Leiden. The controller was uncon-strained, and was updated continuously, allowing arbitrarily closetracking of the desired trajectory. Bottom: the heparin input in nM/sthat generated the output above. Although this input yields perfecttracking, it is obviously unacceptable for any realistic drug-deliverymechanism, operating as it does on a continuous time scale andusing negative inputs (see the text)

thrombin trajectory, but also undershoots the pre-peak tra-jectory, evidently because it was unable to recover fromthe initial input; and, for similar reasons, undershoots thepost-peak thrombin curve slightly more than its counterpartcontinuous controller (Fig. 4).

A smaller controller step yields, as expected, superiorresults (not shown); and raising the input maximum againeliminates the overshoot. In the limit, of course, we shouldbe able to reproduce the results of the continuous controller.However, and for that very reason, the smaller the step, themore unrealistic the controller. Then, in any case, we shallpresumably not be able to eliminate the undershoot that thecontinuous controller exhibits, since that results from theinput minimum; nor, if we take seriously the input maximumof 20 nM/s, shall we be able to eliminate the overshoot.

The ‘clipping’ that the upper constraint imposes on theinput can be eliminated by increasing the input penalty (rel-ative to the state penalty) in the linear-quadratic regulator(producing feedback gains Kp = 0.1414, Kd = 0.5503), butat the cost of some degradation in output tracking. Nor couldadjustment of the relative penalties eliminate clipping at the

IET Control Theory Appl., 2012, Vol. 6, Iss. 17, pp. 2630–2643doi: 10.1049/iet-cta.2011.0475

0 50 100 150 200 250 3000

10

20

30

40

50

60

70

80

90

time (s)

conc

entr

atio

n (n

M)

0 50 100 150 200 250 3000

2

4

6

8

10

12

14

16

18

20

time (s)

conc

entr

atio

n ra

te (

nM/s

)

[IIa] Controlled

[IIa] Uncontrolled

[IIa] Desired

Fig. 4 Graphs showingTop: simulated control of thrombin concentration during a clot-ting event in a patient with factor-V Leiden, again continuouslysampling. Bottom: the input generating the output above, it wasconstrained to the range 0–20 nM/s

lower input bound (u = 0). In Fig. 5 (dashed lines), the inputpenalty has been increased from unity to 50 (while the statepenalty remains the identity matrix).

Whether the controlled trajectories of Fig. 5 present ahypercoagulatory risk is something of an open question:Rapid product formation ‘downstream’ in the cascade, thatis, the formation of fibrin from fibrinogen and activation offactor XIII – requires concentrations of thrombin less than2 nM [19], so it is not clear how important the exact tra-jectory of the thrombin curve is. Nevertheless, it may beimportant, and so we propose a more sophisticated approachbelow.

5.4 Hæmophilia A: ‘impulse’ controller

We turn to ‘moderate’ hæmophilia A, at which native levelsof the zymogen factor VIII are about 2.5% of normal levels.Now, if indeed we have a controller that can be switchedon precisely at the onset of a clotting event (perhaps by therelease of tissue factor), and can apply the control locally –assumptions under which we have been operating so far –then feedback linearisation is overkill: we can simply dumpin the healthy initial concentration of factor VIII in the firstcontroller time step, and then turn the controller off. Usingagain a control sampling rate of 2 Hz, this technique yieldsthe near-perfect results of Fig. 6. Here the non-zero inputat the first time step (not shown) is x(0)/T , where x(0) isthe desired initial concentration of factor VIII, 0.7 nM, and

2639© The Institution of Engineering and Technology 2012

Page 11: ISSN 1751-8644 Real-time control of human coagulation

www.ietdl.org

0 50 100 150 200 250 3000

10

20

30

40

50

60

70

80

90

time (s)

conc

entr

atio

n (n

M)

0 50 100 150 200 250 3000

2

4

6

8

10

12

14

16

18

20

time (s)

conc

entr

atio

n ra

te (

nM/s

)

[IIa] Controlled 2[IIa] Controlled 1[IIa] Uncontrolled[IIa] Desired

Fig. 5 Graphs showingTop: simulated control of thrombin concentration during a clottingevent in a patient with factor-V Leiden, this time using a discretecontroller sampling every 0.5 s. The dashed line shows a controllerwith a high input penalty (relative to the state penalty). Bottom:the discrete input, constrained to lie between 0 and 20 nM/s, thatgenerated the output above. Increasing the relative input penaltyavoids ‘clipping’ at the upper input bound, but degrades tracking

T is the controller time step, 0.5 s. An added boon of thistechnique is that it requires no sampling at all: the controllerwould not require sensors.

5.5 Hæmophilia A: step controller

Now, the simple, step controller discussed in Section 3.3above operates on similar principles, although it makes noassumptions as to the controller operating frequency. Apply-ing the simple learning algorithm of (17) with α = 1 × 10−4

and β = 1 × 10−5 yields the trajectory in Fig. 7. The con-stant input (not shown) of factor VIII is the very low rate of6.92 pM/s. The price we pay for the simplicity of this controlscheme is the overshoot on the back half of the trajectory.

5.6 Factor-V Leiden: step controller

Finally, we ask how well we can manage factor-V Leidenwith the step controller. In fact, we can do quite well. Choos-ing α = 0.05 and β = 0.01 in (17), the input will settle onu = 4.39 nM/s, at which rate the peak concentration and itsoccurrence can be very nearly matched (Fig. 8). The onlyquestion that remains is how significant the undershoot isfrom a clinical perspective.

2640© The Institution of Engineering and Technology 2012

0 50 100 150 200 250 3000

5

10

15

20

25

time (s)

conc

entra

tion

(nM

)

[IIa] Desired

[IIa] Controlled

[IIa] Uncontrolled

Fig. 6 Simulated control of thrombin in a patient with moderatehæmophilia A (see text), where the controller inputs the properinitial concentration of factor VIII at the first time step, then shutsoff

0 50 100 150 200 250 3000

5

10

15

20

25

time (s)

conc

entra

tion

(nM

)[IIa] Desired

[IIa] Controlled

[IIa] Uncontrolled

Fig. 7 Control of thrombin concentration in a hæmophiliac usinga single-step input at time t = 0The input is 6.92 pM/s

0 50 100 150 200 250 3000

10

20

30

40

50

60

70

80

90

time (s)

conc

entra

tion

(nM

)

[IIa] Desired[IIa] Uncontrolled[IIa] Controlled

Fig. 8 Control of thrombin concentration in a patient withfactor-V Leiden by a single-step input at time t = 0The input is 4.39 nM/s

IET Control Theory Appl., 2012, Vol. 6, Iss. 17, pp. 2630–2643doi: 10.1049/iet-cta.2011.0475

Page 12: ISSN 1751-8644 Real-time control of human coagulation

www.ietdl.org

6 Discussion

6.1 Controllability

We have shown via an algebraic criterion that our system ofODEs is not full-state linearisable. As a matter of fact, wecan say more. It follows from the method of transformingchemical equations into ODEs that every reversible reactionwill introduce into C two columns that are additive inversesof each other. Thus, while the rank of C is obviously upper-bounded by the number of uni-directional reactions, that is,the number of arrows in all our chemical equations, or againthe number of rate constants – since this is the number ofcolumns in C, it is also upper-bounded by the (generallysmaller) number of uni- or bidirectional reactions, that is,counting each pair of arrows as just one reaction; hence-forth simply called ‘reactions’. If this number falls short ofthe dimension of the state space, then full-state linearisa-tion requires the deficit to be covered by the control vectorfields – which, in the case of a single input, can provideadditional rank of at most one. So in chemical systems ofmass-action kinetics with no outflow, a single output, anda single input, a necessary (though insufficient) conditionfor full-state linearisation is that there be at least as manyreactions as state variables.

We can translate this result into a statement about the con-trollability of the system. The rank of the strong-accessibilitydistribution (at some point x0) is the dimension of the locallyaccessible manifold (from x0); or alternatively, the differ-ence between this rank and the dimension of the state, n,is the number of uncontrollable modes of the system. Sothe dimension of the locally accessible manifold is upper-bounded by the total number of reactions (again, countingbidirectional reactions just once). More formally:

Theorem 1: Consider a chemical system of mass-actionkinetics with mu unidirectional reactions, mb bidirectionalreactions, and n states (reactants). Under control with pinputs, the rank q of the strong-accessibility distribution iseverywhere less than or equal to the total number of reac-tions – each bidirectional reaction counting just once –, plusp. That is

q ≤ mu + mb + p ≤ n (19)

The proof is immediately from the discussion above. Thatthe result is global (rather than at some particular x0) followsfrom the ‘augmented matrix’ D of Section 3.2 being inde-pendent of the state. Note also that, although p is obviouslyincreased by the addition of an exogenous pharmaceutical,like heparin, this addition will likewise increase n, bothbecause heparin (say) was not part of the original system,and because it will introduce as well additional complexes(e.g. Hep:AT, Hep:IIa, etc.). It follows that, for an exoge-nous pharmaceutical to increase the controllability of thesystem, it must introduce more reactions than complexes.

Considering the present system, we find that without theheparin reactions there are 96 reactions and 98 state vari-ables. Factor-V Leiden knocks out some reactions and somestate variables, leaving 77 and 84, respectively. Now, we areobviously interested in pharmaceutical interventions in addi-tion to heparin (a whole host of pro-coagulants, for instance).However, it can now be claimed that in order even to standa chance of fully linearising the system with one of theseinterventions as the (sole) input, or equivalently of fully con-trolling the system, the drug must interact with the systemvia at least six more reactions than the number of variables

IET Control Theory Appl., 2012, Vol. 6, Iss. 17, pp. 2630–2643doi: 10.1049/iet-cta.2011.0475

that these reactions introduce. (We say six because the con-trol vector field may not be in the span of C and hencecover the remaining dimension; recall (15).) And indeed,heparin introduces only six reactions for its five new statevariables, yielding 83 and 89, respectively (see Table 3).So even though there are more unidirectional reactions (rateconstants, columns of C) than state variables (proteins, rowsof C), we know that there are uncontrollable modes evenwithout constructing C simply by noting that the number ofreactions falls short of the number of state variables.

Finally on the topic of full-state linearisation, it canbe shown that outflow of the blood factors – in particu-lar, heterogeneous outflow – can increase the rank of theaccessibility matrix, and hence of the strong-accessibilitydistribution [10]. For simplicity, the model as it standsneglects outflow, but in fact in vivo coagulation will per-force have some flow (both in and out) of clotting factors,and if it be not negligible over the time scale of interest,this may provide for greater controllability of the system.

6.2 Non-regularity

The coagulation system under application of heparin is, wehave said, non-regular; that is, the relative degree is not con-stant across the state space. What consequences ensue forcontrol?

It appears that the technique of allowing the system todrift away from the singularity has no adverse effects oncontrol. We describe now in detail why this should be so.The feedback non-linearity that shows up in the denominatorof the control law, (10), is in the present case, which hasstrict relative degree (q = 2)

LgLfh(x) = −kon[IIa](t) (20)

inducing a singularity precisely when concentrations ofthrombin (factor IIa) are zero. However, during a clottingevent, thrombin concentrations will always be greater thanzero, so the state is safely bounded away from the singu-larity. We could make this bound precise by choosing someδ in thrombin-concentration space below which we do notcare to regulate the thrombin trajectory. And the simulationshave demonstrated that in the concentration range of inter-est – viz., on the order of nanomolarity – the state is notnear enough to the singularity to introduce numerical errorsinto the control.

Now, on the other hand, some (too large) choice of δwould violate the initial-condition criteria too grossly for thecontroller to be able to recover. That it does recover at allis a consequence (presumably) of the error correction termsadded to the control law in (16). Again, the simulations havedemonstrated that the controller can indeed recover as lateas 60 seconds into the clotting event.

We saw in the Results section above that the uncon-strained (continuous) controller led to numerical errors inthe ODE simulation – which is why the data shown in Fig. 3were produced with input clamped at 100 nM/s. These errorsare indeed the consequence of the non-regularity of the sys-tem, but a peculiar consequence that could not arise in arealistic controller: since the heparin input rate was allowedus to assume negative values, the concentration of heparinitself as well as other state variables could be driven belowzero. And of course pushing thrombin levels through zeromeans operating the controller in the region of the singu-larity. Lower-bounding the input at zero, as we must in

2641© The Institution of Engineering and Technology 2012

Page 13: ISSN 1751-8644 Real-time control of human coagulation

www.ietdl.org

any realistic controller and as we do in the other simula-tions, eliminates this threat to the numerical stability of thecontroller.

Are there not more theoretically sound, less heuristical,bases for avoiding the singularity? There are, but these [12,13] involve variations on the following theme: compute therelative degree of the system at the singularity, and use thecorresponding control law (again (10), but with a higherstrict relative degree, q) in the region of the singularity. Theproblem with this type of technique for our model is that therelative degree at the singularity is higher than the ability ofour machine to compute iterated Lie derivatives; that is, weare back to the same, apparently insuperable, computationalobstacle that prevented direct computation of the rank of thestrong-accessibility distribution. Fortunately, the simulationspresented here evidently demonstrate the superfluity of thesemore sophisticated methods.

6.3 Controller merits and demerits

How precise does our control need to be, after all? Theanswer is not known, but there is some suggestion thatthe required amount of thrombin for clotting is much lessthan the actual peak concentration: Brummel et al. [19]have demonstrated in an in vitro study that less than 2 nMof thrombin is required for rapid product formation down-stream in the cascade. Certainly this suggests that step inputssuffice to steer hæmophilia-A and factor-V-Leiden patientsthrough safe clotting, given the qualitative matches of Figs. 7and 8. In case these matches are not sufficient, however,or more critically in case the upper bound on the inputneeds to be lowered, we propose in the following sectiona model-predictive approach to remedy the defects of thefeedback-linearising controller.

However, the step controllers have another enormousadvantage over the feedback-linearising control: they requireno sampling, only initiation at the inception of a clottingevent. There is not currently a method for measuring blood-protein concentrations in vivo in real time – certainly not atthe required 2 Hz sampling rate. The term L2

f h(x) from thecontrol law, (10), is a function of some 32 proteins. (Theother term in the control law, LgLfh(x), is a function only ofthrombin. And in fact, thrombin can indeed be measured inreal-time ex vivo; see [20].) We interpret our results on feed-back linearisation, then, as showing that if indeed a greaterdegree of accuracy is required for thrombin tracking, then itwould be fruitful to devise the appropriate sensors for real-time sampling, since they would make possible this suitabletechnique. Now, naïvely constraining the input vitiates theaccuracy of this technique (Figs. 4 and 5), hence the needfor the model-predictive controller.

On the other hand, two significant advantages of the feed-back controllers must be stressed. First of all, they require notraining. Patient-to-patient variation and model errors (e.g.in rate constants, which vary fairly widely from study tostudy) make pre-training on a model insufficient for thedevelopment of actual treatments, though it certainly pro-vides a good starting point. So in the limit, the training ofthe step-controller approaches the actual (barbarous) state ofthe art, namely testing the patient once a day and modifyingdoses accordingly. Second, and perhaps more significantly,the feedback controllers operate fairly robustly over a widerange of delays in turning on. So, for example, in the caseof constrained, discrete control (the most plausible scenario),turning on the controller a full minute after the release oftissue factor increases the overshoot by less than 5%, with

2642© The Institution of Engineering and Technology 2012

the rest of the trajectory tracking essentially identical (datanot shown). The issue here is how similar the state at (forexample) t = 60 s is to the desired initial conditions – thecloser, the more easily the controller can recover – and asis clear from the plots above, thrombin concentrations (pre-sumably, inter alia) do not increase significantly until about85 s.

One envisages, then, the following real-world scenario inwhich this type of controller would be particularly useful: anindividual suffering from factor-V Leiden cuts herself, and –within the first minute – applies to the site of injury a devicethat injects heparin, in response to the local concentration ofall (or some subset of) the other blood proteins.

6.4 Significance of model assumptions

Two model assumptions require elaboration. First, both thefeedback and step controllers show some sensitivity to thechoice of rate constants. An earlier version of the model,for instance, estimated both on- and off-rates for the heparinreactions, again setting the former to 0.1 nM/s but also fixingthe latter at 10 s−1, and achieved essentially perfect trackingwith the same input limits and sampling rate. As we havesaid repeatedly, reported rate constants in the blood clot-ting literature vary somewhat widely, so this considerationis quite relevant. However, even if we assume the affinitiesin [14] are correct, as in the present model, there is stillthe matter of our assumption of the on-rates (recall that theunknown heparin on- and off-rates were estimated from theirratio, fixing the on-rate in the middle of the reasonable phys-iological range at 0.1 nM/s and then computing the off-ratefrom their known ratio; see Section 4).

The effect on the controller turns out to be quite small.Similar tracking results (not shown) can be achieved by bothcontrollers for rate-constant values that span three orders ofmagnitude, with kon from 0.01–1 nM/s and koff set to main-tain the desired ratio. Values outside this range are unlikely,but there the step controller can match the thrombin peak’stiming and height only on pain of greatly distorting thesubsequent trajectory.

A related question is how sensitive the control is tounknown deviations in the rate constants. Given the heavydependence of the control law on the state, it will not be sur-prising that this sensitivity is indeed acute. Simulations (notshown for reasons of space) reveal that even minor perturba-tions of the model rate constants – leaving the control-lawconstants at their nominal values – degrades the trackingof thrombin intolerably. Nor is the step controller immunefrom this problem: a suitable step input was found by opti-misation over repeated runs of the model; no such luxuryis available for the patient, so if model and patient (‘plant’)differ, the step controller will issue in poor control.

Nevertheless, the degeneration of tracking can be elim-inated by increasing the gain on the proportional term ofthe feedback-linearisation controller. This is true for rate-constant deviations of up to 5%. (In the simulations, rateconstants were randomly perturbed by multiplicative Gaus-sian noise with unity mean and standard deviation 1/3 ×5%, keeping 99% of the rate constants within 5% of theiroriginal values.) The insensitivity of the model to highproportional-term gains is of itself interesting; we pursueit at length in a companion article.

The final model assumption adverted to at the beginningof this section is that of locality: the supply of unactivated(zymogen) blood factors was treated as limited. This choicewas made to conform to the model of [3] as well as to the

IET Control Theory Appl., 2012, Vol. 6, Iss. 17, pp. 2630–2643doi: 10.1049/iet-cta.2011.0475

Page 14: ISSN 1751-8644 Real-time control of human coagulation

www.ietdl.org

in vitro results of [16]. In reality, however, zymogens arereplenished by circulating blood, and activated proteins aresimilarly removed. On the other hand, the limited amountof lipids also restricts the amount of each zymogen that canever take part in the reaction, since nearly all of the clottingreactions take place on a phospholipid surface. Congruently,simulations (not shown) indicate that even if the zymogensare modelled as inexhaustible, the thrombin curve remainsqualitatively the same. We proceeded then with the localityassumption, in order to avoid the complexities of an addedcirculation model.

7 Conclusions

Regulation of many clotting disorders requires daily visitsto a doctor both for the administration of blood thickenersor thinners, and for assaying blood–protein concentrationsto determine dosages. In exceptional circumstances, con-tinuous intravenous administration is necessary. Neither ofthese alternatives is desirable, and both are in some sensea consequence of taking into account only the most prim-itive knowledge about human coagulation: certain drugsencourage clotting, whereas others discourage it.

This paper has proposed and demonstrated two con-trol techniques that are much more powerful in virtue ofexploiting a mathematical model of a large portion of thecoagulation cascade and the effects of the relevant pharma-ceuticals. Both controllers assume local application and theability to detect a clotting event (release of tissue factor);under which assumptions both have been shown capable ofregulating the time-varying concentrations of thrombin, thekey blood protein which alone of the proteins in this modeldetermines clot formation.

Each controller has its merits and demerits: the step con-troller is more critically dependent on the accuracy of themodel but requires no sensors; whereas the feedback con-troller is robust to model changes but requires real-time(of the order of 2 Hz) sampling of a number of proteinconcentrations (viz., all those involved in the feedbackterms). Feedback control is also (slightly) degraded by con-straining the input rate to lie within a feasible range. Weaddress both of these deficiencies of feedback linearisa-tion in a companion article (although not, unfortunately, atthe same time), demonstrating a feedback controller thatrequires observation of only one state variable, and a dif-ferent (model-predictive) controller, which overcomes someof the constraint-induced tracking degradation.

The significance of both the controllers presented in thispaper for clinical application would be greatly advanced byvarious improvements to the model. First, as lately noted,the rate constants for heparin interactions with the clot-ting process are not known; if a precise controller is tobe constructed along the lines proposed here, these mustbe determined. Second, a circulation model would dispensewith the somewhat dubious locality assumption, as well aspossibly afford more powerful control, since inflow and out-flow can increase the strict relative degree of the system(and hence the number of linearised variables that can becontrolled).

IET Control Theory Appl., 2012, Vol. 6, Iss. 17, pp. 2630–2643doi: 10.1049/iet-cta.2011.0475

Finally, this paper has made some useful general obser-vations about chemical kinetics and controllability. In par-ticular, we have shown that the dimension of the locallyaccessible manifold is upper-bounded by the number of reac-tions in a chemical system, where bi-directional reactionsare counted just once. The authors have not seen this resultelsewhere. We can say then, e.g. that since full-state lineari-sation requires that the dimension of the locally accessiblymanifold be equal to the dimension of the state, in chemicalsystems it requires that the number of chemicals not exceedthe number of reactions.

8 References

1 Makin, J.G., Narayanan, S.: ‘A hybrid-system model of human bloodclotting’. Technical Report, International Computer Science Institute,1947 Center Street, Berkeley, CA 94703, 2008

2 Mitchell, I., Bayen, A.M, Tomlin, C.J.: ‘Validating a Hamilton–Jacobiapproximation to hybrid system reachable sets’, in Benedetto M.D.D.,Sangiovanni-Vincentelli A.L. (Eds.): Fourth Int. Workshop HybridSystems: Computation and Control, HSCC 2001, Rome, Italy, 28–30March 2001, (LNCS, 2034), 2001, pp. 418–432

3 Bungay, S., Gentry, P., Gentry, R.: ‘A mathematical model oflipid mediated thrombin generation’, Math. Med. Biol., 2003, 20,pp. 105–129

4 Hockin, M.F., Cawthern, K.M., Kalafatis, M., Mann, K.G.: ‘A modeldescribing the inactivation of factor Va by APC: bond cleavage, frag-ment dissociation, and product inhibition’, Biochemistry, 1999, 38,pp. 6918–6934

5 Makin, J.G.: ‘A computational model of human blood clotting’. Tech-nical Report, International Computer Science Institute, 1947 CenterStreet, Berkeley, CA, 94703, 2008

6 Sastry, S.S.: ‘Nonlinear systems: analysis, stability, and control’(Springer–Verlag, New York, 1999)

7 Isidori, A.: ‘Nonlinear control systems’ (Springer, 1995, 3rd edn.)8 Adcock, D.M., Jensen, R., Johns, C.S.: ‘Coagulation handbook’ (Eso-

terix Coagulation, 2002)9 Khanin, M.A., Rakov1, D.V., Kogan, A.E.: ‘Mathematical model for

the blood coagulation prothrombin time test’, Thrombosis Res., 1998,89, pp. 227–232

10 Bastin, G, Lévine, J.: ‘On state accessibility in reaction systems’, IEEETrans. Autom. Control, 1993, 38, (5), pp. 733–742

11 Henson, M.A., Seborg, D.E.: ‘Feedback linearizing control’, in SeborgD.E., Henson M.A. (Eds.): Nonlinear Process Control, Prentice-Hall,Upper Saddle River, NJ, 1996, pp. 149–231

12 Tomlin, C.J., Sastry, S.S.: ‘Switching through singularities’. Proc. 36thConf. Decision and Control, 1997, pp. 1–6

13 Hauser, J., Sastry, S., Kokotovic, P.: ‘Nonlinear control via approx-imate input–output linearization: the ball and beam example’. Proc.28th Conf. Decision and Control, Tampa, Florida, 1989, pp. 392–398

14 Olson, S.T.: ‘Transient kinetics of heparin-catalyzed protease inactiva-tion by antithrombin III’, J. Biol. Chem., 1988, 263, (4), pp. 1698–1708

15 Egan, J., Kalafatis, M., Mann, K.G.: ‘The effect of Arg306 → Ala andArg506 → Gln substitutions in the inactivation of recombinant humanfactor VA by activated protein C and protein S’, Protein Sci., 1997,6, pp. 2016–2027

16 Butenas, S., van’t Veer, C., Mann, K.G.: ‘ “Normal” thrombin gener-ation’, Blood, 1999, 94, pp. 2169–2178

17 Fersht, A.: ‘Structure and mechanism in protein science’ (W.H.Freeman and Company, New York, 1999)

18 MATLAB. version 7.1 (R14SP3). Natick, Massachusetts: the Math-Works Inc., 2005

19 Brummel, K., Paradis, S., Butenas, S., Mann, K.: ‘Thrombin functionsduring tissue factor-induced blood coagulation’, Blood, 2002, 100, (1),pp. 148–152

20 Hemker, H.C., Giesen, P., AlDieri1, R., et al. ‘The calibrated auto-mated thrombogram(CAT): a universal routine test for hyper- andhypocoagulability’, Pathophysiol. Haemostasis Thrombosis, 2002, 32,pp. 249–253

2643© The Institution of Engineering and Technology 2012