J Pharmacokinet Pharmacodyn (2013) 40:639–650 DOI 10.1007/s10928-013-9336-y

ORIGINAL PAPER

Solution and implementation of distributed lifespan models Gilbert Koch • Johannes Schropp

Received: 25 July 2013 / Accepted: 8 October 2013 / Published online: 1 November 2013  Springer Science+Business Media New York 2013

Abstract We consider a population where every individual has a unique lifespan. After expiring of its lifespan the individual has to leave the population. A realistic approach to describe these lifespans is by a continuous distribution. Such distributed lifespan models (DLSMs) were introduced earlier in the indirect response context and consist of the mathematical convolution operator to describe the rate of change. Therefore, DLSMs could not directly be implemented in standard PKPD software. In this work we present the solution representation of DLSMs with and without a precursor population and an implementation strategy for DLSMs in ADAPT, NONMEM and MATLAB. We fit hemoglobin measurements from literature and investigate computational properties. Keywords Lifespan  Distribution  Convolution  Population

After expiration of its lifespan the individual has to leave the population and maybe maturate into another population. A realistic approach to describe lifespans is by a continuous distribution. Such models were introduced by Krzyzanski et al. [1] to PKPD modeling in the indirect response context in 2006. In [1, 2] these models where applied to describe the stimulation of reticulocytes. In [3, 4] simpler models where every individual has the same lifespan are used to describe erythropoiesis. Let T be a random variable describing the distribution of the lifespan s of individuals based on a probability density function (PDF) l(s). The only assumption is that negative lifespans are not possible, more precisely l(s) = 0 for s \ 0. The expectation of the lifespan distribution is denoted by Z1 T ¼ E½T  ¼ slðsÞds; 0

Introduction

the variance is

We consider the development of a population N(t) in time t where every individual has its own and unique lifespan s.

r2 ¼ Var½T  ¼

Electronic supplementary material The online version of this article (doi:10.1007/s10928-013-9336-y) contains supplementary material, which is available to authorized users. G. Koch (&) Department of Pharmaceutical Sciences, State University of New York at Buffalo, 403 Kapoor Hall, Buffalo, NY 14214, USA e-mail: [email protected] J. Schropp Department of Mathematics and Statistics, Universita¨t Konstanz, PO Box 195, 78457 Konstanz, Germany

Z1

s2 lðsÞds  T 2 ;

0

and the cumulative distribution function (CDF) is defined by LðsÞ ¼

Zs lðsÞds for

s  0:

0

In the special case of a point distribution (i.e. every individual has the same lifespan T) we have l(s) = d(s - T) where d is the Dirac delta function. A model describing the rate of change of a population reads

123

640

J Pharmacokinet Pharmacodyn (2013) 40:639–650

d NðtÞ ¼ kin ðtÞ  kout ðtÞ; dt

Nð0Þ ¼ N 0 ;

ð1Þ

where the production (birth) rate is described by kin(t) and kout(t) denotes the loss (death) rate. The initial amount of individuals in the population at time point t = 0 is denoted by N0. Please note that only zero order terms appear in (1). The lifespan distribution is included into (1) via the convolution operator Z1 kout ðtÞ ¼ ðkin  lÞðtÞ ¼ kin ðt  sÞlðsÞds; ð2Þ

Fig. 1 The upper scheme is the basic distributed lifespan model, DLSM (5), (6) and the lower scheme shows the distributed lifespan model with precursor, DPLSM (7), (8)

0

see [1] for derivation. Hence, to calculate the outflow kout(t) at a time point t an integral over the lifespan interval I ¼ ½0; 1 has to be solved. Note that in (2) the past t \ 0 of the production kin(t) plays an important role and has to be defined. In this work we admit a non-constant past for the production kin(t) for t \ 0 which is an extension of traditional PKPD modeling and covers more pharmacological situations. Please note that (2) reduces to kout(t) = kin(t - T) in case of a point distributed lifespan, see [1]. The initial value N(0) in (1) describes the amount of born and already died individuals in the past up to the starting time point t = 0 and is calculated by

from that population maturate into a follow-up population M(t). Mathematically, this means that kin and lN from (5), (6) have to be replaced by kin * lN and lM. Hence, for a DPLSM we additionally obtain Z1 d MðtÞ ¼ ðkin  lN ÞðtÞ  ðkin  lN Þðt  sÞlM ðsÞds dt ð7Þ 0

¼ ðkin  lN ÞðtÞ  ðkin  lN  lM ÞðtÞ; with Mð0Þ ¼

Z1

lM ðyÞ

0

Nð0Þ ¼

Z1

Z0 lðxÞ

kin ðsÞdsdx;

ð3Þ

see Appendix 1 for derivation. In case of a constant past of 0 the production kin(t) = kin for t B 0 the initial value (3) results in 0 Nð0Þ ¼ kin T;

ð4Þ

see Appendix 2 for a proof and compare [1]. Summarizing, a distributed lifespan model (DLSM) for a population N(t) with the lifespan PDF lN(s) reads d NðtÞ ¼ kin ðtÞ  ðkin  lN ÞðtÞ; dt

ð5Þ

with Nð0Þ ¼

0

lN ðxÞ

Z0

kin ðsÞdsdx:

ð6Þ

x

We call (5), (6) the rate of change formulation. Please note that in this formulation for every time point t the convolution over the lifespan interval I has to be calculated. In Fig. 1 (5), (6) is schematically visualized. Additionally, we consider DLSMs with a precursor population (DPLSM), as introduced in [1]. Then (5), (6) serves as a precursor population and individuals expiring

123

ðkin  lN ÞðsÞdsdy;

ð8Þ

y

see Appendix 3 for derivation of the initial value. Again for a constant past of the production rate

x

0

Z1

Z0

0 Mð0Þ ¼ kin TM ;

ð9Þ

holds. Please note that M(t) is independent from N(t) and that in (7) the loss term is a double integral. In Fig. 1 (7), (8) is schematically visualized. Standard PKPD software like ADAPT [5] or NONMEM [6] and even general engineering software like MATLAB [7] provide no internal function for the convolution operator. Therefore, the convolution operator has to be approximated by a method provided by the user or an equivalent model formulation is necessary. Krzyzanski et al. [1] presented an approximation of the convolution operator by introducing a large number of additional differential equations. In case of a special variant of the gamma distribution, Nichols et al. [8] constructed an equivalent differential equation system without the convolution but with constraints between the initial values and the model parameters. Freise et al. [2] reported the solution representation of a DLSM without a precursor and applied the less common software WINFUNFIT with additional subroutines. In this paper we explicitly calculate the equivalent integral solution representation of DLSMs without precursor (5), (6) and with precursor (7), (8). The obtained solution could then be simply approximated e.g. by the

J Pharmacokinet Pharmacodyn (2013) 40:639–650

641

Theoretical

Weibull 1 0.9

PDF l(τ) and CDF L(τ)

trapezoidal rule [9]. Therefore, a direct implementation of DLSMs in standard PKPD software is possible. We simulate artificial data to test the approximation of the solution and compare the computational effort with the original rate of change formulation and the approach presented in [1]. Finally, we apply our method to a published model [8] and re-fit hemoglobin (Hgb) data.

0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0

The main idea in the present paper is to exploit the special structure of Eqs. (5)–(8) and to calculate the solution in an explicit manner. Since (5), (6) do not depend on N(t), the integral solution representation of the rate of change formulation (5), (6) reads NðtÞ ¼

Z1

ð1  LN ðsÞÞkin ðt  sÞds;

ð10Þ

0

1  LN ðna hN Þ\a;

ð11Þ

for a given probability significance level a. Condition (11) bounds the probability of large lifespans, see Fig. 2. A reasonable value is a = 0.01. The approximation of (10) reads NðtÞ 

Zsend

ð1  LN ðsÞÞkin ðt  sÞds ð12Þ

0

 hN

na X

ð1  LN ðihN ÞÞkin ðt  ihN Þ:

i¼1

For (7), (8) we obtain the integral solution representation MðtÞ ¼

Z1 0

ð1  LM ðxÞÞ

Z1

kin ðt  x  sÞlN ðsÞdsdx;

ð13Þ

0

see Appendix 6 and e.g. the Riemann sum of (13) reads

4

nh=T

6

Lifespan τ

nα h

8

10

Fig. 2 The solid line is the PDF l(s) and the dashed line is the CDF L(s) of the Weibull distribution with expectation T = 5 and variance r2 = 1. The number of partitions for the lifespan interval [0, T] for a step size h is denoted by n. The minimal number of partitions for the whole lifespan interval up to a significance level a based on condition (11) is denoted by na

0

see Appendix 4 for the proof or the alternative derivation in [2]. The remaining integral over the lifespan interval I in formulation (10) could be simply approximated by standard integration techniques such as the trapezoidal rule (see Appendix 5) or Simpson’s method, see [9]. For illustration purpose, we present the simplest possible approximation (Riemann sum, see [10]) for the realistic lifespan interval eI ¼ ½0; send  where send is a natural bound on the lifespan. Let n be a given number of partitions for the interval [0,T] and hN ¼ Tn the step size. The final number of partitions na [ n is the minimal value which fulfills the condition

2

MðtÞ hN hM

nb X na X

ð1  LM ð jhM ÞÞ

j¼1 i¼1

ð14Þ

 kin ðt  jhM  ihN ÞlN ðihN Þ:

Methods In this work we consider three typical examples of continuous distributions, namely Weibull (WBD), gamma (GAD) and lognormal (LND). The WBD is generally applied to describe the lifespan of individuals in statistics [11]. In [12] a combination of two WBDs was presented to describe human mortality and applied by Korell et al. [13] to model red blood cell (RBC) survival. The GAD is e.g. used to describe the lifespan of reticulocytes [1] and also RBCs [8]. The LND is commonly used in clinical PKPD population analysis to describe the distribution of individual model parameter. Each of these distributions consist of two parameters, denoted by a and b, see Table 1. For WBD the gamma function CðsÞ is necessary to calculate the expectation, for LND the Gauss error function erf(s) is needed for the CDF and for GAD the incomplete gamma function c(s) as well as CðsÞ are necessary, see Table 1. These additional functions could e.g. be implemented with the algorithms presented in [14]. ADAPT 5 is a purely Fortran based PKPD software and could simply be extended with additional subroutines and functions provided by the user, see Appendix 7. In [14] Fortran subroutines for cðsÞ; CðsÞ and erf(s) could be found. In ADAPT the tolerance for the simplex algorithm was set to Reqmin = 1E-8 and for the LSODA solver to

123

642

J Pharmacokinet Pharmacodyn (2013) 40:639–650

Table 1 The PDF, CDF, expectation and variance of the Weibull, lognormal and gamma distribution

PDF l(x, a, b) CDF L(x, a, b) Expectation T Variance r2

Weibull

Lognormal

   a  a x a1 exp  bx b b

  ðlogðxÞbÞ2 1 pffiffiffiffiffiffiffi exp  2 2 2a x 2pa   logðxÞb p ffiffi 0:5 þ 0:5erf 2a   2 exp b þ a2

  a  1  exp  bx   bC 1 þ 1a   b2 C 1 þ 2a  T 2

Gamma ba a1 CðaÞ x

expðbxÞ

cða; bxÞ CðaÞ a b

a ðexpða2 Þ  1Þ expð2b þ a2 Þ b2 R1 Rx The gamma function is CðxÞ ¼ 0 sx1 expðsÞds; the incomplete gamma function is cðs; xÞ ¼ 0 ys1 expðyÞdy and the Gauss error function Rx 2 2 is erfðxÞ ¼ pffiffip 0 expðs Þds

RTOL = ATOL = 1E-8. In NONMEM 7.2 the $PRED environment was applied and Verbatim Code was used to implement additional functionality. The FOCE method was used for estimation. In both software the trapezoidal rule for the DLSM was implemented. In MATLAB (R2012a) the integral solution representation (10) and the rate of change formulation (5), (6) was implemented. We applied the trapezoidal rule and the MATLAB internal QUAD integrator (adaptive Simpson’s method) to integrate (10). Finally, we solved the rate of change formulation (5), (6) where the convolution was calculated with QUAD and the rate of change was solved with ODE45. The gamma, incomplete gamma and Gauss error function were calculated with the internal commands gamma, gammainc and erf. The optimization function LSQCURVEFIT (Levenberg–Marquardt algorithm) and additionally the optimization method FMINSEARCH (Nelder–Mead simplex) were used, see Example 1 in Results section for more details. The multiple dosing was coded as an explicit function because the drug concentration has to be calculated at different time points in the past, compare (12). In all optimizations an unweighted least squares approach was applied. All source codes are available as supplemental material and at the homepage of the authors. Exemplarily, in Appendix 7 we present the modified ADAPT subroutines and the additional functions of the implementation of the trapezoidal rule for the integral solution representation (10).

Results In this section we demonstrate the power of the presented method. In the first example we produce artificial data in MATLAB and fit those in ADAPT and NONMEM. Additionally, we investigate the computational effort of the presented approximation of the integral solution, the rate of change formulation and the approach presented in [1]. In

123

the second example we apply our method to a model for Hgb measurements in patients with renal disease from [8]. Example 1

PKPD test problem.

To test the presented formulations we construct an artificial PKPD experiment. The PK is described by the Bateman function cðtÞ ¼

doseka ðexpðke tÞ  expðka tÞÞ; Vðka  ke Þ

ð15Þ

where the parameter are set to ka = 0.7, ke = 0.5 and V = 1. A multiple dosing situation with the time points t = 0 and 7 is considered and four different dosing groups with dose 2 f10; 50; 100; 500g are constructed. The production kin(t) of the population is set in the context of indirect response and reads ( 0  kin  for t\0; kin ðtÞ ¼ ð16Þ Smax cðtÞ 0 kin 1 þ SC50 þcðtÞ for t  0: We apply (16) to (4), (5) and obtain the test model d NðtÞ ¼ kin ðtÞ  ðkin  lN ÞðtÞ; dt

0 Nð0Þ ¼ kin TN ;

ð17Þ

with the model parameter vector   0 h ¼ Smax ; SC50 ; kin ; a; b ; where (a, b) are the distribution related parameter. The WBD is used and the significance level is set to a = 0.01. We investigate three different scenarios. In scenario I, unperturbed data at times 0, 1, 2,…,25 with the parameters h = (0.5, 10, 1, 5.7974, 5.3999) is simulated. In II, this data is moderately perturbed by an additive distributed error with N ð0; 0:04Þ: Finally, we strongly perturbed scenario I by an additive distributed error with N ð0; 0:25Þ; denoted as III. The data is produced with the solution representation (10) where the integrator QUAD is used. The initial estimates h = (0.7, 5, 0.8, 3.88755, 6.6314) are roughly guessed based on the scenario III data. In Fig. 3 the three different produced data scenarios and the corresponding fits are presented. In Table 2 the fitting

J Pharmacokinet Pharmacodyn (2013) 40:639–650

PK

4

Response (Unperturbed) 8 7.5

2

10

7 6.5

N(t)

Drug concentration c(t)

10

0

10

6 5.5

−2

10

5 4.5

−4

10

0

5

10

15

20

4

25

0

5

7.5

7.5

7

7

6.5

6.5

N(t)

8

6

5.5

5

5

4.5

4.5 0

5

10

15

20

15

20

25

20

25

6

5.5

4

10

Response (Strong)

Response (Moderate) 8

N(t)

Fig. 3 Fit of the artificially produced data with the approximation (trapezoidal rule with n = 500) of the solution (10). The upper left panel shows the PK for dose 2 f10; 50; 100; 500g: The upper right panel is the response N(t) of the DLSM fitted to the unperturbed scenario I. The dose = 10 data is denoted by x, the dose = 50 data by squares, the dose = 100 by plus and circles denotes the dose = 500 data. The lower left panel is the fit of the moderately perturbed data II and the lower right panel shows the fit of the strongly perturbed data III

643

4

25

0

5

10

15

Time t

Table 2 Parameter estimates of the approximation (trapezoidal rule with n = 500) of the DLSM for scenarios I (unperturbed), II (moderate perturbed) and III (strong perturbed) obtain from the ADAPT implementation

Parameters

Initial estimates

Scenario I

Scenario II

Scenario III

Model parameters Smax (CV%)

0.7

0.50 (0.32E-2)

0.50 (3.16)

0.494 (7.80)

SC50 (CV%)

5

9.99 (0.11E-1)

9.12 (11.4)

8.01 (28.3)

0 (CV%) kin

0.8

1.00 (0.43E-2)

1.02 (4.21)

1.06 (10.3)

a (CV%)

3.88

5.77 (0.23E-2)

6.36 (3.09)

6.21 (7.01)

b (CV%)

6.63

5.40 (0.63E-2)

5.24 (7.51)

5.04 (17.3)

6 3

5.00 (0.64E-2) 1.01 (0.99E-2)

4.87 (7.65) 0.80 (10.9)

4.68 (17.6) 0.77 (25.6)

Secondary parameters TN (CV%) r2N

(CV%)

Fitting information SofS SofS objective function value, Iter number of iterations of the simplex algorithm

0.2633E-5

3.26162

20.4013

Iter

165

151

100

Time

39 s

33 s

26 s

results of the three different scenarios from the ADAPT implementation of the trapezoidal rule with n = 500 are listed. The NONMEM results are comparable. To get an understanding of the performance of various formulations, three different groups of methods are compared in MATLAB: (i)

Approximation of (10) with the trapezoidal rule or QUAD

(ii)

Approximation of the rate of change formulation (5), (6) (iii) Approximation due to Krzyzanski et al. [1]

The unperturbed scenario I data was used and therefore, in a successful fit the sum of squares is zero. The results are presented in Table 3 in detail. The outcome is that methods of type (i) are extremely superior to those of types (ii) and (iii) in regard of precision and computational effort.

123

644

J Pharmacokinet Pharmacodyn (2013) 40:639–650

Table 3 Numerical performance of the different MATLAB implementations for the scenario I data Formulations

TOL (ODE)

n

Time

Solution Trapeza

100

18 s

7

0.000011

Solution Trapezc

100

5 min 25 s

500d

0.000012

Solution Trapeza

500

1 min 25 s

7

0.000003

25 min

d

0.000003

7

0.000000

Solution Trapez

TOL (QUAD)

c

500

Iter

500

SofS

Solution QUADa

1E-8

37 s

Solution QUADc

1E-8

11 min

500d

1E-8

1E-8

11 min

6

Rate of changea

1E-10

1E-10

3h

27

0.039566

Rate of changea

1E-11

1E-11

16 h

59

0.000000

Rate of change

Rate of change

a

b

0.000000 23.34637

1E-8

1E-8

1 h 11 min

26

0.000153

Rate of changec Method from [1]a

1E-8 1E-8

1E-8

3 h 10 min 1 h 15 min

500d 10

0.000000 2.202447

Method from [1]a

1E-8

100

3 h 10 min

8

2.202672

Method from [1]a

1E-10

50

5 h 40 min

27

1.348316

Method from [1]b

1E-8

50

5 h 30 min

31

0.006383

Method from [1]b

1E-8

100

26 h 30 min

51

0.000000

Method from [1]c

1E-8

50

10 h 30 min

414

0.000000

50

SofS objective function value, Iter number of iterations of the optimization algorithm, TOL (ODE) tolerance used in ODE45 for RTOL = ATOL, TOL (QUAD) tolerance used for the integrator QUAD, n number of partitions of the lifespan interval up to the expectation T in case of the trapezoidal rule or n denotes the number of additional differential equations for the approach from Krzyzanski et al. [1] a

Levenberg–Marquardt with forward difference quotient

b

Levenberg–Marquardt with central difference quotient

c

Simplex algorithm

d

Maximum number of iterations achieved

Additionally, we remark that the use of the GAD or LND leads to similar performance results.

by the GAD (see Table 1) of second order (a = 2). Hence, we have the PDF lM ðsÞ ¼ b2 s expðbsÞ;

Example 2 Stimulation of Hgb by recombinant human erythropoietin (EPO). EPO is a glycoprotein hormone that stimulates RBCs and is primarily produced by kidneys. Hgb is the ironcontaining oxygen-transport metal protein in these RBCs. In renal disease, endogenously produced EPO is insufficient to maintain normal Hgb concentration. Therefore, recombinant human EPO (rHuEPO) is administered to patients to produce the desired concentration of Hgb. Nichols et al. [8] presented a rich data set of Hgb measurements from renal disease patients over 460 days with about three measurements a week. At every measurement time point rHuEPO was administered to patients. The amount of administered rHuEPO is known but drug concentration profiles were not available. For this example we digitized the data from patient 1. The authors modeled the data by a DLSM with precursor of the form (5)–(8) where the precursor N(t) describes hematopoietic stem cells which maturate after a constant lifespan TN into RBCs denoted by M(t). The lifespan of the RBCs was described

123

ð18Þ

and the CDF could be explicitly calculated and reads LM ðsÞ ¼ 1  ð1 þ bsÞ expðsbÞ:

ð19Þ

EPO concentration in plasma is the sum of endogenous EPO Een, a constant value, and exogenous rHuEPO E(t). The PK of rHuEPO is described by nonlinear elimination d Vmax EðtÞ EðtÞ ¼  ; Eð0Þ ¼ 0; dt Km þ EðtÞ and the total concentration of EPO reads cðtÞ ¼ EðtÞ þ Een : The stimulatory effect is kin ðtÞ ¼

Smax cðtÞ SC50 þ cðtÞ

for

t  0;

Een 0 and kin ðtÞ ¼ kin ¼ SCSmax for t \ 0. The final model from 50 þEen [8] reads

d MðtÞ ¼ kin ðt  TN Þ  ðkin  lM Þðt  TN Þ; dt

ð20Þ

J Pharmacokinet Pharmacodyn (2013) 40:639–650

645

Hemoglobin H(t) (g/dL)

15

Table 4 Parameter estimates of the approximation (trapezoidal rule [n = 500]) of the hemoglobin model (20)

14

Parameters

Units

Estimates

Vmax

IU/day

6,890a

Km

IU

1,042a

13

Model parameters 12

11

10

9 0

50

100

150

200

250

300

350

400

450

500

Time t (day)

Fig. 4 Fit of the hemoglobin data with the approximation (trapezoidal rule with n = 500) of the solution (22) of the distributed lifespan model with precursor (20)

HðtÞ ¼ KH MðtÞ;

ð21Þ

where M(t) describes the RBC and H(t) the Hgb concentration. The value KH = 29.5 g/dL is the average amount of Hgb per RBC. The model parameters are h ¼ ðVmax ; Km ; Hen ; Smax ; SC50 ; TN ; bÞ: where Hen is the endogenous level of Hgb due to Een, see [8]. In [8] a three dimensional ordinary differential equation system for (20) based on (18) was constructed with constraints between initial values and model parameter. This reformulation was applied in [8] to fit the Hgb data. Unfortunately, in their reformulation of (20) the initial value is only formulated as a constraint between the initial values of the additional ODEs and not explicitly given. In DLSMs with known inflow for t B 0, according to (7)–(9) there is no freedom to choose 0 M(0). Therefore, due to (9) we have M(0) = kin TM. In this work we apply the solution representation (13) of the precursor model (7), (8) to (20) and obtain Z1 MðtÞ ¼ ð1  LM ðsÞÞkin ðt  TN  sÞds; ð22Þ 0

where LM(s) is given by (19). The solution (22) is approximated by the trapezoidal rule with n = 500 and a = 0.01. In addition, we apply the PK averaging technique presented in [8] because c(t) has an impulse-like behavior (186 doses over 460 days with rapid clearance). Therefore, Nichols et al. suggested to average the concentration in certain time intervals by   Smax h Km Een kin ðdi Þ ¼ Km  SC50  Ti Vmax SC50 þ Een   SC50 þ Een þ di  ln SC50 þ Een   i Een Smax Een þ 1 ; ð23Þ di þ SC50 þ Een SC50 þ Een

Hen (CV%)

g/dL

10.26 (1.97)

Smax

cell/day

0.207E-1a

SC50 (CV%)

IU

86.71 (83.1)

TN (CV%) b (CV%)

day

6.98 (10.0) 0.605E-1 (10.0)

TM (CV%)

day

33.1 (10.0)

r2M

day2

547 (20.0)

Secondary parameters (CV%)

Fitting information

a

SofS

68.84

Iter

129

fixed to the values presented in [8]

where Ti is a time interval and di the dose at the time point ti. However, in the following some discussion is necessary. In [8] it is not obvious how Ti was set in (23). Here we choose the interval length Ti for a given dose di as the difference between the time point of the next dose di?1 and the time point of di. Also the PK parameters (Vmax,Km) (no PK data available) are estimated just from the Hgb PD data. This could lead to an over-parameterization of the model. Therefore, we fix Vmax, Km and also Smax to the values presented in the original work. In Fig. 4 the fit of the approximation of (22) with the trapezoidal rule (n = 500) is presented and in Table 4 the parameter estimates of (SC50, Hen, TN, b) are shown from the ADAPT implementation. The fit and the estimates do not exactly coincide due to the differences with the setting of the initial value M(0). Unfortunately, in [8] no derivation of their reformulation of (20) is presented and no information about the optimization process is given.

Discussion Previously published population models with a distributed lifespan (see [1]) based on the convolution operator were mathematically observed. The appearance of the convolution in the rate of change formulation makes an implementation extremely difficult in standard PKPD software. In the presented work we calculated the integral solution representation of DLSMs with and without a precursor population for arbitrary distributions and approximated these representations e.g. by the trapezoidal rule. Therefore, these models could now be simply implemented

123

646

J Pharmacokinet Pharmacodyn (2013) 40:639–650

in every PKPD software. In addition, the presented results are formulated in the general case of a non-constant past for the production kin(t). Further we investigated computational properties of different model formulations. We demonstrated that approximating the calculated solution representation instead of solving the rate of change formulation or the approximation presented in [1], results in a tremendous computational time advantage. Moreover, it was observed that our approach avoids numerical optimization problems. Finally, we applied our method to a published Hgb model and data from [8]. For the sake of completeness we mention that if the production kin(t) is not explicitly given, the advantage of our method will decrease because an additional ODE has to be simultaneously solved. This leads to a relevant increase in computational time. To overcome this issue other approximation techniques are necessary. For the special case of lifespan models with a point distribution it is already proven that an approximation with transit compartments exists, see [15, 16]. However, for lifespan models with an arbitrary continuous distribution an approximation by transit compartments is topic of our next research.

Then lim NðtÞ ¼ 0

t!1

, Nð0Þ ¼

Z1

lN ðxÞ

Z1

0

kf in ðs  xÞdsdx:

0

Using (24) Nð0Þ ¼

¼

Z1

lN ðxÞ

Zx

0

0

Z1

Z0

lN ðxÞ

kin ðs  xÞdsdx

kin ðsÞdsdx:

x

0

Appendix 2 0 We consider a constant past kin(t) = kin for t B 0 and obtain for the initial values

Nð0Þ ¼

Acknowledgments The present project is supported by the National Research Fund, Luxembourg, and cofunded under the Marie Curie Actions of the European Commission (FP7-COFUND).

Z1

Z0

lN ðxÞ

0 kin dsdx

x

0 0 ¼ kin

Z1

0 lN ðxÞxdx ¼ kin TN ;

0

Appendices

and

Appendix 1 Mð0Þ ¼ To estimate the amount of individuals at the initial population state N(0) for the DLSM (5), (6) we define the production term kin ðtÞ t  0; f kin ðtÞ ¼ ð24Þ 0 t [ 0: Then the appropriate initial value N(0) is determined by the claim limt!1 NðtÞ ¼ 0: Substituting (24) into (5) results in d NðtÞ ¼ kf in ðtÞ  dt

Z1

kf in ðt  xÞlN ðxÞdx:

ð25Þ

¼

¼

Z1

lM ðyÞ

Z0

0

y

Z1

Z0

lM ðyÞ

0

y

Z1

Z0

lM ðyÞ

 0  kin  lN ðsÞdsdy

0 kin

Z1

lN ðxÞdxdsdy

0 0 0 kin dsdy ¼ kin TM :

y

0

Appendix 3

0

To estimate the amount M(0) of the follow-up population for the DPLSM (7), (8) a similar approach as in Appendix 1 is applied. We define the production

The general solution of (25) reads NðtÞ ¼ Nð0Þ þ

¼ Nð0Þ 

Zt 0 Z1 0

123

kf in ðsÞds 

Z t Z1 0

lN ðxÞ

Zt 0

kf in ðs  xÞlN ðxÞdxds

0

kf in ðs  xÞdsdx:

kf in ðtÞ ¼



ðkin  lÞðtÞ 0

t  0; t [ 0;

ð26Þ

and use the condition limt!1 MðtÞ ¼ 0 to determine M(0). Substituting (26) into (7) results in

J Pharmacokinet Pharmacodyn (2013) 40:639–650

MðtÞ ¼ Mð0Þ þ

¼ Mð0Þ 

Zt

kf in ðsÞds 

0 Z1

Z t Z1 0

lM ðyÞ

0

Zt

647

Appendix 5 kf in ðs  yÞlM ðyÞdyds Generally, for an arbitrary continuous function f, the trapezoidal rule reads

0

kf in ðs  yÞdsdy;

Zsend

0

f ðsÞds 

s0

and therefore, Mð0Þ ¼

¼

Z1

lM ðyÞ

Z1

0

0

Z1

Z0

lM ðyÞ

with the equidistant partition sj = s0 ? (j - 1)h with h ¼ for j = 1,…,m ? 1 of the lifespan interval [s0, send].

kf in ðs  yÞdsdy ð27Þ ðkin  lN ÞðsÞdsdy:

Appendix 4

Appendix 6 Differentiation of (13) with respect to time t gives Z1 Z1 d 0 MðtÞ ¼ ð1  LM ðxÞÞ kin ðt  x  sÞlN ðsÞdsdx dt 0

Differentiation of (10) with respect to time t gives d NðtÞ ¼ dt

Z1

¼

0

Z1

lN ðsÞ

0

ð1 

0 LN ðsÞÞkin ðt

ð28Þ

send s0 m

y

0

m hX ð f ðsiþ1 Þ þ f ðsi ÞÞ; 2 i¼1

Z1

0

ð1  LM ðxÞÞkin ðt  x  sÞdxds:

0

ð29Þ

 sÞds

0

¼ ½ð1  LN ðsÞÞkin ðt  sÞs¼1 s¼0 Z1  ðlN ðsÞÞðkin ðt  sÞÞds

The inner integral in (29) could be written as Z1

0

ð1  LM ðxÞÞkin ðt  x  sÞdx

0 0

¼ kin ðtÞ 

Z1

lN ðsÞkin ðt  sÞds

0

With (10) we obtain for the initial value Nð0Þ ¼

2 ¼ 4ð1  LN ðsÞÞ

Zs

Z1

ðlN ðsÞÞ

0

¼

0

kin ðt  x  sÞlM ðxÞdx:

ð31Þ

lN ðsÞ

Zs

Substitution of (31) into (29) gives 3s¼1

kin ðxÞdx5

0

Z1

¼ kin ðt  sÞ 

Z1 0

ð1  LN ðsÞÞkin ðsÞds

0



ð30Þ

0

¼ kin ðtÞ  ðkin  lN ÞðtÞ:

Z1

¼ ½ð1  LM ðxÞÞkin ðt  x  sÞx¼1 x¼0 Z1  kin ðt  x  sÞlM ðxÞdx

d MðtÞ ¼ dt

Z1 0

s¼0

 kin ðxÞdxds

lN ðsÞkin ðt  sÞds

Z1 0

lN ðsÞ

Z1

kin ðt  x  sÞlM ðxÞdxds

0

0

Z0 s

¼ ðkin  lN ÞðtÞ  kin ðxÞdxds:

Z1

lN ðsÞððkin  lM Þðt  sÞÞds

0

¼ ðkin  lN ÞðtÞ  ðkin  lN  lM ÞðtÞ:

123

648

J Pharmacokinet Pharmacodyn (2013) 40:639–650

For the initial value we obtain Mð0Þ ¼

Z1

ð1  LM ðxÞÞ

Z1

0

Z x Z1

¼ 4ð1  LM ðxÞÞ

0

þ

Z1

lM ðxÞ

0

¼

¼

kin ðx  sÞlN ðsÞdsdx

0

2

Z1

Z x Z1 0

lM ðxÞ

Zx

kin ðy  sÞlN ðsÞdsdy5

0

x¼0

kin ðy  sÞlN ðsÞdsdydx

ðkin  lN ÞðyÞdydx

0

0

Z0

0

3x¼1

0

Z1

lM ðxÞ

CDF L(x) of the WBD:

ðkin  lN ÞðyÞdydx:

x

Appendix 7 In the following we present the implementation of the trapezoidal rule for (10). Only additional functions and modified ADAPT subroutines are presented. For the complete code see the supplemental material. Logarithm of the gamma function CðxÞ based on [14]:

123

Multiple dosing function for c(t):

J Pharmacokinet Pharmacodyn (2013) 40:639–650

Production kin(t):

649

The ADAPT subroutine OUTPUT calls the trapezoidal rule for the different dosing groups.

Integrand in (10):

Trapezoidal rule:

123

650

References 1. Krzyzanski W, Woo S, Jusko WJ (2006) Pharmacodynamic models for agents that alter production of natural cells with various distributions of lifespans. J Pharmacokinet Pharmacodyn 33(2):125–166

123

J Pharmacokinet Pharmacodyn (2013) 40:639–650 2. Freise K, Widness J, Schmidt R, Veng-Pedersen P (2008) Modeling time variant distributions of cellular lifespans: increases in circulating reticulocyte lifespans following double phlebotomies in sheep. J Pharmacokinet Pharmacodyn 35(3):285–323 3. Ramakrishnan R, Cheung WK, Wacholtz MC, Minton N, Jusko WJ (2004) Pharmacokinetic and pharmacodynamic modeling of recombinant human erythropoietin after single and multiple doses in healthy volunteers. J Clin Pharmacol 44:991–1002 4. Ait-Oudhia S, Scherrmann JM, Krzyzanski W (2010) Simultaneous pharmacokinetics/pharmacodynamics modeling of recombinant human erythropoietin upon multiple intravenous dosing in rats. J Pharmacol Exp Ther 334:897–910 5. D’Argenio DZ, Schumitzky A, Wang X (2009) ADAPT 5 user’s guide: pharmacokinetic/pharmacodynamic systems analysis software. Biomedical Simulations Resource, Los Angeles 6. Beal S, Sheiner LB, Boeckmann A, Bauer RJ (2009) NONMEM user’s guides. Icon Development Solutions, Ellicott City 7. MATLAB Release (2012) The MathWorks, Inc., Natick 8. Nichols B, Shrestha RP, Horowitz J, Hollot CV, Germain MJ, Gaweda AE, Chait Y (2011) Simplification of an erythropoiesis model for design of anemia management protocols in end stage renal disease. In: 33rd Annual international conference of the IEEE EMBS Boston, Massachusetts, USA 9. Atkinson K (1989) An introduction to numerical analysis, 2nd edn. Wiley, New York 10. Thomas GB, Finney RL (1996) Calculus and analytic geometry, 9th edn. Addison-Wesley, Reading 11. Marshall AW, Olkin I (2007) Life distributions. Springer, New York 12. Bebbinton M, Lai CD, Zitikis R (2007) Modeling human mortality using mixtures of bathtub shaped failure distributions. J Theor Biol 245(3):528–538 13. Korell J, Coulter CV, Duffull SB (2011) A statistical model for red blood cell survival. J Theor Biol 268(1):39–49 14. Press WH, Teukolsky SA, Vetterling WT, Flannery BP (1992) Numerical recipes in Fortran 77, the art of scientific computing, 2nd edn. Cambridge University Press, New York 15. Krzyzanski W (2011) Interpretation of transit compartments pharmacodynamic models as lifespan based indirect response models. J Pharmacokinet Pharmacodyn 38(2):179–204 16. Koch G, Schropp J (2012) General relationship between transit compartments and lifespan models. J Pharmacokinet Pharmacodyn 39(4):343–355

Solution and implementation of distributed lifespan models.

We consider a population where every individual has a unique lifespan. After expiring of its lifespan the individual has to leave the population. A re...
1MB Sizes 0 Downloads 0 Views