Sie sind auf Seite 1von 11

Pediatric Anesthesia ISSN 1155-5645

EDUCATIONAL REVIEW

An introduction to physiologically-based pharmacokinetic


models
Richard N. Upton, David J.R. Foster & Ahmad Y. Abuhewla
Australian Centre for Pharmacometrics and Sansom Institute, School of Pharmacy and Medical Sciences, University of South Australia,
Adelaide, SA, Australia

Keywords Summary
pharmacokinetics; physiology; regional
blood flow; anesthesia; pediatrics; computer Physiologically-based pharmacokinetic (PBPK) models represent drug
simulation kinetics in one or more ‘real’ organs (and hence require submodels of organs/
tissues) and they describe ‘whole-body’ kinetics by joining together submod-
Correspondence els with drug transport by blood flow as dictated by anatomy. They attempt
Prof. Richard N. Upton, Australian Centre for
to reproduce ‘measureable’ physiological and/or pharmacokinetic processes
Pharmacometrics, School of Pharmacy and
rather than more abstract rate constants and volumes. PBPK models may be
Medical Sciences, University of South
Australia, Adelaide, SA 5000, Australia built using a ‘bottom-up’ approach, where parameters are chosen from first
Email: richard.upton@unisa.edu.au principles, literature, or in vitro data as opposed to a ‘top-down’ approach,
where all parameters are estimated from data. The basic principles of PBPK
Section Editor: Brian Anderson models are described, focusing on the equations for three individual organs: a
single flow-limited compartment describing distribution only, a membrane-
Accepted 24 July 2016
limited compartment describing distribution, and a single flow-limited com-
partment with elimination. These organ models are linked to make a basic
doi:10.1111/pan.12995
three-compartment physiological model of the whole body. PBPK models are
particularly suited to scaling kinetics across body size (e.g., adult to neonate)
and species (e.g., animal to first-in-man) as physiology and pharmacology can
be represented by independent parameters. Maturation models can be incor-
porated as for compartmental models. PBPK models are now available in
commercial software packages, and are perhaps now more accessible than
ever. Alternatively, even complex PBPK models can be represented in generic
differential equation-solving software using the simple principles described
here. The relative ease of constructing the code for PBPK models belies the
most difficult aspect of their implementation—collecting, collating, and justi-
fying the data used to parameterize the model.

does a hyper- or hypodynamic circulation mean for drug


Introduction
disposition and effect? How should the dose regimen be
Anesthesiologists, as ‘applied pharmacologists’, have modified? These questions are not readily answered
long had an interest in pharmacokinetic (and pharmaco- using compartmental models.
dynamic) models. The majority of pharmacokinetic There has been a long-standing interest, particularly
models in the literature are empirical mammillary com- by anesthesiologists, in physiologically-based pharma-
partmental (‘compartmental’) models. While these have cokinetic (PBPK) models. PBPK models represent drug
been reliable servants in many fields of pharmacology, kinetics in one or more ‘real’ organs (and hence require
many anesthesiologists find the connection with patient a submodel of the organ/tissue). They describe ‘whole-
physiology (arguably at its most variable in the pre-, body’ kinetics by joining together organ submodels.
peri- and postoperative setting) is not well represented. Drug transport is by blood flow as dictated by anatomy.
For example, an anesthesiologist may ask why do arte- They attempt to reproduce real, measureable physiologi-
rial and venous blood concentrations have different cal and/or pharmacokinetic processes rather than more
time-courses, and what does the difference mean? What abstract rate constants and volumes. Model parameters

© 2016 John Wiley & Sons Ltd 1


Introduction to PBPK models R.N. Upton et al.

may be ‘intrinsic’ covariates (e.g., weight) rather than a patients described by a one-compartment model and
posthoc explanation that improves model fit. In this some by a two-compartment model for the same drug?
review, the basic principles of PBPK models are exam- How do physiological processes relate to the model?
ined. The aim is to help anesthesiologists interpret and Addressing these limitations has seen PBPK models
critique the PBPK anesthetic literature, and perhaps evolve alongside compartmental models.
provide a catalyst for their wider use in anesthetic
research.
History and applications of PBPK models
Some of the earliest PBPK models appear in the anes-
Modeling strategies
thetic literature, including the work of Eger (6) and
Compartmental models are traditionally developed using Mapleson (7) on volatile anesthetics. Price (8) used a
a ‘top-down’ approach where all of the information for PBPK model to argue that recovery from thiopental
the model comes from a given pharmacokinetic and was due to redistribution rather than elimination.
covariate data set. In contrast, PBPK models can be built Benowitz et al. (9,10) used a PBPK model to describe
using a ‘bottom-up’ approach, where the information for the effect of hemorrhage on the disposition of lido-
the model comes from a priori physiological and pharma- caine in monkeys. Davis and Mapleson showed that
cological mechanisms and in vitro data. At its most the kinetics of pethidine in organs could be estimated
extreme, this approach could be used to predict the kinet- from first principles (11). More recently, recirculatory
ics of a new chemical entity based only on physiology and PBPK models have been used to describe the disposi-
physicochemical properties (1). A ‘middle-out’ approach tion of propofol (12) and fentanyl (13) in clinical pop-
may also be possible (and sometimes desirable) that mixes ulations.
new data, known physiology and pharmacological princi-
ples, as necessary, to construct a model.
Commercially available physiologically-based
Full PBPK models have representations of the major
pharmacokinetic models
organs of the body. In contrast, Semi-PBPK or ‘Recir-
culatory’ models (2) may use pooling or lumping where Commercially available PBPK models have increased
different regions or organs of the body are represented the accessibility of PBPK methods. Software packages
as a single kinetic system. This may be useful where a include GastroPlusTM (Simulation Plus, Inc., Lancaster,
particular organ of the body is of particular interest CA), SimcypÒ simulator (Simcyp Ltd, Sheffield, United
(e.g., the brain for anesthetics (3)) but kinetics in the Kingdom), and PK-SimÒ (Bayer Technology Services,
remainder of the body can be represented as one, two, Leverkusen, Germany). For orally administered drugs,
or three lumped compartments. The lumping of model commercially available PBPK models have been used
compartments need not be ad hoc, but can potentially be predict food effects (14), formulation effects (15), and to
done using formal approaches (4). identify causes of poor oral bioavailability (16). PBPK
models were also used to study the effect of organ dys-
function (17), and drug–drug interactions (18). Pediatric
PBPK vs compartment models
modules are also available in these software using
Physiologically-based pharmacokinetic and compart- known information on age-related changes in blood
mental models are both made from ‘compartments’, but flows, organ volumes, binding protein concentrations,
differ in the way the compartments are joined together. and drug metabolism capacity.
A compartment is a region assumed to have uniform GastroPlusTM was the first commercially available
distribution, such that a sample from the compartment whole-body PBPK software initially based on the Com-
is representative of the rest of the compartment. The partmental Absorption and Transit (CAT) model (19),
analytical solutions of compartmental models become which was further expanded by Simulation Plus
intractable once more than three compartments are (www.simulations-plus.com) to the Advanced CAT
used, and hence, models with more than three compart- model (ACAT) (16). The ACAT model divides the GI
ments are rare. Population pharmacokinetic methods tract into nine compartments in sequence representing
(5) have allowed compartmental models to describe stomach, duodenum, jejunum (two compartments),
complex clinical populations. However, the limitations ileum (three compartments), cecum, and ascending
of compartmental models have been widely discussed: colon. It accounts for drug dissolution, precipitation,
they are not necessarily identifiable (more than one efflux or influx transporters, chemical stability, and first-
model can describe the same data set equally well). What pass metabolism. Formulation and drug-specific input
do the compartments actually represent? Why are some data, obtained from in vitro, in vivo, or in silico estimates

2 © 2016 John Wiley & Sons Ltd


R.N. Upton et al. Introduction to PBPK models

are fed into the dissolution and the absorption model. but the time-course of tissue equilibration is required to
Physiological parameters and characteristics for each estimate the distribution volume of the tissue.
gastrointestinal (GI) compartment are built in the model
based on literature data or user input. The cumulative
Regional venous concentrations
absorption is estimated and then convoluted with the
disposition kinetics of the drug to predict plasma con- The kinetics of drugs in individual organs can also be
centration-time profiles. inferred from measurements of concurrent arterial and
SimcypÒ was initially designed for mechanistic pre- regional venous plasma drug concentrations (i.e., the
diction of drug metabolism then expanded to predict drug concentration in blood entering and leaving the
pharmacokinetics including drug absorption (20). It organ). Cannulae can be placed acutely in man (26) or
uses the Advanced Dissolution, Absorption, and Meta- chronically in animals (27). This method potentially
bolism (ADAM) model itself structured on the CAT allows the collection of dense data with many time
model; however, it uses a more sophisticated model to points per subject and many sites per subject, even in
describe dissolution (21). SimcypÒ incorporates a priori man (28).
estimates of interindividual variability (IIV) on physio-
logic parameters. For example, it considers a popula-
In silico
tion mean and IIV of GI pH and bile acid
concentrations in the fed and fasted states and models Extensive experimentation has allowed the development
the effects of these factors on the drug’s solubility and of in silico methods for estimating the steady-state tis-
dissolution rate (20). sue:plasma partition coefficients of drugs based on their
The PK-SimÒ absorption model was based on the dis- physicochemical properties, plasma binding, and the
persion model (22) which considers the GI tract as a lipid and water content of the tissue in question (29,30).
one-dimensional cylindrical tube with spatially different These partition coefficients can be used in ‘flow-limited’
properties (e.g., pH, length, surface area) rather than a PBPK models.
series of transit compartments.
Getting blood-flow data
Getting pharmacokinetic data
The anatomical structure of PBPK models dictates that
Compartmental models are usually parameterized by fit- drugs are transported to and from organs by blood flow;
ting plasma concentration data from a single site in the hence the knowledge of blood flow values is integral to
body (e.g., arm venous blood). However, such a data set PBPK models and can be obtained from a number of
is generally not sufficient to estimate all parameters of sources.
PBPK models, and kinetic data for other sites in the
body are needed. Whether obtained by new experimen-
Databases and publications
tation or from the literature, data sources could include:
Documented values for blood flows (and organ size) first
came from the idea of a ‘Standard Man’ in radiation biol-
Tissue concentrations
ogy (http://en.wikipedia.org/wiki/Standard_person) who
Direct measurement of drug concentrations in tissue is 20–30 years of age, 70 kg and 170 cm tall. These values
homogenates is possible (23) but this method is generally are useful starting points, but do not account for the
only available for animals, and tissue samples are most between subject variability in body size and composition.
often taken postmortem. Therefore, only one time-point More usefully, Price et al. (31) extracted data from the
is available per animal (but with the collection of many NHANES database and developed software that can
tissues at that time). Defining the time-course of tissue return the organ size and blood flows for ‘virtual’ popula-
concentrations requires the sacrifice of a number of ani- tions (http://www.thelifelinegroup.org/p3m). Data for
mals at different time points. Steady-state tissue:plasma animals (albeit not accounting for variability in size) are
partition coefficients can be estimated and may be suffi- available for rats, mice, dogs, monkeys, sheep, and pigs
cient if the mechanisms of tissue:plasma equilibrium are (32–34).
well understood (e.g., perfusion limited). Tissue homoge-
nate data may not distinguish between intravascular,
Direct measurement
extracellular, and intracellular locations of the drug.
Microdialysis provides an alternative method for estimat- In animals, microspheres (35), flow probes (36), or tracer
ing free extracellular concentrations in tissues (24,25), kinetics (37) can be used to measure blood flow. In man,

© 2016 John Wiley & Sons Ltd 3


Introduction to PBPK models R.N. Upton et al.

Z t
tracer kinetics, ultrasound methods (e.g., Transcranial
Doppler (26)) and imaging methods are possible (38). It amountðtÞ ¼ ðQ:Cart  Q:Cven Þ ð4Þ
0
is not yet clear under what circumstances measuring
blood flow is warranted over using estimated or litera- The rate of change in the amount (A) of drug in the
ture blood flow values. It is almost certainly important tissue can be obtained by differentiation with respect to
for drugs or diseases associated with acute changes in time:
blood flow. dA
¼ Q:Cart  Q:Cven ð5Þ
dt
The basic maths Which can then be further expressed as the rate of
change in the venous (effluent) concentration by divid-
Consider the transport of drug in a blood vessel (e.g.,
ing both sides by the apparent volume of the tissue (V):
the blood vessels entering or leaving an organ). The rate
(or flux) of drug transport rate (J) in the vessel is a func- dCven Q:Cart  Q:Cven
¼ ð6Þ
tion of the blood flow (Q) and drug concentration in the dt V
blood (C): The apparent volume of the tissue is, in turn, the pro-
J ¼ Q:C ð1Þ duct of the real physical volume of the tissue (Vreal) and
the partition coefficient of the drug between blood and
Alternatively, if drug is added to the flowing blood at tissue (represented by convention as Kp):
a constant rate (e.g. as a constant rate infusion), the con-
centration achieved in the blood is proportional to the V ¼ Vreal :Kp ð7Þ
infusion rate (J) and inversely proportional to the blood Thus, the real physical volume of the tissue may be
flow: based upon known values, while the partition coefficient
J may be determined experimentally in vitro from tissue
C¼ ð2Þ
Q homogenates if a ‘bottom-up’ approach is employed, or
as an estimated parameter when fitting to blood and tis-
As an inverse relationship, low blood flows are associ-
sue concentration-time data. As can be seen in Equa-
ated with relatively higher blood concentrations for a
tion 6, during a constant infusion of drug into the
given infusion rate—a fundamental concept underlying
arterial blood, the venous drug concentrations will
the relationship between pharmacokinetics and the cir-
increase to a steady-state value equal to that in the arte-
culation.
rial blood. Importantly, the rate of change in concentra-
tion will be affected by both the apparent volume of the
Organs as a differential equations organ and the blood flow, as illustrated in Figure 2.
Ultimately a PBPK model connects representations of
kinetics in organs in the body, but these can be consid- Membrane-limited distribution
ered in isolation.
The next most complex situation for distribution only
occurs when the drug does not distribute within the tis-
Flow-limited distribution sue instantaneously due to a diffusion barrier which
The simplest model of an organ considers the case where restricts the drug’s movement into a component of the
a drug is constantly infused into a blood vessel which tissue (Figure 1). In this case, the tissue can be divided
perfuses a single noneliminating organ, and where there into two (or more) ‘compartments’ with a permeability
is no restriction to the distribution of the drug within term (represented by convention as PS) used to describe
the organ (Figure 1). The amount of drug in the organ the movement of drug across the barrier separating the
at any time ‘t’ is equal to the integral of the difference initial (V1) and the ‘deeper’ (V2) tissue spaces.
between the rate of entry and the rate of exit from the The diffusion barrier is typically some form of mem-
organ: brane, although the nature of the ‘membrane’ will be
Z t specific to the drug and tissue, as will any physical inter-
amountðtÞ ¼ ðJin  Jout Þ ð3Þ pretation of the two compartments. For example, the
0 capillary wall may present a substantial barrier. In
Substitution of Equation 1 into Equation 3 then which case the initial volume (V1) may represent the
allows for a description in terms of concentrations and vascular space of the tissue, while the ‘deeper’ tissue vol-
blood flows: ume (Vtissue) may represent the total of extracellular and

4 © 2016 John Wiley & Sons Ltd


R.N. Upton et al. Introduction to PBPK models

Figure 1 Simple models of individual organs. Upper: flow-limited


distribution of a drug into a single tissue perfused by a single blood
vessel. Middle: membrane-limited distribution in a single tissue per- Figure 2 Flow-limited distribution of a drug into a single tissue per-
fused by a single blood vessel. Lower: flow-limited distribution with fused by a single blood vessel. The impact of changes in blood flow
elimination of a drug from a single tissue perfused by a single blood (upper panel) and apparent tissue volume (lower panel) on the con-
vessel. Cart and Cven are the arterial and venous blood concentrations centration-time profile resulting from a 10 min 5 mgmin1 continu-
of drug, respectively, Q is the blood flow through the tissue and V is ous arterial infusion. Arterial concentrations are shown in red and
the apparent volume of the tissue, CL is the clearance of the drug by venous concentrations are shown in blue, with line type grouping the
the tissue. For the membrane-limited model, C2 is the deeper tissue data. In the upper panel, the apparent volume of the tissue remained
concentration of the drug, and PS is the permeability term. constant at 10 l while flow was changed. In the lower panel, flow
remained constant at 5 lmin1 while apparent tissue volume was
changed, and so arterial concentrations are identical for all three con-
ditions.
intracellular spaces. Similarly, if the cell membrane pre-
sents a barrier to drug distribution, then V1 would
approach extracellular volume. There may be other bar-
compartment based on distribution rate can be an
riers to drug distribution such as the nuclear or other
empirical decision equivalent to the way that compart-
unknown membranes. Assuming that the distribution
mental model compartments are identified, or based on
across the membrane is reversible, then a mathematical
an understanding of physicochemical properties (e.g.,
description of this system is described below:
lipid solubility).
dCven Q:Cart  Q:Cven  PS:Cven þ PS:Ctissue
¼ ð8Þ
dt V1
Flow-limited with elimination
dCtissue PS:Cven  PS:Ctissue
¼ ð9Þ In the case where an organ has a capacity to eliminate
dt Vtissue the drug, this can be accommodated by incorporating a
While the term ‘membrane-limited’ implies that only clearance (CL) term. For simplicity consider a simple
permeability across the membrane is a limiting factor, flow-limited tissue with clearance added (Figure 1). For
this is not necessarily the case and more generally it is a this model CL should be linked to the arterial concen-
mixture of both flow and membrane limitation trations (39) and cannot exceed flow (Q) for obvious
(Figure 3). The division of a tissue into more than one reasons.

© 2016 John Wiley & Sons Ltd 5


Introduction to PBPK models R.N. Upton et al.

dCven Q:Cart  Q:Cven  CL:Cart


¼ ð10Þ
dt V
Alternatively, elimination can be represented as an
extraction ratio (E) which takes values ranging from 0
to 1. In this case, the clearance of the tissue is the pro-
duct of flow and extraction ratio:
CL ¼ Q:E ð11Þ
Extraction ratio can be calculated from the arterial
and venous concentrations and is a function of blood
flow and intrinsic clearance (CLint) of the organ:
Cart  Cven CLint
E¼ ¼ ð12Þ
Cart Q þ CLint
Extraction ratio, rather than CL, may be used in a
model and determined from perfused tissue experiments
if a ‘bottom-up’ approach is employed, or as an esti-
mated parameter when fitting to blood and tissue con-
centration-time data. Importantly changes in CL do not
affect the rate of equilibration of the tissue. However,
during a constant infusion of drug into the arterial
blood, the venous drug concentrations of the eliminating
organ will increase to a steady-state equal to (1-E) of
those in the arterial blood (Figure 4).

Figure 3 Membrane-limited distribution of a drug into a single tissue


perfused by a single blood vessel. The impact of changes in perme-
ability across a membrane barrier on the concentration-time profile
resulting from a 10-min 5 mgmin1 continuous arterial infusion.
Figure 4 Flow-limited distribution with elimination of a drug in a sin-
Arterial concentrations are shown in red, venous concentrations are
gle tissue perfused by a single blood vessel following an intravenous
shown in blue, and tissue concentrations are shown in green, with
infusion. The impact of changes in clearance on the concentration-
line type grouping the data. In all panels, the apparent initial volume,
time profile resulting from a 10-min 5 mgmin1 continuous arterial
deeper tissue volume and flow remained constant at 10 l, 50 l and
infusion in a simple flow-limited tissue. Arterial concentrations are
5 lmin1, respectively, while the permeability was changed from
shown in red and venous concentrations are shown in blue, with line
100 to 10 to 1 lmin1. In the lower panel, flow remained constant at
type grouping the data. The apparent volume of the tissue and blood
5 lmin1 while apparent tissue volume was changed, and so arterial
flow (Q) remained constant at 10 l and 5 lmin1, respectively, while
concentrations were identical for all three conditions. Note the mark-
clearance (CL) was changed. The extraction ratio (E) is equal to CL/Q.
edly slower equilibration of the tissue concentrations as permeability
Note that the venous concentrations approach a steady-state equal
decreases.
to (1-E) multiplied by the arterial concentration.

6 © 2016 John Wiley & Sons Ltd


R.N. Upton et al. Introduction to PBPK models

More complicated organ models


The three organ models described above are the simplest
conceivable, and could describe the kinetics of many
drugs. More complex models such as the well-stirred
hepatic model (40) may be needed for orally adminis-
tered drugs, where first-pass effects are important and
there is a more complex interplay between protein bind-
ing, elimination capacity, and blood flow (41). The well-
stirred model also allows useful descriptions of drug
interactions (42).

Joining organs together


The complexity of the whole-body model will be dic-
tated by the availability of organ data in a top-down
approach, the organs of particular interest or impor-
tance in a bottom-up approach, or a combination of
both in a middle-out approach. It is also common to
‘pool’ or ‘lump’ organs which, although anatomically
distinct, may be considered as a single kinetic unit.
For example, lumping all muscle tissue, bone, fat, vis-
cera, or skin is often employed, and quite often these
may also be lumped together and termed the ‘body’ or
Figure 5 A simple whole-body physiologically-based pharmacoki-
‘carcass’. The appropriateness and nature of such netic model with hepatic elimination. Cart and Cven are the arterial and
lumping will be drug-dependent and should be guided venous blood concentrations of the drug, respectively, QCO, Qhep are
by careful consideration of the known physicochemical cardiac output and hepatic blood flow, respectively. Qbody is equal to
and pharmacological properties of the drug. Most QCO minus Qhep. Vlung, Vhep, and Vbody are the apparent volume of
organs are connected in parallel with several notable the lung, liver, and body, respectively. CLhep is the clearance of the
exceptions. The lung is connected in series with all drug by the liver.

other organs, and is arguably a key organ to include


in any whole-body model, particularly for intravenous
administration where it can be an important first-pass dClung Rateinf  QCO :Cart þ Qhep :Chep þ Qbody :Cbody
organ. Similarly, blood flow from the gut, spleen, and ¼
dt Vlung
pancreas are in parallel and lead to the liver via the
ð13Þ
portal vein. Bjorkman et al. (23,43,44) provide an
excellent example for the opioid analgesics, fentanyl dChep Qhep :Cart  Qhep :Chep  CLhep :Cart
and alfentanil. In order to construct a model of a ¼ ð14Þ
dt Vhep
whole body, models describing the inputs and outputs
of the individual organs were anatomically connected dCbody Qbody :Cart  Qbody :Cbody
via blood flows, with blood flow through any lumped ¼ ð15Þ
dt Vbody
organs equaling cardiac output minus the sum of all
blood flows elsewhere defined. Figure 6 illustrates the concentration-time course in
the arterial blood (Cart), body (Cbody), and liver tissue
(Chep) during a constant infusion of drug (Rateinf) into
A simple whole-body model
the arterial blood at a rate of 100 mgmin1 for
A simple case of a whole-body, semi-PBPK model incor- 200 min, with cardiac output (QCO) at varying values.
porates three compartments: the lungs, the liver as an Checking the coding of the PBPK model is important.
eliminating organ, and the rest of the body (Figure 5). For example, in Figure 6, the arterial concentrations
In this example all organs are assumed to display flow- reach an eventual steady-state of 100 mgl1, corre-
limited distribution with the drug administered as an sponding to the expected value calculated as dose rate/
intravenous infusion. The equations that describe this clearance (100 mgmin1/1 lh1). Also note the slowing
system are outlined below: of the rise in concentrations as cardiac output decreases

© 2016 John Wiley & Sons Ltd 7


Introduction to PBPK models R.N. Upton et al.

to arterial concentrations, remained constant at 50%


regardless of cardiac output as neither hepatic clearance
nor liver blood flow were changed, which is in agreement
with the ratio of CLhep/Qhep; an alternative expression
of extraction ratio.

Adjusting for size and body composition


Physiologically-based pharmacokinetic models are often
structured so that the physiological parameters (e.g.,
organ size and blood flow) and pharmacological param-
eters (e.g., free fraction, extraction ratio, intrinsic clear-
ance) are distinct. This allows the physiological
parameters to be adjusted for differences in body size or
composition, both within and across species.

Within species
Allowing for differences in body size within a species
(e.g., with age) follows the standard principles of allome-
try. The general allometric equation is y = a + log(wt)
*b where b is the allometric coefficient. Organ size can
be scaled as a fixed fraction of body weight (b = 1),
while blood flow can be scaled with an allometric coeffi-
cient of 0.75 (45). The later value (implying higher rela-
tive perfusion (flow per g) in smaller subjects), reflects
the geometry of oxygen delivery to cells (46). The con-
cept of maturation models (45,47,48) accounting for the
early development of metabolic and renal capacity can
be implemented directly in PBPK models as per com-
partmental models.

Across species
The PBPK models are amenable to scaling kinetics
across species, where they have found a useful role in
predicting ‘first in man’ kinetics from animal data (49).
The task is to account not only for the difference in
body size between the species (via allometric principles)
Figure 6 A simple whole-body physiologically-based pharmacoki- but also account for the sometimes substantial differ-
netic model with hepatic elimination following an intravenous infu-
ence in body composition between species (34). Allow-
sion. The impact of changes in cardiac output (3, 6, and 9 lmin1) on
ing for any differences in plasma protein binding is also
the concentration-time profile resulting from a 200 min
100 mgmin1 continuous infusion in a simple three-compartment,
important (13).
whole-body model (see Figure 5). Arterial concentrations are shown
in red and concentrations in the body and liver are in blue and green,
Software
respectively. The apparent volume of the body, liver, and lungs
remained constant at 25 l, 2 l, and 1 l, respectively, liver blood flow Any general-purpose software platform that can solve
and hepatic clearance also remained constant at 6 lmin1 and
systems of differential equations can implement a
1 lmin1, respectively.
bespoke PBPK model. Software can be classified by
whether they can be used for simulation only
with no gradient at steady state between the body and [STELLAÒ(Cognitus Ltd., North Yorkshire, United
arterial blood. In contrast, the hepatic extraction ratio, Kingdom), Berkeley MadonnaTM (www.berkeley-
illustrated by steady-state liver concentrations compared madonna.com)], for fitting single subject data (R,

8 © 2016 John Wiley & Sons Ltd


R.N. Upton et al. Introduction to PBPK models

MATLAB) or fitting population data (NONMEM, S- most difficult aspect of their implementation—collecting,
Adapt, Monolix, WinBugs, Stan). Examples of fitted collating, and justifying the data used to parameterize the
population PBPK are uncommon as the size of the mod- model. Furthermore, for complex PBPK models the task
els make fit times intractable, but examples do appear in of validating that the model is ‘fit for purpose’ is not triv-
the literature (13,41,50–53). ial (49). Recent developments in data archiving and pub-
lishing (55) and the standardization of model
terminology (http://www.ddmore.eu/) will surely make
Pharmacodynamic models
these tasks easier in the future.
The general principles for linking of pharmacodynamic How have PBPK models helped address some of the
models to PBPK models are the same as those for mam- questions posed at the beginning of the review? Why do
millary PK models (54). For PBPK models, drug con- arterial and venous blood concentrations have different
centrations in organs or specific sites in the vasculature time-courses? PBPK-based analysis has shown that
(e.g., arterial or venous) can be linked to drug effect venous blood represents systemic kinetics ‘convolved’
rather than linking to the more ambiguously defined with the kinetics in the tissue drained by the vein in
central compartment concentration of compartmental question. Maximum concentrations in venous blood
PK models. This can be advantageous from some classes after intravenous bolus administration can be lower and
of drugs (such as anesthetics) where the site of drug later than in arterial blood, and this has the potential to
action is well understood (e.g., the CNS). Using drug confound PD analyses in anesthetic settings (13). What
concentrations in the target organ to drive pharmacody- does a hyper- or hypodynamic circulation mean for drug
namic effects may obviate the need for an empirical disposition and effect? PBPK-based analysis has shown
effect compartment (13), with the advantage of an ‘in- that for intravenous anesthetics, peak concentrations
built’ representation of the factors (such as organ blood after intravenous bolus injection are inversely related to
flow) affecting target organ equilibration. cardiac output, and potentially dangerous high peak
concentrations are possible in the hypodynamic patient,
while efficacy may be reduced in the hyperdynamic
Conclusions
patient (27,56,57). While these analyses have only begun
The PBPK models come in a variety of shapes and sizes. to explore the potential uses of PBPK models in anes-
Most are built using compartments and differential equa- thetic research, we are very confident that the trend for
tions and as such a few basic building blocks such as increasing use of PBPK models will allow valuable
those outlined here can be used to build relatively large insight into optimal drug use in pediatric anesthesia.
and sophisticated models. It is prudent and time-efficient
to adapt the level of complexity of a PBPK model to the
Ethical approval
requirements of the project at hand. Models can expand
or develop more complexity as more data or understand- Not applicable.
ing becomes available, or as it becomes necessary for the
model to address more complex questions. Despite the
Funding
growing acceptance and commercial support of PBPK,
we suspect the greatest potential of the PBPK approach The study received no external funding.
will be realized when there is increased acceptance of
adapting commonly used mammillary models with mod-
Conflict of interest
els with physiological features as needed. The relative
ease of constructing the code for PBPK models belies the The authors report no conflicts of interest.

References
1 Edginton AN. Knowledge-driven 3 Upton RN, Ludbrook GL. A model of the pharmacokinetic-pharmacodynamic data.
approaches for the guidance of first-in- kinetics and dynamics of induction of anaes- CPT Pharmacomet Syst Pharmacol 2014; 3:
children dosing. Pediatr Anesth 2011; 21: thesia in sheep: variable estimation for e90.
206–213. thiopental and comparison with propofol. 5 Mould DR, Upton RN. Basic concepts in
2 Krejcie TC, Avram MJ. Recirculatory phar- Br J Anaesth 1999; 82: 890–899. population modeling, simulation, and
macokinetic modeling: what goes around, 4 Gulati A, Isbister GK, Duffull SB. Scale model-based drug development. CPT
comes around. Anesth Analg 2012; 115: reduction of a systems coagulation model Pharmacomet Syst Pharmacol 2012; 1:
223–226. with an application to modeling e6.

© 2016 John Wiley & Sons Ltd 9


Introduction to PBPK models R.N. Upton et al.

6 Eger EII. A mathematical model of uptake 19 Yu LX, Lipka E, Crison JR et al. Transport 31 Price PS, Conolly RB, Chaisson CF et al.
and distribution. In: Papper EM, Kitz RJ, approaches to the biopharmaceutical design Modeling interindividual variation in physio-
eds. Uptake and Distribution of Anesthetic of oral drug delivery systems: prediction of logical factors used in PBPK models of
Agents. New York: McGraw-Hill, 1963: 72– intestinal absorption. Adv Drug Deliv Rev humans. Crit Rev Toxicol 2003; 33: 469–503.
87. 1996; 19: 359–376. 32 Brown RP, Delp MD, Lindstedt SL et al.
7 Mapleson WW. Uptake and distribution of 20 Jamei M, Dickinson GL, Rostami-Hodjegan Physiological parameter values for physio-
anesthetic agents. In: Papper E, Kitz RJ, eds. A. A framework for assessing inter-indivi- logically based pharmacokinetic models.
Uptake and Distribution of Anesthetic dual variability in pharmacokinetics using Toxicol Ind Health 1997; 13: 407–484.
Agents. New York: McGraw-Hill, 1963: virtual human populations and integrating 33 Clewell RA, Clewell HJ. Development and
104–119. general knowledge of physical chemistry, specification of physiologically based phar-
8 Price HL. A dynamic concept of the distribu- biology, anatomy, physiology and genetics: a macokinetic models for use in risk assess-
tion of thiopental in the human body. Anes- tale of ‘bottom-up’ vs ‘top-down’ recognition ment. Regul Toxicol Pharmacol RTP 2008;
thesiology 1960; 21: 40–45. of covariates. Drug Metab Pharmacokinet 50: 129–143.
9 Benowitz N, Forsyth FP, Melmon KL et al. 2009; 24: 53–75. 34 Upton RN. Organ weights and blood flows
Lidocaine disposition kinetics in monkey 21 Kostewicz ES, Aarons L, Bergstrand M et al. of sheep and pig for physiological pharma-
and man. I. Prediction by a perfusion model. PBPK models for the prediction of in vivo cokinetic modelling. J Pharmacol Toxicol
Clin Pharmacol Ther 1974; 16: 87–98. performance of oral dosage forms. Eur J Methods 2008; 58: 198–205.
10 Benowitz N, Forsyth RP, Melmon KL et al. Pharm Sci Off J Eur Fed Pharm Sci 2014; 16: 35 Robertson HT, Hlastala MP. Microsphere
Lidocaine disposition kinetics in monkey and 300–321. maps of regional blood flow and regional
man. II. Effects of hemorrhage and sympath- 22 Ni PF, Ho NFH, Fox JL et al. Theoretical ventilation. J Appl Physiol Bethesda Md 1985
omimetic drug administration. Clin Pharma- model studies of intestinal drug absorption 2007; 102: 1265–1272.
col Ther 1974; 16: 99–109. V. Non-steady-state fluid flow and absorp- 36 Upton R, Grant C, Ludbrook G. An ultra-
11 Davis NR, Mapleson WW. Structure and tion. Int J Pharm 1980; 5: 33–47. sonic Doppler venous outflow method for
quantification of a physiological model of the 23 Bj€orkman S, Stanski DR, Verotta D et al. the continuous measurement of cerebral
distribution of injected agents and inhaled Comparative tissue concentration profiles of blood flow in conscious sheep. J Cereb Blood
anaesthetics. Br J Anaesth 1981; 53: 399–405. fentanyl and alfentanil in humans predicted Flow Metab Off J Int Soc Cereb Blood Flow
12 Masui K, Upton RN, Doufas AG et al. The from tissue/blood partition data obtained in Metab 1994; 14: 680–688.
performance of compartmental and physio- rats. Anesthesiology 1990; 72: 865–873. 37 Doolette DJ, Upton RN, Grant C. Agree-
logically based recirculatory pharmacokinetic 24 Bouw R, Ederoth P, Lundberg J et al. ment between ultrasonic Doppler venous
models for propofol: a comparison using Increased blood-brain barrier permeability of outflow and Kety and Schmidt estimates of
bolus, continuous, and target-controlled morphine in a patient with severe brain cerebral blood flow. Clin Exp Pharmacol
infusion data. Anesth Analg 2010; 111: 368– lesions as determined by microdialysis. Acta Physiol 1999; 26: 736–740.
379. Anaesthesiol Scand 2001; 45: 390–392. 38 Cosgrove D, Eckersley R, Blomley M et al.
13 Upton RN, Foster DJR, Christrup LL et al. 25 Ederoth P, Tunblad K, Bouw R et al. Blood- Quantification of blood flow. Eur Radiol
A physiologically-based recirculatory meta- brain barrier transport of morphine in 2001; 11: 1338–1344.
model for nasal fentanyl in man. J Pharma- patients with severe brain trauma. Br J Clin 39 Upton RN, Foster DJR. Representing elimi-
cokinet Pharmacodyn 2012; 39: 561–576. Pharmacol 2004; 57: 427–435. nating organs in physiological pharmacoki-
14 Parrott N, Lukacova V, Fraczkiewicz G 26 Ludbrook GL, Visco E, Lam AM. Propofol: netic models. World Congress of
et al. Predicting pharmacokinetics of drugs relation between brain concentrations, elec- Pharmacometrics. In Seoul, Korea; 2012.
using physiologically based modeling–appli- troencephalogram, middle cerebral artery 40 Baker M, Parton T. Kinetic determinants of
cation to food effects. AAPS J 2009; 11: 45– blood flow velocity, and cerebral oxygen hepatic clearance: plasma protein binding
53. extraction during induction of anesthesia. and hepatic uptake. Xenobiotica Fate Foreign
15 Lukacova V, Woltosz WS, Bolger MB. Pre- Anesthesiology 2002; 97: 1363–1370. Compd Biol Syst 2007; 37: 1110–1134.
diction of modified release pharmacokinetics 27 Upton RN, Ludbrook GL. A physiological 41 Hopkins AM, Wiese MD, Proudman SM
and pharmacodynamics from in vitro, imme- model of induction of anaesthesia with propo- et al. Semiphysiologically based pharmacoki-
diate release, and intravenous data. AAPS J fol in sheep. 1. Structure and estimation of netic model of leflunomide disposition in
2009; 11: 323–334. variables. Br J Anaesth 1997; 79: 497–504. rheumatoid arthritis patients. CPT Pharma-
16 Agoram B, Woltosz WS, Bolger MB. Pre- 28 Mather LE, Runciman WB, Ilsley AH et al. comet Syst Pharmacol 2015; 4: 362–371.
dicting the impact of physiological and bio- Direct measurement of chlormethiazole 42 Zhang D, Zhu M, Humphreys WG. Drug
chemical processes on oral drug extraction by liver, lung and kidney in man. Metabolism in Drug Design and Develop-
bioavailability. Adv Drug Deliv Rev 2001; 1 Br J Clin Pharmacol 1981; 12: 319–325. ment. Hoboken, NJ, USA: John Wiley &
(50 Suppl 1): S41–S67. 29 Small H, Gardner I, Jones HM et al. Mea- Sons; 2007. 634 p.
17 Edginton AN, Willmann S. Physiology-based surement of binding of basic drugs to acidic 43 Bj€
orkman S, Stanski DR, Harashima H et al.
simulations of a pathological condition: pre- phospholipids using surface plasmon reso- Tissue distribution of fentanyl and alfentanil
diction of pharmacokinetics in patients with nance and incorporation of the data into in the rat cannot be described by a blood flow
liver cirrhosis. Clin Pharmacokinet 2008; 47: mechanistic tissue composition equations to limited model. J Pharmacokinet Biopharm
743–752. predict steady-state volume of distribution. 1993; 21: 255–279.
18 Yeo KR, Jamei M, Rostami-Hodjegan A. Drug Metab Dispos Biol Fate Chem 2011; 39: 44 Bj€
orkman S, Wada DR, Stanski DR. Appli-
Predicting drug-drug interactions: applica- 1789–1793. cation of physiologic models to predict the
tion of physiologically based pharmacoki- 30 Rodgers T, Rowland M. Mechanistic influence of changes in body composition
netic models under a systems biology approaches to volume of distribution predic- and blood flows on the pharmacokinetics of
approach. Expert Rev Clin Pharmacol 2013; tions: understanding the processes. Pharm fentanyl and alfentanil in patients. Anesthesi-
6: 143–157. Res 2007; 24: 918–933. ology 1998; 88: 657–667.

10 © 2016 John Wiley & Sons Ltd


R.N. Upton et al. Introduction to PBPK models

45 Anderson BJ, Holford NHG. Tips and traps 50 Langdon G, Gueorguieva I, Aarons L model to stabilise the Bayesian analysis
analyzing pediatric PK data. Pediatr Anesth et al. Linking preclinical and clinical of clinical data. AAPS J 2016; 18:
2011; 21: 222–237. whole-body physiologically based pharma- 196–209.
46 West GB, Brown JH. The origin of allomet- cokinetic models with prior distributions in 54 Upton RN, Mould DR. Basic concepts in
ric scaling laws in biology from genomes to NONMEM. Eur J Clin Pharmacol 2007; population modeling, simulation, and model-
ecosystems: towards a quantitative unifying 63: 485–498. based drug development: part 3-introduction
theory of biological structure and organiza- 51 Gueorguieva I, Aarons L, Rowland M. Dia- to pharmacodynamic modeling methods.
tion. J Exp Biol 2005; 208: 1575–1592. zepam pharamacokinetics from preclinical to CPT Pharmacomet Syst Pharmacol 2014; 3:
47 Anderson BJ, Holford NHG. Mechanism- phase I using a Bayesian population physio- e88.
based concepts of size and maturity in phar- logically based pharmacokinetic model with 55 Kratz J, Strasser C. Data publication consen-
macokinetics. Annu Rev Pharmacol Toxicol informative prior distributions in WinBUGS. sus and controversies. F1000Research 2014;
2008; 48: 303–332. J Pharmacokinet Pharmacodyn 2006; 33: 3: 94.
48 Rhodin MM, Anderson BJ, Peters AM et al. 571–594. 56 Ludbrook GL, Upton RN. A physiological
Human renal function maturation: a quantita- 52 Wendling T, Dumitras S, Ogungbenro K model of induction of anaesthesia with
tive description using weight and postmenstrual et al. Application of a Bayesian approach to propofol in sheep. 2. Model analysis and
age. Pediatr Nephrol Berl Ger 2009; 24: 67–76. physiological modelling of mavoglurant pop- implications for dose requirements. Br J
49 Rowland M, Lesko LJ, Rostami-Hodjegan ulation pharmacokinetics. J Pharmacokinet Anaesth 1997; 79: 505–513.
A. Physiologically based pharmacokinetics is Pharmacodyn 2015; 42: 639–657. 57 Upton RN, Ludbrook G. A physiologically
impacting drug development and regulatory 53 Wendling T, Tsamandouras N, Dumitras based, recirculatory model of the kinetics
decision making. CPT Pharmacomet Syst S et al. Reduction of a whole-body and dynamics of propofol in man. Anesthesi-
Pharmacol 2015; 4: 313–315. physiologically based pharmacokinetic ology 2005; 103: 344–352.

© 2016 John Wiley & Sons Ltd 11

Das könnte Ihnen auch gefallen