Physiologically based pharmacokinetic modelling

Last updated
Graphic representation of a physiologically based whole body model. Here, it is dissected into seven tissue/organ compartments: brain, lungs and heart, pancreas, liver, gut, kidney and adipose/muscle tissue. Blood flows, Q, and concentration, [X], of a substance of interest are depicted. WholeBody wiki.svg
Graphic representation of a physiologically based whole body model. Here, it is dissected into seven tissue/organ compartments: brain, lungs and heart, pancreas, liver, gut, kidney and adipose/muscle tissue. Blood flows, Q, and concentration, [X], of a substance of interest are depicted.

Physiologically based pharmacokinetic (PBPK) modeling is a mathematical modeling technique for predicting the absorption, distribution, metabolism and excretion (ADME) of synthetic or natural chemical substances in humans and other animal species. PBPK modeling is used in pharmaceutical research and drug development, and in health risk assessment for cosmetics or general chemicals.

Contents

PBPK models strive to be mechanistic by mathematically transcribing anatomical, physiological, physical, and chemical descriptions of the phenomena involved in the complex ADME processes. A large degree of residual simplification and empiricism is still present in those models, but they have an extended domain of applicability compared to that of classical, empirical function based, pharmacokinetic models. PBPK models may have purely predictive uses, but other uses, such as statistical inference, have been made possible by the development of Bayesian statistical tools able to deal with complex models. [1] That is true for both toxicity risk assessment and therapeutic drug development.

PBPK models try to rely a priori on the anatomical and physiological structure of the body, and to a certain extent, on biochemistry. They are usually multi-compartment models, with compartments corresponding to predefined organs or tissues, with interconnections corresponding to blood or lymph flows (more rarely to diffusions). A system of differential equations for concentration or quantity of substance on each compartment can be written, and its parameters represent blood flows, pulmonary ventilation rate, organ volumes etc., for which information is available in scientific publications. Indeed, the description they make of the body is simplified and a balance needs to be struck between complexity and simplicity. Besides the advantage of allowing the recruitment of a priori information about parameter values, these models also facilitate inter-species transpositions or extrapolation from one mode of administration to another (e.g., inhalation to oral). An example of a 7-compartment PBPK model, suitable to describe the fate of many solvents in the mammalian body, is given in the Figure on the right.

History

The first pharmacokinetic model described in the scientific literature [2] was in fact a PBPK model. It led, however, to computations intractable at that time. The focus shifted then to simpler models, [3] for which analytical solutions could be obtained (such solutions were sums of exponential terms, which led to further simplifications.) The availability of computers and numerical integration algorithms marked a renewed interest in physiological models in the early 1970s. [4] [5] For substances with complex kinetics, or when inter-species extrapolations were required, simple models were insufficient and research continued on physiological models. [6] [7] [8] By 2010, hundreds of scientific publications had described and used PBPK models, and at least two private companies have based their business on their expertise in this area.

Building a PBPK model

The model equations follow the principles of mass transport, fluid dynamics, and biochemistry in order to simulate the fate of a substance in the body. [9] Compartments are usually defined by grouping organs or tissues with similar blood perfusion rate and lipid content (i.e. organs for which chemicals' concentration vs. time profiles will be similar). Ports of entry (lung, skin, intestinal tract...), ports of exit (kidney, liver...) and target organs for therapeutic effect or toxicity are often left separate. Bone can be excluded from the model if the substance of interest does not distribute to it. Connections between compartment follow physiology (e.g., blood flow in exit of the gut goes to liver, etc.)

Basic transport equations

Drug distribution into a tissue can be rate-limited by either perfusion or permeability. [10] [11] Perfusion-rate-limited kinetics apply when the tissue membranes present no barrier to diffusion. Blood flow, assuming that the drug is transported mainly by blood, as is often the case, is then the limiting factor to distribution in the various cells of the body. That is usually true for small lipophilic drugs. Under perfusion limitation, the instantaneous rate of entry for the quantity of drug in a compartment is simply equal to (blood) volumetric flow rate through the organ times the incoming blood concentration. In that case; for a generic compartment i, the differential equation for the quantity Qi of substance, which defines the rate of change in this quantity, is:

where Fi is blood flow (noted Q in the Figure above), Cart incoming arterial blood concentration, Pi the tissue over blood partition coefficient and Vi the volume of compartment i.

A complete set of differential equations for the 7-compartment model shown above could therefore be given by the following table:

TissueDifferential Equation
Gut
Kidney
Poorly-perfused tissues (muscle and skin)
Brain
Heart and lung
Pancreas
Liver

The above equations include only transport terms and do not account for inputs or outputs. Those can be modeled with specific terms, as in the following.

Modeling inputs

Modeling inputs is necessary to come up with a meaningful description of a chemical's pharmacokinetics. The following examples show how to write the corresponding equations.

Ingestion

When dealing with an oral bolus dose (e.g. ingestion of a tablet), first order absorption is a very common assumption. In that case the gut equation is augmented with an input term, with an absorption rate constant Ka:

That requires defining an equation for the quantity ingested and present in the gut lumen:

In the absence of a gut compartment, input can be made directly in the liver. However, in that case local metabolism in the gut may not be correctly described. The case of approximately continuous absorption (e.g. via drinking water) can be modeled by a zero-order absorption rate (here Ring in units of mass over time):

More sophisticated gut absorption model can be used. In those models, additional compartments describe the various sections of the gut lumen and tissue. Intestinal pH, transit times and presence of active transporters can be taken into account . [12]

Skin depot

The absorption of a chemical deposited on skin can also be modeled using first order terms. It is best in that case to separate the skin from the other tissues, to further differentiate exposed skin and non-exposed skin, and differentiate viable skin (dermis and epidermis) from the stratum corneum (the actual skin upper layer exposed). This is the approach taken in [Bois F., Diaz Ochoa J.G. Gajewska M., Kovarich S., Mauch K., Paini A., Péry A., Sala Benito J.V., Teng S., Worth A., in press, Multiscale modelling approaches for assessing cosmetic ingredients safety, Toxicology. doi: 10.1016/j.tox.2016.05.026]

Unexposed stratum corneum simply exchanges with the underlying viable skin by diffusion:

where is the partition coefficient, is the total skin surface area, the fraction of skin surface area exposed, ...

For the viable skin unexposed:

For the skin stratum corneum exposed:

for the viable skin exposed:

dt(QSkin_u) and dt(QSkin_e) feed from arterial blood and back to venous blood.

More complex diffusion models have been published [reference to add].

Intra-venous injection

Intravenous injection is a common clinical route of administration. (to be completed)

Inhalation

Inhalation occurs through the lung and is hardly dissociable from exhalation (to be completed)

Modelling metabolism

There are several ways metabolism can be modeled. For some models, a linear excretion rate is preferred. This can be accomplished with a simple differential equation. Otherwise a Michaelis-Menten equation, as follows, is generally appropriate for a more accurate result.

.

Uses of PBPK modeling

Simulated drug plasma concentration over time curves following IV infusion and multiple oral doses. The drug has an elimination half-life of 4 hours, and an apparent volume of distribution of 10 liters. DrugConcOverTimeIVOral.png
Simulated drug plasma concentration over time curves following IV infusion and multiple oral doses. The drug has an elimination half-life of 4 hours, and an apparent volume of distribution of 10 liters.
Pharmacokinetic model of drugs entering a tumor. (A) Schematic illustration of a tumor vessel illustrating loss of smooth muscle cells, local degradation of the extracellular matrix, and increased permeability of the endothelium. (B) Illustration of the pharmacokinetic model taking into account the EPR effect. The rate constants kp and kd describe exchange with the peripheral volume. The rate constants kepr and kb describe extravasation from circulation into the tumor, and intravasation back into the circulation, respectively. The rate constant kel represents clearance by the kidneys, MPS, and any other non-tumor elimination processes, such that when kb = 0, k10 = kepr + kel where kel is the elimination rate constant. (C) Standard two compartment model with central and peripheral compartments. c1 and c2 represent the drug concentration in blood (central compartment) and normal tissue (peripheral compartment), respectively. The first order rate constant k10 describes all elimination pathways, including clearance by the kidneys, uptake by the MPS, and tumor accumulation. The first order rate constants k12 and k21 describe exchange between the two compartments. Note that kp = k12, kd = k21. (D) Two compartment model defined in terms of the drug amount, where Nbl is the amount of drug in blood (mg), and Np is the amount in peripheral tissue (mg). (E) Three compartment model with the addition of a tumor "compartment" where Nt is the amount of drug in the tumor. Exchange with the tumor is described by the rate constants kepr and kb, respectively. The rate constant kel describes elimination pathways including clearance by the kidneys and uptake by the MPS, but does not include tumor accumulation. Pharmacokinetic model of drugs entering a tumor.svg
Pharmacokinetic model of drugs entering a tumor. (A) Schematic illustration of a tumor vessel illustrating loss of smooth muscle cells, local degradation of the extracellular matrix, and increased permeability of the endothelium. (B) Illustration of the pharmacokinetic model taking into account the EPR effect. The rate constants kp and kd describe exchange with the peripheral volume. The rate constants kepr and kb describe extravasation from circulation into the tumor, and intravasation back into the circulation, respectively. The rate constant kel represents clearance by the kidneys, MPS, and any other non-tumor elimination processes, such that when kb = 0, k10 = kepr + kel where kel is the elimination rate constant. (C) Standard two compartment model with central and peripheral compartments. c1 and c2 represent the drug concentration in blood (central compartment) and normal tissue (peripheral compartment), respectively. The first order rate constant k10 describes all elimination pathways, including clearance by the kidneys, uptake by the MPS, and tumor accumulation. The first order rate constants k12 and k21 describe exchange between the two compartments. Note that kp = k12, kd = k21. (D) Two compartment model defined in terms of the drug amount, where Nbl is the amount of drug in blood (mg), and Np is the amount in peripheral tissue (mg). (E) Three compartment model with the addition of a tumor “compartment” where Nt is the amount of drug in the tumor. Exchange with the tumor is described by the rate constants kepr and kb, respectively. The rate constant kel describes elimination pathways including clearance by the kidneys and uptake by the MPS, but does not include tumor accumulation.

PBPK models are compartmental models like many others, but they have a few advantages over so-called "classical" pharmacokinetic models, which are less grounded in physiology. PBPK models can first be used to abstract and eventually reconcile disparate data (from physicochemical or biochemical experiments, in vitro or in vivo pharmacological or toxicological experiments, etc.) They give also access to internal body concentrations of chemicals or their metabolites, and in particular at the site of their effects, be it therapeutic or toxic. Finally they also help interpolation and extrapolation of knowledge between:

Some of these extrapolations are "parametric" : only changes in input or parameter values are needed to achieve the extrapolation (this is usually the case for dose and time extrapolations). Others are "nonparametric" in the sense that a change in the model structure itself is needed (e.g., when extrapolating to a pregnant female, equations for the foetus should be added).

Owing to the mechanistic basis of PBPK models, another potential use of PBPK modeling is hypothesis testing. For example, if a drug compound showed lower-than-expected oral bioavailability, various model structures (i.e., hypotheses) and parameter values can be evaluated to determine which models and/or parameters provide the best fit to the observed data. If the hypothesis that metabolism in the intestines was responsibility for the low bioavailability yielded the best fit, then the PBPK modeling results support this hypothesis over the other hypotheses evaluated.

As such, PBPK modeling can be used, inter alia, to evaluate the involvement of carrier-mediated transport, clearance saturation, enterohepatic recirculation of the parent compound, extra-hepatic/extra-gut elimination; higher in vivo solubility than predicted in vitro; drug-induced gastric emptying delays; gut loss and regional variation in gut absorption. [15]

Limits and extensions of PBPK modeling

Each type of modeling technique has its strengths and limitations. PBPK modeling is no exception. One limitation is the potential for a large number of parameters, some of which may be correlated. This can lead to the issues of parameter identifiability and redundancy. However, it is possible (and commonly done) to model explicitly the correlations between parameters (for example, the non-linear relationships between age, body-mass, organ volumes and blood flows).

After numerical values are assigned to each PBPK model parameter, specialized or general computer software is typically used to numerically integrate a set of ordinary differential equations like those described above, in order to calculate the numerical value of each compartment at specified values of time (see Software). However, if such equations involve only linear functions of each compartmental value, or under limiting conditions (e.g., when input values remain very small) that guarantee such linearity is closely approximated, such equations may be solved analytically to yield explicit equations (or, under those limiting conditions, very accurate approximations) for the time-weighted average (TWA) value of each compartment as a function of the TWA value of each specified input (see, e.g., [16] [17] ).

PBPK models can rely on chemical property prediction models (QSAR models or predictive chemistry models) on one hand. For example, QSAR models can be used to estimate partition coefficients. They also extend into, but are not destined to supplant, systems biology models of metabolic pathways. They are also parallel to physiome models, but do not aim at modelling physiological functions beyond fluid circulation in detail. In fact the above four types of models can reinforce each other when integrated. [18]

Related Research Articles

<span class="mw-page-title-main">Blood alcohol content</span> Metric of alcohol intoxication

Blood alcohol content (BAC), also called blood alcohol concentration or blood alcohol level, is a measurement of alcohol intoxication used for legal or medical purposes.

<i>In vitro</i> Latin term meaning outside a natural biological environment

In vitro studies are performed with microorganisms, cells, or biological molecules outside their normal biological context. Colloquially called "test-tube experiments", these studies in biology and its subdisciplines are traditionally done in labware such as test tubes, flasks, Petri dishes, and microtiter plates. Studies conducted using components of an organism that have been isolated from their usual biological surroundings permit a more detailed or more convenient analysis than can be done with whole organisms; however, results obtained from in vitro experiments may not fully or accurately predict the effects on a whole organism. In contrast to in vitro experiments, in vivo studies are those conducted in living organisms, including humans, known as clinical trials, and whole plants.

<span class="mw-page-title-main">Pharmacodynamics</span> Area of Academic Study

Pharmacodynamics (PD) is the study of the biochemical and physiologic effects of drugs. The effects can include those manifested within animals, microorganisms, or combinations of organisms.

In pharmacology, bioavailability is a subcategory of absorption and is the fraction (%) of an administered drug that reaches the systemic circulation.

In the physical sciences, a partition coefficient (P) or distribution coefficient (D) is the ratio of concentrations of a compound in a mixture of two immiscible solvents at equilibrium. This ratio is therefore a comparison of the solubilities of the solute in these two liquids. The partition coefficient generally refers to the concentration ratio of un-ionized species of compound, whereas the distribution coefficient refers to the concentration ratio of all species of the compound.

Compartmental models are a very general modelling technique. They are often applied to the mathematical modelling of infectious diseases. The population is assigned to compartments with labels – for example, S, I, or R,. People may progress between compartments. The order of the labels usually shows the flow patterns between the compartments; for example SEIS means susceptible, exposed, infectious, then susceptible again.

In pharmacology, the volume of distribution is the theoretical volume that would be necessary to contain the total amount of an administered drug at the same concentration that it is observed in the blood plasma. In other words, it is the ratio of amount of drug in a body (dose) to concentration of the drug that is measured in blood, plasma, and un-bound in interstitial fluid.

In pharmacology, clearance is a pharmacokinetic parameter representing the efficiency of drug elimination. This is the rate of elimination of a substance divided by its concentration. The parameter also indicates the theoretical volume of plasma from which a substance would be completely removed per unit time. Usually, clearance is measured in L/h or mL/min. The quantity reflects the rate of drug elimination divided by plasma concentration. Excretion, on the other hand, is a measurement of the amount of a substance removed from the body per unit time. While clearance and excretion of a substance are related, they are not the same thing. The concept of clearance was described by Thomas Addis, a graduate of the University of Edinburgh Medical School.

<span class="mw-page-title-main">Biological half-life</span> Time taken for a drug to halve its concentration in blood plasma

Biological half-life is the time taken for concentration of a biological substance to decrease from its maximum concentration (Cmax) to half of Cmax in the blood plasma. It is denoted by the abbreviation .

Absorption is the journey of a drug travelling from the site of administration to the site of action.

<span class="mw-page-title-main">Pharmacokinetics</span> Branch of pharmacology

Pharmacokinetics, sometimes abbreviated as PK, is a branch of pharmacology dedicated to describing how the body affects a specific substance after administration. The substances of interest include any chemical xenobiotic such as pharmaceutical drugs, pesticides, food additives, cosmetics, etc. It attempts to analyze chemical metabolism and to discover the fate of a chemical from the moment that it is administered up to the point at which it is completely eliminated from the body. Pharmacokinetics is based on mathematical modeling that places great emphasis on the relationship between drug plasma concentration and the time elapsed since the drug's administration. Pharmacokinetics is the study of how an organism affects the drug, whereas pharmacodynamics (PD) is the study of how the drug affects the organism. Both together influence dosing, benefit, and adverse effects, as seen in PK/PD models.

GNU MCSim is a suite of simulation software. It allows one to design one's own statistical or simulation models, perform Monte Carlo simulations, and Bayesian inference through (tempered) Markov chain Monte Carlo simulations. The latest version allows parallel computing of Monte Carlo or MCMC simulations.

Cmin is a term used in pharmacokinetics for the minimum blood plasma concentration reached by a drug during a dosing interval, which is the time interval between administration of two doses. This definition is slightly different from Ctrough, the concentration immediately prior to administration of the next dose. Cmin is the opposite of Cmax, the maximum concentration that the drug reaches. Cmin must be above certain thresholds, such as the minimum inhibitory concentration (MIC), to achieve a therapeutic effect.

The plateau principle is a mathematical model or scientific law originally developed to explain the time course of drug action (pharmacokinetics). The principle has wide applicability in pharmacology, physiology, nutrition, biochemistry, and system dynamics. It applies whenever a drug or nutrient is infused or ingested at a relatively constant rate and when a constant fraction is eliminated during each time interval. Under these conditions, any change in the rate of infusion leads to an exponential increase or decrease until a new level is achieved. This behavior is also called an approach to steady state because rather than causing an indefinite increase or decrease, a natural balance is achieved when the rate of infusion or production is balanced by the rate of loss.

An organ-on-a-chip (OOC) is a multi-channel 3-D microfluidic cell culture, integrated circuit (chip) that simulates the activities, mechanics and physiological response of an entire organ or an organ system. It constitutes the subject matter of significant biomedical engineering research, more precisely in bio-MEMS. The convergence of labs-on-chips (LOCs) and cell biology has permitted the study of human physiology in an organ-specific context. By acting as a more sophisticated in vitro approximation of complex tissues than standard cell culture, they provide the potential as an alternative to animal models for drug development and toxin testing.

A Logan plot is a graphical analysis technique based on the compartment model that uses linear regression to analyze pharmacokinetics of tracers involving reversible uptake. It is mainly used for the evaluation of nuclear medicine imaging data after the injection of a labeled ligand that binds reversibly to specific receptor or enzyme.

In vitro to in vivo extrapolation (IVIVE) refers to the qualitative or quantitative transposition of experimental results or observations made in vitro to predict phenomena in vivo, biological organisms.

Leon Aarons is an Australian chemist who researches and teaches in the areas of pharmacodynamics and pharmacokinetics. He lives in the United Kingdom and from 1976 has been a professor of pharmacometrics at the University of Manchester. In the interest of promoting the effective development of drugs, the main focus of his work is optimizing pharmacological models, the design of clinical studies, and data analysis and interpretation in the field of population pharmacokinetics. From 1985 to 2010 Aarons was an editor emeritus of the Journal of Pharmacokinetics and Pharmacodynamics and is a former executive editor of the British Journal of Clinical Pharmacology.

<span class="mw-page-title-main">Lumped parameter model for the cardiovascular system</span>

A lumped parameter cardiovascular model is a zero-dimensional mathematical model used to describe the hemodynamics of the cardiovascular system. Given a set of parameters that have a physical meaning, it allows to study the changes in blood pressures or flow rates throughout the cardiovascular system. Modifying the parameters, it is possible to study the effects of a specific disease. For example, arterial hypertension is modeled increasing the arterial resistances of the model.

In medicinal chemistry, Drug Permeability is an empirical parameter that indicates how quickly a chemical entity or an active pharmaceutical ingredient crosses a biological membrane or another biological barrier to become bioavailable in the body. Drug permeability, together with drug aqueous solubility are the two parameters which defines the fate of the active ingredient after oral administration that ultimately defines its bioavailability. When drug permeability is empirically measured in vitro, it is generally called apparent permeability (Papp) as its absolute value varies according to the method selected for its measurement. Papp is measured in vitro utilizing cellular based barriers such as the Caco-2 model or utilizing artificial biomimetic barriers, such as the Parallel Artificial Membrane Permeation Assay (PAMPA) or the PermeaPad. All these methods are built on an acceptor compartment where the drug solution is placed, a biomimetic barrier and an acceptor compartment, where the drug concentration is quantified over time. By maintaining sink condition, a steady state is reached after a lag time.

References

  1. Gelman, A.; Bois, F.Y.; Jiang, J. (1996). "Physiological pharmacokinetic analysis using population modeling and informative prior distributions". Journal of the American Statistical Association. 91 (436): 1400–1412. CiteSeerX   10.1.1.44.2288 . doi:10.1080/01621459.1996.10476708.
  2. Teorell, T. (1937). "Kinetics of distribution of substances administered to the body". Archives Internationales de Pharmacodynamie et de Thérapie. 57: 205–240.
  3. Gibaldi M., and Perrier D. (1982). Pharmacokinetics, , Second Edition, revised and expanded ed. New-York: Marcel Dekker.
  4. Bischoff, K.B.; Dedrick, R.L.; Zaharko, D.S.; Longstreth, J.A. (1971). "Methotrexate pharmacokinetics". Journal of Pharmaceutical Sciences. 60 (8): 1128–1133. doi: 10.1002/jps.2600600803 . PMID   5127083.
  5. Dedrick, R.L.; Forrester, D.D.; Cannon, J.N.; El Dareer, S.M.; Mellett, L.B. (1973). "Pharmacokinetics of 1-beta-D-arabinofuranosylcytosine (ARA-C) deamination in several species". Biochemical Pharmacology . 22 (19): 2405–2417. doi:10.1016/0006-2952(73)90342-0. PMID   4200888.
  6. Gerlowski, L.E.; Jain, R.K. (1983). "Physiologically based pharmacokinetic modeling: principles and applications". Journal of Pharmaceutical Sciences. 72 (10): 1103–1127. doi:10.1002/jps.2600721003. PMID   6358460.
  7. Droz, P.O.; Guillemin, M. P. (1983). "Human styrene exposure - V. Development of a model for biological monitoring". International Archives of Occupational and Environmental Health. 53 (1): 19–36. doi:10.1007/bf00406174. PMID   6654499. S2CID   21135553.
  8. Lutz, R.J.; Dedrick, R.L.; Tuey, D.; Sipes, G.; Anderson, M.W.; Matthews, H.B. (1984). "Comparison of the pharmacokinetics of several polychlorinated biphenyls in mouse, rat, dog, and monkey by means of a physiological pharmacokinetic model". Drug Metabolism and Disposition. 12 (5): 527–535. PMID   6149901.
  9. Campbell JL Jr, Clewell RA, Gentry PR, Andersen ME, Clewell HJ 3rd (2012). "Physiologically based pharmacokinetic/toxicokinetic modeling". Computational Toxicology. Methods in Molecular Biology. Vol. 929. pp. 439–499. doi:10.1007/978-1-62703-050-2_18. ISBN   978-1-62703-049-6. PMID   23007440.
  10. Jeong, Yoo-Seong; Yim, Chang-Soon; Ryu, Heon-Min; Noh, Chi-Kyoung; Song, Yoo-Kyung; Chung, Suk-Jae (2017). "Estimation of the minimum permeability coefficient in rats for perfusion-limited tissue distribution in whole-body physiologically-based pharmacokinetics". European Journal of Pharmaceutics and Biopharmaceutics. 115: 1–17. doi:10.1016/j.ejpb.2017.01.026. PMID   28215648. S2CID   3455193.
  11. Espié, Pascal; Tytgat, Dominique; Sargentini-Maier, Maria-Laura; Poggesi, Italo; Watelet, Jean-Baptiste (2009). "Physiologically based pharmacokinetics (PBPK)". Drug Metabolism Reviews. 41 (3): 391–407. doi:10.1080/10837450902891360. ISSN   1097-9883. PMID   19601719. S2CID   20219523.
  12. Darwich, AS; Pade D; Ammori BJ; Jamei M; Ashcroft DM; Rostami-Hodjegan A (2012). "A mechanistic pharmacokinetic model to assess modified oral drug bioavailability post bariatric surgery in morbidly obese patients: interplay between CYP3A gut wall metabolism, permeability and dissolution". Journal of Pharmacy and Pharmacology. 64 (7): 1008–1024. doi: 10.1111/j.2042-7158.2012.01538.x . PMID   22686346. S2CID   46066754.
  13. Creative Commons by small.svg  This article incorporates D. Wong, Mao Ye, Martin B. Ulmschneider, Peter C. Searson text available under the CC BY 4.0 license.
  14. Bouvier d'Yvoire, M.; Prieto P.; Blaauboer B.J.; Bois F.; Boobis A.; Brochot C.; Coecke S.; Freidig A.; Gundert-Remy U.; Hartung T.; Jacobs M. N.; Lavé T.; Leahy D.E.; Lennernäs H.; Loizou G.D.; Meek B.; Pease C.; Rowland M.; Spendiff M.; Yang J.; Zeilmaker M. (2007). "Physiologically-based kinetic modelling (PBK modelling): meeting the 3Rs agenda - The report and recommendations of ECVAM Workshop 63a". Alternatives to Laboratory Animals. 35 (6): 661–671. doi: 10.1177/026119290703500606 . PMID   18186671. S2CID   1247463.
  15. Peters, SA (2008). "Evaluation of a generic physiologically based pharmacokinetic model for lineshape analysis". Clin Pharmacokinet. 47 (4): 261–75. doi:10.2165/00003088-200847040-00004. PMID   18336055. S2CID   10598918.
  16. Bogen, K.T. (1988). "Pharmacokinetics for regulatory risk analysis: The case of trichloroethylene". Regulatory Toxicology and Pharmacology . 8 (4): 447–466. doi:10.1016/0273-2300(88)90045-1. PMID   3222486.
  17. Bogen, K.T.; Hall, L.C. (1989). "Pharmacokinetics for regulatory risk analysis: The case of 1,1,1-trichloroethane (methyl chloroform)". Regulatory Toxicology and Pharmacology. 10 (1): 26–50. doi:10.1016/0273-2300(89)90011-1. PMID   2672126.
  18. Bois, F.Y. (2009). "Physiologically-based modelling and prediction of drug interactions". Basic & Clinical Pharmacology & Toxicology. 106 (3): 154–161. doi: 10.1111/j.1742-7843.2009.00488.x . PMID   19912164.

Further references:

Forums

Software

Dedicated software:

General software: