Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2015 Jan 1.
Published in final edited form as: Biopharm Drug Dispos. 2013 Dec 10;35(1):15–32. doi: 10.1002/bdd.1879

Computational Approaches to Analyze and Predict Small Molecule Transport and Distribution at Cellular and Subcellular Levels

Kyoung Ah Min 1, Xinyuan Zhang 1, Jing-yu Yu 1, Gus R Rosania 1,*
PMCID: PMC3947293  NIHMSID: NIHMS538727  PMID: 24218242

Abstract

Quantitative structure-activity relationship (QSAR) studies and mechanistic mathematical modeling approaches have been independently employed for analyzing and predicting the transport and distribution of small molecule chemical agents in living organisms. Both of these computational approaches have been useful to interpret experiments measuring the transport properties of small molecule chemical agents, in vitro and in vivo. Nevertheless, mechanistic cell-based pharmacokinetic models have been especially useful to guide the design of experiments probing the molecular pathways underlying small molecule transport phenomena. Unlike QSAR models, mechanistic models can be integrated from microscopic to macroscopic levels, to analyze the spatiotemporal dynamics of small molecule chemical agents from intracellular organelles to whole organs, well beyond the experiments and training data sets upon which the models are based. Based on differential equations, mechanistic models can also be integrated with other differential equations-based systems biology models of biochemical networks or signaling pathways. Although the origin and evolution of mathematical modeling approaches aimed at predicting drug transport and distribution has occurred independently from systems biology, we propose that the incorporation of mechanistic cell-based computational models of drug transport and distribution into a systems biology modeling framework is a logical next-step for the advancement of systems pharmacology research.

Keywords: Cellular pharmacokinetics, Computational modeling, Drug Transport, Systems pharmacology

The Impact of Cellular Pharmacokinetics on Pharmaceutical Research

Cell-based computational modeling of drug transport and distribution is an active area of research in pharmaceutical sciences, because the subcellular location of drug targets within the three dimensional architecture of cells, tissues and organs ultimately defines the site of action of drug molecules in the body. In multicellular organisms, cell monolayers constitute physical boundaries that separate one tissue or anatomical compartment from another, and determine the movement of molecules between compartments. For example, in the gastrointestinal tract, the monolayer of epithelial cells that lines the lumen of the gut also controls the absorption of nutrients and xenobiotics into the body (Figure 1A). After absorption, nutrients and xenobiotics also have to make their way through hepatocytes in the liver where specialized efflux mechanisms and metabolizing enzymes act to keep xenobiotics from reaching the systemic circulation (1) (Figure 1B). Therefore, the ability to model, analyze and predict the subcellular transport properties of drug-like molecules is an increasingly important area of research in pharmaceutical sciences. Similarly, to the extent that systems pharmacology intends to analyze and predict the spatiotemporal dynamics of the effects of drugs on living systems, modeling the transport and distribution phenomena affecting the concentrations of drug molecules in different sites of an organism may be as important as modeling the effects of drug molecules on biochemical reactions, signal transduction pathways, or other regulatory molecular control networks that are currently the subject of systems biology research.

Figure 1.

Figure 1

Cellular pharmacokinetic models can be used to study the transport and distribution of small molecules in the presence of a transcellular concentration gradient. As drugs are absorbed across the lining of the gastrointestinal tract (A) or as they are metabolized in the hepatocytes (B), cells normally experience a transcellular drug concentration gradient. Outside the cells, drug molecules can exist in free or bound form. Only freely soluble molecules can diffuse across the membranes. Inside the cells, drug molecules can be metabolized, partition into cellular lipids, bind to cellular macromolecules, or become sequestered in organelles. All these mechanisms are influenced by chemical parameters such as the lipophilicity (as reflected in the octanol:water partition coefficient); the ionization states of the molecules (pKa); physiological parameters such as the pH of the different compartments and the electrical potentials across their bound membranes; and, biological parameters such as the volume and number of organelles, member surface areas and the presence of binding, sorption, and active transport mechanisms, as well as chemical transformations catalyzed by metabolic enzymes.

At the level of the whole organism, the development and application of computational models to predict drug transport phenomena has been an active area of pharmacokinetics research for many years. For example, physiologically based pharmacokinetic (PBPK) models are mechanism-based mathematical models that have been used to analyze the concentration and distribution of drugs from the circulation to the different organs, based on the perfusion and permeability characteristics of each organ. In drug development, PBPK models allow predicting drug distribution from one species of organism to another (2-4). Since their origin in the 1970s, a variety of PBPK modeling approaches have been developed to analyze the spatiotemporal dynamics of a wide range of drug absorption, distribution, metabolism and elimination phenomena affecting drug activity and toxicity (5-13). Nevertheless, local concentrations of drug molecules at the microscopic sites where drug molecules interact with their intended (or unintended) biological targets may not be necessarily the same as the concentrations of drug molecules in macroscopic bulk compartments. This is because cells and subcellular organelles are delimited by membranes, and drug transport across these membranes is determined by active and passive transport mechanisms that can facilitate the accumulation of drug molecules on one side of the membrane relative to the other. Beyond PBPK modeling, molecular mechanisms affecting the active transport of drugs into subcellular compartments could be subject of systems pharmacology research. These mechanisms include protein “pumps” that translocate drug molecules from one side of the plasma membrane to the other by using energy derived from ATP hydrolysis. In addition, the passive distribution of drugs across plasma and organelle membranes can be influenced by differences in the local microenvironment of the two compartments separated by membranes. These differences lead to transmembrane pH gradients and electrical potentials affecting the transport and distribution of molecules possessing charged or ionizable functional groups. From the perspective of systems pharmacology, microscopic transmembrane transport phenomena can be considered as the key determinants of whether or not small molecule chemical agents can interact with the molecular circuitry that establishes and controls the structure and function of the organism. If the site of the action of a drug molecule is located in a specific organelle inside a specific cell type, the rate and extent to which the drug may accumulate at that target site can be influenced by active and passive transport mechanisms that may facilitate or prevent a drug from reaching that site. Similarly, if the drug target is localized on an extracellular membrane surface, mechanisms that drive the sequestration of drug molecules inside cells or that facilitate their excretion from the organism can effectively reduce extracellular drug concentrations to the point where the drug may be unable to influence the function of an extracellular target (14-19).

Nevertheless, at this microscopic level, pharmaceutical researchers have also been elaborating computational models to predict the transcellular permeability of small molecule chemical agents and develop small molecule chemical agents targeted to specific sites of action inside cells (5, 7, 8, 20-23). In general, there have been two kinds of mathematical approaches applied towards predicting the subcellular transport and distribution properties of small drug-like molecules: statistics-based QSAR studies and differential equations-based mechanistic modeling. QSAR involves the application of correlation analyses to determine associations between the chemical structures and the subcellular transport and distribution properties of small molecules (8, 21-28). Mechanistic modeling involves using differential equations to represent the transport of molecules as a change in the concentration of molecules between compartments separated by cellular membranes (5, 7, 8). In the following sections, we will discuss how these two different computational approaches originated and evolved, and we will evaluate their potential impact on the field of systems pharmacology. Furthermore, we will also point readers to references of relevant research studies analyzing drug-drug interactions and the interplay between efflux transporters, metabolic enzymes and drug targets while integrating pharmacokinetics and pharmacodynamics at cellular and subcellular levels.

QSAR Studies of Drug Transport and Distribution in Single Cells

The QSAR approach to predictive cellular pharmacokinetics can be considered an extension of classical QSAR studies which have been widely used in pharmaceutical and medicinal chemistry research since the 1960s (29). For classical QSAR studies, multivariate, statistical regression techniques are used to link the results of experimental assays to the chemical structures of large sets of chemical agents tested with these assays. The regression techniques yield an equation that specifies the relationship between a physicochemical or biological property of interest (the dependent variable of the equation) to different input parameters that represent the chemical structures, atomic composition or physicochemical properties of the assayed molecules (the independent variables of the equation). Using a training set of molecules, the resulting regression equations can be used to predict the behavior of untested sets of molecules using the molecule’s chemical structure or physicochemical properties as input (29). Examples of QSAR studies range from the prediction of in vitro solubility (30-34) to the prediction of complex biological properties, such as in vivo bioavailability (35-37).

While several other different QSAR approaches have been developed to link the chemical structure with the physicochemical properties or pharmacological activities of small molecules (38-47), there are fewer examples of QSAR approaches linking the structure of small molecules with their cellular transport and distribution properties (Table 1) (8, 25-27, 48-50). Interestingly, some of the first QSAR models ever published were aimed at predicting drug concentrations in cells. These QSAR models date all the way back to the 1960s. Indeed some of the original QSAR models were elaborated to describe the relationship between hydrophobicity and the biological activity of chemical agents, related to the ability of lipophilic molecules to cross cellular membranes but without partitioning into those membranes (29, 51, 52). Subsequently, these regression-based QSAR models were elaborated and applied in an increasingly sophisticated manner, to study drug disposition and activity in multicompartment (aqueous and lipid) biosystems (28, 53, 54) and to analyze concentration-time profiles in bacterial or mammalian cells (55-59).

Table 1.

Examples of empirical QSAR models for predicting subcellular distribution.

Method Localization Descriptors Number of
compounds
References
QSAR mitochondria logP, Z 41
QSAR lysosome logP, pKa, CBN, Z 50 (27)
QSAR nuclei logP, pKa, Z, CBN,
AI, LCF,
LCF / CBN ratio
44 (25)
QSAR ER logP, pKa, Z, CBN,
AI, LCF
37 (26)
QSAR mitochondria /
non-mitochondria
logP, pKa, Z, CBN,
AI, LCF
109 (60)
MLR
(multiple
linear
regression)
mitochondria logD, Z, α,
MW, PSA
20 (50)
Descriptor mitochondria /
lysosome / nuclei /
cytosol / ER
/ Golgi body /
plasma membrane
/ multiple localization
483 2D and 3D
descriptors
967 (66)

logP: logarithm of the octanol/water partition coefficient

logD: logarithm of the octanol/water partition coefficient at pH 7.4

pKa: negative logarithm of the acidic associate constant

Z: electrical charge

CBN: conjugated bond number

AI: amphilicity index

LCF: the largest conjugated fragment

α: polarizability

MW: molecular weight

PSA: polar surface area

To predict the distribution of small molecules in different organelles of eukaryotic cells, a different kind of QSAR modeling approach was introducedin the 1990s, based on the construction of nested if/then rules (60). With this approach, a simple decision tree-like model, named the Chinese box model, was used to describe the distribution of small drug-like molecules inside cells (27, 48). Initially, this model was applied to 41 cationic probes to study their mitochondrial localization as a function of the logarithm of octanol/water partition coefficient (logP) (48). Based on this model, cationic molecules with logP between 0 and 5 were expected to accumulate in the mitochondrial inner membrane. Cationic molecules with logP < 0 would be excluded outside cells, and with logP > +5 would accumulate preferentially in the plasma membrane (48). A few years later, a similar decision tree-like model was built to analyze the accumulation of small molecules in lysosomes (27). Based on this model, probes were classified into three major categories: 1) those that had 0 < logP < 5 were able to cross the plasma membrane. 2) those that had logP > 15 could also enter the cell and accumulate in lysosomes by adsorptive pinocytosis; and 3) those that had logP < 0, z < 0 could also enter cells and accumulate in lysosomes by fluid phase pinocytosis. Probes in group 1) were classified into two sub-categories: 1) Probes accumulating in lysosomes by ion trapping mechanisms; and, 2) Probes comprised of hydrolysable lipophilic esters, usually weak acids, which were metabolized into free acids by lysosomal esterases and trapped in lysosomes by precipitation in the low pH microenvironment (27). To further extend this decision-tree based approach, additional studies with molecules localizing to other organelles (endoplasmic reticulum, nuclear chromatin, and plasma membrane) have been performed (26, 60, 61).

More recently, an additive factorial logistic regression modeling approach was developed to analyze the subcellular localization of a combinatorial library of organelle-targeting cationic styryl dyes (21, 62). Each styryl dye was synthesized using a chemical conjugation reaction to link two chemical building blocks: an aldehyde building blocks combined with a pyridinium or quinolinium building blocks. By combining 168 aldehyde with 8 pyridinium or quinolinium building blocks, the contributions of each building block to the peak excitation and emission wavelength of 1344 molecules in could the library wascalculated using least squares to minimize an additive, multivariate regression function over all compounds having experimental data (21, 62). Most importantly, using a factorial logistic regression approach, the binary (mitochondrial vs. nonmitochondrial) localization data obtained from live cells incubated with these compounds was also related to quantitative contributions of each aldehyde and pyridinium or quinolinium building blocks. Cross-validation was carried out for both spectral and localization data, to obtain unbiased estimates of prediction performance.

While most QSAR studies have focused on analyzing the steady state distribution of small molecule chemical agents inside cells (for recent reviews, see (63-65)), complementary QSAR approaches have been developed to study the cellular and subcellular transport kinetics of small molecules. In one study, time of exposure was integrated with a QSAR model to capture the kinetics of the drug-receptor interaction based on the law of mass action (28). This “QSTAR” model was developed by applying non-linear regression to a data set of 36 compounds (28). Subsequently, this QSTAR modeling approach was further elaborated by incorporating enzymatic activity, membrane accumulation, non-covalent protein binding, and excretion (56). In a different study, using a combinatorial library of fluorescently-tagged cell-permeant small molecule chemical agents, a nested multi-compartment model was used to analyze the subcellular transport and accumulation kinetics of the compounds between extracellular medium, cytosol and intracellular vesicles (22). Using kinetic image data acquired from cells incubated with the fluorescent compounds at different times after the probes were added to cells, changes in the total intensity and coefficients of variation of pixel intensities were linked to changes in the intracellular concentration of the probes. Using this approach, the partition coefficients from the extracellular medium to the cytosol, and from the cytosol to the intracellular vesicles was inferred and related to the chemical structures of the compounds (22).

Most recently, a database of 967 molecules with published subcellular localization features derived from a survey of the scientific literature was used as a starting point for identifying relationships between the chemical space occupied by small molecules and their reported intracellular localization features (66). Because this is a diverse group of compounds whose localization was determined by a large number of different investigators using a wide variety of different experimental techniques, a rigorous, quantitative structure-localization relationship study was not performed. Nevertheless, by studying the localization of compounds to multiple different organelles, this study identified many interesting, candidate chemical property-subcellular localization relationships that had not been previously noted in QSAR studies focusing on a single target organelle. Amongst the most interesting trends, molecular weight of the compounds was identified as a key variable associated with differences in the subcellular distribution properties of small molecules. More specifically, the chemical property-subcellular localization relationships of the compounds was very different for molecule >500 Daltons as compared to those that were <500 Daltons.

Mechanistic Models of Drug Transport and Distribution in Single Cells

The development of mechanistic models of drug transport and distribution at the single cell level began in the 1970s, with the incorporation of permeability as a parameter variable in physiologically based pharmacokinetic (PBPK) model. In these models, permeability was used to capture differences in the rate of accumulation of drug molecules in the different organs. However, it was not until the mid-1990s that attempts were made to relate the abstract, pharmacokinetic permeability parameter to the physical cell permeability. The realization of the importance of cell permeability in pharmaceutical discovery was spurred by the development of the advanced compartmental absorption and transit model (ACAT) (67). The development of the ACAT model effectively connected the results of in vitro cell-based assays measuring the transport rates of small molecule drugs across cell monolayers to the fraction of an oral dose of drug absorbed in the gastrointestinal tract.

To analyze the results of in vitro cell-based transport assays, compartmental cellular pharmacokinetic models began to be constructed to analyze cellular transport and metabolism properties of small molecules across cell monolayers, in the presence of a transcellular concentration gradient (6-8, 11, 54, 68). By incorporating the Goldman-Hodgkin-Katz equation, differential equation-based cellular pharmacokinetic models have been converted into predictive cell-based transport and distribution models. The Goldman-Hodgkin-Katz equation is derived from the Nernst-Planck equation (equation 1).

J=D[dC(x)dx+C(x)zFRTdV(x)dx] (1)

The first term corresponds to the flux of the neutral species captured by Fick’s law of diffusion. The second term corresponds to the flux of the ionized species which is influenced by the transmembrane electrical potential. D is the diffusion coefficient (area per time unit). F, R, T, and z are the Faraday’s constant, molar gas constant, temperature (in Kelvin) and electric charge, respectively. Assuming the transport direction (denoted by x) being perpendicular to the membrane, then dC(x)/dx reflects the concentration change along the membrane, and dV(x)/dx reflects the voltage change along the membrane. C(x) indicates the concentration at point x. If transmembrane electrical potential is assumed to be constant along the membrane, and the membrane thickness is d, equation 1 is rewritten as equation 2.

J=D[dC(x)dx+C(x)zFRTVd] (2)

Rearranging the terms in equation 2 leads to equation 3:

1=dC(x)dxJDC(x)zFRTVd (3)

Letting N = zFV/RT, and integrating from x = 0 to d yields equation 4:

0ddx=0ddC(x)dxJD+NC(x)ddxd=dN[inNdC(d)+JDinNdC(0)+JD] (4)

Accordingly, the flux of a charged molecules across a biomembrane with a transmembrane electrical potential and concentration gradient is captured by equation 5:

J=DdNeN1(C(0)eNC(d)) (5)

where D/d captures the permeability of the molecule across the membrane (length per time units). C(0) is the concentration at outer membrane surface, and C(d) is the concentration at inner membrane surface.

To facilitate predictions, the flux of a molecule across a membrane can be expressed with Fick’s equation (equation 6).

Jn=Pn(Co,nCi,n) (6)

where Pn is the membrane permeability which can be estimated as a function of the octanol:water partition coefficient of the neutral form of the molecule, Cn is the concentration of neutral form of the molecule with the subscripts o and i indicating the directions of the flux, J, from outside to inside compartment. Similarly, for charged molecules, the net fluxes of passive diffusion across membranes can be expressed using Pd as the membrane permeability of the ionized form of the molecule which can also be estimated as a function of the octanol:water partition coefficient , with equation 7 (derived from equation 5)

Jd=PdNeN1(Co,dCi,deN) (7)

For equation 7, subscript d indicates the ionized form of the molecule. Co,d and Ci,d are the concentration of ionized forms of the molecule outside and inside, respectively. Combining equations 6 and 7 (Fick’s and Goldman-Hodkin-Katz equation) and considering as the thermodynamic activity (a) of the different protonated states of a molecule may differ from their concentration depending on the pH, ionic strength, and binding to macromolecules and lipids present in the local microenvironment, the net fluxes of molecule weakly basic or weakly acidic molecule across lipid membranes delimiting cells and the various intracellular organelles can be described with equation 8.

J=Pn(ao,nai,n)+PdNeN1(ao,dai,deN) (8)

This equation directly captures the transmembrane fluxes of neutral and ionized species of monovalent weak acids and bases and can be further elaborated to capture dibasic, diacidic and zwitterionic small molecules (5, 66, 69). For simulating subcellular transport phenomena with this equation, input parameters can be systematically varied, to capture the behavior of molecules with varying physicochemical properties, as well as to study the effect of variations in cell morphology and physiology on cellular pharmacokinetics. This equation has been parameterized to simulate the subcellular transport and disposition properties of small molecules in single cells suspended in a homogeneous extracellular drug concentration (Figure 2A) as well as in attached cell monolayers surrounded by a transcellular concentration gradient (Figure 2B) (5, 7, 8, 69, 70).

Figure 2.

Figure 2

Passive transport mechanisms can be modeled to predict the accumulation and distribution of the weakly basic small molecule inside cells surrounded by a homogeneous drug concentration (A) or in the presence of a transcellular concentration gradient (B). For mechanistic modeling, the different ionization states of the molecule is calculated using the Henderson-Hasselbalch equation (1). Transport of the neutral or charged (ionized) species of the molecule from the surrounding medium or the apical compartment into the cells (2) can be modeled with the Fick and Goldman-Hodgkin-Katz equations. From the cytosol, transport into lysosomes (3) or mitochondria (4), as well as transport from the cytosol into the basolateral compartment can also be modeled with the Goldman-Hodgkin-Katz equation.

Role of Mechanistic Cellular Pharmacokinetic Modeling in Pharmacology

From an empirical pharmacological perspective, cellular pharmacokinetic models have been mostly used to analyze the interaction of small molecule chemical agents with specific intracellular drug targets, in the context of passive and active transport mechanisms affecting the transport of the molecules between extracellular and intracellular compartments (Table 2). One excellent example of such an application involved studying the effect of P-glycoprotein (P-gp) on the intracellular binding of paclitaxel to microtubules (20, 71, 72). The model took account into saturable binding to extracellular proteins, saturable and nonsaturable binding to intracellular components, cell density variation, and changes in tubulin concentration as a function of paclitaxel concentration. First, the model was validated in human breast MCF7 tumor cells, which had negligible P-glycoprotein expression, after which the effect of P-glycoprotein mediated efflux was added into the model and validated in human breast carcinoma cell line derived from MCF7 cells transfected with mdr1 (72). As a follow up study, a parametric analysis was performed to study the differential effects of extracellular drug concentration, intracellular drug binding capacity and affinity, and P-glycoprotein expression level on the intracellular drug accumulation (20). The study showed that the four biological factors determined paclitaxel intracellular concentration interdependently. Among the four factors, extracellular concentration was the most sensitive factor, followed by intracellular binding capacity and affinity. The effect of P-glycoprotein expression was relatively minor, suggesting that to improve clinical efficacy, effective delivery of paclitaxel to tumor cells was more important than other factors, such as inhibition of P-glycoprotein efflux.

Table 2.

Examples of mechanistic cellular pharmacokinetic models

Drugs
/ molecules
Model Components Cell Type Relation
to
Systemic
PK/PD
References
Monovalent
small
molecules
Passive transcellular
transport and
Subcellular
organelles
Epithelial and
round shaped
non-polarized
cells
Absorption,
tissue
distribution
(5, 7, 8, 69)
Paclitaxel P-gp efflux,
extracellular /
intracellular binding
Human breast
cancer cell lines
MCF7 and
BC19
Uptake to
cancer cells
(20, 71, 72)
Substrates of
multiple
transporters
Active uptake,
passive diffusion,
nonspecific binding
Chinese hamster
ovary (CHO)
cells
overexpressing
OATP1A1 or
OATP1B1 and
rat hepatocytes
Liver
clearance
(74)
Ranitidine Uptake and efflux
transporters,
paracellular and
Transcellular transport
Caco-2 Absorption (75)

Baicalein
Passive diffusion,
cellular binding,
transporters and
enzymes

Caco-2 or other
similar in vitro
system

Absorption,
metabolism

(6)
GCSF Endosomal
trafficking, PK/PD
GCSF-
dependent
human
suspension cell
line: OCI/AML1
Cell-
mediated
clearance,
link with
PD modeling
(138)

Another important application of empirical, mechanistic cellular pharmacokinetic modeling involved the determination of Vmax and Km values of drug transporters and enzymes with an intracellular site of action, using results of experiments performed with intact cells (73-75). Based on extracellular drug concentrations, the mechanism-based empirical modeling approach can facilitate an estimation of intracellular Vmax and Km values, or elementary rate constants of enzymes including drug transporters in their natural, intracellular microenvironments (76-78). Using a mechanistic cell-based transport and distribution model to guide interpretation of experimental measurements, parameters associated with the effect of an enzyme substrate or inhibitor can be estimated by fitting the experimental data with a model, using nonlinear least-squares regression. For modeling drug metabolism, a catenary model based on the compartmental analysis was developed to analyze the activity of intracellular, drug metabolizing enzymes (6). This model captured the mechanisms of passive diffusion, cellular binding, carrier-mediated and efflux transporter-mediated transport, in addition to metabolic activity. The model was applied to study the transport and metabolism of baicalein inside cells (6). This empirical, compartmental modeling approach has also been applied to facilitate the design and interpretation of experiments exploring the role of transporters and metabolic enzymes in limiting drug absorption and has helped explore the molecular mechanisms responsible for drug-drug interactions (11, 79-84). Empirical, compartmental, cellular pharmacokinetic modeling approaches are increasingly being used to help interpret the interplay between drug molecules, drug transporters and metabolic enzymes, from a molecular, mechanistic perspective (85-87).

Arguably, for the future development of systems pharmacology, application of differential equations-based, mechanistic mathematical models will become increasingly important, because many biochemical and signaling networks are localized at specific organelles inside the cell. For example, mitochondria are involved in the regulation of apoptosis, and thus are considered as important target sites for anticancer agents (88-90). Mitochondria are also the sites of energy production, with the electron transport chain located in the mitochondrial inner membrane generating a marked electrical potential and pH gradient (91, 92). Many lipophilic cations have been observed accumulating in mitochondria as a function of the transmembrane electrical potential, which can be predicted with compartmental, cellular pharmacokinetic model using the Goldman-Hodgkin-Katz equation (21, 91, 93-95). Examples of these molecules include rhodamine 123 (91, 93, 96, 97), F16 (88, 89), and styryl molecules (98-101). Differences in mitochondrial membrane potential and pH gradient explain differences in the toxicity of cationic compounds on different cell lines from different origins (91).

Like mitochondria, lysosomes are another example of an organelle that is interesting from a systems pharmacology perspective. Biologically, mutations in genes affecting lysosomal function lead to a range of protein and membrane accumulation defects that are characteristics of lysosomal storage diseases (102). Unlike mitochondria, the intralumenal pH of lysosomes is acidic (103) while the cytosolic pH is near neutrality (18). The low pH of lysosomes is caused by the activity of a proton ATPase (103). As a result, weakly basic molecules tend to accumulate inside lysosomes by pH-dependent ion-trapping mechanism (104): for weakly basic molecules with pKa close to physiological pH, they exist predominantly as neutral species in cytosol (pH ~ 7.2). After neutral molecules enter the acidic subcellular organelles, they become protonated due to the acidic environment. Generally, the lipophilicity may greatly differ between neutral species and ionized species (8). Therefore, after entering the acidic compartment, transmembrane permeability of the molecules is reduced due to the protonation, and accumulation is induced. Because of the ion-trapping mechanism, many clinically-useful drugs accumulate in lysosomes, including as antimalarial drug chloroquine (105, 106), as well as many antidepressant drugs (107).

Furthermore, while mechanistic cellular pharmacokinetic models have been useful to describe organelle targeting, drug bioaccumulation inside cells, and also predict the effect of microenvironments on drug transport and distribution inside cells (5, 7, 8, 69, 70), these models can also be used as building blocks to study drug transport in higher order cellular organizations, from the tissue to the organ level (108, 109). The ability to model the transport properties of small molecules at the level of single cells, and predict the pharmacokinetics and biodistribution properties of drug from organelles to cells to tissues to organisms could be used to predict systemic pharmacokinetic parameters, such as volume of distribution. Indeed, correlative in vitro and in vivo studies suggest that the volume of distribution and the intracellular accumulation of small molecules are related to each other (110-114): Propranolol, a drug with a high volume of distribution, mostly accumulates in association with mitochondria (110). Mefloquine, another drug with very large volume of distribution, extensively accumulates in lysosomes (114). Other basic drugs with large volumes of distribution that accumulate in lysosomes are chlorpromazine, imipramine, and biperiden (111, 112, 115). Extensive distribution in the lung has been observed for many lipophilic bases (115-120) and lysosomal ion trapping has been proposed as a mechanism contributing to high drug accumulation in the lung (113, 115, 119, 120). Interestingly, lysosomal volume changes have been independently implicated in various drug-drug interactions involving lysosomotropic compounds (121, 122).

Illustrating the application of mechanistic cellular pharmacokinetic models for predicting small molecule transport and distribution from the microscopic to the macroscopic levels, the passive transport of small drug-like molecules in the lung was modeled based on differences in the physiological, anatomical, and histological organization of airways and alveoli (Figure 3A). In this model, the transmembrane transport was modeled using the Fick and Goldman-Hodgkin-Katz equations to describe the transport of small molecules across the various membrane bound compartments separating the lumen of the airway from the circulation. This model was used to predict the rate of absorption of passively diffusing molecules from the airway surface lining liquid, across the epithelial cells, the interstitium, the capillary endothelial cells and ultimately into the blood (123). Furthermore, the predictions of the model were tested after intravenous vs. intratracheal injection of fluorescent compounds (124) (Figure 3B), reflected in differences in the local distribution behavior and systemic absorption profiles of the fluorescent compounds with different physicochemical properties in airways and alveoli (Figure 4). These results illustrate how mechanistic cellular pharmacokinetic models can be used to analyze local difference in the transport and distribution properties of small molecule chemical agents from the individual organelles to a whole organ.

Figure 3.

Figure 3

Mechanistic modeling facilitates predicting the transport and distribution properties of small molecules from the microscopic to the macroscopic level. At the macroscopic level (A) the anatomy of an organ can be modeled geometrically as surfaces comprised of cylinders. Histologically, the walls of these cylinders can be modeled as concentric layers of different cell types. By integrating the transport of small molecules across these multilayered organizations, it is possible to calculate the changes in drug concentrations over time at both local and systemic levels. (B) Using fluorescent molecules as probes of transport phenomena, the cellular labeling patterns of the probes can be related to predictions based on the models. In this case, the micrograph shows the distribution of Mitotracker Red (red), Hoechst 33342 (blue) and Rhodamine 1,2,3 (green) in the lungs, after the dyes have been administered directly into the airways. Relative to the other two fluorescent probes, Mitotracker Red was retained inside the epithelial cells that line the airways, as predicted by its transmembrane transport properties. Hypotheses are used to guide the design of experiments and build models. The models are used to make predictions which may or may not be validated by experiments, leading to formulation of new hypotheses leading to further development of the models.

Figure 4.

Figure 4

Transmembrane mass transport equations can be used to build increasingly complex models from single cells to multicellular organizations. The same equations can be used to capture the transport into single cells and intracellular, membrane-bound organelles, as can be used to capture the transport across multicellular tissue layers. In essence, the mass of drug that diffuses from one side of the membrane to the other over time is calculated by multiplying the flux of molecules across the membrane (J) times the area of the membrane (A). For passively diffusing molecules, the flux can be calculated based on the ionization states of the molecules, the transcellular concentration gradients of freely soluble molecule species, and the membrane permeability of each ionized or neutral molecular species. Membrane surface areas are biologically-defined, measurable parameters that reflect the overall number, volume, shape, spatial arrangement and membrane organization of the cells, as well as the number, volume, shape spatial arrangement and membrane organization of organelles.

Conclusions and Future Outlook

An important consideration to evaluate the scientific value of a computational approach involves assessing the predictive accuracy of the approach. Comparing QSAR with differential equations-based mechanistic modeling approaches (60) in terms of predicting mitochondrial accumulation, both methods have been found to predict the mitochondrial localization of lipophilic cations and lipophilic weak acids with some accuracy (60). However, for electrically neutral species, including zwitterions, predictions with the empirical QSAR model were better than with the mechanistic model (60). For lipophilic cations of partially ionized bases, the mechanistic model failed to predict mitochondriotropic behavior in eight of nine cases, while QSAR model successfully predicted all nine cases (60). Thus, the QSAR approach may be advantageous when there is a good training set of molecules and a very specific target organelle. In a different study, the prediction performances of mechanistic and empirical QSAR models were compared using a dataset of toxicities against Tetrahymena pyriformis (125). Based on this toxicity study, the mechanistic model had slightly higher predictive accuracy than the empirical models (based on a leave-one-out cross-validation and two types of leave-several-out cross-validation approaches) (125). More importantly, this second study suggests that the mechanism-based model performed better than the empirical models for compounds falling outside the parameters space of the training set (125).

Nevertheless, as a tool to aid in experimental design and data interpretation, differential equations-based cellular pharmacokinetic models can be more useful than statistical QSAR models. First, mechanistic predictive cellular pharmacokinetic models that are based on differential equations can be readily integrated with differential equation-based systems biology models capturing the interaction of biological components inside cells. Using a mechanistic cellular pharmacokinetic model, concentrations of small molecule drugs can be calculated inside cells, and these concentrations can be incorporated into a systems biology model to calculate the effect on a biochemical reaction or signal transduction pathway. Second, as we have argued before, mechanistic models can be integrated across multiple scales (126), to make predictions about differences in drug transport and drug action at different sites within the same organ. Lastly, from a biological perspective, mechanistic pharmacokinetic models can be integrated with systems biology models and used to frame quantitative hypothesis about the effects of exogenous, small molecule activators or inhibitors on cell structure and functionfrom the local control of signal transduction pathways (127) to the global control of cell population dynamics. (128-131). Furthermore, the inability of a mechanistic model to capture a molecule’s behavior is often an indication of an unknown mechanism affecting drug transport and distribution (132-136). This can point to additional experimental studies to identify new biological phenomena affecting drug transport behavior (66, 137).

To conclude, a variety of computational approaches have been developed to analyze, interpret and predict the transport and distribution properties of small molecule chemical agents inside cells. These approaches can be broadly classified either as statistics-based QSAR or as differential equations-based mechanistic models. While statistics-based, predictive QSAR models are simple and straightforward in terms of pointing to chemical modifications that may be useful for drug development purposes, mechanistic models based on differential equations have many additional advantages in the context of systems pharmacology. Mechanistic models that are based on differential equations can be readily integrated with other differential equation-based models of drug-target interactions. Furthermore, these integrated models can be subsequently extended by adding new mechanism, and can be used to model compounds outside the training dataset, and can be used as a starting point for discovering new mechanisms. In line with one of the primary objectives of systems pharmacology which is to predict the effects of exogenous chemical agents on living systems, many mechanistic cell-based pharmacokinetic models are already being used to help analyze and predict the transport, distribution, metabolism and excretion properties of exogenous chemical agents from the subcellular to the organism levels. Therefore, further integration of differential equations-based cellular pharmacokinetics with mechanistic systems biology modeling research seems like a natural next-step in the advancement of systems pharmacology.

Acknowledgments

This work was supported by NIH grant R01 GM078200 to G.R.R. None of the authors have any conflicts of interest.

Sponsor: This work was supported by funding from the National Institute of General Medical Sciences (grant R01 GM078200 to G.R.R.).

References

  • 1.Chu X, Korzekwa K, Elsby R, Fenner K, Galetin A, Lai Y, et al. Intracellular drug concentrations and transporters: measurement, modeling, and implications for the liver. Clin Pharmacol Ther. 2013;94:126–141. doi: 10.1038/clpt.2013.78. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Theil FP, Guentert TW, Haddad S, Poulin P. Utility of physiologically based pharmacokinetic models to drug development and rational drug discovery candidate selection. Toxicol Lett. 2003;138:29–49. doi: 10.1016/s0378-4274(02)00374-0. [DOI] [PubMed] [Google Scholar]
  • 3.Lave T, Parrott N, Grimm HP, Fleury A, Reddy M. Challenges and opportunities with modelling and simulation in drug discovery and drug development. Xenobiotica. 2007;37:1295–1310. doi: 10.1080/00498250701534885. [DOI] [PubMed] [Google Scholar]
  • 4.Jones HM, Gardner IB, Watson KJ. Modelling and PBPK simulation in drug discovery. Aaps Journal. 2009;11:155–166. doi: 10.1208/s12248-009-9088-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Zhang X, Shedden K, Rosania GR. A cell-based molecular transport simulator for pharmacokinetic prediction and cheminformatic exploration. Mol Pharm. 2006;3:704–716. doi: 10.1021/mp060046k. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Sun H, Zhang L, Chow EC, Lin G, Zuo Z, Pang KS. A catenary model to study transport and conjugation of baicalein, a bioactive flavonoid, in the Caco-2 cell monolayer: demonstration of substrate inhibition. J Pharmacol Exp Ther. 2008;326:117–126. doi: 10.1124/jpet.108.137463. [DOI] [PubMed] [Google Scholar]
  • 7.Trapp S, Rosania GR, Horobin RW, Kornhuber J. Quantitative modeling of selective lysosomal targeting for drug design. Eur Biophys J. 2008;37:1317–1328. doi: 10.1007/s00249-008-0338-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Trapp S, Horobin RW. A predictive model for the selective accumulation of chemicals in tumor cells. Eur Biophys J. 2005;34:959–966. doi: 10.1007/s00249-005-0472-1. [DOI] [PubMed] [Google Scholar]
  • 9.Watanabe T, Kusuhara H, Maeda K, Shitara Y, Sugiyama Y. Physiologically based pharmacokinetic modeling to predict transporter-mediated clearance and distribution of pravastatin in humans. J Pharmacol Exp Ther. 2009;328:652–662. doi: 10.1124/jpet.108.146647. [DOI] [PubMed] [Google Scholar]
  • 10.Kato M, Shitara Y, Sato H, Yoshisue K, Hirano M, Ikeda T, et al. The quantitative prediction of CYP-mediated drug interaction by physiologically based pharmacokinetic modeling. Pharm Res. 2008;25:1891–1901. doi: 10.1007/s11095-008-9607-2. [DOI] [PubMed] [Google Scholar]
  • 11.Liu L, Pang KS. An integrated approach to model hepatic drug clearance. European Journal of Pharmaceutical Sciences. 2006;29:215–230. doi: 10.1016/j.ejps.2006.05.007. [DOI] [PubMed] [Google Scholar]
  • 12.Rodgers T, Leahy D, Rowland M. Physiologically based pharmacokinetic modeling 1: predicting the tissue distribution of moderate-to-strong bases. J Pharm Sci. 2005;94:1259–1276. doi: 10.1002/jps.20322. [DOI] [PubMed] [Google Scholar]
  • 13.Rodgers T, Rowland M. Physiologically based pharmacokinetic modelling 2: predicting the tissue distribution of acids, very weak bases, neutrals and zwitterions. J Pharm Sci. 2006;95:1238–1257. doi: 10.1002/jps.20502. [DOI] [PubMed] [Google Scholar]
  • 14.Honegger UE, Zuehlke RD, Scuntaro I, Schaefer MH, Toplak H, Wiesmann UN. Cellular accumulation of amiodarone and desethylamiodarone in cultured human cells. Consequences of drug accumulation on cellular lipid metabolism and plasma membrane properties of chronically exposed cells. Biochem Pharmacol. 1993;45:349–356. doi: 10.1016/0006-2952(93)90070-d. [DOI] [PubMed] [Google Scholar]
  • 15.Reasor MJ, McCloud CM, Beard TL, Ebert DC, Kacew S, Gardner MF, et al. Comparative evaluation of amiodarone-induced phospholipidosis and drug accumulation in Fischer-344 and Sprague-Dawley rats. Toxicology. 1996;106:139–147. doi: 10.1016/0300-483x(95)03175-f. [DOI] [PubMed] [Google Scholar]
  • 16.Chen VY, Rosania GR. The great multidrug-resistance paradox. ACS Chem Biol. 2006;1:271–273. doi: 10.1021/cb600215q. [DOI] [PubMed] [Google Scholar]
  • 17.Duvvuri M, Konkar S, Hong KH, Blagg BS, Krise JP. A new approach for enhancing differential selectivity of drugs to cancer cells. ACS Chem Biol. 2006;1:309–315. doi: 10.1021/cb6001202. [DOI] [PubMed] [Google Scholar]
  • 18.Duvvuri M, Krise JP. Intracellular drug sequestration events associated with the emergence of multidrug resistance: a mechanistic review. Front Biosci. 2005;10:1499–1509. doi: 10.2741/1634. [DOI] [PubMed] [Google Scholar]
  • 19.Chen VY, Posada MM, Zhao L, Rosania GR. Rapid doxorubicin efflux from the nucleus of drug-resistant cancer cells following extracellular drug clearance. Pharm Res. 2007;24:2156–2167. doi: 10.1007/s11095-007-9369-2. [DOI] [PubMed] [Google Scholar]
  • 20.Jang SH, Wientjes MG, Au JL. Interdependent effect of P-glycoprotein-mediated drug efflux and intracellular drug binding on intracellular paclitaxel pharmacokinetics: application of computational modeling. J Pharmacol Exp Ther. 2003;304:773–780. doi: 10.1124/jpet.102.044172. [DOI] [PubMed] [Google Scholar]
  • 21.Shedden K, Brumer J, Chang YT, Rosania GR. Chemoinformatic analysis of a supertargeted combinatorial library of styryl molecules. J Chem Inf Comput Sci. 2003;43:2068–2080. doi: 10.1021/ci0341215. [DOI] [PubMed] [Google Scholar]
  • 22.Chen VY, Khersonsky SM, Shedden K, Chang YT, Rosania GR. System dynamics of subcellular transport. Mol Pharm. 2004;1:414–425. doi: 10.1021/mp049916t. [DOI] [PubMed] [Google Scholar]
  • 23.Balaz S. Modeling kinetics of subcellular disposition of chemicals. Chem Rev. 2009;109:1793–1899. doi: 10.1021/cr030440j. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Shedden K, Li Q, Liu F, Chang YT, Rosania GR. Machine vision-assisted analysis of structure-localization relationships in a combinatorial library of prospective bioimaging probes. Cytometry A. 2009;75:482–493. doi: 10.1002/cyto.a.20713. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Horobin RW, Stockert JC, Rashid-Doubell F. Fluorescent cationic probes for nuclei of living cells: why are they selective? A quantitative structure-activity relations analysis. Histochemistry and Cell Biology. 2006;126:165–175. doi: 10.1007/s00418-006-0156-7. [DOI] [PubMed] [Google Scholar]
  • 26.Colston J, Horobin RW, Rashid-Doubell F, Pediani J, Johal KK. Why fluorescent probes for endoplasmic reticulum are selective: an experimental and QSAR-modelling study. Biotech Histochem. 2003;78:323–332. doi: 10.1080/10520290310001646659. [DOI] [PubMed] [Google Scholar]
  • 27.Rashid F, Horobin RW. Accumulation of fluorescent non-cationic probes in mitochondria of cultured cells: observations, a proposed mechanism, and some implications. J Microsc. 1991;163:233–241. doi: 10.1111/j.1365-2818.1991.tb03175.x. [DOI] [PubMed] [Google Scholar]
  • 28.Balaz S, Sturdik E, Rosenberg M, Augustin J, Skara B. Kinetics of drug activities as influenced by their physico-chemical properties: antibacterial effects of alkylating 2-furylethylenes. J Theor Biol. 1988;131:115–134. doi: 10.1016/s0022-5193(88)80125-5. [DOI] [PubMed] [Google Scholar]
  • 29.Hansch C, Leo A, Exploring Q. Fundamentals and Applications in Chemistry and Biology, ACS Professional Reference Book. American Chemical Society; Washington, DC: 1995. [Google Scholar]
  • 30.Goller AH, Hennemann M, Keldenich J, Clark T. In silico prediction of buffer solubility based on quantum-mechanical and HQSAR- and topology-based descriptors. Journal of Chemical Information and Modeling. 2006;46:648–658. doi: 10.1021/ci0503210. [DOI] [PubMed] [Google Scholar]
  • 31.Huuskonen J, Salo M, Taskinen J. Aqueous solubility prediction of drugs based on molecular topology and neural network modeling. J Chem Inf Comput Sci. 1998;38:450–456. doi: 10.1021/ci970100x. [DOI] [PubMed] [Google Scholar]
  • 32.Delaney JS. Predicting aqueous solubility from structure. Drug Discov Today. 2005;10:289–295. doi: 10.1016/S1359-6446(04)03365-3. [DOI] [PubMed] [Google Scholar]
  • 33.Chen XQ, Cho SJ, Li Y, Venkatesh S. Prediction of aqueous solubility of organic compounds using a quantitative structure-property relationship. J Pharm Sci. 2002;91:1838–1852. doi: 10.1002/jps.10178. [DOI] [PubMed] [Google Scholar]
  • 34.Tantishaiyakul V. Prediction of the aqueous solubility of benzylamine salts using QSPR model. J Pharm Biomed Anal. 2005;37:411–415. doi: 10.1016/j.jpba.2004.11.005. [DOI] [PubMed] [Google Scholar]
  • 35.Brendel K, Comets E, Laffont C, Laveille C, Mentre F. Metrics for external model evaluation with an application to the population pharmacokinetics of gliclazide. Pharm Res. 2006;23:2036–2049. doi: 10.1007/s11095-006-9067-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Ekins S, Rose J. In silico ADME/Tox: the state of the art. J Mol Graph Model. 2002;20:305–309. doi: 10.1016/s1093-3263(01)00127-9. [DOI] [PubMed] [Google Scholar]
  • 37.Bai JP, Utis A, Crippen G, He HD, Fischer V, Tullman R, et al. Use of classification regression tree in predicting oral absorption in humans. J Chem Inf Comput Sci. 2004;44:2061–2069. doi: 10.1021/ci040023n. [DOI] [PubMed] [Google Scholar]
  • 38.McFarland JW, Berger CM, Froshauer SA, Hayashi SF, Hecker SJ, Jaynes BH, et al. Quantitative structure-activity relationships among macrolide antibacterial agents: in vitro and in vivo potency against Pasteurella multocida. J Med Chem. 1997;40:1340–1346. doi: 10.1021/jm960436i. [DOI] [PubMed] [Google Scholar]
  • 39.Blake JF. Chemoinformatics - predicting the physicochemical properties of 'drug-like' molecules. Curr Opin Biotechnol. 2000;11:104–107. doi: 10.1016/s0958-1669(99)00062-2. [DOI] [PubMed] [Google Scholar]
  • 40.Ghose AK, Crippen GM. Quantitative structure-activity relationship by distance geometry: quinazolines as dihydrofolate reductase inhibitors. J Med Chem. 1982;25:892–899. doi: 10.1021/jm00350a003. [DOI] [PubMed] [Google Scholar]
  • 41.Ghose AK, Crippen GM. Use of physicochemical parameters in distance geometry and related three-dimensional quantitative structure-activity relationships: a demonstration using Escherichia coli dihydrofolate reductase inhibitors. J Med Chem. 1985;28:333–346. doi: 10.1021/jm00381a013. [DOI] [PubMed] [Google Scholar]
  • 42.Marrero Ponce Y, Cabrera Perez MA, Romero Zaldivar V, Gonzalez Diaz H, Torrens F. A new topological descriptors based model for predicting intestinal epithelial transport of drugs in Caco-2 cell culture. J Pharm Pharm Sci. 2004;7:186–199. [PubMed] [Google Scholar]
  • 43.Ghose AK, Crippen GM. Modeling the benzodiazepine receptor binding site by the general three-dimensional structure-directed quantitative structure-activity relationship method REMOTEDISC. Mol Pharmacol. 1990;37:725–734. [PubMed] [Google Scholar]
  • 44.Svetnik V, Liaw A, Tong C, Culberson JC, Sheridan RP, Feuston BP. Random forest: a classification and regression tool for compound classification and QSAR modeling. J Chem Inf Comput Sci. 2003;43:1947–1958. doi: 10.1021/ci034160g. [DOI] [PubMed] [Google Scholar]
  • 45.Wildman SA, Crippen GM. Three-dimensional molecular descriptors and a novel QSAR method. J Mol Graph Model. 2002;21:161–170. doi: 10.1016/s1093-3263(02)00147-x. [DOI] [PubMed] [Google Scholar]
  • 46.Hansch C. The QSAR paradigm in the design of less toxic molecules. Drug Metabolism Reviews. 1984;15:1279–1294. doi: 10.3109/03602538409029960. [DOI] [PubMed] [Google Scholar]
  • 47.Leo A, Hansch C, Church C. Comparison of parameters currently used in the study of structure-activity relationships. J Med Chem. 1969;12:766–771. doi: 10.1021/jm00305a010. [DOI] [PubMed] [Google Scholar]
  • 48.Rashid F, Horobin RW. Interaction of molecular probes with living cells and tissues. Part 2. A structure-activity analysis of mitochondrial staining by cationic probes, and a discussion of the synergistic nature of image-based and biochemical approaches. Histochemistry. 1990;94:303–308. doi: 10.1007/BF00266632. [DOI] [PubMed] [Google Scholar]
  • 49.Horobin RW, Flemming L. Structure-staining relationships in histochemistry and biological staining. II. Mechanistic and practical aspects of the staining of elastic fibres. J Microsc. 1980;119:357–372. doi: 10.1111/j.1365-2818.1980.tb04107.x. [DOI] [PubMed] [Google Scholar]
  • 50.Durazo SA, Kadam RS, Drechsel D, Patel M, Kompella UB. Brain mitochondrial drug delivery: influence of drug physicochemical properties. Pharm Res. 2011;28:2833–2847. doi: 10.1007/s11095-011-0532-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Hansch C, Maloney PP, Fujita T, Muir RM. Correlation of biological activity of phenoxyacetic acids with Hammett substituent constants and partition coefficients. 1962 [Google Scholar]
  • 52.Fujita T, Iwasa J, Hansch C. A new substituent constant, π, derived from partition coefficients. Journal of the American Chemical Society. 1964;86:5175–5180. [Google Scholar]
  • 53.McFarland JW. On the parabolic relationship between drug potency and hydrophobicity. J Med Chem. 1970;13:1192–1196. doi: 10.1021/jm00300a040. [DOI] [PubMed] [Google Scholar]
  • 54.Baláž Š , Šturdík E, Tichý M. Hansch approach and kinetics of drug activities. Quantitative Structure-Activity Relationships. 1985;4:77–81. [Google Scholar]
  • 55.Baláz S, Wiese M, Dubovský P, Baricic P, Seydel J. Quantitative and explicit structure-time-activity relations. Progress in clinical and biological research. 1989;291:37. [PubMed] [Google Scholar]
  • 56.Balaz S, Wiese M, Seydel JK. A kinetic description of the fate of chemicals in biosystems. Sci Total Environ. 1991;109-110:357–375. doi: 10.1016/0048-9697(91)90190-p. [DOI] [PubMed] [Google Scholar]
  • 57.Balaz S, Wiese M, Seydel JK. A time hierarchy-based model for kinetics of drug disposition and its use in quantitative structure-activity relationships. J Pharm Sci. 1992;81:849–857. doi: 10.1002/jps.2600810902. [DOI] [PubMed] [Google Scholar]
  • 58.Balaz S. Model-based description of distribution of chemicals in biosystems for the continuous dose. SAR QSAR Environ Res. 1995;4:177–187. doi: 10.1080/10629369508029915. [DOI] [PubMed] [Google Scholar]
  • 59.Schultz T, Hermens J. Kinetics of subcellular distribution of multiply ionizable compounds: a mathematical description and its use in QSAR. Journal of theoretical biology. 1996;178:7–16. [Google Scholar]
  • 60.Horobin RW, Trapp S, Weissig V. Mitochondriotropics: a review of their mode of action, and their applications for drug and DNA delivery to mammalian mitochondria. J Control Release. 2007;121:125–136. doi: 10.1016/j.jconrel.2007.05.040. [DOI] [PubMed] [Google Scholar]
  • 61.Horobin RW. Biological staining: mechanisms and theory. Biotech Histochem. 2002;77:3–13. [PubMed] [Google Scholar]
  • 62.Shedden K, Rosania GR. Chemical address tags of fluorescent bioimaging probes. Cytometry A. 2010;77:429–438. doi: 10.1002/cyto.a.20847. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Horobin R, Rashid-Doubell F, Pediani J, Milligan G. Predicting small molecule fluorescent probe localization in living cells using QSAR modeling. 1. Overview and models for probes of structure, properties and function in single cells. Biotech Histochem. 2013 doi: 10.3109/10520295.2013.780634. [DOI] [PubMed] [Google Scholar]
  • 64.Horobin R, Rashid-Doubell F. Predicting small molecule fluorescent probe localization in living cells using QSAR modeling. 2. Specifying probe, protocol and cell factors; selecting QSAR models; predicting entry and localization. Biotech Histochem. 2013 doi: 10.3109/10520295.2013.780635. [DOI] [PubMed] [Google Scholar]
  • 65.Horobin RW, Stockert JC, Rashid-Doubell F. Uptake and localisation of small-molecule fluorescent probes in living cells: a critical appraisal of QSAR models and a case study concerning probes for DNA and RNA. Histochem Cell Biol. 2013;139:623–637. doi: 10.1007/s00418-013-1090-0. [DOI] [PubMed] [Google Scholar]
  • 66.Zheng N, Zhang X, Rosania GR. Effect of phospholipidosis on the cellular pharmacokinetics of chloroquine. J Pharmacol Exp Ther. 2011;336:661–671. doi: 10.1124/jpet.110.175679. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Amidon GL, Lennernäs H, Shah VP, Crison JR. A theoretical basis for a biopharmaceutic drug classification: the correlation of in vitro drug product dissolution and in vivo bioavailability. Pharmaceutical Research. 1995;12:413–420. doi: 10.1023/a:1016212804288. [DOI] [PubMed] [Google Scholar]
  • 68.Balaz S. Lipophilicity in trans-bilayer transport and subcellular pharmacokinetics. Perspectives in drug discovery and design. 2000;19:157–177. [Google Scholar]
  • 69.Zhang X, Zheng N, Zou P, Zhu H, Hinestroza JP, Rosania GR. Cells on pores: a simulation-driven analysis of transcellular small molecule transport. Molecular Pharmaceutics. 2010;7:456–467. doi: 10.1021/mp9001969. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Trapp S. Plant uptake and transport models for neutral and ionic chemicals. Environmental Science and Pollution Research. 2004;11:33–39. doi: 10.1065/espr2003.08.169. [DOI] [PubMed] [Google Scholar]
  • 71.Kuh HJ, Jang SH, Wientjes MG, Au JL. Computational model of intracellular pharmacokinetics of paclitaxel. J Pharmacol Exp Ther. 2000;293:761–770. [PubMed] [Google Scholar]
  • 72.Jang SH, Wientjes MG, Au JL. Kinetics of P-glycoprotein-mediated efflux of paclitaxel. J Pharmacol Exp Ther. 2001;298:1236–1242. [PubMed] [Google Scholar]
  • 73.Tran TT, Mittal A, Aldinger T, Polli JW, Ayrton A, Ellens H, et al. The elementary mass action rate constants of P-gp transport for a confluent monolayer of MDCKII-hMDR1 cells. Biophys J. 2005;88:715–738. doi: 10.1529/biophysj.104.045633. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Poirier A, Lave T, Portmann R, Brun ME, Senner F, Kansy M, et al. Design, data analysis, and simulation of in vitro drug transport kinetic experiments using a mechanistic in vitro model. Drug Metab Dispos. 2008;36:2434–2444. doi: 10.1124/dmd.108.020750. [DOI] [PubMed] [Google Scholar]
  • 75.Bourdet DL, Pollack GM, Thakker DR. Intestinal absorptive transport of the hydrophilic cation ranitidine: a kinetic modeling approach to elucidate the role of uptake and efflux transporters and paracellular vs. transcellular transport in Caco-2 cells. Pharm Res. 2006;23:1178–1187. doi: 10.1007/s11095-006-0204-y. [DOI] [PubMed] [Google Scholar]
  • 76.Lumen AA, Acharya P, Polli JW, Ayrton A, Ellens H, Bentz J. If the KI is defined by the free energy of binding to P-glycoprotein, which kinetic parameters define the IC50 for the Madin-Darby canine kidney II cell line overexpressing human multidrug resistance 1 confluent cell monolayer? Drug Metab Dispos. 2010;38:260–269. doi: 10.1124/dmd.109.029843. [DOI] [PubMed] [Google Scholar]
  • 77.Kalvass JC, Pollack GM. Kinetic considerations for the quantitative assessment of efflux activity and inhibition: implications for understanding and predicting the effects of efflux inhibition. Pharm Res. 2007;24:265–276. doi: 10.1007/s11095-006-9135-x. [DOI] [PubMed] [Google Scholar]
  • 78.Agnani D, Acharya P, Martinez E, Tran TT, Abraham F, Tobin F, et al. Fitting the elementary rate constants of the P-gp transporter network in the hMDR1-MDCK confluent cell monolayer using a particle swarm algorithm. PLoS One. 2011;6:e25086. doi: 10.1371/journal.pone.0025086. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Pang KS, Rowland M. Hepatic clearance of drugs. I. Theoretical considerations of a "well-stirred" model and a "parallel tube" model. Influence of hepatic blood flow, plasma and blood cell binding, and the hepatocellular enzymatic activity on hepatic drug clearance. J Pharmacokinet Biopharm. 1977;5:625–653. doi: 10.1007/BF01059688. [DOI] [PubMed] [Google Scholar]
  • 80.Pang KS, Gillette JR. Kinetics of metabolite formation and elimination in the perfused rat liver preparation: differences between the elimination of preformed acetaminophen and acetaminophen formed from phenacetin. J Pharmacol Exp Ther. 1978;207:178–194. [PubMed] [Google Scholar]
  • 81.Igari Y, Sugiyama Y, Sawada Y, Iga T, Hanano M. Prediction of diazepam disposition in the rat and man by a physiologically based pharmacokinetic model. J Pharmacokinet Biopharm. 1983;11:577–593. doi: 10.1007/BF01059058. [DOI] [PubMed] [Google Scholar]
  • 82.Naritomi Y, Terashita S, Kimura S, Suzuki A, Kagayama A, Sugiyama Y. Prediction of human hepatic clearance from in vivo animal experiments and in vitro metabolic studies with liver microsomes from animals and humans. Drug Metab Dispos. 2001;29:1316–1324. [PubMed] [Google Scholar]
  • 83.Yamazaki M, Suzuki H, Sugiyama Y. Recent advances in carrier-mediated hepatic uptake and biliary excretion of xenobiotics. Pharm Res. 1996;13:497–513. doi: 10.1023/a:1016077517241. [DOI] [PubMed] [Google Scholar]
  • 84.Turncliff RZ, Hoffmaster KA, Kalvass JC, Pollack GM, Brouwer KL. Hepatobiliary disposition of a drug/metabolite pair: Comprehensive pharmacokinetic modeling in sandwich-cultured rat hepatocytes. J Pharmacol Exp Ther. 2006;318:881–889. doi: 10.1124/jpet.106.102616. [DOI] [PubMed] [Google Scholar]
  • 85.Yeo KR, Jamei M, Rostami-Hodjegan A. Predicting drug-drug interactions: application of physiologically based pharmacokinetic models under a systems biology approach. Expert Rev Clin Pharmacol. 2013;6:143–157. doi: 10.1586/ecp.13.4. [DOI] [PubMed] [Google Scholar]
  • 86.Jones H, Rowland-Yeo K. Basic concepts in physiologically based pharmacokinetic modeling in drug discovery and development. CPT Pharmacometrics Syst Pharmacol. 2013;2:e63. doi: 10.1038/psp.2013.41. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Neuhoff S, Yeo KR, Barter Z, Jamei M, Turner DB, Rostami-Hodjegan A. Application of permeability-limited physiologically-based pharmacokinetic models: Part I-digoxin pharmacokinetics incorporating P-glycoprotein-mediated efflux. J Pharm Sci. 2013;102:3145–3160. doi: 10.1002/jps.23594. [DOI] [PubMed] [Google Scholar]
  • 88.Fantin VR, Berardi MJ, Scorrano L, Korsmeyer SJ, Leder P. A novel mitochondriotoxic small molecule that selectively inhibits tumor cell growth. Cancer Cell. 2002;2:29–42. doi: 10.1016/s1535-6108(02)00082-x. [DOI] [PubMed] [Google Scholar]
  • 89.Fantin VR, Leder P. F16, a mitochondriotoxic compound, triggers apoptosis or necrosis depending on the genetic background of the target carcinoma cell. Cancer Res. 2004;64:329–336. doi: 10.1158/0008-5472.can-03-0899. [DOI] [PubMed] [Google Scholar]
  • 90.Kroemer G, Galluzzi L, Brenner C. Mitochondrial membrane permeabilization in cell death. Physiol Rev. 2007;87:99–163. doi: 10.1152/physrev.00013.2006. [DOI] [PubMed] [Google Scholar]
  • 91.Chen LB. Mitochondrial membrane potential in living cells. Annu Rev Cell Biol. 1988;4:155–181. doi: 10.1146/annurev.cb.04.110188.001103. [DOI] [PubMed] [Google Scholar]
  • 92.Mitchell P. Keilin's respiratory chain concept and its chemiosmotic consequences. Science. 1979;206:1148–1159. doi: 10.1126/science.388618. [DOI] [PubMed] [Google Scholar]
  • 93.Johnson LV, Walsh ML, Bockus BJ, Chen LB. Monitoring of relative mitochondrial membrane potential in living cells by fluorescence microscopy. Journal of Cell Biology. 1981;88:526–535. doi: 10.1083/jcb.88.3.526. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Smiley ST, Reers M, Mottola-Hartshorn C, Lin M, Chen A, Smith TW, et al. Intracellular heterogeneity in mitochondrial membrane potentials revealed by a J-aggregate-forming lipophilic cation JC-1. Proc Natl Acad Sci U S A. 1991;88:3671–3675. doi: 10.1073/pnas.88.9.3671. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Modica-Napolitano JS, Aprille JR. Delocalized lipophilic cations selectively target the mitochondria of carcinoma cells. Adv Drug Deliv Rev. 2001;49:63–70. doi: 10.1016/s0169-409x(01)00125-9. [DOI] [PubMed] [Google Scholar]
  • 96.Lampidis TJ, Salet C, Moreno G, Chen LB. Effects of the mitochondrial probe rhodamine 123 and related analogs on the function and viability of pulsating myocardial cells in culture. Agents Actions. 1984;14:751–757. doi: 10.1007/BF01978920. [DOI] [PubMed] [Google Scholar]
  • 97.Johnson LV, Walsh ML, Chen LB. Localization of mitochondria in living cells with rhodamine 123. Proc Natl Acad Sci U S A. 1980;77:990–994. doi: 10.1073/pnas.77.2.990. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Mewes HW, Rafael J. The 2-(dimethylaminostyryl)-1-methylpyridinium cation as indicator of the mitochondrial membrane potential. FEBS Lett. 1981;131:7–10. doi: 10.1016/0014-5793(81)80875-7. [DOI] [PubMed] [Google Scholar]
  • 99.Bereiter-Hahn J. Dimethylaminostyrylmethylpyridiniumiodine (daspmi) as a fluorescent probe for mitochondria in situ. Biochim Biophys Acta. 1976;423:1–14. doi: 10.1016/0005-2728(76)90096-7. [DOI] [PubMed] [Google Scholar]
  • 100.Bereiter-Hahn J, Seipel KH, Voth M, Ploem JS. Fluorimetry of mitochondria in cells vitally stained with DASPMI or rhodamine 6 GO. Cell Biochem Funct. 1983;1:147–155. doi: 10.1002/cbf.290010306. [DOI] [PubMed] [Google Scholar]
  • 101.Snyder DS, Small PL. Staining of cellular mitochondria with LDS-751. J Immunol Methods. 2001;257:35–40. doi: 10.1016/s0022-1759(01)00440-9. [DOI] [PubMed] [Google Scholar]
  • 102.Cooper G, Hausman R. The Cell-A Molecular Approach. Third ASM Press; Washington, DC: [Google Scholar]
  • 103.Mellman I, Fuchs R, Helenius A. Acidification of the endocytic and exocytic pathways. Annu Rev Biochem. 1986;55:663–700. doi: 10.1146/annurev.bi.55.070186.003311. [DOI] [PubMed] [Google Scholar]
  • 104.de Duve C, de Barsy T, Poole B, Trouet A, Tulkens P, Van Hoof F. Commentary. Lysosomotropic agents. Biochem Pharmacol. 1974;23:2495–2531. doi: 10.1016/0006-2952(74)90174-9. [DOI] [PubMed] [Google Scholar]
  • 105.Slater AF. Chloroquine: mechanism of drug action and resistance in Plasmodium falciparum. Pharmacol Ther. 1993;57:203–235. doi: 10.1016/0163-7258(93)90056-j. [DOI] [PubMed] [Google Scholar]
  • 106.Zhang J, Krugliak M, Ginsburg H. The fate of ferriprotorphyrin IX in malaria infected erythrocytes in conjunction with the mode of action of antimalarial drugs. Mol Biochem Parasitol. 1999;99:129–141. doi: 10.1016/s0166-6851(99)00008-0. [DOI] [PubMed] [Google Scholar]
  • 107.Kornhuber J, Tripal P, Reichel M, Terfloth L, Bleich S, Wiltfang J, et al. Identification of new functional inhibitors of acid sphingomyelinase using a structure-property-activity relation model. J Med Chem. 2008;51:219–237. doi: 10.1021/jm070524a. [DOI] [PubMed] [Google Scholar]
  • 108.Bhattacharya S, Shoda LK, Zhang Q, Woods CG, Howell BA, Siler SQ, et al. Modeling drug- and chemical-induced hepatotoxicity with systems biology approaches. Front Physiol. 2012;3:462. doi: 10.3389/fphys.2012.00462. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Burrowes KS, Swan AJ, Warren NJ, Tawhai MH. Towards a virtual lung: multi-scale, multi-physics modelling of the pulmonary system. Philos Trans A Math Phys Eng Sci. 2008;366:3247–3263. doi: 10.1098/rsta.2008.0073. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Schneck DW, Pritchard JF, Hayes AH., Jr Studies on the uptake and binding of propranolol by rat tissues. J Pharmacol Exp Ther. 1977;203:621–629. [PubMed] [Google Scholar]
  • 111.Ishizaki J, Yokogawa K, Hirano M, Nakashima E, Sai Y, Ohkuma S, et al. Contribution of lysosomes to the subcellular distribution of basic drugs in the rat liver. Pharm Res. 1996;13:902–906. doi: 10.1023/a:1016061330387. [DOI] [PubMed] [Google Scholar]
  • 112.Ishizaki J, Yokogawa K, Ichimura F, Ohkuma S. Uptake of imipramine in rat liver lysosomes in vitro and its inhibition by basic drugs. J Pharmacol Exp Ther. 2000;294:1088–1098. [PubMed] [Google Scholar]
  • 113.Yokogawa K, Ishizaki J, Ohkuma S, Miyamoto K. Influence of lipophilicity and lysosomal accumulation on tissue distribution kinetics of basic drugs: a physiologically based pharmacokinetic model. Methods Find Exp Clin Pharmacol. 2002;24:81–93. doi: 10.1358/mf.2002.24.2.677131. [DOI] [PubMed] [Google Scholar]
  • 114.Glaumann H, Motakefi AM, Jansson H. Intracellular distribution and effect of the antimalarial drug mefloquine on lysosomes of rat liver. Liver. 1992;12:183–190. doi: 10.1111/j.1600-0676.1992.tb01045.x. [DOI] [PubMed] [Google Scholar]
  • 115.Ishizaki J, Yokogawa K, Nakashima E, Ohkuma S, Ichimura F. Characteristic subcellular distribution, in brain, heart and lung, of biperiden, trihexyphenidyl, and (-)-quinuclidinyl benzylate in rats. Biol Pharm Bull. 1998;21:67–71. doi: 10.1248/bpb.21.67. [DOI] [PubMed] [Google Scholar]
  • 116.Yata N, Toyoda T, Murakami T, Nishiura A, Higashi Y. Phosphatidylserine as a determinant for the tissue distribution of weakly basic drugs in rats. Pharm Res. 1990;7:1019–1025. doi: 10.1023/a:1015935031933. [DOI] [PubMed] [Google Scholar]
  • 117.Hayes A, Cooper RG. Studies on the absorption, distribution and excretion of propranolol in rat, dog and monkey. J Pharmacol Exp Ther. 1971;176:302–311. [PubMed] [Google Scholar]
  • 118.Street JA, Hemsworth BA, Roach AG, Day MD. Tissue levels of several radiolabelled beta-adrenoceptor antagonists after intravenous administration in rats. Arch Int Pharmacodyn Ther. 1979;237:180–190. [PubMed] [Google Scholar]
  • 119.Wilson AG, Pickett RD, Eling TE, Anderson MW. Studies on the persistence of basic amines in the rabbit lung. Drug Metab Dispos. 1979;7:420–424. [PubMed] [Google Scholar]
  • 120.Seydel JK, Wassermann O. NMR-studies on the molecular basis of drug-induced phospholipidosis. Interaction between chlorphentermine and phosphatidylcholine. Naunyn Schmiedebergs Arch Pharmacol. 1973;279:207–210. doi: 10.1007/BF00503985. [DOI] [PubMed] [Google Scholar]
  • 121.Kaufmann AM, Krise JP. Lysosomal sequestration of amine-containing drugs: analysis and therapeutic implications. J Pharm Sci. 2007;96:729–746. doi: 10.1002/jps.20792. [DOI] [PubMed] [Google Scholar]
  • 122.Funk RS, Krise JP. Cationic amphiphilic drugs cause a marked expansion of apparent lysosomal volume: implications for an intracellular distribution-based drug interaction. Mol Pharm. 2012;9:1384–1395. doi: 10.1021/mp200641e. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Yu J-y, Rosania GR. Cell-based multiscale computational modeling of small molecule absorption and retention in the lungs. Pharmaceutical Research. 2010;27:457–467. doi: 10.1007/s11095-009-0034-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Yu JY, Zheng N, Mane G, Min KA, Hinestroza JP, Zhu H, et al. A cell-based computational modeling approach for developing site-directed molecular probes. Plos Computational Biology. 2012;8:e1002378. doi: 10.1371/journal.pcbi.1002378. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Balaz S, Lukacova V. Subcellular pharmacokinetics and its potential for library focusing. J Mol Graph Model. 2002;20:479–490. doi: 10.1016/s1093-3263(01)00149-8. [DOI] [PubMed] [Google Scholar]
  • 126.Poirier A, Funk C, Scherrmann JM, Lave T. Mechanistic modeling of hepatic transport from cells to whole body: application to napsagatran and fexofenadine. Mol Pharm. 2009;6:1716–1733. doi: 10.1021/mp8002495. [DOI] [PubMed] [Google Scholar]
  • 127.Baik J, Rosania GR. Modeling and Simulation of Intracellular Drug Transport and Disposition Pathways with Virtual Cell. J Pharmaceu Pharmacol. 2013;1:8. doi: 10.13188/2327-204X.1000004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Sanga S, Sinek JP, Frieboes HB, Ferrari M, Fruehauf JP, Cristini V. Mathematical modeling of cancer progression and response to chemotherapy. Expert Rev Anticancer Ther. 2006;6:1361–1376. doi: 10.1586/14737140.6.10.1361. [DOI] [PubMed] [Google Scholar]
  • 129.Krzyzanski W, Woo S, Jusko WJ. Pharmacodynamic models for agents that alter production of natural cells with various distributions of lifespans. J Pharmacokinet Pharmacodyn. 2006;33:125–166. doi: 10.1007/s10928-006-9007-3. [DOI] [PubMed] [Google Scholar]
  • 130.Budha NR, Kovar A, Meibohm B. Comparative performance of cell life span and cell transit models for describing erythropoietic drug effects. AAPS J. 2011;13:650–661. doi: 10.1208/s12248-011-9302-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.El-Kareh AW, Secomb TW. A mathematical model for cisplatin cellular pharmacodynamics. Neoplasia. 2003;5:161–169. doi: 10.1016/s1476-5586(03)80008-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Baik J, Rosania GR. Molecular imaging of intracellular drug-membrane aggregate formation. Mol Pharm. 2011;8:1742–1749. doi: 10.1021/mp200101b. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Morissette G, Lodge R, Marceau F. Intense pseudotransport of a cationic drug mediated by vacuolar ATPase: procainamide-induced autophagic cell vacuolization. Toxicol Appl Pharmacol. 2008;228:364–377. doi: 10.1016/j.taap.2007.12.031. [DOI] [PubMed] [Google Scholar]
  • 134.Morissette G, Ammoury A, Rusu D, Marguery MC, Lodge R, Poubelle PE, et al. Intracellular sequestration of amiodarone: role of vacuolar ATPase and macroautophagic transition of the resulting vacuolar cytopathology. Br J Pharmacol. 2009;157:1531–1540. doi: 10.1111/j.1476-5381.2009.00320.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Baik J, Rosania GR. Macrophages sequester clofazimine in an intracellular liquid crystal-like supramolecular organization. PLoS One. 2012;7:e47494. doi: 10.1371/journal.pone.0047494. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Baik J, Stringer KA, Mane G, Rosania GR. Multiscale distribution and bioaccumulation analysis of clofazimine reveals a massive immune system-mediated xenobiotic sequestration response. Antimicrob Agents Chemother. 2013;57:1218–1230. doi: 10.1128/AAC.01731-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Min KA, Talattof A, Tsume Y, Stringer KA, Yu JY, Lim DH, et al. The Extracellular Microenvironment Explains Variations in Passive Drug Transport Across Different Airway Epithelial Cell Types. Pharm Res. 2013 doi: 10.1007/s11095-013-1069-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Sarkar CA, Lauffenburger DA. Cell-level pharmacokinetic model of granulocyte colony-stimulating factor: implications for ligand lifetime and potency in vivo. Mol Pharmacol. 2003;63:147–158. doi: 10.1124/mol.63.1.147. [DOI] [PubMed] [Google Scholar]

RESOURCES