in vitro s for druguu.diva-portal.org/smash/get/diva2:166491/fulltext01.pdf · bioavailability oral...

84
in vitro s for Drug in the Caco-2 Cell Model

Upload: others

Post on 11-Oct-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

in vitro sforDrug

in the Caco-2 Cell Model

Page 2: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery
Page 3: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

… finally!

Page 4: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

ContentsContents iv

Glossary vi

Papers Discussed viii

1. Introduction 91.1. Gastrointestinal (GI) drug absorption and bioavailability 10

1.1.1. Function of the GI tract 111.1.2. Factors affecting GI absorption of drugs 12

1.2. Mechanisms of intestinal drug transport 131.2.1. Paracellular transport 141.2.2. Passive transcellular transport 151.2.3. Active transcellular transport 16

1.3. Permeability screening in drug discovery 171.3.1. Factors affecting data from cell culture assays 19

1.4. The Caco-2 cell 201.5. pH and pH-gradients 21

1.5.1. Henderson-Hasselbalch equation 211.5.2. pH partitioning hypothesis 221.5.3. pH and pH-gradients in the GI tract 221.5.4. Impact on transporters 24

1.6. Protein binding 261.6.1. Importance of protein binding in vivo 261.6.2. Protein composition 261.6.3. Protein binding in in vitro experiments 27

1.7. Surfactants 271.7.1. Cremophor® EL 291.7.2. Surfactants in in vitro assays 29

1.8. Clinical relevance of drug-drug interactions 30

2. AIMS OF THE THESIS 32

3. MATERIALS AND METHODS 333.1. Materials 33

3.1.1. Drugs and radiolabelled compounds 333.1.2. Bovine serum albumin (BSA) 333.1.3. Other materials 33

3.2. Cell culture conditions 333.2.1. Caco-2 cell culture 333.2.2. MDCK cell culture 34

3.3. Transport studies 343.3.1. pH Dependency in transport studies 35

Page 5: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

3.3.2. Inhibition studies 353.3.3. Temperature dependency in transport studies 363.3.4. Transport studies using extracellular additives 36

3.4. Cell monolayer integrity 373.4.1. Electrophysiological measurements 373.4.2. Mannitol transport 37

3.5. Drug concentration in monolayer and filter support 373.7. Sample analysis 373.8. Calculations of permeability coefficients and efflux ratios 383.9. Accounting for low recovery 393.10. Arrhenius equation and the activation energy 403.11. Protein binding determination 403.12. Accounting for protein binding 413.13. Theoretical solubility calculation 413.14. Critical micelle concentration determination 423.15. Statistical analysis 42

4. RESULTS and DISCUSSION 434.1. pH gradient systems across the Caco-2 monolayers 43

4.1.1. Effect on weakly basic drugs 434.1.2. Effect on weakly acidic drugs 454.1.3. Effect on concentration dependency 474.1.4. Effect on temperature dependency 504.1.5. pH Dependency of drug-drug interactions 514.1.6. The paracellular pathway 52

4.2. BSA-gradient systems across Caco-2 monolayers 534.2.1. Accounting for the maximal effect of protein binding 544.2.2. Impact of basolateral BSA on transport in the absorptive direction 554.2.3. Impact of basolateral BSA on the efflux/uptake ratios 57

4.3. Cremophor® EL and Caco-2 cells 60

5. SUMMARY AND CONCLUSIONS 64

6. REMAINING ISSUES 66

7. APPENDIX 67

8. ACKNOWLEDGEMENTS 70

9. REFERENCES 71

Page 6: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

Glossarya-b apical-to-basolateral, transport in the absorptive direction ABC ATP-binding cassette ABL aqueous boundary layer ADME(T) absorption, distribution, metabolism, elimination and toxicity ATP adenosine triphosphate ATCC American Type Culture Collection b-a basolateral-to-apical, transport in the secretory direction BBB blood-brain barrier BCRP breast cancer resistance protein BCS Biopharmaceutics Classification System BDDCS Biopharmaceutics Drug Disposition Classification System BSA bovine serum albumin BSP sulfobromophthalein Caco-2 cells Human colon adenocarcinoma cells clone 2 CEL Cremophor® EL Cmax maximum plasma concentration CMC critical micelle concentration Cu concentration of unbound drug CYP cytochrome P450 enzyme system DDI drug-drug interaction DMEM Dulbecco’s Modified Eagle Medium DMSO dimethyl sulphoxide DPHT 1,6-Diphenyl-1,3,5-hexatriene Efflux transport out of the cell F bioavailability fa fraction of drug absorbed after oral administration to humans fu fraction of unbound drug as percent of total concentration FCS foetal calf serum GI gastrointestinal HBSS Hank’s balanced salt solution HEPES 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid HTS high-throughput screening MDCK Madin Darby canine kidney cells MDR multidrug resistance MES 2-morpholino-ethanesulfonic acid monohydrate MRP multidrug resistance associated protein MW molecular weight OAT organic anion transporter OATP organic anion transporting polypeptide OCT organic cation transporter PAMPA parallel artificial membrane permeation assay Papp apparent permeability coefficient PB protein binding PBS phosphate-buffered saline P-gp MDR1 gene product; Permeability-affecting glycoprotein pHi intracellular pH RT-PCR reverse transcription polymerase chain reaction

Page 7: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

SLC solute carrier SD standard deviation TER transepithelial electrical resistance Uptake transport into the cell Å Ångström (1 x 10-10 m)

Page 8: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

Papers Discussed This thesis includes the following papers, which will be referred to in the text by their Roman numerals assigned below:

I. Neuhoff, S., A.-L. Ungell, I. Zamora, and P. Artursson. pH-Dependent bidirectional transport of weakly basic drugs across Caco-2 monolayers: Implication for drug-drug interactions. Pharm. Res. 20(8): 1141-1148 (2003). Reprinted with permission from Springer Science and Business Media. © 2003 Klu-wer

II. Neuhoff, S., A.-L. Ungell, I. Zamora, and P. Artursson. pH-Dependent passive and active transport of acidic drugs across Caco-2 monolayers. Eur. J. Pharm. Sci. In press (2005). Reprinted with permission from Elsevier. © 2005 Elsevier.

III. Neuhoff, S., P. Artursson, I. Zamora, and A.-L. Ungell. Impact of extracellular protein binding on passive and active drug trans-port across Caco-2 cells. Submitted.

IV. Neuhoff, S., P. Artursson, and A.-L. Ungell. Advantages and disadvantages of using bovine serum albumin and/or Cremo-phor® EL as extracellular additives during transport studies of lipophilic compounds across Caco-2 monolayers. In manuscript.

Page 9: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

9

1. Introduction

There are several routes by which drugs are commonly administered. However, oral administration is usually preferred by patients and can encourage compliance. In addition, oral formulations normally cost less to produce. Therefore, the majority of therapeutic agents today are administered orally, even when the formulation might be associated with problems such as poor solubility of the drug or acidic or enzy-matic degradation of the compound in the gastrointestinal (GI) tract. Only when the drug is sufficiently soluble and stable in the GI fluids can it be absorbed.

The absorptive processes after oral administration can be summarized as the permea-tion of a compound across tissue membranes (Chap. 1.1.). The rate and extent of absorption determines the rate and extent of distribution by the systemic circulation to the various tissues of the body (Chap. 1.1.1.). The rate of absorption (permeation) is, in turn, determined by the physicochemical properties of the drug molecule, the specificity of the drug for the intestinal transport systems, and the environmental, anatomical and physiological state of the GI tract (Chap. 1.1.2.). In order to predict the absorption of a drug (e.g., permeation through the intestinal epithelium) and to distinguish between the different mechanisms for absorption (Chap. 1.2), several invitro models have been developed (Chap. 1.3), one of which is the Caco-2 cell model (Chap. 1.4.). Currently, the fraction of a drug dose absorbed by humans is frequently predicted from permeability values obtained from Caco-2 cell monolayer systems (Artursson and Karlsson, 1991; Lennernäs, 1998). These systems are re-garded as reasonably reliable tool in the area of drug discovery. However, several issues limit the performance of Caco-2 cell monolayers and further research is es-sential to increase the reliability and predictive ability of these in vitro systems.

The work presented in this thesis is part of an effort to refine in vitro models for the prediction of intestinal drug transport. One major theme of this thesis is the investi-gation of the impact of physiological pH gradients in the intestine on the active and passive transport of drugs across the membrane (Chap. 1.5.). Although this approach has been partly examined in the literature already (Palm et al., 1999; Yamashita et al., 2000), a more systematic investigation of the effects on various types of drugs was not available at the beginning of this project.

Another approach to improving the physiological basis of in vitro systems for per-meability screening is the inclusion of extracellular proteins (Chap. 1.6.). The baso-lateral application of bovine serum albumin (BSA) in cell culture transport assays had been described in the literature (Raub et al., 1993), but also in this case more systematic studies on different types of drugs had not been performed. During the course of the work presented here, some work has now been published on alterations to drug transport rates due to the inclusion of BSA in cell culture transport assays

Page 10: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

10

(Aungst et al., 2000; Saha and Kou, 2002; Taub et al., 2002; Walgren and Walle, 1999; Yamashita et al., 2000). However, the extent of these changes have not, to my knowledge, been sufficiently evaluated in the Caco-2 cell model to date.

Structurally different surfactants have been used to enhance drug absorption across various epithelia (van Hoogdalem et al., 1989) and the addition of bile acids or mi-celle-building solubility enhancers has been studied in Caco-2 cell monolayers (Anderberg et al., 1992; Ingels et al., 2002) (Chap. 1.7.). The application of apical Cremophor® EL (CEL) in combination with basolateral BSA to Caco-2 cell monolayers was a new approach used in this thesis to develop an in vitro system for determining the permeation of poorly soluble compounds and predicting the fraction of the dose absorbed. This method might have the advantage of both increasing throughput capacity and improving the quality of permeability data.

Finally, the interaction of drugs with other drugs (and food) is a current issue in drug development. For instance, multidrug resistance (MDR) plays an important role in cancer treatment, and MDR1-encoded P-glycoprotein (P-gp) in combination with CYP3A4 has been found to be responsible for several drug-drug and food-drug interactions (DDI) (Deferme and Augustijns, 2003; Rau et al., 1997; Rodrigues, 2002) (Chap. 1.8.). However, at the onset of this thesis, there were no publications on the pH-dependence of DDIs and, therefore, this subject was also studied in vitroin the present work.

1.1. Gastrointestinal (GI) drug absorption and bioavailabilityOral administration is currently considered the most convenient and safest drug delivery route. However, an orally administered drug needs to reach the systemic circulation in order to arrive at the site of action in adequate quantities, as deter-mined by the potency of the drug.

The fraction of an oral dose that reaches the systemic circulation in its parent (origi-nal) form is called the oral bioavailability F, which can be described by the follow-ing equation:

)1()1( HGaHGa EEffffF Eq. 1.

where fa is the fraction of the administered dose that is absorbed across the apical membrane of the enterocytes, fG is the fraction of the absorbed dose that is not me-tabolized in the gut wall (fG = 1-EG) and fH is the fraction that escapes metabolism and biliary excretion in the liver (fH = 1-EH). There are many reasons for low oral bioavailability, and some of the losses occur in the GI lumen. For example, the lu-minal pH can affect the disintegration, dissolution and/or the stability of the dosage form and the drug. Moreover, the intestinal permeability (as a measure of the ability of a compound to cross the rate-limiting barrier) varies according to the specific site in the GI tract. Although first pass metabolism generally occurs in the liver, it has become increasingly recognized that metabolism may also occur within the entero-cyte during absorption and should be considered (Fisher et al., 1999; Wu et al.,

Page 11: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

11

1995). The main processes, which affect the bioavailability of a drug are schematically illustrated in Figure 1.

Figure 1: Schematic illustration of factors influencing the route to the systemic circulation after oral drug administration of a solid dosage form. After disintegra-tion, dissolution, enzymatic and chemical degradation and complexation in the GI tract, permeation of the epithelium may influence absorption of drug molecules. Apical efflux and basolateral uptake together with enzymatic degradation in the intestinal epithelium (EG) and in the liver (EH) may decrease the fraction of drug molecules entering the systemic circulation (F). This figure is based on figure 2-2. in reference: Rowland and Tozer (1989).

1.1.1. Function of the GI tract The small intestine absorbs the largest amount of digestive water. It also absorbs nutrients, i.e. electrolytes, fat, monosaccharides, amino acids, di- and tripeptides and vitamins, while the major physiological functions of the colon are to store and con-centrate feces by absorbing water and electrolytes. The colon lacks digestive en-zymes, but the lumen contains hundreds of different microorganism species which are involved in the reductive reactions of food digestion.

The GI tract is constantly exposed to chemicals, either man-made (e.g. drugs, indus-trial chemicals, pesticides, pollutants) or natural (e.g. alkaloids, secondary plant metabolites, and toxins produced by moulds, plants and animals). Thus, while the intestine must absorb nutrients, it must at the same time avoid absorption of poten-

Gut Wall

Liver

PortalVein

Intact drugavailable

to systemiccirculation (F)

Metabolism (EH),biliary excretion

Metabolism (EG)

GutLumen

MetabolismEnzymatic/chemical degradation

Adsorption/Complexation

To feces

Disintegration

DissolutionPermeability

Given dose

fa * fGfa fa * fG * fH

Gut Wall

Liver

PortalVein

Intact drugavailable

to systemiccirculation (F)

Metabolism (EH),biliary excretion

Metabolism (EG)

GutLumen

MetabolismEnzymatic/chemical degradation

Adsorption/Complexation

To feces

Disintegration

DissolutionPermeability

Given dose

fa * fGfa fa * fG * fH

Page 12: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

12

tially harmful compounds and microorganisms. These dual requirements of absorp-tion of nutrients on the one hand and exclusion of harmful substances on the other hand result in the necessity for a selective permeability barrier for transport both into and out of the intestinal lumen.

1.1.2. Factors affecting GI absorption of drugs The factors that affect oral drug absorption can be divided into two categories: phys-icochemical and physiological.

The physicochemical properties of the compound, such as particle or molecular size, hydrogen bonding, conformation or partition coefficients (logP, logD), direct the pathway of a compound across the intestinal tissue, i.e. whether the compound will be transported by paracellular or transcellular routes, or transported at all (Goodwin et al., 1999; Horter and Dressman, 2001; Kulkarni et al., 2002; Lennernäs, 1998; Ungell, 1997). Properties such as hydrophilicity and lipophilicity, electrostatic po-tential, and polarizability may influence affinities to transporters and enzymes. Thus, they direct the degree of transport and metabolism of the substance. Additionally, the susceptibility of a drug to chemical degradation and stability will determine whether the compound is intact long enough for absorption to occur (Kerns, 2001).

The composition of intestinal contents, including food, changes in gastric emptying and intestinal transit can significantly alter the amount of time that compounds re-side at favorable absorptive sites (Norberg et al., 2003). The composition of intesti-nal content can provide significant opportunities for the binding or complexation of drugs, altering the amount of drug available for absorption (Davis et al., 1986; Galia et al., 1998). The extent of complexation or binding may also depend on the disease state because of up- or downregulation of protein levels along the intestine. RegionalpH differences can prevent dissolution of the drug or provide environments for in-creased degradation (Fallingborg et al., 1989; Galia et al., 1998). Depending on the pKa of the drug, the acid microclimate could act either as a barrier to or an enhancer of drug transport.

The type of delivery system chosen for drug administration will influence where and when the compound will be available for absorption. Dissolution and release rates are factors affecting absorption, since they determine the amount of the drug avail-able for absorption and indirectly at the site of action. Additives, e.g. in the formula-tion enhancing solubility, can influence passive and active transport processes.

The rate-limiting step in the absorption of many orally administered drugs is the permeation of the drug across the intestinal wall (Jackson, 1987; Powell, 1981), a single continuous monolayer of columnar cells forming the so called epithelium of the small intestine (Mackay et al., 1991; Madara and Trier, 1994). Absorptive cells make up over 90% of the epithelial cell population that lines the gut lumen (Karali, 1989). They are formed in the crypts and differentiate during migration to the tips of the villi. The most distinct feature of these cells is their apical brush border mem-brane. The apical membrane of an absorptive cell is enlarged by microvilli, increas-ing the surface area of the apical membrane (Madara and Trier, 1994). The absorp-tive surface area of the colon is lower than the surface area of the small intestine,

Page 13: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

13

since the absorptive cells of the large intestine process microvilli but are not organ-ized on a villus (Mackay et al., 1991). Compared to the colon, the large surface area and the leakier epithelium make the small intestine the major site of drug absorption.

In general, GI absorption of a compound is often predominantly limited by the dis-solution rate/solubility of the drug or the permeability of the apical membrane of the enterocyte to the drug. The Biopharmaceutics Classification System (BCS), which scales these two parameters, provides a basis for predicting oral drug absorption and for in vitro/in vivo correlations (Amidon et al., 1995; Food and Drug Administration, 2000).

The Biopharmaceutics Drug Disposition Classification System (BDDCS) is a modi-fied version of the BCS, which aims to predict overall drug disposition by including the major route of elimination, described by the relative amount of metabolites to the parental drug, and the effect of transporters on oral drug absorption (Wu and Benet, 2005).

This thesis is focused on permeability to drugs as the rate-limiting step to enter the systemic circulation. Hence, a fundamental understanding of drug transport across the main rate-limiting barrier, the intestinal mucosa, is required. The mechanisms of intestinal drug transport are, therefore, discussed next.

1.2. Mechanisms of intestinal drug transport The aqueous boundary layer (ABL, also called ‘unstirred water layer’) is the aque-ous region adjacent to the membrane surface which acts as a diffusion barrier. The permeation of rapidly transported drugs is dependent on the barrier properties of the ABL while the permeation of more slowly permeating drugs is unaffected (Karlsson and Artursson, 1991). The mucus layer contains more than 90% water and usually 0.5-5% mucin (Carlstedt et al., 1985). After the drug has passed the ABL and the mucus, it is exposed to the acid microclimate at the luminal surface of the epithe-lium. The drug molecule can then be transported across the intestinal epithelium. Figure 2 schematically shows the drug absorption routes across the intestinal epithe-lium and the intestinal barriers for drug absorption in vivo. In general, there are two distinct pathways by which molecules, such as drugs, can cross the intestinal epithe-lium; either through absorptive cells (transcellular, case 1) or between cells via tight junctions (paracellular, case 2). The drug can, in the process, be metabolized (extra- and/or intracellularly, case 3) and can be exposed to carrier-mediated transport proc-esses (uptake, efflux, antiport, symport, cases 4 and 5) or transcytosis and endocyto-sis (variations of the passive transcellular pathway, case 6). Transport by the tran-scytosis route is very limited and is only relevant for some macromolecules (Gullberg, 2005). This pathway has not been studied in this thesis and therefore is not discussed further.

Page 14: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

14

Figure 2: Scheme of intestinal barriers (A-H) to drug absorption and the different transport mechanisms (1-6) through the intestinal epithelial membrane. (1) Passive transcellular transport. (2) Passive paracellular transport. (3) Passive transcellular diffusion affected by (3*) intracellular and/or extracellular metabolism. (4) Passive transcellular diffusion modified by an apical and/or basolateral uptake and/or efflux transporter. (5) Passive transcellular diffusion modified by apical and/or baso-lateral exchangers, e.g. antiporters. (6) Transcellular vesicular transcytosis.

1.2.1. Paracellular transportIn general, the tightness of the tight junctional complex and the concentration gradi-ent of the solution are the rate-limiting parameters which regulate passive diffusion via the paracellular pathway. It is believed that the narrow pores of the tight junc-tional complex carry a net negative charge and, thus, it appears to be more perme-able to cationic drugs than to anionic and neutral species (Adson et al., 1994; Karls-son et al., 1999; Pade and Stavchansky, 1997; Powell, 1981). The tight junction permeability can be influenced by factors such as intra- and extracellular Ca2+-ionconcentration, osmolarity, protein kinase activators and inhibitors, Na+-ion linked nutrient absorption and peptide hormones (Hochman and Artursson, 1994). Thus, transport via the paracellular pathway depends mainly on the size (molecular weight (MW) or volume) and the net charge of the compound (Tavelin, 2003).

The paracellular transport route is dynamically regulated. It varies not only along the intestine, but also along the crypt-villus axis (Figure 3). In general, the paracellular route is leakier in the upper part of the small intestine than in the lower parts of the small intestine and colon (Artursson et al., 1993; Taylor et al., 1985). The paracellu-lar pathway constitutes less than 0.1% (estimated 0.01-0.1%) of the total surface area of the intestinal tract and is mainly used by small molecules of hydrophilic character, e.g. mannitol. Drugs transported by this route are often incompletely ab-sorbed (Lennernäs, 1998; Nellans, 1991).

Although the paracellular transport route is generally considered to be passive, the pathway has been shown to be saturable (Gan et al., 1998; Lee and Thakker, 1999) and indirectly linked to intracellular processes (Zhou et al., 1999) or processes in-

1 2 3 54 6

3*

3 4

4

1 2 3 4 5 6

Unstirred water layer (A)

Mucus layer (B)Acidic microclimate (C)Apical membrane (D)

Cytosol (E)

Basolateral membrane (F)

Basement membrane (G)Lamina propria (H)

3*1 2 3 54 6

3*

3 4

4

1 2 3 4 5 6

Unstirred water layer (A)

Mucus layer (B)Acidic microclimate (C)Apical membrane (D)

Cytosol (E)

Basolateral membrane (F)

Basement membrane (G)Lamina propria (H)

3*

Page 15: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

15

volving the tight junction complex (Lee et al., 2002). Thus, the complexity and the dynamics of regulation of this pathway are still under investigation.

Compounds which are believed to be significantly absorbed by the paracellular route include: furosemide (Flanagan et al., 2002), atenolol (Adson et al., 1995), cimetidine (Nagahara et al., 2004; Zhou et al., 1999), glycinamide (Hubatsch et al., 2004), ranitidine and famotidine (Lee and Thakker, 1999). These drugs are all of moderate MW and are relatively hydrophilic.

1.2.2. Passive transcellular transport Molecules using the transcellular pathway can be transported either passively or by carrier-mediated (active) transport. Generally the passive transcellular pathway is restricted to molecules with a lipophilic character and the concentration gradient is the driving force of this pathway. Although a degree of lipophilicity is necessary to enter the lipid membrane, compounds that are too lipophilic can be trapped in the membrane. Thus, the membrane composition as well as the physicochemical proper-ties of the drug characterize the importance of this transport route for drug absorp-tion.

The membrane composition (phospholipids, proteins, cholesterol) varies between different cell types and is altered during the cell cycle. It may, therefore, influence the membrane properties. Even the apical membrane contains, for instance, lower levels of phosphatidylcholine and higher levels of glycosphingolipids than the baso-lateral membrane of epithelial cells (van Meer, 1989). The apical membrane is be-lieved to be less permeable than the basolateral membrane (Fromm and Hierholzer, 2000; Hayton, 1980) due to less membrane fluidity and increased thickness as a result of different membrane compositions (Lande et al., 1994). In general, the api-cal membrane is considered the rate-limiting barrier to passive transcellular transport in polarized cells.

Since it is assumed that the apical membrane is the rate-limiting barrier, a satisfac-tory description of the process can often be generated by only considering passive transport across this membrane. Thus, experimental and theoretical models describ-ing drug transport only by this mechanism have recently gained attention, e.g. PAMPA (Avdeef et al., 2005) and in silico predictions (Palm, 1998).

The structural requirements of the transcellular route can be simplified by the use of physicochemical properties of the compound such as logP or logD, MW etc.. Also other properties like polar surface area and hydrogen bonding have lately been used to describe the molecule. The simplest prediction model for passive transcellular transport is probably the ‘Rule of 5’ (Lipinski, 2003; Lipinski et al., 2001). Famous outliers for the ‘Rule of 5’ are, for instance, cyclosporine A and digoxin, both com-pounds known for their relevant interactions with active transport systems.

The anatomical and physiological state at the site of absorption will affect drug absorption. For a drug with slow passive transcellular permeation, the absorptive surface area is the rate-limiting step (Figure 3). The longer residence time in the intestinal fluids permits the drug to diffuse down the crypt-villus axis (Artursson et

Page 16: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

16

al., 1996). This not only increases the surface area, but also changes the cell compo-sition, which can in turn lead to an altered route of absorption for the drug.

Figure 3: Light micrograph of rat jejunum, approximately x50 magnification (Åsa Sjöberg, AstraZeneca AB, Mölndal). The absorptive surface area is more accessible for compounds with a long residence time. In vitro models have been developed to reflect the different regions of the small intestine: Caco-2 cells mimic the villus tip and 2/4/A1 cells mimic the crypt base, as seen in the difference of the average pore radius (Tavelin, 2003). Inset: Caco-2 cells grown on filter support for 21 days. The tissue and the cell monolayer were stained with Periodic-acid Schiff (purpur mem-brane, mucin, glycogen etc.) and Mayer’s hematoxylin solution (blue nuclei).

1.2.3. Active transcellular transport Carrier-mediated transport processes can be specific and highly selective for a num-ber of chemical structures (e.g. D-glucose, vitamins) and they operate in parallel with passive transcellular transport. The transport system can be saturated with high substrate concentrations, which occupy all binding sites. Additionally, other struc-turally similar compounds, which bind to the carrier proteins, may cause competitive inhibition. Mechanism-based inhibition, e.g. via allosteric inhibition or energy de-pletion, is also possible.

Figure 4 gives an overview of passive absorptive transport and active transport across a cell monolayer combined to the overall transcellular transport of a com-pound. The active transporter can work as an uptake (influx) transporter, which transports the compound into the cell (white circle with grey triangle; cases 1 and 2), or an efflux transporter, which transports the compound out of the cell (grey circle with white triangle; cases 3 and 4). However, unless a compound can passively gain intracellular access, it is not possible to simply investigate whether it is a substrate for efflux transporters. Transporters can be located in the apical and/or basolateral membrane. Their location will determine the effect on the overall absorption. For instance, apical uptake and a basolateral efflux transporter can work in serial in the direction of the concentration gradient in the absorptive direction (case 5) and apical

Short residence time Long residence time

Caco-25 Å

2/4/A110-15 Å

Page 17: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

17

efflux or basolateral uptake mechanisms work against this concentration gradient (case 6).

Figure 4: Scheme of the passive (a) and active (b) transcellular transport mecha-nisms (1-8) through the intestinal epithelial membrane, leading to the total (c) ob-served transport. (1) Passive transcellular diffusion supported by apical uptake. (2) Passive transcellular diffusion reduced by basolateral uptake. (3) Passive transcel-lular diffusion reduced by apical efflux. (4) Passive transcellular diffusion followed by basolateral efflux. (5) Passive transcellular diffusion supported by apical uptake and basolateral efflux. (6) Passive transcellular diffusion reduced by apical efflux and basolateral uptake. (7) Apical and basolateral efflux. (8) Apical and basolateral uptake. Key: White circles demonstrate uptake transporter, grey circles demonstrate efflux transporter. The triangle inside the circle indicates the working direction of the transporter.

Gut epithelial metabolism can also affect the transport of a compound. The product of such phase I and phase II metabolism can be transported out of the cell by apical and basolateral efflux transporters (case 7). Finally, apical and basolateral uptake transporters can work together, for example in transporting compounds which the cell needs to have fast access to and concentration control over (case 8). Similarly, as for the absorptive transport direction, active transporters could have the opposite effect on transport in the secretory direction.

Active transport against the concentration gradient of the compound requires energy. This energy can be supplied in two ways: (i) By hydrolysis of a high-energy com-pound such as adenosine triphosphate (ATP). Transporters that use this energy source have an ATP-binding site and are therefore called ABC-transporters. (ii) Via the coupled transport of another molecule, e.g. proton or sodium-ion. These symport or antiport transporters use the concentration gradient of the co-transported com-pound as a driving force for their own transport and belong in general to the group of solute carriers (SLC). Thus, some exchangers (Figure 2) are able to change their transport direction with an altered driving force.

Since bulk pH and temperature can, for instance, affect membrane fluidity, mem-brane composition and ion diffusion, all active transport processes are dependent on these variables as well.

1.3. Permeability screening in drug discovery High-throughput combinatorial chemistry has led to an increased number of possible new drug candidates and has spurred the development of high-throughput screening (HTS) methods for profiling pharmacokinetic drug properties such as absorption,

1 2 3 4 5 6 7 8

ba c ba c ba c ba c ba c ba c ba c ba ca: passiveb: activec: total

1 2 3 4 5 6 7 8

ba cba c ba cba c ba cba c ba cba c ba cba c ba cba c ba cba c ba cba ca: passiveb: activec: total

Page 18: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

18

distribution, metabolism, elimination and toxicity (ADMET). Several HTS assays for the assessment of membrane permeability to drugs are based on cell lines such as Caco-2 and MDCK (Artursson, 1990; Balimane and Chong, 2005; Irvine et al., 1999; Lentz et al., 2000; Li, 2001; Liang et al., 2000; Yamashita et al., 2002).

The methods for studying the epithelial absorption mechanisms for a drug should be sufficiently simple to gain results that are easy to interpret in order to obtain, for instance, clear cut-off values and, thus, to create guidelines and decision trees for the development process. At the same time, however, they should represent the impor-tant features which the investigated drug would have to face at the in vivo barrier under physiological conditions, e.g. pH gradients and protein binding. In addition, the methods should require small amounts of the test compound and other expensive materials. The methods should also be reproducible, rapid and not labour intensive.

In vitro permeability methods can be developed as screening tools for low through-put as well as automated high throughput systems. Cell-based permeability methods can help not only to select a drug with a particular property from a library of drugs, but also to gain information on the mechanisms for structurally closely related com-pounds. Thus, depending on the question to be answered in the stage of the devel-opment of the drug, a useful in vitro system has to be adaptable (Quaroni and Hochman, 1996).

Some immortalized epithelial cell lines can be cultured on porous filter supports to form confluent monolayers with reproducible (relative) protein expression and po-larity (Artursson, 1990; Pinto et al., 1983) (Figure 5). Hence, these cells can be used for automatic screening assays as well as manual and mechanistic studies. A general aim of this thesis was to improve the performance of in vitro cell culture models for mechanistic studies. Such more demanding studies can be used to follow up unclear results from HTS studies of drug permeation. They can also be used as guidelines for improvement of the experimental set-up and, thus, the outcome (increased qual-ity and throughput) of HTS.

Figure 5: Schematic illustra-tion of an epithelial monolayer grown on a permeable sup-port, e.g. Transwell®.

The species and the organ from which the cell line originates is important. While a human cell line originating from colon carcinomas like Caco-2, HT29 and T84 can be an obvious first choice for predicting absorption mechanisms in the GI tract, a transformed canine cell line from the kidneys such as MDCK II, which has been transfected with a human gene encoding for instance an efflux transporter such as P-gp (MDCKII-MDR1), can be an appropriate alternative in studies of P-gp efflux. However, the most reliable conclusions might be drawn from combined results of both assays.

Lower basolateral compartment

Upper apical compartmentTranswell

12-well cluster plate

Cell monolayer onmicroporous membrane

MicrovilliTight junction

NucleusFilter supportFilter pores

Lower basolateral compartment

Upper apical compartmentTranswell

12-well cluster plate

Cell monolayer onmicroporous membrane

MicrovilliTight junction

NucleusFilter supportFilter pores

Page 19: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

19

Since the human anatomy is very complex, a simple system will always only repre-sent and open a small window to the overall picture. The results from in vitro ex-periments have to be put in reasonable perspective; they should never be seen as a replacement for in vivo experiments. However, in vitro systems significantly reduce the number and increase the quality of in vivo experiments in drug discovery.

1.3.1. Factors affecting data from cell culture assays The assay conditions are crucial, independent of which cell line is used, since they will determine if for instance a mechanism is functional or not. When investigating a pH gradient-dependent transporter or an energy-dependent transporter, we have to supply the cells with a pH gradient and the necessary energy. This might be prob-lematic using cells in suspension (membrane vesicles, oocytes) or cells grown on plastic, e.g. in culture flasks.

The assay conditions also determine the drug concentration, which can be altered by parameters like temperature, pH, co-solvent and metabolism. Since the rate of per-meation is determined by the rate of transport, i.e. amount of compound transported per time, across a defined surface area relative to this concentration (see Equation 6), all calculated permeability values depend on this donor concentration and the accu-racy of its determination.

The accuracy of the determination of the drug concentration will also affect the calculated recovery for the experiment and the calculated mass balance during a time slot in the experiment. Loss in recovery can be due to low solubility of the compound, binding to the equipment (plastic tips, filter material, cell culture dishes etc.) or to the lipid membrane, and metabolism, which can often be experienced for lipophilic compounds. Parameters like pH, temperature and co-solvents, which might alter (increase) the solubility by changing the hydration shell of the molecule, can also affect the cells, the buffer capacity and even the equipment. In order to obtain a better overview of the transport of the compound by time, it can be useful to investigate changes in the mass balance at different time intervals. This is particu-larly important if the compound concentration is altered during the experiment, e.g. in metabolism studies.

The stirring rate will determine the thickness of the aqueous boundary layer adjacent to the epithelial membrane and, for mucus-producing cell lines, the stability of the mucus layer.

The ‘sink’ conditions will determine how close the system is to the equilibrium situation, which all closed systems strive for. Thus, sampling time points and sample volumes alter the ‘sink’ conditions. For efflux experiments, the cells will be loaded under equilibrium conditions in order to define an accurate (‘start’) donor concentra-tion, i.e. concentration in the cell. However, the experiment itself should be per-formed preferably under ‘sink’ conditions to avoid back-diffusion (in this case into the cell) and to maintain a constant applied concentration gradient. Also, in order to mimic the in vivo situation of drug transport, ‘sink’ conditions have to be maintained since, in vivo, the blood circulation represents a perfect sink, i.e. an open system.

Page 20: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

20

In addition, differences in the cell culture conditions, such as seeding density, pas-sage, degree of differentiation and expression level of the major proteins, can result in a lack of reproducibility between laboratories and need to be exposed for accurate comparison of results (Ungell, 2004; Ungell and Karlsson, 2003).

As long as there is a high rate of variability between assay conditions, a substantial variability in the results, e.g. drug transport rates, will be obtained. Hence, a pro-found understanding of the underlying mechanisms of drug transport is desirable and standardized assay conditions are necessary (Ungell, 2004).

1.4. The Caco-2 cell The closer an in vitro model describes the in vivo bioavailability, the more complex it becomes. The models used in early drug development simulate as closely as pos-sible the physiology of the intestinal epithelium in order to predict oral bioavailabil-ity, which is basically influenced by the lack of absorption from the GI tract, first-pass metabolism by enzymes in the gut wall, and first-pass metabolism by the liver.

Caco-2 cells are the most common cell models used to investigate intestinal absorp-tion processes in vitro (Balimane and Chong, 2005; Ungell, 2004). The cell line was derived from a moderately differentiated colon adenocarcinoma in a 72-year-old patient (Fogh et al., 1977). It is well characterized with respect to its cyto- and bio-chemical properties (Lu et al., 1996; Wilson et al., 1990). Caco-2 cell line clones are highly standardizable; however, carriers for intestinal transport may be over- or under-expressed compared to the protein expression of the various parts of the hu-man intestine (Sun et al., 2002).

The cells differentiate spontaneously under normal culture conditions and form tight monolayers of well-polarized absorptive cells (Artursson, 1990; Chantret et al., 1988; Hidalgo et al., 1989; Hilgers et al., 1990; Pinto et al., 1983). The cells show similarities in membrane function with human colon, but the carrier systems resem-ble those of the small intestine more closely (Ungell and Karlsson, 2003; Zweibaum et al., 1984). Their apical surface is equipped with the typical brush-border micro-villi (Artursson, 1991; Artursson and Magnusson, 1990; Hidalgo et al., 1989; Wil-son et al., 1990). Caco-2 membrane proteins include transport systems and enzymes typical of those in intestinal epithelial enterocytes in vivo (Sun et al., 2002) (Appen-dix, Figure 19). The transport of amino acids, bile acids, biotin, peptides and sugars has, for instance, been studied in this model (Bissonnette et al., 1996; Dantzig and Bergin, 1990; Hidalgo and Borchardt, 1990a; Hidalgo and Borchardt, 1990b; Ng and Borchardt, 1993). In addition, they express organic anion and cation transporters, mainly as uptake transporters and exchangers. They also express ATP-dependent efflux pumps such as P-gp, breast cancer resistance protein (BCRP) and several multidrug resistance-associated proteins (MRPs) (Taipalensuu et al., 2001). The activity and amount of the transporters can depend on the average age of the cells and the culture condition, as seen for P-gp and dipeptide transporters (Anderle et al., 1998; Herrera-Ruiz et al., 2001; Hosoya et al., 1996; Seithel et al., 2004).

Caco-2 cells also express various phase I (e.g., CYP1A1) and phase II (e.g., glu-tathione-S-transferases, and sulfotransferase) metabolizing enzymes. The culture

Page 21: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

21

conditions are also important for the development (amount and expression) of these proteins. For instance, Caco-2 cells grown in the presence of 1- -25-dihydroxy-vitamin D3 express enzymatically active CYP3A4 (Engman et al., 2003; Schmiedlin-Ren et al., 1997). The metabolic activity of Caco-2 cells is generally lower or between those obtained in the small intestine and colonic mucosae of ani-mals or humans (Jumarie and Malo, 1991; Madara and Trier, 1994; Quaroni and Hochman, 1996; Sun et al., 2002). Caco-2 cells are representative of columnar ab-sorptive cells, and, thus, they lack the complexity of the small intestinal wall, which contains several epithelial cell types, including mucin producing goblet cells. Thus, Caco-2 cells lack a mucus layer in culture and the cells do not maintain the acid microclimate. Instead, the selected buffer under standard experimental conditions maintains the pH.

In drug discovery and development settings, Caco-2 cells are widely used to: rapidly assess compound permeation across the intestinal epithelium of potential drug candidates and consequently determine suitable physicochemical properties for passive diffusion of drugs,study the function of intestinal epithelial cells (e.g. passive versus carrier-mediated pathways of drug transport),investigate formulation strategies in order to enhance membrane permeability, investigate potential toxic effects of compounds, pharmaceutical additives or formulations, and study presystemic drug metabolism.

1.5. pH and pH-gradientsThe pH can affect the degree of dissociation and in turn the membrane permeability and, thus, the potency of a compound to interact with proteins in the membrane and behind the membrane barrier. Additionally, the pH varies in the different compart-ments of the body from the macroscopic to the microscopic level, i.e. in the body fluids and organs, as well as cell cytosol, membranes and transporter levels.

1.5.1. Henderson-Hasselbalch equation Many drugs are either weakly acidic or weakly basic. When these drugs are in solu-tion equilibria exist between undissociated molecules and their ions. Thus, in a solu-tion of a weakly basic drug B (or a weakly acidic drug HA) the equilibrium can be described by the following equations:

Base: BHHB Eq. 2a

Acid: AHHA Eq. 2b.

In solutions of most salts of strong basic or acidic compounds in water, such equilib-ria are shifted to the right, i.e. to the ionized species (A- and BH+). The ionized con-stant (or dissociation constant) Ka of a weakly basic or acidic compound can be written as:

Page 22: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

22

Base:BH

BHKa Eq. 3a

Acid:HA

AHK a Eq. 3b.

When calculating the negative logarithm of Ka (pKa) and of the proton concentration (pH), the Henderson-Hasselbalch equation can be obtained (Hasselbalch, 1916):

Base:B

BHpHpKa log Eq. 4a

Acid:AHApHpK a log Eq. 4b.

The degree of ionization of a weakly basic or acidic compound in a solution can be calculated from the Henderson-Hasselbalch equation, if the pKa values of the drug and the pH of the solution are known.

1.5.2. pH partitioning hypothesis The pH-partitioning hypothesis states that only the uncharged species of a com-pound can permeate the membrane and that the permeation of a drug is therefore a function of the concentration of the uncharged species (Shore et al., 1957). The pHof the solution at the site of absorption, i.e. close to the membrane on the luminal side, will affect the degree of ionization of weak acids and bases according to the pHpartition hypothesis. Acids will diffuse more slowly and bases more rapidly across the membrane as pH increases. Deviations from the pH-partitioning theory for pas-sively transported compounds have been attributed to the presence of an acid micro-climate adjacent to the apical surface of the intestinal epithelium and to effects of the aqueous boundary layer (Lucas et al., 1978; Lucas et al., 1975; Tsuji et al., 1978).

1.5.3. pH and pH-gradients in the GI tract The intestine has to have the capacity to regulate the acidic compounds released by gastric secretion and by the activity of the intestinal microbes. Additionally, the HCO3

--ion secretions from the pancreas, the changes in CO2 pressure and the pH of the blood have to be accounted for.

The mean bulk pH along the human intestine in vivo ranges between 5.5 and 7.5 (Fallingborg et al., 1989). The bulk pH varies throughout the GI-tract in the fasted state from about 1.5-2.0 in the stomach, to 5.5-6.5 in the duodenum, 6.0-6.5 in the jejunum, 7.0-7.4 in the ileum and 5.5-6.5 in the colon (Dressman et al., 1990; Evans et al., 1988; Fallingborg et al., 1989). Following a meal, the gastric pH is increased to a median peak value of 6.7, then declining gradually back to the fasted state value during less than 2 hours (Dressman et al., 1990). In contrast to the pH changes ob-served in the stomach, the pH in the duodenum decreases in the fed state to a median

Page 23: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

23

value of 5.4. The pH in the GI tract can influence the absorption processes on sev-eral levels. Firstly, the pH of the GI fluids determines the degree of dissociation, as described by the Henderson-Hasselbalch equation. The degree of dissociation in turn can affect the solubility and, thus, the dissolution rate of a compound. Thus, the bulk pH affects solubility and release from the dosage form and the amount of compound interacting with the acid microclimate. Secondly, the pH of the acid microclimate can trigger the activity of active transport processes as described in the next chapter. Hence, the degree of dissociation can also affect the passive diffusion of the com-pound, according to the pH-partition hypothesis and, thus, the passive paracellular and transcellular pathways.

In 1959, Hogben and co-workers suggested that the metabolism of the epithelium may lead to a formation of an acid microclimate at the luminal surface and that this acid microclimate is responsible for deviations from the pH-partition hypothesis in observation on the intestinal transport of drugs (Hogben et al., 1959). In 1978, Lucas and co-workers measured an in vitro jejunal surface pH of 5.93 0.05 in human proximal jejunum biopsy samples (Lucas et al., 1978). In another study, the surface pHs of five rectal biopsy specimens obtained from a normal human mucosa were between 6.23 and 6.80, while the buffer pH varied between 5.96 and 7.51 (McNeil et al., 1987; McNeil and Ling, 1984). Independently of the bulk pH, the acid microcli-mate on the apical surface of the intestinal epithelium (approximately 20 µm thick) is relatively constant, when measured, for example, at the tip of the villi of rat je-junal tissue (Figure 6). Thus, the acid microclimate offers a permanent natural pHgradient across the absorptive cell membrane in vivo.

Figure 6: Relationship between microclimate pH and bulk pH in the rat jejunum. Individual symbols represent results of separate publications, sometimes including several studies (Daniel et al., 1985; Lucas, 1983; Shiau et al., 1985). The measured microclimate pH was between 5.0 and 6.7, while the bulk pH ranged between pH 4.0 and 10.0. However, beyond this range, there was a steep drop and rise, respectively, in the microclimate pH. The alterations in microclimate pH were relatively depend-ent on the distance to the tissue and, thus, on the probe rather than on the bulk pH. The dashed line is the line of identity.

Further down the crypt-villus axis, the pH increases from approximately 6.0 to 8.0 (Figure 7) (Daniel et al., 1985; Hogerle and Winne, 1983). Thus, while the bulk pH

0

3

6

9

12

0 3 6 9 12

microclimate pH

bulk

pH

Shiau et al., 1985

Lucas, 1983

Daniel et al., 1985

Page 24: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

24

gradient alters along the GI tract, the microclimate pH gradient across the epithelium alters along the crypt-villus axis.

Figure 7: The acid microclimate as a pH-profile illustrated for the rat small intes-tine (Light micrograph of jejunum; approximately x50 magnification; Åsa Sjöberg, AstraZeneca AB, Mölndal). Values of the proton concentration in relation to the distance are from reference: Daniel et al. (1985).

The acid microclimate exists because of the mucus, an ampholyte at the surface lining of the intestine. Several HCO3

--ion exchangers (e.g. Cl-/HCO3--exchanger) in

combination with sodium-proton exchangers (NHEs) and other transporters maintain the buffer system (Chang and Rao, 1994); however, the formation of the acid micro-climate is not only dependent on HCO3

--ion and hydrogen ion secretion. The expres-sion and activity of the transporters along the GI tract, and also along the crypt-villus axis, will determine the buffer capacity at the particular site.

1.5.4. Impact on transporters In order to distinguish between the possible pH-gradient systems, the concept of a 3-compartment model can be used (Jackson et al., 1974; Lucas and Whitehead, 1994) (Figure 8). A Caco-2 cell monolayer represents such a 3-compartment model. The pH of the buffer solutions on both sides of the monolayers can vary, while the cyto-solic pH is assumed to be around 7.4. Caco-2 cells express at least three different sodium proton exchangers (NHE1, NHE2, and NHE3), which help to stabilize the intracellular pH (pHi) (Cavet et al., 2001; Janecki et al., 1999; Thwaites et al., 1999).

6.5 7.4 8.0pH

[H+] [nmol.l-1]

Distance fromthe villus tip [µm]

bulk phase

Mucus layer (120 µm)

crypt

Aqueaous boundary layer (40 µm)

villi

Acid microclimate (20 µm)

-400

-300

-200

-100

0

100

200

300

400

500

-5050150250

6.5 7.4 8.0pH

[H+] [nmol.l-1]

Distance fromthe villus tip [µm]

bulk phase

Mucus layer (120 µm)

crypt

Aqueaous boundary layer (40 µm)

villi

Acid microclimate (20 µm)

-400

-300

-200

-100

0

100

200

300

400

500

-5050150250

Page 25: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

25

Figure 8: Scheme of different pH-gradient systems in a 3- compartment model. Key:’7.4’ represents all pH values above and below 7.4.

Transporters can be directly and indirectly affected by the bulk pH and the acid microclimate pH:

Antiporters or symporters can require a proton gradient in order to function. These transporters need a proton gradient either over one membrane (e.g. apical pH 6.0; intracellular pH 7.4) or over the entire cell (e.g. apical pH/intracellularpH/basolateral pH: 6.0/7.4/7.4). Transporters which have been reported to be proton-dependent include OCTN1 (Yabuuchi et al., 1999), MCT1 (Tamai et al., 1999), PepT1 (Thwaites et al., 2002), and OATP-B (Kobayashi et al., 2003). Three of them are thought to be apically located in the GI tract: MCT1 (Cuff et al., 2002), PepT1 (Rubio-Aliaga and Daniel, 2002) and OATP-B (Kobayashi et al., 2003).

It has been suggested that PepT1 and NHE3 share a common location (Thwaites et al., 2002) since absorption via PepT1 is dependent on maintenance of a proton gra-dient over the apical membrane alone, which is, in part, established by NHE3.

In contrast, the MCT1 transporter may be counter-ion dependent. This means that either a proton is transported with the substrate (symport; MCT1 (Halestrap and Price, 1999)) into the cell or an HCO3

--ion is transported out of the cell while the substrate is transported into the cell (antiport; SCFA-/HCO3

--antiporter (Stein et al., 2000)). In both cases, the intracellular compartment gains one positive charge (pro-ton) per transport cycle. Thus, the proton gradient over the membrane could influ-ence the choice of counter-ion.

Transporters can favor one species of a compound, i.e. the charged or uncharged form. They can also favor one enantiomer over the other. For instance, L-lactic acid is a substrate of the MCT1 (Stein et al., 2000; Tamai et al., 1999; Tamai et al., 1995), while its enantiomer D-lactic acid is not. Flupentixol is an example for an allosteric inhibitor of P-gp. Both cis and trans isomers of the dopamine receptor antagonist flupentixol inhibit drug transport and reverse drug resistance mediated by P-gp with a stereoselective potency. A stimulatory effect of cis(Z)-flupentixol and the inhibitory effect of trans(E)-flupentixol on ATP hydrolysis and radioligand la-beling were reported (Dey et al., 1999).

As indicated in Figure 4, the working direction and, thus, the location (apical or basolateral membrane) of the transporters relative to the concentration gradient of the compound, which determines the passive diffusion of the compound, is impor-tant. The expression of the same transporters in the various organs is also important in this context. It is expected that apical efflux transporters such as P-gp in the GI

7.4 7.4

7.4

7.4

apical(compartment #1)

basolateral(compartment #3)

cytosol(compartment #2)

7.47.4

7.4

7.4

7.4

7.4 7.4

7.4

7.4

apical(compartment #1)

basolateral(compartment #3)

cytosol(compartment #2)

7.47.4

7.4

7.4

7.4

Page 26: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

26

tract can reduce the absorption of a compound before metabolism in the cell, while P-gp in the liver will affect transport after the metabolism step (Benet et al., 2003). The ease of use of relatively new methods such as RT-polymerase chain reaction (RT-PCR) technology has allowed determination of the transporters and their ex-pression in relation to each other; however, this technique is quite expensive and comprehensive published data are currently limited (Seithel et al., 2004; Taipalensuu et al., 2001).

According to the pH-partitioning theory, the uncharged compound is more likely to permeate the membrane, and the extent of permeation, in turn, will determine the amount of compound available at the binding site of the efflux transporters. Thus, the impact of the transporter is indirectly affected by the bulk pH. For example, although efflux of drugs by P-gp is not directly affected by extracellular or intracel-lular pH (pHi) (Altenberg et al., 1993; Goda et al., 1996), an indirect connection between pHi regulation and the expression of functional P-gp has been suggested because of the acidification of the external medium observed in cells overexpressing P-gp. In addition, Johnstone and co-workers suggest, in an explanation of the regula-tion of apoptosis by P-gp, that expression of functional P-gp may increase the intra-cellular pH (2000). Thus, an active transporter can also change the intracellular and/or extracellular pH close to its location (Landwojtowicz et al., 2002).

If a drug is highly lipophilic and also has other physicochemical properties that favor a strong interaction with the cell membrane, it can become trapped in the apical membrane. After diffusion across the apical membrane, the intracellular pH will determine the fate of the compound. While weakly basic compounds can be trapped in the lysosomes (pHi: 4), acidic compounds can be trapped in the cytosol of epithe-lial cells (pHi: 7.2-7.4). All these processes lead to a loss of available compound at the site of action and result in poor or variable bioavailability. The intracellular drug concentration will also be increased, which could trigger a change in the transport direction of an exchanger.

1.6. Protein binding

1.6.1. Importance of protein binding in vivoProtein binding of drugs to human plasma and tissue in vivo leads to several effects in the body. By affecting the effective concentration of the drug and the equilibrium levels between tissue and plasma, it can alter the storage (when acceptors are in-volved), the metabolic transformation (when enzymes are involved), the transloca-tion (when transporters are involved) and the pharmacological effects (when recep-tors are involved) of the drug. When over 95% of the compound is plasma protein bound, a compound is regarded as a highly protein bound compound in vivo.

1.6.2. Protein composition Albumin is the most abundant protein component of plasma, with a concentration of about 40 mg.ml-1 (0.6 mM); it accounts for approximately 60% of the total protein content (Colmenarejo, 2003; Peters, 1996). The binding properties and the various

Page 27: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

27

binding sites for drugs on BSA and HSA (human serum albumin) are similar but are not identical (Kosa et al., 1997). BSA is less expensive then HSA and is, therefore, commonly used as a replacement in in vitro studies (Saha and Kou, 2002; Taub et al., 2002; Yamashita et al., 2000). There are also additional proteins, e.g. 1-acid glycoprotein, low-density lipoprotein (LDL) and high-density lipoprotein (HDL), in plasma (and intestinal fluid) which can affect the permeation of drugs (Shen et al., 2004). In humans, serum contains 56 g.l-1 of total protein, 35-50 g.l-1 (500-700 µM) of albumin, 0.4-1.0 g.l-1 (9-23 µM) of 1-acid-glycoprotein, about 2.4 g.l-1 (10.4 µM) of HDL, and about 2.4 g.l-1 (0.96 µM) of LDL (Bree and Tillement, 1986; Rowland and Tozer, 1989).

1.6.3. Protein binding in in vitro experimentsBSA can be applied basolaterally, i.e. on the serosal side, in the Caco-2 cell system in order to more closely approximate physiological conditions (Saha and Kou, 2002; Yamashita et al., 2000). This method of mimicking the extracellular tissue protein levels may improve the mass balance for lipophilic drugs, which will bind to the protein instead of to the plastic wells (Artursson, 1990; Wilson et al., 1990). The addition of extracellular protein will also help to maintain sink conditions during screening experiments (Krishna et al., 2001; Saha and Kou, 2002; Taub et al., 2002; Yamashita et al., 2000). In studies of the secretory (basolateral-to-apical; b-a) trans-port direction, however, applying BSA only to the basolateral side could reduce the concentration of unbound drug (Cu) that is then available to permeate the membrane. Nonetheless, it would still reflect the in vivo situation more closely than the tradi-tional Caco-2 set-up (without BSA).

Although also depending on the extent of protein binding and the degree to which the compound permeates the membrane passively, addition of BSA basolaterally will generally increase transport in the absorptive (apical-to-basolateral; a-b) direc-tion (Saha and Kou, 2002) and reduce transport in the secretory direction (Walgren and Walle, 1999). This can result in false positive (apical influx or basolateral ef-flux) or false negative (apical efflux or basolateral influx) interpretations of the in-volvement of a carrier-mediated transport system. Since apical efflux systems such as P-gp, BCRP and MRP2 influence the transport of drugs from the site of admini-stration to the site of action and thus influence the maintenance of adequate drug concentrations at the site(s) of action, it is necessary to develop predictive models for the transport of drugs in the presence of extracellular protein. Thus, it is impor-tant to understand the effect of protein binding on the efflux of model drugs. The validity of commonly used HTS methods in predicting efflux-related pharmacoki-netic behaviour in vivo in humans must therefore be revisited under improved ex-perimental conditions.

1.7. Surfactants The complexity and likelihood of error in calculating transport rates across biologi-cal membranes are substantially increased for poorly soluble compounds. Aqueous media without any solubility enhancing surfactant are preferred in screening studies, and limits for solubility and recovery data often have to be set when conducting invitro permeability studies, in order to ensure reasonably high throughput and an

Page 28: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

28

acceptable (minimum) quality of the results. However, when screening for new drugs, there are often few poorly soluble compounds that fulfil the minimum crite-ria. For this reason, solubility enhancers have been investigated as additives for adequate increase in solubility.

The common use of aqueous-organic solvent mixtures as a buffer medium is ques-tionable, since organic solvents such as DMSO and ethanol are known to interact with biological material. Thus, the concentration of these organic solvents in the final buffer solution is restricted, which in turn limits their solubility enhancing capacity.

Surfactant, amphiphile and tenside are all terms used to describe the same class of compound. The word surfactant is an abridgement of the term surface-active agent, i.e. a chemical species that is adsorbed at an interface. Common ionic and non-ionic surfactants such as polyoxyethylene sorbitan monolaurate (Tween), polyoxyl castor oil (Cremophor) and taurocholic acid can be used as solvent systems (Aungst, 2000; Ingels and Augustijns, 2003; van Hoogdalem et al., 1989).

The concentration of surfactant needed to increase the solubility of the drug depends on the critical micelle concentration (CMC) of the surfactant and the degree to which the compound partitions into the micelle formed by the surfactant. The CMC is the concentration (concentration range) which reflects the transition from a monomer solution of surfactant to a colloidal solution of surfactant micelles. At very low con-centrations of surfactant, no aggregates are formed. The surfactant molecules are present in the solution as monomers or, at somewhat higher concentrations, as dimers. When the concentration is raised to a certain value, specific for each surfac-tant, the surfactant molecules form aggregates, which are called micelles. This spe-cific concentration is called the CMC. Addition of surfactant will lead to a further increase in micelles, while the concentration of the monomer will stay constant (Figure 9). The CMC of the surfactant depends on the type of surfactant and the buffer solution (pH, ionic strength, type of ions, co-solvents) and the temperature. The partitioning of the compound into the micelle depends on factors such as the degree of ionization of the compound and the size of the molecule (or particle). In general, a linear relationship between solubility and surfactant concentration can be assumed above the CMC.

Figure 9: Concentration of micelles and mono-mers in relation to the total concentration of surfactant.

Page 29: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

29

1.7.1. Cremophor® ELCEL is a non-ionic surfactant and a derivative (polyether) of Castor oil and ethylene oxide. The polarity of the head group is created by polyether or polyhydroxy moie-ties. The hydrophilic-lipophilic-balance value (Griffin, 1950) for CEL lies between 12 and 14 (BASF AG, July 1997); thus, CEL can be used as an oil-in-water emulsi-fier as well as as a wetting or solubilizing agent. CEL is used to emulsify and solubi-lize oils and water-insoluble substances. It is used as a vehicle for water-insoluble vitamins and drugs such as cyclosporine A. The CMC for CEL in HBSS at pH 7.4 and 25 C is 0.0095% (w/v) (0.0073-0.0091% (w/v) range), assuming a molecular weight for the monomer of 2515 (Nerurkar et al., 1997). In general, the CMC is lower for a non-ionic surfactant, than for an ionic surfactant because of the lack of repulsive electrostatic forces between the head groups.

1.7.2. Surfactants in in vitro assays Surfactants used in pharmaceutical formulations can modulate drug absorption by many mechanisms. Of these only solubility enhancement is an acceptable applica-tion for in vitro screening of drug permeability. Other effects, such as membrane solubilization, tight junction modulation and interaction with transport proteins should be avoided (Aungst, 2000; van Hoogdalem et al., 1989). Experiments using surfactants in whole-tissue models offer complex information, which can be difficult to interpret (Florence, 1981). In contrast, in simpler models such as brush border membrane vesicle (BBMV) and cell monolayers the cell membrane, active transport systems and in the cell monolayers also the tight junctions can be studied under more controllable conditions (Anderberg and Artursson, 1993; Aungst, 2000; Ne-rurkar et al., 1996; Nerurkar et al., 1997).

While some apical additives seem to improve the experimental set-up, others should be used with caution, since surfactants can affect both the solubility and the permea-tion of a drug (Rege et al., 2002; Saha and Kou, 2000; Yu et al., 1999). At low con-centrations, the surfactants are incorporated into the lipid bilayer and this can change the properties of the cell membranes. After the lipid bilayer is saturated, mixed mi-celles could be formed, which results in removal of phospholipids from the cell membranes and a solubilization of the membrane. This, in turn, can change the flu-idity of the membrane, which could affect the functioning of membrane proteins, as has been suggested for P-gp (Dudeja et al., 1995; Woodcock et al., 1992). However, membrane fluidity modulators, such as cholesterol and benzylalcohol did not change the P-gp-associated drug transport of rhodamine 123 in Caco-2 cells (Rege et al., 2002). Controlled inhibition of P-gp or other apical efflux/basolateral uptake trans-porters by a surfactant (preferably one that is already used in pharmaceutical formu-lations, such as TPGS or CEL) might be useful for increasing the intestinal absorp-tion (and bioavailability) of substrates for these transporters. This approach has been followed recently by several laboratories (Bogman et al., 2005; Yu et al., 1999).

Thus, the potential positive effects of inclusion of a surfactant (solubility enhance-ment; improved mass balance) should be compared with the potential negative ef-fects (formation of mixed micelles; uncontrolled inhibition of active membrane transporters).

Page 30: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

30

1.8. Clinical relevance of drug-drug interactions In the clinic, a patient might obtain up to 21 different drugs concomitantly (Englund et al., 2004). All effective compounds have the potential to produce both benefits and risks associated with desired and undesired effects. Both types of effects are concentration dependent. Changes in the concentration of a drug can be triggered by a second effective compound (a drug (Lin and Lu, 2001), food (Deferme and Au-gustijns, 2003), etc.). Thus, one drug interaction can be considered as clinically significant when it is linked with a change in plasma drug concentration and when this change has the potential to alter the drug response.

DDIs can occur at a variety of sites, from the initial site of dissolution of the drug in the GI fluids to an effective site distant from the site of administration. Thus, a drug can affect the dissolution of another drug directly after oral administration, its plasma protein binding after both have reached the systemic circulation, or its me-tabolism (Lin and Lu, 2001). However, DDIs at the transporter level are still unre-solved. It could be speculated that transporter-mediated DDIs are most likely to occur where variations in transporter genes have been observed e.g., in rheumatoid arthritis and Crohn’s disease (OCTN1, SLC22A4; (Grundemann et al., 2005)), cho-lestatic liver disease (BSEP, ABCB11; (Kullak-Ublick et al., 2004) and the human Dubin-Johnson syndrome (MRP2, ABCC2; (Keppler and Konig, 1997)).

It is possible that transporter interactions have not yet been fully elucidated because of the difficulties in separating transport and metabolism in vivo, bearing in mind that 56% of a group of 315 marketed drugs were found to be primarily cleared by CYP (Bertz and Granneman, 1997) and that an estimated 30% of clinically marketed drugs are targeted to membrane transporters or channels (AAPS-Workshop, 2003). Not only might a compound have several different metabolic pathways but it might also interact with several different transporters. Thus, DDIs based solely on trans-porters can only be shown with two non-metabolizable compounds which share the same transporter. For example, digoxin and talinolol share P-gp (Westphal et al., 2000), and oral administration of talinolol increased the area under the concentra-tion-time curve of co-administered oral digoxin significantly, increasing the maxi-mum serum levels by 45% (Westphal et al., 2000). In this study the renal clearance and half-life of digoxin remained unchanged. However, infusion of talinolol con-comitantly with oral digoxin had no significant effects on digoxin pharmacokinetics. Digoxin did not affect the disposition of talinolol after either oral or intravenous administration. Westphal and co-workers observed significant increases in the bioavailability of digoxin only with oral co-administration of talinolol, and they concluded that this is most likely caused by competition for intestinal P-gp (Westphal et al., 2000).

A negative effect of co-administration of multiple inhibitors has been described recently (Englund et al., 2004). Multiple DDIs at the same site, e.g. P-gp, may be the result of multiple drug prescriptions. For example, S-digoxin levels increased step-wise with an increased number of co-administered P-gp inhibitors, when S-digoxinand the P-gp inhibitors were taken at the same time (Englund et al., 2004). In un-usual cases, intestinal transporter-based DDIs can be used to improve therapy, as suggested for the HIV-protease inhibitor and Vitamin E-TPGS (Yu et al., 1999). As

Page 31: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

31

proof of concept, TPGS, which inhibits P-gp, was co-administered with talinolol. TPGS increased the area under the plasma concentration-time curve (0- infinity) of talinolol by 39% and the maximum plasma concentration of talinolol by 100% (Bogman et al., 2005). This study showed the intestinal relevance of the drug-formulation interaction; however, it remains to be shown that TPGS is also useful in the clinic.

An additional obstacle in determining the degree of a DDI is the fact that the extent and appearance of DDIs will be altered depending on the organ and, thus, the expect order of which the compound would be exposed to the proteins (e.g. GI tract: P-gp then CYP 3A; liver: CYP3A then P-gp) (Benet et al., 2003; Suzuki and Sugiyama, 2000; Wacher et al., 1998). The properties of the organ should also be considered. For instance, since the brain comprises less than 1% of the total body weight, changes in the overall distribution of a drug within the body will not be observed. However, inhibiting apically located efflux transporters in the blood-brain barrier (BBB) (e.g. P-gp, MRPs or BCRP) can result in significantly increased brain levels of compounds like cyclosporin A and asimadoline, without an observable change in total body distribution. Understanding this type of transporter-based DDI could be useful as a concept for new drug delivery systems, such as those involving co-transport of efflux inhibitor and drug by nanoparticles across the BBB (Gulyaev et al., 1999).

Thus, DDIs are a reflection of the interplay between two compounds and their ef-fects upon each other. These effects are concentration dependent, i.e. dependent on the amount of ionized and un-ionized species of the compounds. Understanding the effect of various concentrations on active and passive transport processes might help to establish an experimental set-up that mimics some of the causes leading to DDIs, i.e. mimicking alterations in concentration at the site of an interaction.

Page 32: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

32

2. AIMS OF THE THESIS The objective of this thesis was to refine studies of the intestinal drug absorption of orally administered drugs using the in vitro Caco-2 cell model.

The specific aims were:

To investigate the passive and active transport processes of weakly basic, weakly acidic, and neutral drugs under the influence of the physiological pH gradient of the intestine using inhibitors and concentration and tem-perature dependency.

To investigate, using a mechanistic approach, DDIs at the level of a trans-port protein under the influence of the physiological pH gradient of the in-testine.

To investigate the mechanism behind the alterations in passive and active transport of drugs in the presence of drug binding to extracellular serum albumin.

To investigate the effects of the solubilizing agent CEL on Caco-2 perme-ability to both actively and passively transported lipophilic, BCS class II (low solubility/high permeation) compounds.

Page 33: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

33

3. MATERIALS AND METHODS

3.1. Materials

3.1.1. Drugs and radiolabelled compounds [3H]-Digoxin (629 GBq/mmol), S-[3H]-propranolol (703 GBq/mmol), [14C]-indomethacin (740 MBq/mmol), [14C]-salicylic acid (2.1 GBq/mmol), and [14C]-mannitol (1.9 GBq/mmol) were purchased from PerkinElmerTM (former NEN™) Life Science Products Inc. (Boston, MA, USA). Quinidine [9-3H] (740 GBq/mmol) was purchased from American Radiolabeled Chemicals Inc. (St. Louis, USA). [14C]-Doxorubicin (2.04 GBq/mmol) was purchased from Amersham Pharmacia Biotech UK Limited (Buckinghamshire, UK). [3H]-Metoprolol (31.5 GBq/mmol), metoprolol, [3H]-felodipine (605 GBq/mmol), felodipine, [3H]-inogatran (58.5 GBq/mmol), inogatran, [3H]-asimadoline (1514 GBq/mmol), asimadoline, [3H]-atenolol (229 GBq/mmol), atenolol, [3H]-talinolol (639 GBq/mmol) and talinolol were synthesized at AstraZeneca (Mölndal, Sweden). L-[1-14C]-Lactic acid (1.85 GBq/mmol) was purchased from American Radiolabeled Chemicals Inc. (St. Louis, MO, USA). 4,4´-Diisothiocyanostilbene-2,2´-disulfonic acid (DIDS), rifamycin SV, and sulfobromophthalein (BSP) were purchased from Sigma-Aldrich Co. (Stock-holm, Sweden). Pravastatin sodium was obtained from LKT laboratories, Inc. (St. Paul, MN, USA).

3.1.2. Bovine serum albumin (BSA) The quality of the BSA determines the binding capacity as well as the binding affin-ity to a drug. The BSA used in these studies was obtained from Sigma-Aldrich Co. (Stockholm, Sweden). The BSA was a Fraction V (Cat. No.: A-8022) of the follow-ing Lot No’s.: 78H1090; 107H0558, and 109H1051.

3.1.3. Other materials The non-ionic surfactant CEL was obtained from Fluka Biochemika (Buchs, Swit-zerland). SOLUENE®-350 tissue solubilizer solution and OptiPhase ‘Highsafe’ 3 were purchased from PerkinElmerTM (Boston, MA, USA) and Wallac (Loughbor-ough, UK), respectively, and 2-morpholino-ethanesulfonic acid monohydrate (MES) was obtained from Fluka (Gothenburg, Sweden). All other chemicals were obtained from Sigma-Aldrich Co. (Stockholm, Sweden).

3.2. Cell culture conditions

3.2.1. Caco-2 cell culture Caco-2 cells, obtained from ATCC (Rockville, MD, USA) at passage 18, were maintained as described previously (Kamm et al., 2001) using heat-inactivated fetal calf serum. All Caco-2 cells used for this thesis were treated under the same condi-tions as those used for screening at AstraZeneca (Mölndal, Sweden): i.e. they were

Page 34: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

34

used at passage 27-41 and routinely seeded at a density of 2 x 106 cells per 175 cm2-flasks. The cells were grown to 90% confluence and harvested by regular trypsiniza-tion using a trypsin (0.05%) - EDTA (0.02%) solution.

The Caco-2 cells were cultivated at 37 C in Dulbecco´s modified Eagle medium (DMEM) supplemented with 8% fetal calf serum (FCS), 1% Minimum Essential Medium of non essential amino acids and 2 mM L-glutamine. The medium was changed every second day. For transport experiments, 2.56 x 105 cells were seeded onto each 12 mm polycarbonate cell culture insert – 12 well plates with a filter area of 1.13 cm2 and a pore size of 0.4 µm (Transwell®, Cat. No. 3401, Corning Costar®

Corporation, Cambridge, MA, USA). The cells on the plates were maintained, using culture medium containing antibiotics (100 U ml-1 penicillin, 100 µg ml-1 streptomy-cin), in an atmosphere of 5% CO2. The relative humidity was 90%. All tissue culture media were obtained from GIBCO, Life Technologies (Paisley, Scotland).

3.2.2. MDCK cell culture MDCK II wild type cells (MDCK-wt) and MDCK II cells transfected with the hu-man MDR1 gene (MDCK-MDR1) were obtained from the Netherland Cancer Insti-tute (Amsterdam, The Netherlands). All MDCK type II cells were cultured in DMEM supplemented with 8% FCS, 1% Minimum Essential Medium of non essen-tial amino acids and 1.5% L-glutamine (200 mM) at 37 C, 90% humidity, and 5% CO2. The final concentration in the medium was 7 mM L-glutamine.

Cells were trypsinized once weekly and seeded at a density of 1 million cells per 185 cm2 flask. For transport studies, cells were seeded onto polycarbonate Transwell®

filter supports (Cat. No. 3401) at a density of 150.000 cells per filter (filter area: 1.13 cm2). The monolayers were ready for studies 3 days later. The transport experiments were performed as described for the Caco-2 cell monolayers (Transport studies – section 3.3).

3.3. Transport studies Transport studies were performed at AstraZeneca (Mölndal, Sweden) as previously described (Palm et al., 1999) with slight modifications. Briefly, bidirectional trans-port rates (apical-to-basolateral: a-b; basolateral-to-apical: b-a) of the test com-pounds were measured using Caco-2 monolayer cultures grown in the Transwell®

system for 23 1 days (Paper I and Paper II) or for 17 1 days (Paper III and Paper IV). Each experiment was performed at least in triplicate. Monolayer permeability to the paracellular marker mannitol was determined simultaneously.

The pH values and the osmolarity of the donor and receiver solutions were con-trolled at the beginning of the experiment as well as at the end of the experiments. The cell monolayers were preincubated for 25 to 30 minutes to stabilize the desired pH-gradient and BSA-gradient system, respectively.

The volumes of the appropriate incubation medium used were 0.5 or 1.5 ml on the apical and basolateral side, respectively. Samples were withdrawn from the donor side of the monolayers to determine the initial and final donor concentration of the

Page 35: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

35

compound and for calculation of the recovery. The following sampling times were in general used for the receiver sample: 0, 5, 15, 25, 50, 80, and 120 minutes. For a-b transport, 1.5 ml of medium was removed from the basolateral side, whereas for b-a transport, 400 µl of the initial volume (500 µl) was removed from the apical side and replaced with fresh buffer solution.

3.3.1. pH Dependency in transport studies The incubation medium was Hank’s balanced salt solution (HBSS) supplemented either with 10 mM MES (pH 5.0 – 6.5), or with 25 mM HEPES (pH 7.0 – 8.0). In general, the apically applied pH was altered between pH 5.0 and 8.0, while the baso-lateral pH was kept at pH 7.4. However, for some mechanistic studies, the pH was also altered (between pH 5.0 and 8.0) on the basolateral side of the monolayers. The pH values of the donor and receiver solutions were verified at the beginning of the experiment as well as at the end of the experiments.

3.3.2. Inhibition studies Inhibition studies can be performed (i) as competitive inhibition, when inhibitor and test compound are simultaneously added, (ii) as ‘total’ inhibition, when the inhibitor is added in advance, i.e. independent from the test compound, (iii) as concentration-dependent inhibition, when the test compound is added at increasing concentrations, until the transport protein is saturated.

Competitive inhibition studies (Paper I) For the DDI studies in Paper I, the culture medium was removed from both sides of the monolayers and they were washed with the desired HBSS-buffer. The cell monolayers were then preincubated for 30 min without the test compounds, and then incubated for 25 min at 37 C with the two test compounds for the DDI. The pHvalues used in the apical compartment were 5.0, 5.5, 6.0, 6.5, 7.0, 7.4, 7.7 or 8.0; while the pH used in the basolateral compartment was 7.4. The receiver medium was replaced after 5, 15, and 25 minutes with fresh buffer solution.

Inhibition of proton-dependent transporters (Paper II) The aim of a ‘total’ inhibition study is to obtain a defined inhibition of one or sev-eral transport proteins. The inhibition can be the result of saturation of the protein or mechanism-based inhibition, e.g. allosteric inhibition. Different pH-gradient systems were applied in order to investigate the inhibition of proton-dependent transport. The cell monolayers were preincubated for 25 min at 37 C in the presence of the inhibi-tor on both sides without the test compound. The buffer was removed from both sides of the monolayers, which were then incubated for another 25 min in the pres-ence of inhibitor at 37 C with fresh buffer containing the test compound.

Phloretin (nominal concentration: 1 mM), which is a nonspecific inhibitor of several active transport systems, was used as inhibitor. Phloretin inhibits glucose transport in Caco-2 cells (Bissonnette et al., 1996) and is thus a potent indirect but reversible inhibitor of the proton-dependent MCT1 (Stein et al., 2000). L-Lactic acid (10 mM) was included as a more specific and enantioselective competitive inhibitor of MCT1 (Stein et al., 2000; Tamai et al., 1995), and D-lactic acid (10 mM) was included as a

Page 36: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

36

non-substrate of MCT1. In addition, pravastatin (5 mM), rifamycin SV (100 µM) and BSP (25 µM and 50 µM) were used as inhibitors of the proton-dependent OATP-B. Positive controls of transporter function, e.g. L-lactic acid for MCT1, were included in each experiment.

Concentration dependency in transport studies (Paper II and Paper III) In Paper II, salicylic acid and indomethacin were investigated in the concentration range 0.025 to 33 mM and 0.025-0.1 mM, respectively.

In Paper III, each compound was investigated at two different donor concentrations. In order to cover a wide concentration range, the lowest concentrations to be inves-tigated were determined by the specific activity of the radiolabeled compound, and the highest concentrations were in general determined by the drug solubility or tox-icity to the monolayers. For quinidine, the highest concentration was that at which the bidirectional transport rates were not significantly different, i.e. 10 µM.

3.3.3. Temperature dependency in transport studies To investigate the dependence of the permeability of the system on temperature, the transport rate for the compounds was measured at 4, 7, 17, 27 and/or 37 C in trans-port buffer after preincubation of the cell monolayers for 25 to 30 minutes at the desired temperature. The experiments at 4 C and 7 C were performed in a refrigera-tor room and a cold storage room, respectively. The experiments at 17 C were per-formed in a cooled incubator (Sanyo - Biomedical Gallenkamp, Loughborough, UK). The experiments at 27 C and 37 C were performed in a laboratory. In all cases, the plates with the monolayers were placed onto a calibrated plate shaker (BMG LabTechnologies GmbH, Offenburg, Germany), which was set at a high stirring rate (450 rpm) throughout the experiment in order to minimize the influence of the aqueous boundary layer, and was adjusted for heat if necessary, i.e. during the experiments at 27 C and 37 C.

3.3.4. Transport studies using extracellular additives

Transport studies in the presence of BSATransport studies for Paper III were performed as described above (chapter 3.3) using the following sampling times: 0, 5, 15, 25, 50, 80, and 120 minutes. The incu-bation medium was HBSS-HEPES supplemented with 0 to 4% BSA (w/v). All ex-periments were performed at a basolateral pH of 7.4, in order to take the possible effect of pH in the blood on the binding rate into account. To avoid the impact of a pH gradient on drug transport, the apical pH was also set to 7.4. Influence of baso-laterally applied BSA (1, 2, and 4%) on the permeability of the monolayer to the drug was determined bidirectionally.

In the initial experiments in Paper IV, BSA was applied apically, basolaterally or on both sides of the monolayers. The receiver medium was replaced after 5, 15, and 25 minutes.

Page 37: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

37

Transport studies in the presence of CEL and BSA The cell monolayers were preincubated for 30 min without the test compound, and thereafter incubated for 25 min at 37 C with the test compound. The CEL concentra-tions used in the apical compartment were between 0.00001 – 10% (w/v), whereas no CEL was added to the basolateral compartment. Four percent BSA was added to the basolateral compartment.

3.4. Cell monolayer integrity

3.4.1. Electrophysiological measurements The initial and final values of transepithelial electrical resistance (TER) of the Caco-2 cell monolayers grown on permeable supports were measured with an epithelial volt-ohmmeter (EVOM), equipped with an SX-2 electrode (World Precision Instru-ments Inc., Sarasota, Florida, USA). The monolayers were equilibrated for 25-35 minutes before TER measurements. Only cell monolayers with TER values over 220

.cm2 were used. When inhibitors were used, the effect of the inhibitors on TERwas measured after preincubation with the inhibitor.

3.4.2. Mannitol transport The lower threshold for the permeability of the monolayer to the paracellular marker, mannitol, was set at < 0.5 x 10-6 cm.s-1.

3.5. Drug concentration in monolayer and filter support After the final receiver and donor samples were collected, the filters containing the cell monolayers were washed twice with ice-cold HBSS-HEPES buffer (pH 7.4). The filter supports were then cut out with a scalpel and transferred into 20 ml glass vials. The cell monolayers were dissolved with the filter support in 2 ml SOLU-ENE®-350 tissue solubilizer solution and, thereafter, 15 ml OptiPhase ‘Highsafe’ 3 solution was added to each vial. The radioactivity content of the well mixed samples was measured using a liquid scintillation counter (Wallac, Turku, Finland) and the amount of drug in the sample was then calculated. The radioactivity associated with the cells or adsorbed onto the filter was determined.

The binding of compound to the filter supports was determined in control wells without cells. The filter supports were preincubated at 37 C for 10 minutes in cell culture medium, washed twice with transport buffer and then incubated for 25 min-utes in transport buffer including the compound. Afterwards, the filter supports were washed twice with ice-cold HBSS-HEPES buffer (pH 7.4) and treated like the filters containing the cell monolayers.

3.7. Sample analysis The amount of drug transported was determined from the radioactivity content in the samples using a scintillation counter (Wallac, Turku, Finland), with a quench curve for BSA when necessary.

Page 38: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

38

3.8. Calculations of permeability coefficients and efflux ratiosThe transport rate is described by the following equation obtained from Fick’s law

)()()( tCtCAPVdttdCdtdQ rdapprr Eq. 5

where Q [mg] is the amount transported over the cell monolayer, Cr(t) [mg/ml] the receiver concentration, Cd(t) [mg.ml-1] the donor concentration, Vr [ml] the receiver volume, Papp [cm.s-1] the apparent permeability coefficient, A [cm2] the area of the filter and t [s] the time.

For all experiments that were performed under 'sink' conditions, the apparent perme-ability coefficients were calculated according to the following equation:

)0(dapp CA

dtdQP Eq. 6

where the transport rate (dQ/dt) was determined as the slope obtained by linear re-gression of cumulative fraction absorbed-time profiles. Equation 6 is obtained from Equation 5, where Cd(t) have been approximated constant and Cr(t) on the left side of Equation 5 has been approximated as zero.

When the sink conditions could not be maintained, Papp was determined using the following equation:

tVVAP

rdr

rdr

drappeVV

MCVV

MtC 11)0()( Eq. 7 a.

Where Cr(0) is the initial receiver concentration Vd [ml], the donor volume and Mthe total amount compound in the system. Equation 7 a is the exact solution of the differential equation (Equation 5), i.e. Cd(t) and Cr(t) is no longer considered con-stant and Cd(t) has been recalculated in terms of M and Cr(t).

The removing of compound in sampling procedure was considered by the recalcula-tion of the total amount of compound in the system and initial receiver concentration for each sample interval. With these considerations the theoretical value for the receiver concentration Cr(ti) for a certain Papp at each sampling time point, ti, is ex-pressed by:

1111 )()( iidrapp ttVVAP

rd

iir

rd

iir e

VVM

ftCVV

MtC Eq. 7 b.

Where Mi is the amount substance in the system at the sampling interval and f is the dilution factor caused by the sampling and replacement procedure. The Papp was determined by curve fitting, using the Levenberg-Marquardt algorithm when mini-

Page 39: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

39

mizing the sum of squared residuals [ (Cr(ti)obs - Cr(ti)) 2], where Cr(ti)obs is the ob-served receiver concentration and Cr(ti) is the corresponding calculated concentra-tion at the end of the interval (i).

The ratios of the transport in the b-a direction to that in the a-b direction, and viceversa, were calculated in order to obtain information regarding any asymmetry in the transport of the drug. Thus, (apical) drug efflux ratios and uptake ratios were calcu-lated using the following equations:

)(

)(

baapp

abapp

PP

oEffluxRati Eq. 8 a

)(

)(

abapp

baapp

PP

oUptakeRati Eq. 8 b.

The relative contribution of the P-gp transport to the overall transport in MDCK-MDR1 monolayers were determined by calculating the efflux ratio obtained in MDR1 transfected cells and dividing this value by the calculated efflux ratio ob-tained from the transport rates across MDCK-wt cells.

3.9. Accounting for low recoveryThe recovery R [%] is the percentage of original drug mass accounted for at the end of the experiment (the sum of the amounts on the apical and basolateral sides). The recovery was calculated using the following equation:

dd

bbaa

VCVtCVtCR

)0(100)()( Eq. 9

where Ca(t) and Cb(t) are the drug concentrations on the apical and basolateral sides of the monolayer at time t, Cd(0) is the concentration of the donor at time zero (t = 0), Va and Vb are the volumes of the apical and basolateral compartments, and Vd is the volume of the donor solution added to the appropriate side of the monolayer.

It was assumed that reductions in recovery were due to a process that occurred im-mediately, thus affecting the initial concentration on the donor side, and the follow-ing equation was used to account for low recovery:

RP

correctederyreP appapp

100)cov( Eq. 10.

Note, that this equation cannot be used when time-dependent processes, e.g. metabo-lism, are the reason for the low recovery.

Page 40: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

40

3.10. Arrhenius equation and the activation energy Transporters are temperature sensitive. The movement of the transporters within the lipid bilayer, the movement of counterions in the buffer solution and, thus, the activ-ity of the transporters are all decreased, when the temperature is decreased because of the reduced Brownian molecular movement with temperature. The diffusion of the compound which might interact with the transporter in the membrane, is also affected.

Arrhenius plots of the logarithmic permeability values at the same donor concentra-tion versus the reciprocal of the absolute temperature within one pH-gradient system were prepared. Values for the activation energy Ea were determined from the slope of the linear regression lines of the Arrhenius plots, where the rate constant k is correlated with the temperature T [ K] and were calculated in units of [kJ.mol-1]according to the Arrhenius equation:

TREa

eAk Eq. 11 a.

TRE

Ak a 1303.2

loglog Eq. 11 b.

were R is the gas constant with a value of 8.31J.mol-1.K-1 and A is the frequency factor for the process.

3.11. Protein binding determination To determine protein binding an ultrafiltration method was used. Standardized con-ditions for binding experiments were used to mimic physiological conditions. The pH was 7.4, and the maximal albumin concentration was 4% (w/v), i.e. 40 g.l-1. Cen-trifree® micropartition devices from Millipore (Bredford, MA, USA) were pre-warmed to 37 C and a 400 µl sample volume was loaded into the reservoir. Each protein binding estimation was determined twice using two micropartition devices. After centrifuging in a Hettich Rotixa/KS centrifuge (Tutlingen, Germany) with a fixed-angle bucket at 3,080 rpm (2,000 g) for 20 minutes, radioactivities in 100-µl ultrafiltrate were determined using a liquid scintillation counter (Wallac, Turku, Finland). The initial total concentration (Cinitial) was also determined from the radio-activity content in the initial samples. The fraction of the compound that was un-bound (fu) was then determined according to the following equation:

initial

uu C

Cf Eq. 12

where Cu is the concentration of unbound drug available for permeation. An assump-tion was made that only the unbound drug is able to move across membranes. Low recovery from either protein-free ultrafiltrate or buffer could indicate adsorp-tive losses and/or membrane rejection of ligand. Thus, comparable concentrations of

Page 41: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

41

drug in HBSS-HEPES buffer (pH 7.4) without protein were analyzed to ensure that binding to the filter material did not occur.

3.12. Accounting for protein binding The values for Cu were used in the appropriate equation (Equations 5 to 7) to replace the values of the measured initial concentration on the donor side.

3.13. Theoretical solubility calculation Prediction models for aqueous solubility are still under development; however, rough estimations are often possible based on the existing models and the Hender-son-Hasselbalch equation. The following three linear correlations by Hansch and co-workers (1968), Yalkowsky and Valvani (1980) and Meylan and co-workers (1996) were used to estimate the solubility of the compounds.

The equation by Hansch and co-workers is a linear correlation between logP and solubility (r2 = 0.87) based on organic liquids (1968).

98.0log34.11log0

Pcs Eq.13

The estimated solubilities for asimadoline and felodipine based on Equation 17 are 4.6 µM and 2.4 µM, respectively.

Yalkowsky and Valvani combined the lipophilicity term (logP) with the melting point, a solid state characteristic, in order to predict the solubility of solids.

PcTs m log2501.05.0log 0 Eq. 14

The estimated solubilities for asimadoline and felodipine based on Equation 18 are 3.6 µM and 2.6 µM, respectively.

The predicted solubility of un-ionized asimadoline at 25 C was estimated as 0.5 µM, using the following equation according to Meylan and co-workers (1996):

MWTPcs m 00468.0)25(0082.0log935.0978.0log 0 Eq. 15

where s0 is the water solubility in mol.l-1 of the un-ionized species, Tm is the melting temperature of 222 C (van Osdol and Watanabe, 1999) and MW is the molecular weight of 414.5 g.l-1. Values for clogP (ACDlogP) and pKa (ACDpKa) were calcu-lated using the ACDlabs databases version 4.56 (Advanced Chemistry Development Inc., Toronto, Canada). The predicted solubility at a given buffer pHd (s(pHd); pHd = pH in the donor compartment) can then be estimated using the Henderson-Hasselbalch equation:

Page 42: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

42

)(

)(

0)( 10101

pKapH

pKapH

pH d

d

dss Eq. 16.

For asimadoline the predicted solubility increased by a factor of 54 (from 0.54 µM to 29.4 µM) when the pH was reduced from pH 8.0 to 5.0. In contrast, felodipine does not ionize over the pH range in the GI tract; therefore, pH-dependent variations in its solubility are not likely. The corresponding predicted solubility of felodipine was 0.4 µM (25 C, pH 7.4, Tm = 142 C, pKa = 3.97). This value is in reasonably good agreement with the measured aqueous solubility of felodipine, which is 2.6 µM at 37 C (Scholz et al., 2002).

In order to ensure that the compounds were in solution in HBSS (pH 7.4, 37 C), an initial donor concentrations below the (predicted) solubility limit, i.e. asimadoline at 25 nM and felodipine at 60 nM, was used in Paper IV.

3.14. Critical micelle concentration determination For the approximation of the CMC of CEL the following method was used. 1,6-Diphenyl-1,3,5-hexatriene (DPHT) is a fluorescence dye; the intensity of fluores-cence increases greatly above the CMC of CEL because of incorporation of the compound into the hydrophobic interior of the micelles. The abrupt change in inten-sity represents the CMC value (Zhang et al., 1996). A DPHT stock solution was added to various concentrations of CEL solution prepared in HBSS-HEPES buffer (pH 7.4) in a 96-well plate. The final DPHT concentration was ¼ of a saturated solution (pH 7.4) in all wells and the final volume per well was 200 µl. These sam-ples were equilibrated for 24 h in the dark at 37 C. A SPECTRAmax® GeminiXS spectrofluorophotometer from Molecular Devices (Sunnyvale, California, U.S.A.) was used to measure DPHT fluorescence. Each surfactant concentration was meas-ured in quadruplicate. The wavelengths of excitation and emission were 355 nm and 428 nm, respectively. The temperature was controlled at 37 C.

3.15. Statistical analysis Values are expressed as the means standard deviation (SD). All cell culture ex-periments were performed at least in triplicate. The protein binding was determined in duplicate.

The statistical difference between the permeabilities of the cells to the drugs at dif-ferent pH gradients or in the absence or presence of 4% BSA was calculated using an unpaired t-test with a two-tailed distribution for comparison of two mean values.

ANOVA was used when more than two mean values were compared; i.e. when the statistical difference between the permeabilities of the cells to the drug at different BSA concentrations was calculated (0, 1, 2, and 4%). A p-value of less then 0.05 was considered statistically significant.

Page 43: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

43

4. RESULTS and DISCUSSION

4.1. pH gradient systems across the Caco-2 monolayers To investigate the passive and active transport processes under the influence of the physiological pH gradient of the intestine, the pH-dependent transport rates of weakly basic (Paper I), weakly acidic (Paper II) and neutral drugs (Paper I) were studied bidirectionally across Caco-2 monolayers (pH 5.0-8.0).

4.1.1. Effect on weakly basic drugs Weakly basic compounds, which were chemically and metabolically stable under the in experimental conditions, were selected for Paper I. The compounds were classi-fied as either poorly permeable or highly permeable using calculated physicochemi-cal properties such as polar surface area and lipophilicity (Palm et al., 1996). At-enolol, a poorly permeable compound that is predominantly passively transported across the cell membrane, and metoprolol, a compound that is highly permeable, were selected. Two more compounds that are actively effluxed from cells via P-gp were also selected: talinolol, which is poorly permeable, and quinidine, which is highly permeable (Table I).

Table I: Oral fraction absorbed (fa) and pKa values of the investigated drugs __________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

Permeability Efflux Compound pKa(a) fa(b) [%]

____________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

Low No Atenolol 9.6 51 22 High No Metoprolol 9.7 98 41 ____________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

Low Yes Talinolol 9.17(c) 52 14 High Yes Quinidine 4.2; 7.9 78 12 ____________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

(a) Experimentally determined pKa values (Craig, 1990). (b) S-atenolol (Mehvar et al., 1990); metoprolol (Feliciano et al., 1990; Ungell and Karlsson, 2003), S-talinolol (Bogman et al., 2005; Zschiesche et al., 2002), quinidine (Rakhit et al., 1984; Ungell and Karls-son, 2003) (c) Value for pKa was calculated using the ACDlabs databases version 4.56 (Advanced Chemistry Devel-opment Inc. Toronto, Canada).

At equal buffer pH values of 6.5 or 7.4 on both sides of the monolayers, the bidirec-tional permeabilities for metoprolol were not significantly different (Figure 10). In contrast, the transport rates in the absorptive direction were lower, and those in the secretory direction were higher in the presence of a pH gradient (apical pH < baso-lateral pH, Figure 10). This behavior was also observed for the poorly permeating compound atenolol (Paper I) and is in accordance with the pH partition hypothesis. In Paper I, the efflux ratio (apical pH 6.5, basolateral pH 7.4) of metoprolol was determined as 4.5 0.4. A comparable efflux ratio (4.8 0.2), displaying one log unit in difference between the two buffer pH values, was seen with an apical pH 5.5 and basolateral pH 6.5 (Figure 10). These efflux ratios, obtained under different pHconditions, but with the same relative change in the ratios of uncharged to charged amounts of compound in both compartments, were in good agreement and, thus, could be taken to reflect passive efflux. These results illustrated that passive (pseu-do-) efflux can occur across Caco-2 cell monolayers under certain experimental

Page 44: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

44

conditions. Thus, the asymmetric transport was indeed the result of the pH-dependent concentration of un-ionized species of the drug in the apical compart-ment.

Figure 10: Bidirectional pH-dependent transport of the rapidly transported weakly basic drug, metoprolol, across Caco-2 monolayers. When the basolateral pH was kept at 6.5 (A) or 7.4 (B) and the apical pH was increased, the apparent permeabil-ity coefficient (Papp) in the absorptive direction (black bars) increased and the Pappin the secretory direction (white bars) decreased. Each bar indicates mean SD (n

3). The values used in Figure B are published in Paper I, for comparison the scale is modified.

The transport rates in the absorptive direction were in general higher at a basolateral pH of 6.5 (Figure 10A) than those at pH 7.4 (Figure 10B; data published in Paper I), when values at the same apical pH were compared. In contrast, the transport rates in the secretory direction were consistently lower at a basolateral pH of 6.5 than at 7.4. These results were in agreement with the pH-partitioning hypothesis. The fraction of a basic compound that was ionized decreased as the pH changed from 6.5 to 7.4, and the concentration of un-ionized compound was lower at basolateral pH 6.5, thus offering an increased sink for the absorptive transport direction. In the secretory transport direction, however, the un-ionized species, which drove the passive trans-membrane transport of the compound, was lower at pH 6.5 than at 7.4 for a weakly basic drug, leading consequently to lower transport rates at basolateral pH 6.5.

For the actively effluxed weakly basic compounds talinolol and quinidine, the trans-port rates in the absorptive direction increased and those in the secretory direction decreased as the apical pH increased. However, at equal buffer pH values on both sides of the monolayers, the impact of the apical efflux transporter P-gp was still reflected by the significant efflux ratios of 6.0 and 4.0 at concentrations of 58 nM (talinolol) and 50 nM (quinidine), respectively (Paper I).

Several approaches for improving the predictive power of the Caco-2 cell assay have been investigated. Firstly, as suggested by Yamashita and co-workers (Yamashita et al., 2000) and Yee (Yee, 1997), an optimized permeability model for predicting fashould include transport rates obtained in a pH-gradient system using an apical pHof 6.0 or 6.5, respectively, and a basolateral pH of 7.4. These pH-gradient systems mimic the acid microclimate in vivo. Secondly, a smaller pH gradient of 6.8/7.4 (apical pH/basolateral pH), as suggested by Volpe (Volpe, 2004), might be not effi-cient enough to establish a good functioning of proton-dependent transporters (Fig-ures 14 and 16). Finally, extreme pH gradients (e.g. 5.0/7.4 and 5.5/7.4) can be a

0

100

200

300

5 5.5 6 6.5 7 7.5 7.7 8pH on the apical side

P app [

cm s

-1] x

10-6

absorptive (a-b) transport

secretory (b-a) transport

0

100

200

300

5 5.5 6 6.5 7 7.4 7.7 8pH on the apical side

P app [

cm s

-1] x

10-6

absorptive (a-b) transportsecretory (b-a) transport

BA

Page 45: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

45

challenge for the cells and the investigator. (The proportion of cell monolayers re-taining good integrity (TER, Mannitol) is generally lower than at apical pH 6.0 or 6.5.) However, the data presented here and in Paper I show that a system with dif-ferent pH values on the apical and basolateral sides will lead to different concentra-tions of uncharged drug species, which in turn will lead to asymmetry in bidirec-tional transport. This asymmetry in the transport rate over the membrane can occur independently of active transport. Therefore, it is not advisable to use, as suggested in the literature (Volpe, 2004), an efflux ratio obtained in a pH-gradient system for predictions of active transport involvement.

While drug efflux ratios obtained from a pH-gradient system can falsely indicate active apical efflux for weakly basic drugs in in vitro systems, these pH-gradienteffects may influence drug absorption in vivo. Therefore, a refined screening model for the prediction of fa, assuming that the jejunum is the major site of absorption for the drug, passive permeation of a compound should be measured at pH 6.0/7.4 in the absorptive transport direction only. This pH gradient mimics the acid microclimate and supplies a proton gradient over the entire monolayer as well as over the apical membrane.

4.1.2. Effect on weakly acidic drugsThe model acidic compounds indomethacin and salicylic acid were selected for Paper II since they are sufficiently soluble for the purpose of this study (Yazdanian et al., 2004), chemically and metabolically stable and ionizable within the physio-logical pH range of the GI tract. While indomethacin is predominantly passively transported across the cell membrane, salicylic acid is both passively and actively transported (Table II).

Table II: pKa values of the investigated drugs __________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

Permeability Main transport Compound pKa(a)

____________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

High Passive Indomethacin 4.5 High Active Salicylic acid 2.97; 13.4 ____________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

(a) Experimentally determined pKa values (Craig, 1990).

In contrast to the results obtained for weakly basic compounds, transport of the acidic compounds was higher in the absorptive direction than in the secretory direc-tion when the apical pH was lower than the basolateral pH. Figure 12 shows the bidirectional pH-dependent transport of indomethacin when the apical pH was var-ied between 5.0 and 8.0, while the basolateral pH was kept at 6.5 or at 7.4. The val-ues for Figure 11B are published in Paper II. The transport rates of indomethacin decreased 22-fold and 14-fold for the basolateral pH of 6.5 and 7.4, respectively, in the absorptive direction with increasing apical pH, while transport rates in the secre-tory direction increased 10-fold and 20-fold for the basolateral pH of 6.5 and 7.4, respectively (Figure 11).

Indomethacin is practically insoluble in water. The predicted solubility of the un-ionized species at 25 C is, according to Equation 13, estimated as 1.8 µM, Equation 14 as 8.5 µM, and Equation 15 as 2.3 µM. The dissolved radioligand was supplied in ethanol. The final concentration of ethanol in the donor solution was no more than 1%.

Page 46: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

46

Figure 11: Bidirectional pH-dependent transport of the rapidly transported weakly acidic drug, indomethacin, across Caco-2 monolayers. When the basolateral pH was kept at 6.5 (A) or 7.4 (B) and the apical pH was increased, the apparent per-meability coefficient (Papp) in the absorptive direction (black bars) decreased and the Papp in the secretory direction (white bars) increased. Each bar indicates mean SD (n 3). The values used in Figure B are published in Paper II.

This was expected, according to the pH-partitioning hypoyhesis, since the amount of uncharged species that drove the passive permeation of the drug, in the apical com-partments decreased with increased buffer pH from 5.0 to 8.0.

Using data presented in Paper II, the efflux ratio (apical pH 6.5, basolateral pH 7.4) for indomethacin was determined as 7.4 0.5 by calculating the permeability in absorptive direction in the non-sink equation (Equation 7; Papp (a-b) = 257 8 cm.s-1 x 10-6 and Papp (b-a) = 34.6 2 cm.s-1 x 10-6). A comparable efflux ratio for indometha-cin (apical pH 5.5, basolateral pH 6.5) of 7.7 0.4 was obtained using the transport rates shown in Figure 11. These values were one indication of the mainly passive transport of indomethacin across Caco-2 cell monolayers.

In addition, indomethacin transport was not inhibited by MCT1-, OATP-B-, or glu-cose-transport inhibitors (L-Lactic acid, pravastatin, phloretin), or by salicylic acid (Paper II). Most importantly, there was no concentration dependency of indometha-cin observed in any pH-gradient system, i.e. in a system where the apical pH was lower than the basolateral pH and in a system where a pH of 7.4 was applied in both compartments. Hence, the asymmetric transport of indomethacin was not caused by real carrier-mediated uptake in the Caco-2 cell system . As with weakly basic com-pounds in Paper I, it was shown in Paper II that asymmetric transport of weakly acidic compounds, observed using a pH-gradient system, can falsely indicate active transport in in vitro systems. However, physiological pH-gradient effects of at least one pH unit may influence drug absorption in vivo as they do for weakly basic com-pounds (Paper II).

Salicylic acid was the second compound investigated in Paper II. The results for salicylic acid indicated interplay of both pH-dependent active transport and passive diffusion. General inhibitors were used to inhibit the active transport of salicylic

It should be noted that indomethacin is actively transported in vivo by OATs. However, these transporters are probably not expressed in the Caco-2 cell clone used (Seithel et al., 2004).

0

100

200

300

400

500

5 5.5 6 6.5 7 7.4 7.7 8pH on the apical side

P app [

cm s

-1] x

10-6

absorptive (a-b) transportsecretory (b-a) transport

0

100

200

300

400

500

5 5.5 6 6.5 7 7.4 7.7 8pH on the apical side

P app [

cm s

-1] x

10-6

absorptive (a-b) transportsecretory (b-a) transport

A B

Page 47: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

47

acid. The data indicated that OATP-B and MCT1, two proton-dependent transport-ers in human intestine, which are also expressed in Caco-2 cells (see appendix), are involved in the uptake of salicylic acid. These data also left room for the speculation that a third transporter is involved in the transport process. This transporter can be either an apical uptake or a basolateral efflux transporter, which seems to respond only to a pH gradient over the whole monolayer. Since no accumulation of salicylic acid in the cell monolayer was observed, a basolateral efflux transporter that trans-ports salicylic acid actively out of the cell was suggested. A possible candidate is the heterodimeric organic solute transporter (OST) / for which mRNA was recently detected in kidney and intestine (Dawson et al., 2005; Hubbert et al., 2004). In Paper II, an extreme pH gradient (apical pH < 6.0) was used, which can be helpful for mechanistic studies, investigating proton-dependent transport, to mimic a high pro-ton concentration located close to the transporter or to inhibit the NHE3 (Thwaites et al., 2002). Therefore it can further be concluded that the proposed additional trans-porter for salicylic acid should be directly or indirectly pH-gradient dependent but should be independently working of the NHE3, since the NHE3 is inhibited under the investigated pH-gradient conditions (5.0/7.4/7.4).

The results of Papers II indicated that a pH-gradient over the monolayer, rather than over the apical membrane alone, is important for the function of some transporters. As a refinement it can be suggested that weakly acidic compounds should be studied in both a pH-gradient system across cell monolayers (apical pH 5.0-6.0 and baso-lateral pH 7.4) and without a pH gradient, when investigating active uptake (baso-lateral efflux) systems.

4.1.3. Effect on concentration dependency The transport of the neutral P-gp substrate, digoxin, was unaffected by changes in buffer pH, but was affected by changes in donor concentration (Papers I and III). Constant efflux ratios (8.9 0.8; n = 90) were obtained over the physiological pHrange, reflecting the efflux of the neutral drug digoxin by a transporter, with an ac-tivity that is independent of the pH gradient. However, the efflux ratios were re-duced when the donor concentration was increased. This was mainly a consequence of a decrease in the transport rates in secretory direction, caused by saturation of the transporter. Thus, the function of P-gp seems not to depend on pH, but the efficiency of the transport depends on the amount of compound available at the binding site. Thus, any pH change along the GI tract would not affect digoxin directly. This result is in agreement with the decreased permeation of digoxin further down the intestine, which is consistent and can be correlated with an increase in P-gp expression further down the GI tract (Stephens et al., 2001; Zimmermann et al., 2005).

Concentration-dependent transport and inhibition studies of compounds are nor-mally investigated with the same buffer pH on both sides of the monolayers, e.g. taurocolic acid (Wilson et al., 1990), vinblastine (Hunter et al., 1993), biotin (Ng and Borchardt, 1993) and fexofenadine (Petri et al., 2004). However, proton-dependent transporters have also been investigated in pH-gradient systems (apical pH < baso-lateral pH), e.g. salicylic acid (Takanaga et al., 1994), benzoic acid (Tsuji et al., 1994) and Gly-Sar (Bravo et al., 2004).

Page 48: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

48

For weakly basic compounds, the pH gradient used has to be accounted for (Paper I). This can be illustrated by the change in the efflux ratios at different concentra-tions and in different pH-gradient systems. Figure 12 shows the calculated pH-dependent efflux ratios for quinidine at three different donor concentrations .

Figure 12: Efflux ratios calculated from the bidirectional pH-dependent transport rates of the weakly basic drug, quinidine, across Caco-2 monolayers at three concentrations. When the basolateral pH was kept at 7.4 and the apical pH was increased, the ap-parent permeability coefficient (Papp) in the absorptive direction increased (n 3) and the Papp in the secretory direction decreased (n 3). Hence, with increased apical pH the efflux ratios decreased. Within each pH-gradient system, an increased concentration of quinidine gave a reduction in the efflux ratio. Each bar indicates mean SD.

The dramatic change in efflux ratios between the various pH-gradient systems re-flected the combined alterations in the transport rates of the various active and pas-sive transport processes that were involved. The alterations within a pH-gradientsystem can be assigned to the concentration-dependent part of the transport, which is generally assumed to be the active part. For quinidine, the most important trans-porter is probably P-gp (Emi et al., 1998). Since quinidine is a weakly basic com-pound, only the concentration-dependency observed, when an equal pH of 7.4 was applied on both sides of the monolayer, reflects the active efflux. The concentration-dependency in all other pH-gradient systems is a mixture of active and passive ef-flux. The functioning of P-gp is independent of intracellular and extracellular pH(Altenberg et al., 1993) and, if P-gp would be the only transporter involved in the transport of quinidine, comparable Km values within the different pH-gradient sys-tems should be obtained, when calculating only with the un-ionized amount of the drug.

As with the weakly basic compounds, the concentration-dependency relationship within one particular pH-gradient system does not necessarily contain the informa-tion for distinguishing between active and passive transport of a weakly acidic com-pound. Figure 13 shows the concentration-dependent and pH-gradient dependent bidirectional transport rates for salicylic acid.

The efflux ratios of the lowest concentration (50 nM) are calculated from data of the pH-dependent bidirectional transport rates for quinidine published in Paper I.

0

5

10

15

20

25

30

35

40

45

50

6 6.5 7 7.4 7.7 8

pH on the apical side

Efflu

x ra

tio

50 nM50 µM150 µM

Page 49: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

49

Figure 13: Bidirectional pH-dependent transport of the rapidly transported weakly acidic drug, salicylic acid, across Caco-2 monolayers at three concentrations. When the basolateral pH was kept at 7.4 and the apical pH was increased, the apparent permeability coefficient (Papp) in the absorptive direction (black bars) decreased and the Papp in the secretory direction (white bars) increased. The extent of the change in permeability was concentration and pH dependent. Each bar indicates mean SD (n 3).

No alterations in the transport rates for salicylic acid were observed when the apical and basolateral pHs were equal at 7.4. However, transport rates in the absorptive direction increased and those in the secretory direction decreased with decreased apical pH and within a pH-gradient system with decreased concentration of salicylic acid .

When the pH-gradient system 5.0/7.4/7.4 (apical pH/intracellular pH/basolateral pH)was used, a Km value of 5.4 0.7 mM was calculated (Paper II). Takanaga and co-workers reported a Km value, probably measured at apical pH 6.0 and basolateral pH7.4 across Caco-2 monolayers, of 5.28 0.72 mM (Takanaga et al., 1994). However, at pH 7.4 on both sides of the monolayers, a Km value cannot be obtained. The ap-parent Km values calculated for salicylic acid will decrease with increased apical pHas indicated by these results (Figure 13). It can therefore be argued that the use of apparent Km values obtained from a pH-dependent process can be questioned espe-cially when several transporters are working simultaneously.

The pH-dependent bidirectional transport rates for salicylic acid at 25 µM are published in Paper II.

0

200

400

600

800

25 50 100 25 50 100 25 50 100 25 50 100

Papp

[cm

s-1

] x 1

0-6absorptive (a-b) transport

secretory (b-a) transport

Concentration [µM] :

apical pH: 5.0 5.5 6.5 7.4 basolateral pH: 7.4 7.4 7.4 7.4

Page 50: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

50

4.1.4. Effect on temperature dependency It has been suggested that alterations in drug transport across tissue and monolayers due to temperature changes can be used to identify active transport processes (Hidalgo and Borchardt, 1990a). Hidalgo and Borchardt state that ‘the normal range for Ea values associated with enzymatic reactions or carrier-mediated processes lies between 7 and 25 kcal.mol-1 (30-107 kJ.mol-1), while Ea values associated with sim-ple diffusion processes are less than 4 kcal.mol-1 (17 kJ.mol-1)’. In Paper II, the ef-fect of temperature change on the passive and active transport of two acidic com-pounds, indomethacin and salicylic acid, was investigated. It was found that the transport rates of both compounds were highly affected by temperature and pH as illustrated for salicylic acid in Figure 14A. The Ea values were obtained from the slope of the corresponding Arrhenius plot (Figure 14B, Table III).

Figure 14: (A) Temperature dependence of the pH-dependent transport of salicylic acid (25 µM) across Caco-2 monolayers. (B) Arrhenius-plot. The apparent perme-ability coefficients (Papp) in the absorptive direction were measured at 4, 7, 17, 27 and 37°C, using the non-gradient systems (pH 7.4) on both sides of the monolayer, or in the presence of a pH-gradient (apical pH 5.0, 5.5 or 6.5 and basolateral pH 7.4), or in the presence of a double pH-gradient system (pH 5.0, 5.5 or 6.5 on both sides of the monolayer, while the pHi was always assumed to be around 7.3). Each point indicates mean ± SD, (n 4).

The results presented here and in Paper II do not support the hypothesis that passive and active transport of - weakly acidic or weakly basic - compounds across a lipid membrane are distinguishable by change in temperature. In contrast, Ea values for passively and actively transported compounds significantly overlap (Table III); precluding their use as a reliable criterion for the identification of active transport.

permeability values were obtained at the same initial donor concentration

A B

-0.5

0.5

1.5

2.5

3.53.1 3.3 3.5 3.7

1/Temperature [°K] x 10-3

Log

P app

[cm

.s-1

] x 1

0-6

5.0/5.0 5.5/5.5 6.5/6.5 7.4/7.45.0/7.4 5.5/7.4 6.5/7.4 7.4/7.4

0

200

400

600

800

0 10 20 30 40Temperature [°C]

P app

[cm

.s-1] x

10-6

5.0/5.0 5.5/5.5 6.5/6.5 7.4/7.45.0/7.4 5.5/7.4 6.5/7.4 7.4/7.4

Page 51: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

51

Table III: Activation energies ____________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

Compound Main transport pH on the apical side 5.0 5.5 6.5 7.4 ____________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

Indomethacin Passive 35(a) 38 n.d. 48 Salicylic acid Active 50 58 69 70 Metoprolol Passive n.d. 19 46 75 ____________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

n.d. not determined. The basolateral pH was always kept at 7.4. (a) Activation energies were expressed in units of [kJ.mol-1].

4.1.5. pH Dependency of drug-drug interactions Apart from digoxin, several compounds that are clinically often co-administered with digoxin also use P-gp as a transporter, thus, there is potential for DDIs (Englund et al., 2004). In the screening phase, the final dose of the drug is often not known; however, data presented here and in Paper I indicate the importance of a good estimate for this parameter in designing a relevant DDI study in vitro.

In Paper I, the dependency of the clinically relevant digoxin-quinidine interaction on pH was investigated. The concentrations used were estimated from (i) the given dose, (ii) the relative relationship of the doses of each drug in vivo, (iii) the concen-tration causing toxicity to the cells and (iv) the Cmax of the compound from in vivostudies. The transport rates of the two compounds both, in combination and sepa-rately, were measured across Caco-2 cell monolayers. The digoxin-quinidine inter-action was clearly dependent on pH, although digoxin transport itself is not pH de-pendent (Paper I), indicating that the interaction is dependent on the amount of charged and uncharged species of quinidine available at the binding site of P-gp.

By using different ratios of the dose of quinidine and digoxin, the relevance of this interaction becomes even more apparent (Figure 15). If the quinidine concentration is low, the interaction in vitro is not significant (Figure 15A) and from this result, the relevance in vivo would be predicted to be insignificant; however, if the quinidine concentration is high in comparison to the relative dose ratio, then the interaction in vitro is dramatic (Figure 15B). Thus, conclusions relevant to the invivo situation could be drawn from these data.

Figure 15: pH-Dependent permeability values for bidirectional transport of the neutral P-gp substrate, digoxin (2 nM), in the presence of (A) 10 µM quinidine or (B) 100 µM quinidine. Values are means SD (n 3).

A B

0

25

50

5 5.5 6 6.5 7 7.4 7.7 8pH on the apical side

P app [

cm s

-1] x

10-6

absorptive (a-b) transportsecretory (b-a) transport

0

25

50

5 5.5 6 6.5 7 7.4 7.7 8pH on the apical side

P app [

cm s

-1] x

10-6

absorptive (a-b) transportsecretory (b-a) transport

Page 52: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

52

It should be borne in mind that the in vitro experiments should help to predict and explain in vivo observations. The concentration-dependency relationship of the competitive inhibition studies shown in Figure 15, indicate that the relative dose ratio will determine the extent of the DDI. Thus, the concentration of free drug in its un-ionized form will determine the amount of drug that enters the cell via the apical membrane. This amount has to be high enough to reach a concentration that affects the transporter (above the apparent Km of the drug on P-gp) and thus inhibit the transport of the competing drug. Thus, the dependence on pH in combination with the relative concentration ratio has clinically relevant repercussions for P-gp-dependent DDI and should be considered in the further design of in vitro models for DDIs.

Independently of whether the inhibition is competitive or non-competitive, the pH-dependency of the inhibitor has to be taken into account, since the pKa of the inhibi-tor will determine the amount of inhibitor available at the inhibitor-binding site of the transporters. If the pKa value of the compound is known, the appropriate condi-tions can be applied in the permeability assay. If the pKa value is not known, the concentration-dependency and inhibition studies might have to be performed twice, with and without a pH gradient.

4.1.6. The paracellular pathway In vivo pharmacokinetics indicate that atenolol is slowly and incompletely absorbed (Mehvar et al., 1990). This is in agreement with in vitro data that suggest that at-enolol slowly permeates the GI membrane, thus resulting in a low fraction absorbed. Consistent with the pH-partition hypothesis, basic molecules are less ionized in the ileum (pH = 7.3) than in the jejunum (pH = 6.5) and are therefore more quickly absorbed in the lower parts of the GI tract (Taylor et al., 1985). This pH-dependenttransport of bases was confirmed in the in vitro experiments published in Paper I using Caco-2 cell monolayers. However, the hydrophilic and poorly permeating basic -blocker atenolol shows slightly faster transport in the upper GI tract (Taylor et al., 1985), indicating a more significant role of the paracellular route in the jeju-num.

The paracellular pathway cannot be ignored in the Caco-2 cell model; in fact, paracellular diffusion has been investigated previously in this model (Adson et al., 1995; Adson et al., 1994; Karlsson et al., 1999; Nagahara et al., 2004; Pade and Stavchansky, 1997; Tavelin, 2003). The impact of the paracellular pathway can be seen in the different efflux ratios obtained for atenolol and metoprolol in the pres-ence of a pH gradient (apical pH 6.5; basolateral pH 7.4) (Paper I). Although at-enolol and metoprolol have approximately the same pKa values and should, accord-ing to pH-partitioning theory, be equally affected by the changes in the buffer pH,the efflux ratio for atenolol is lower (1.7 0.2) than for metoprolol (4.5 0.4). The contribution of the paracellular pathway affects the ratio more for the poorly perme-ating drug, atenolol, than for the rapidly permeating drug, metoprolol. Adaptation of the calculations in the publication by Palm et al. (Palm et al., 1999) reveal that the

Page 53: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

53

paracellular route contributes less than 20% to the overall transport of atenolol, compared with around 0% for metoprolol.

At low donor concentrations (25 µM) and in a pH gradient of 5.0/7.4/7.4 (apical pH/intracellular pH/basolateral pH), over 80% of the salicylic acid transport is via active carrier-mediated transport, 17% is via pure passive diffusion and less than 3% is via the paracellular pathway as estimated from the data presented in Paper II. In the non-gradient system, only passive transcellular diffusion and the paracellular pathway are involved. Pade and Stavchansky estimated the contribution of the paracellular pathway to the overall transport of salicylic acid as 1% at apical pH 7.4 and 0% at apical pH 5.4 (Pade and Stavchansky, 1997). Thus, the extent to which the various active and passive transport processes are involved in drug transport is affected by the fraction of ionized and un-ionized species according to the pH-partitioning theory.

4.2. BSA-gradient systems across Caco-2 monolayers Two groups of drugs were investigated in Paper III. Group 1 included compounds transported by passive diffusion only and group 2 comprised compounds whose transport is influenced by active apical efflux systems such as P-gp and/or BCRP (Table IV). Each group contained compounds with either high or low protein bind-ing and high or low fa, as ascertained from the literature (Table IV). In order to cover a wide variety of compounds and to fulfill the above criteria for the in vitro experi-ments, the ‘low’ and ‘high’ limits for both protein binding and fa were set as 0-30% and 70-100%, respectively (Table IV). This was despite the general view that a compound is highly protein bound in vivo when over 95% of the compound is plasma protein bound and that a compound is considered to be easily absorbed when the extent of absorption in humans is 90% or more of an administered dose.

Table IV: Oral fraction absorbed (fa) and protein binding (PB) of the investigated drugs.__________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

PB Permeability Efflux Compound PB [%] fa(a) [%]

___________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

Low Low No Mannitol - 16-26 Low High No Metoprolol 11 80-100 High High No Propranolol 87 90-100 High Low No n. a. n. a. n. a. ____________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

Low Low Yes Inogatran 10-20 < 5 Low High Yes Digoxin 25 81 High High Yes Quinidine 87 66-90 High Low Yes Doxorubicin 76 > 5 ____________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

n. a. = not available.(a) Values from Artursson and Karlsson, 1991, Ungell and Karlsson, 2003, Benet et al., 1996 and refer-ences therein.

For the purpose of the study described in Paper I atenolol was selected as a poorly permeat-ing drug that is predominantly passively transported across the cell membrane, since the paracellular contribution (< 20%) does not affect the general interpretation of the data. How-ever, a discussion of the paracellular transport in Paper I was avoided, since the aim was to emphasize the impact of a pH-gradient on the transcellular pathway.

Page 54: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

54

The results presented in Paper III show that basolaterally applied BSA, with the aim of mimicking in vivo protein binding, modified both passive and active drug trans-port in the Caco-2 cell model. The inclusion of BSA helped to maintain sink condi-tions during drug transport in the Caco-2 cell model. Basolateral BSA increased transport in the absorptive direction and decreased transport in the secretory direc-tion. Thus, the inclusion of BSA in the receiver compartment increased the transport rates, while inclusion of BSA in the donor compartment decreased the rates. This was also observed in Paper IV, where BSA was applied either on one side or on both sides of the monolayers. In order to account for the unbound substance in the pres-ence of BSA, the unbound fraction (fu) was determined and the unbound concentra-tion (Cu) of each of the compounds was calculated, according to Equation 16. De-termination of Cu provides an opportunity to elucidate the different mechanisms behind the alterations.

4.2.1. Accounting for the maximal effect of protein binding If the recovery of the compound is around 100% in the absence of BSA and the compound is highly protein bound (99.9%; fu = 0.1%), then according to Equation 17, the expected maximal increase in permeation due to extracellular BSA would be approximately twice that obtained without BSA, assuming that the BSA represents a perfect sink which reduces the back flux of the compound across the basolateral membrane to zero. Thereby also reducing the intracellular concentration to such a low level that no compound passively effluxed over the apical membrane.

)%4(100

)%4(100100)%0( )()( BSAP

BSAfBSAP baapp

ubaapp

Eq. 17

where Papp(a-b) (0% BSA) is the apparent permeability in the absorptive direction in the absence of BSA, Papp(a-b) (4% BSA) is the apparent permeability in the absorptive direction in the presence of 4% BSA, and fu(4% BSA) is the percentage of unbound compound in the presence of 4% BSA.

This equation can be used for compounds that have poor recovery in traditional studies of in vitro permeability (without BSA), often seen for lipophilic BCS class II compounds, which are per definition poorly soluble and highly permeating. For these compounds a maximal theoretical transport rate in the absorptive direction can be estimated with Equation 17. Felodipine and asimadoline are two highly lipophilic BCS class II compounds which can be easily trapped in the lipid membrane (Paper IV). Inclusion of BSA reduced the amount of felodipine and asimadoline associated with the cell monolayer, the filter support and to the plastic of the equipment (Figure 16). These results explain the observed increase in the recovery of these two com-pounds at the end of the experiments in the presence of BSA compared to experi-ments without BSA.

Page 55: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

55

Figure 16: Effect of increased con-centrations of basolaterally applied bovine serum albumin (BSA, w/v) on the intracellular concentration of asimadoline at an initial donor con-centration of 25 nM. Values are means SD (n = 3).

Using Equation 17, a theoretical maximum transport rate for the passively trans-ported felodipine in the absorptive direction would be 75 x 10-6 cm.s-1. For the ac-tively transported asimadoline, the corresponding theoretical maximum transport rate in absorptive direction would be 62 x 10-6 cm.s-1 at the investigated initial donor concentration (20 nM). This theoretical transport rate (for 100% recovery of the compound) can then be compared to permeation values obtained in the traditional Caco-2 cell model (without BSA), for instance, for the prediction of fa. This conver-sion facilitates the use of BSA in the Caco-2 cell model as a secondary screening tool.

4.2.2. Impact of basolateral BSA on transport in the absorptive directionFor passively transported compounds, the relative increase in the permeation of the compound across the Caco-2 cell monolayers in the absorptive direction was directly proportional to the extent of protein binding for each compound (Figure 17A). Man-nitol showed no change in permeability as expected for a compound without PB (Figure 17A). The slope of 0.78 can be used to estimate the maximal expected im-pact of 78% change if a protein binding of 100% is assumed using the experimental set-up with basolaterally applied BSA and the limited set of compounds (mannitol, metoprolol and propranolol) (Figure 17A). The difference between the theoretical maximal change of 100% and the experimentally obtained maximal change of 78% can probably be explained with the amount bound to the monolayer (as representa-tive of the intracellular concentration, Figure 16) that was only reduced in the pres-ence of BSA but not totally abolished. Thus, the sink was improved in the presence of BSA, but not perfect.

0

100000

200000

300000

0 1 2 4Basolateral BSA [%, w/v]

Com

poun

d in

filte

r + m

onol

ayer

[d

pm]

drug added to the apical side of themonolayerdrug added to the basolateral sideof the monolayer

Page 56: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

56

Figure 17: The relative change in the apparent permeability coefficients (Papp) in the absorptive direction caused by the sink conditions created by basolaterally applied 1, 2 or 4% BSA is correlated to the fraction of compound bound to BSA. A good correlation for the passively transported compounds ( mannitol, metoprolol, propranolol) was obtained (r2 = 0.9) (A), while a very weak correlation was observed for the actively effluxed compounds ( inogatran, doxorubicin, digoxin, quinidine) (r2 = 0.4) (B), probably due to the dose dependency of the active transport component. Values are for each compound from two initial donor concentrations. The trendlines for the passively (solid line) and actively (dashed line) transported compounds are shown.

Figure 17B shows that the correlation coefficient for the actively effluxed com-pounds was r2 = 0.4 and the slope was slightly lower (0.74) than the slope of the correlation for the passively transported compounds (Figure 17A). For the actively transported compounds, binding to BSA led to a scattered and compound-specific deviation from the trendline obtained from the passively transported compounds. One reason for this could be that the relative change in permeation of actively ef-fluxed compounds in the absorptive direction was not only increased proportionally to the extent of protein binding for each compound, but at the same time additionally decreased by active transport in secretory direction. The total effect was then ob-served in the transport rates. The impact of active transport in the secretory direction on the overall transport in the absorptive direction was concentration dependent and more efficient at initial donor concentrations around the Km of the compound for the active transporter. The concentration at the transporter binding site was reduced because of the presence of BSA which, at least in the case of doxorubicin, reduced the amount of compound in the intracellular compartment (Paper III).

When Saha and Kou investigated transport in the absorptive direction, they con-cluded that the inclusion of 4% BSA in the receiver chamber during transport studies can dramatically affect the estimated permeation across Caco-2 monolayers and the

A B

y = 0.7424xR2 = 0.4168

-40

-20

0

20

40

60

0 30 60

Fraction [%] bound to BSA

Rel

ativ

e ch

ange

in P

app [%

]

actively effluxed compounds

y = 0.7825xR2 = 0.9187

-40

-20

0

20

40

60

0 30 60

Fraction [%] bound to BSA

Rel

ativ

e ch

ange

in P

app

[%]

passively transported compounds

Page 57: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

57

BCS permeability ranking of highly lipophilic compounds. This is presumably due to improved sink conditions and/or a reduction in non-specific drug adsorption to plastic wells (Saha and Kou, 2002) and the lipid bilayer as shown for asimadoline (Paper IV). They also suggested that the permeability classification of a compound with a log P less than 2.4 will not be affected by the presence of receiver BSA (Saha and Kou, 2002). However, as shown for doxorubicin (clogP = - 1.42; BioByte MacLogP V. 2.0 Software, measured logP = 0.1), the impact of BSA is strictly re-lated to the extent of protein binding and not to the lipophilicity of the compound (Figure 17B).

Inclusion of BSA can increase the recovery, and will affect all compounds in pro-portion to the extent of their protein binding under the experimental conditions. In an ideal case, it would be possible to account for the unbound concentration, e.g. by determining the free concentration via an ultracentrifugation method (Paper III). However, even when Cu cannot be measured, an estimate for purely passively trans-ported compounds can be arrived at, as discussed for felodipine (Paper IV). Inclu-sion of basolateral BSA without correction for protein binding probably gives a more physiologically sound model of absorption, since it mimics both protein bind-ing in the blood and the sink effect better than the aqueous buffer solution normally used in in vitro systems. However, this hypothesis remains to be shown on an in-creased number of compounds.

4.2.3. Impact of basolateral BSA on the efflux/uptake ratios Efflux and uptake ratios are used to identify any asymmetry in the bidirectional transport of a compound, which might be due to active transport. Ratios from com-parable transport rates, i.e. good recovery (< 95%) calculated for the same cut-off sink, at the same initial donor concentration and without a pH gradient (Chapter 4.1.), can be used to estimate the in vivo relevance of asymmetrical transport.

An uptake ratio is observed for passively transported compounds in the presence of basolateral BSA. Similarly, transport rates for the actively transported compounds increased in the absorptive direction and decreased in the secretory direction in the presence of basolateral BSA. Thus, the efflux ratios of the actively transported com-pounds were reduced in the presence of basolaterally applied BSA. This change suggests that active secretory transport for highly protein bound compounds might have less effect on in vivo absorption than predicted from traditional Caco-2 cell models (without BSA).

Under controlled experimental conditions, the transport rates in the absorptive and secretory directions are affected to the same degree by BSA. Since extracellular BSA can reduce the amount of compound at the binding site of active transporters, as observed for doxorubicin in Paper III and asimadoline in Paper IV, their impact on the observed overall transport will be altered in a concentration dependent man-ner. However, the presence of BSA does not directly affect the functionality of ac-tive transport systems (such as P-gp) as shown in Paper III.

Scheme 1 illustrates this deviation. In the presence of BSA a lower Papp is obtained in secretory direction (step 1), but this Papp actually represents a values for a lower

Page 58: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

58

concentration of drug, Cu, (step 2). Calculating a new Papp for this lower donor con-centration, Cu, will lead for a passively transported compound to a Papp, which is not different from the original Papp (Scheme 1A), but for an actively transported com-pound the Papp can increase (Scheme 1B).

Scheme 1: Basolaterally applied BSA reduces the unbound concentration of the compound in proportion to the protein binding of the compound. Thus, all the ap-parent permeability coefficients (Papp) in the secretory direction obtained in the pres-ence of BSA are lower than those obtained without BSA. Due to the different dose dependency of the passive (A) and active (B) transport component (dotted lines), Papp calculated with the unbound concentration as donor concentration can be higher than the Papp observed in the absence of BSA for actively transported com-pounds.

Thus, the involvement of active transport processes can be estimated, when the ef-flux ratios re-calculated from permeabilities considering the Cu for a compound differ from the efflux ratios obtained from the traditional model (without BSA) at the same initial donor concentration (Scheme 1). The greatest increase should be observed around the concentration of the Km value. This principle can be used to distinguish after re-calculation for fu between actively and passively transported compounds as shown in Paper III.

It has been argued that the role of apical efflux transporters as an intestinal barrier to oral bioavailability may have been overemphasized, particularly when evaluated in in vitro systems such as the Caco-2 cell model (Lin and Yamazaki, 2003). Signifi-cant transport in the secretory direction across cell monolayers does not necessarily correspond to unsatisfactory oral bioavailability in vivo, as seen for compounds like digoxin and cyclosporin A. Walgren and Walle observed in their study that the ef-flux of the investigated compounds in the presence of basolateral BSA was signifi-cantly reduced, mainly due to a reduction in the transport rates in the secretory direc-tion. They concluded that, if the effect of plasma binding is not considered, overes-timation of secretory transport could result in a misleading net flux calculation (Walgren and Walle, 1999). The inclusion of 4% BSA basolaterally significantly

A B

1.

2.

3.

concentration

Perm

eabi

li ty

(b-a

)

passively transportedcompounds

Measured valuein the presence of BSA

Theoretical valueaccounted for Cu

1.2.

3.

concentration

Perm

eabi

lity

(b-a

)

actively effluxedcompounds Measured value

in the absence of BSA

Theoretical valuenot accounted for Cu

1.

2.

3.

concentration

Perm

eabi

li ty

(b-a

)

passively transportedcompounds

Measured valuein the presence of BSA

Theoretical valueaccounted for Cu

1.2.

3.

concentration

Perm

eabi

lity

(b-a

)

actively effluxedcompounds Measured value

in the absence of BSA

Theoretical valuenot accounted for Cu

Page 59: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

59

reduced the efflux ratio for digoxin (30 nM) from 7.73 0.49 to 5.86 0.46, as reported in Paper III, indicating that lower amounts of the compound interacting with the apical efflux pump. Thus, in the presence of BSA the efflux will be esti-mated to be lower than in the absence of BSA.

Thus, efflux or uptake ratios calculated from transport rates obtained in the presence of BSA applied on the basolateral side of the monolayers can be used to determine the degree of overestimation of the apical efflux component in comparison with data obtained without BSA. This approach can be used as a selection tool in a small set of compounds, since the compounds can be ranked according to the change in their efflux or uptake ratios. Additionally, by accounting for Cu actively and passively transported compounds can be distinguished (Paper III), as illustrated in Scheme 1.

Table V lists the advantages and disadvantages of the possible models that can be studied by including BSA in the Caco-2 monolayer assays and the ’traditional sys-tem’ without BSA. In Paper III the ‘GI-mimicking model’ (basolateral application of BSA) was investigated. In Paper IV, BSA was applied either on both sides of the monolayer (‘protein model’) or only on the receiver side (‘receiver-side model’).

The data presented in Paper IV indicate that a compound with a high intrinsic (pas-sive) permeability like asimadoline can show a small, but ‘relevant’ asymmetry (efflux ratio) across in-vitro cell monolayers. While in the GI-mimicking model an uptake ratio was obtained, in the receiver-side model a small but significant efflux ratio was calculated (in agreement with calculated efflux ratios obtained from MDCK-MDR1 experiments), indicating the possible relevance of efflux transporter for asimadoline in the brain, but not in the GI tract. This interpretation is in agree-ment with in vivo studies of asimadoline in the rat (Jonker et al., 1999). However, the binding to BSA should in a blood-brain barrier (BBB) model be considered also on the apical side (blood side of a BBB-model), the sink will be provided by BSA in the blood, i.e. at the ‘donor’ side, and for a protein-bound drug, such as asimadoline, an additional effect of P-gp efflux and the BSA sink will result. Thus as a sugges-tion, the apical addition of 4% BSA and an addition of 1% BSA basolaterally, to mimic the protein content in the brain, should be considered in in vitro models of the BBB. This may improve predictions of in vivo efflux transport (‘BBB-mimicking-model’).

Page 60: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

60

Table V. Advantages and disadvantages of the absence or presence of bovine serum albumin (BSA) on either or both sides of the cell monolayers for lipophilic com-pounds.____________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

System[transport direction (+) Advantages; (apical / basolateral) % BSA] (-) Disadvantages ____________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

Traditional system (-) poor recovery [a-b (0/0); b-a (0/0)] (-) high adsorption to plastic, high cellular binding (+) equal initial donor concentrations

(-) poor data quality for permeability a-b and b-a (+) requires no determination of protein binding ____________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

‘GI-mimicking model’ (+) low adsorption to plastic, low cellular binding [a-b (0/4); b-a (0/4)] (+) good recovery (+) good data quality for permeability a-b

(-) poor data quality for permeability b-a, since the initial donor concentration is affected (-) free drug concentration has to be measured; compound-specific ____________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

‘Protein model’ (+) low adsorption to plastic, low cellular binding [a-b (4/4); b-a (4/4)] (+) good recovery (+) good data quality for efflux

(-) poor data quality for permeability a-b and b-a, since the initial donor concentration is affected (+) equal initial donor concentrations on both sides (-) free drug concentration has to be measured; compound-specific ____________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

‘Receiver-side model’ (+) low adsorption to plastic, low cellular binding [a-b (0/4); b-a (4/0)] (+) good recovery

(+) good data quality for permeability a-b and b-a (-) free drug concentration has to be measured in the receiver solution; compound-specific (+) equal initial donor concentrations ____________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

‘BBB-mimicking model’ (+) low adsorption to plastic, low cellular binding [a-b (4/1); b-a (4/1)] (+) good recovery

(-) poor quality permeability data a-b and b-a, since the initial donor concentration is affected

(-) free drug concentration has to be measured at different protein concentrations (+) equal initial total donor concentrations

4.3. Cremophor® EL and Caco-2 cells Two compounds were chosen as representatives of BCS class II drugs (low solubil-ity/high permeation, (Amidon et al., 1995)). The model drugs, felodipine and asi-madoline, were selected because of their known (felodipine; (Scholz et al., 2002)) and predicted (asimadoline) low solubility within the physiological pH range (5.0-8.0) of the GI tract. In order to study the effect of CEL on passive transport only,

Page 61: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

61

felodipine was chosen, since felodipine is permeating highly and passively across the lipid membrane. To study the effect of CEL also on active transport, asimadoline was selected, since it is known that asimadoline is effectively transported via the apically located efflux transporter P-gp (Bender and Dasenbrock, 1998). Asimado-line is a compound which is also completely absorbed over the intestinal mem-branes.

Table VI: Oral Fraction Absorbed (fa), Solubility and Protein Binding (PB) of the Investigated BCS class II Drugs.____________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

PB Permeability Solubility Efflux Compound PB fa [%] [%] ____________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

High High Low No Felodipine 99.6(a) 100(a)

High High Low Yes Asimadoline 97(b) 100(c)____________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

(a) Data for oral absorption and protein binding are taken from (Edgar et al., 1985) and (Artursson and Karlsson, 1991). (b) Data for plasma protein binding in rats are taken from (Bender and Dasenbrock, 1998) and (Jonker et al., 1999). (c) Data for oral absorption was predicted from permeation data across Caco-2 cell monolayers.

BSA was applied on the basolateral side and CEL on the apical side. This approach mimicked the in vivo situation in which BSA is in the blood/plasma of the capillaries and CEL is delivered with the drug. In addition, this design separated the two addi-tives, thus avoiding direct interactions between them (Vidanovic et al., 2003). The CMC for CEL was approximated in the concentration range of 0.0050 to 0.0075% (w/v) in HBSS-HEPES buffer. This estimated CMC range is comparable to that reported for CEL by Nerurkar and co-workers (0.0070-0.0088%, w/v) (Nerurkar et al., 1997).

Some surfactants are capable of incorporating themselves into the lipid bilayer. This can alter the membrane permeability. Mannitol and TER measurements were fol-lowed to monitor the monolayer integrity (Anderberg et al., 1992). No significant alterations in TER values or mannitol permeation were observed for concentrations up to 10% (w/v) CEL on the apical side and 4% BSA on the basolateral side in the study presented in Paper IV. Thus, the paracellular pathway and the electrical pa-rameters were not affected by CEL.

In contrast to the result of the paracellular marker, the transport rates for felodipine, a transcellularly transported compound, were affected above the estimated CMC of CEL (Figure 18, Paper IV). For instance, in the presence of 0.1% (w/v) CEL, a 3-fold reduction in felodipine permeation in the absorptive direction was observed, suggesting that micelle solubilization of felodipine decreased the fraction of free drug in the donor chamber. This, in turn, reduces the available amount of compound for transport across the monolayers. As a consequence, apparent permeability values based on the initial total felodipine concentration were substantially lower than without CEL.

Page 62: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

62

Figure 18: Transport of the transcellularly transported compound, felodipine, across Caco-2 monolayers in the absorptive (a-b) (black bars) and secretory (b-a) (white bars) directions in the presence of various Cremophor® EL concentrations (0 – 0.1%; w/v) on the apical side and bovine serum albumin (4%; w/v) on the baso-lateral side. Values are means SD (n 3).

If the free concentration of felodipine, thus the solubility in the aqueous solution in the donor compartment, could be determined, then it should be possible to correct the measured transport rates for felodipine in a similar way, as shown for BSA (Page 50). Since the micellar interaction is a fragile system the determination of the free concentration of felodipine is not made easily. However, the use of CEL would give a possibility to include compounds for which the dissolution rate is the limiting step to absorption, since the micelles are serving as a reservoir that is easily accessible.

Non-ionic surfactants such as polysorbate 80 and CEL are among the substances reported to inhibit the activity of apically polarized efflux systems such as P-gp (Nerurkar et al., 1996). However, the bidirectional transport rates of asimadoline were unaffected by the presence of CEL (Paper IV). Thus, it can be speculated that any increase in asimadoline transport as a result of the inhibition of P-gp efflux by monomers of the CEL or decrease in transport due to the effect of surfactant mi-celles was insignificant beside the much larger effect of the high capacity for bind-ing to BSA in the basolateral compartment. This efficient BSA binding is supported by the high lipophilicity of the compound. Thus, for an efflux compound with high lipophilicity, inclusion of CEL will be difficult to evaluate since several effects are present concurrently.

Inclusion of BSA on either or both sides of the monolayers is generally sufficient to increase the recovery of lipophilic BCS class II compounds. The overall effect of CEL on the permeation of a drug is more compound specific and less reliable and, therefore, less predictable for these drugs. Thus, the inclusion of BSA is recom-mended for BCS class II drugs, and the apical addition of a solubility enhancer such as CEL cannot be recommended.

0

25

50

75

100

125

150

175

0 0.00001 0.0001 0.001 0.0075 0.01 0.025 0.05 0.1

Cremophor® EL concentration on the apical side [%; w/v]

Papp

[cm

.s-1

] x 1

0-6 a-bb-a

Critical micelle concentration = 0.0050-0-0075%

Page 63: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

63

Taub and co-worker suggested a pH gradient (apical pH 6.0, basolateral pH 7.4), together with basolateral BSA (0.25%) and apically applied sodium taurocholate (5 mM) as an optimized system for MDCK permeability determinations (Taub et al., 2002). This approach combines the use of pH-gradient and the use of additives. As shown in this thesis, each parameter changed in the Caco-2 cell system had its own impact on the results. In several cases the effects were working in opposing direc-tions. A combination of all parameters can lead to unpredictable changes as ob-served for CEL in Paper IV and may obscure the interpretation of the data.

Page 64: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

64

5. SUMMARY AND CONCLUSIONS The results of the first part of this thesis (Papers I and II) show that the pH-dependent bidirectional transport of weakly basic and weakly acidic drugs may be biased by a false (passive) efflux and influx component, respectively. The false efflux can often be eliminated when the fraction of un-ionized drug species is equal in all compartments. In contrast, investigations on proton-dependent influx should be undertaken in both the presence and absence of a pH gradient. Concentration dependency and inhibition studies can be used to distinguish between active and passive transport processes, while temperature dependency failed to differentiate between these processes. The pKa of a compound can be used to choose the pH-gradient system for concentration-dependency and inhibition studies and to interpret the obtained permeability data. For predictive models of fa for the jejunum, ileum and colon transport a pH gradient of 6.0/7.4, 7.4/7.4 and 7.4/7.4 (apical pH/ baso-lateral pH) should be applied, respectively. Furthermore, the possibility of pH de-pendence should be considered in studies of DDIs involving efflux transporters such as P-gp, even when the functionalities of the transporters do not depend on the ex-tracellular pH.

The results of the second part of this thesis (Papers III and IV) show that BSA alters both passive and active drug transport in the Caco-2 cell model. Using the unbound concentration in the calculations of the transport rate allowed elucidation of the mechanisms behind these alterations. Firstly, the inclusion of basolateral BSA will improve the mass balance and in combination with Cu allows the permeability of the system to the compound to be predicted (for 100% recovery of the compound) and, thus, prediction of fa values for poorly recovered compounds in the traditional Caco-2 cell model (without BSA). Secondly, BSA does not affect the functionality of active transport systems such as P-gp. However, the amount of compound available at the binding sites of the efflux transporters may be reduced in the presence of BSA, which, in turn, will alter the transport rates of actively transported compounds. This observation may also be used to distinguish between actively and passively transported compounds, after re-calculating with Cu. Finally, the data suggest that active secretory transport for highly protein-bound compounds might have less ef-fect in vivo than predicted from traditional Caco-2 cell models. BSA in the Caco-2 cell model can be used as a secondary screening tool. Ranking of a limited subset of compounds based on efflux (or uptake) ratios obtained by inclusion of basolateral BSA might be used as a model in the CD (Candidate Drug) selection process. Thus, the inclusion of BSA in Caco-2 experiments results in a physiologically sound im-provement of the model and can be recommended; however, the inclusion of CEL affects both passive and active drug transport in the Caco-2 cell model and, since the results are unpredictable, CEL cannot be recommended as an additive in permeabil-ity assays, as long as a determination and accounting of the free drug concentration is not possible.

Page 65: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

65

Table VII is a summary of the refined experimental systems derived from Papers I-IV.

Table VII: Refined models discussed in Papers I- IV ____________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

Paper Model for… Refinement (+)/(-) Advantages/Disadvantages ____________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

I apical efflux a) pH 7.4 (+) study of active efflux, e.g. P-gp b) pH 7.4 (+) mimics physiological conditions

in the ileum and colon (-) proton-gradient dependent

transport is inactive ____________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

II apical uptake a) pH 6.0 (+) study of active uptake, e.g., (basolateral efflux) b) pH 7.4 MCT1, OATP-B, PepT1 (+) mimics physiological conditions

in the jejunum (-) false efflux and uptake compo-

nents____________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

III low recovery a) pH 7.4 (+) increased recovery +IV compounds b) pH 7.4, 4% BSA (+) mimics physiological conditions (-) free drug concentration has to

be measured compound-specific ____________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

IV sparingly soluble a) pH 7.4; CEL (-) compound-specific micellar compounds b) pH 7.4, 4% BSA incorporation ____________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

a) = apical; b) = basolateral

Page 66: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

66

6. REMAINING ISSUES The fraction absorbed in vivo is the sum of the permeation of the compound through the cell membranes along the GI tract. This permeation is affected by the changes in the pH gradients in vivo. The effect of these various pH changes on predictions of fa is an issue that remains to be resolved.

The relative contribution of active and passive transport processes in the various pH-gradient systems of the GI tract is variable and complex. A remaining issue is to understand these alterations further, and additional mechanistic studies of the effect of pH on the function of transporters are needed, e.g. of the interplay be-tween NHEs and H+-cotransporters.

Efflux ratios are concentration dependent for actively transported compounds and, therefore, the efflux ratios should be determined at relevant drug concentra-tions. Approaches of estimating relevant in vivo concentrations for an in vitro as-say are warranted.

Another issue is the standardization of methods of accounting for reduced com-pound recovery. In order to do this, the reasons for poor recovery must first be identified. For instance, a compound may be trapped in the cell cytosol because of changes in the pH (in this compartment) and this may be interpreted as a loss of recovery. It is necessary to account for this loss in order to interpret the trans-port data and evaluate transport mechanisms.

It remains to be shown, on a subset of compounds, covering a wide range of protein binding and ease of permeation, that basolaterally applied BSA will im-prove the prediction for fa.

Page 67: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

67

7. APPENDIX

Figure 19: Schematic illustration of some intestinal epithelial transporters verified to be expressed in the Caco-2 cell line. (A) Transporters shown as a square demon-strate active or facilitated transport. Ion-exchangers are shown as oval shapes. The name of the corresponding transporter for the process is shown within the square or oval shape. For active transporters, the black triangles represent an efflux trans-porter, the white triangles an uptake transporter and the grey triangles represent symport or antiport of the substrate and the driving force. Grey marked squares represent an unverified location for the transporter. (B) The corresponding gene names of the transporters is shown in the right illustration. SLC: solute carrier, ABC: adenosine triphosphate binding cassette. The illustration is based on informa-tion collected from references listed in Table VIII.

A B

Blood

mucosalmembrane

serosalmembrane

Lumen Intestinalepithelial cell

BloodIntestinalepithelial cell

mucosalmembrane

serosalmembrane

MDR1

MRP2

SGLT1

OCT1

OCT3

MRP1

ATP-aseK+Na+

Na+

HCO3-

BCRP

MDR3

GLUT5

MRP3

GLUT1

Lumen

PepT1

OCTN2

OATP-B

MCT1

NBCn1

NHE1 H+Na+

NHE 2H+ Na+

NHE 3H+ Na+

NHE 1H+ Na+

H+

Na+

H+

Na+

H+

HCO3-

PAT1 H+

ABCB1

ABCC2

SLC5A1

SLC22A1

SLC22A3

ABCC1

ATP-aseK+Na+

Na+

HCO3-

ABCG2

ABCB3

SLC2A5

ABCC3

SLC2A1

SLC15A1

SLC22A5

SLC02B1

SLC16A1

SLC4A7

SLC9A1 H+Na+

SLC9A2H+ Na+

SLC9A3H+ Na+

SLC9A1H+ Na+

H+

Na+

H+

Na+

H+

HCO3-

SLC36A1 H+

GLUT3 SLC2A3

Blood

mucosalmembrane

serosalmembrane

Lumen Intestinalepithelial cell

BloodIntestinalepithelial cell

mucosalmembrane

serosalmembrane

MDR1

MRP2

SGLT1

OCT1

OCT3

MRP1

ATP-aseK+Na+

Na+

HCO3-

BCRP

MDR3

GLUT5

MRP3

GLUT1

Lumen

PepT1

OCTN2

OATP-B

MCT1

NBCn1

NHE1 H+Na+

NHE 2H+ Na+

NHE 3H+ Na+

NHE 1H+ Na+

H+H+

Na+

H+H+

Na+Na+

H+H+

HCO3-

PAT1 H+H+

ABCB1

ABCC2

SLC5A1

SLC22A1

SLC22A3

ABCC1

ATP-aseK+Na+

Na+

HCO3-

ABCG2

ABCB3

SLC2A5

ABCC3

SLC2A1

SLC15A1

SLC22A5

SLC02B1

SLC16A1

SLC4A7

SLC9A1 H+Na+

SLC9A2H+ Na+

SLC9A3H+ Na+

SLC9A1H+ Na+

H+H+

Na+

H+H+

Na+Na+

H+H+

HCO3-

SLC36A1 H+H+

GLUT3 SLC2A3

Page 68: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

68

Table VIII: Some transporters that have been identified in Caco-2 cells and that are relevant for drug transport ___________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

Transporter (Gene) Protein or Gene identified (References) ___________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

Efflux transporter: MDR1 (ABCB1) by immunoblotting with P-gp C219 antibody (Wils et al., 1994); by RT-PCR (Taipalensuu et al., 2001). MDR3 (ABCB3) by RT-PCR (Taipalensuu et al., 2001). MRP1 (ABCC1) by RT-PCR (Taipalensuu et al., 2001). MRP2 (ABCC2) by RT-PCR (Taipalensuu et al., 2001). MRP3 (ABCC3) by RT-PCR (Taipalensuu et al., 2001). MRP4 (ABCC4) by RT-PCR (Taipalensuu et al., 2001). MRP5 (ABCC5) by RT-PCR (Taipalensuu et al., 2001). MRP6 (ABCC6) by RT-PCR (Taipalensuu et al., 2001). BCRP (ABCG2) by RT-PCR (Taipalensuu et al., 2001). ______________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________--------__________________________________________________________________________________________________________________

Uptake transporter: OCT1 (SLC22A1) by RT-PCR (Hayer-Zillgen et al., 2002; Martel et al., 2001; Zhang et al., 1999). OCT2 (SLC22A2) by RT-PCR (Hayer-Zillgen et al., 2002). OCT3 (SLC22A3) by RT-PCR (Hayer-Zillgen et al., 2002). OCTN2 (SLC22A5) by Western blot analysis and by immunofluorescence (Elimrani et al., 2003); by RT-PCR (Seithel et al., 2004). OATP-B (SLC02B1) by RT-PCR (Seithel et al., 2004). PepT1 (SLC15A1) by RT-PCR and Nothern blot analysis (PepT2 / SLC15A2 was negative)(Ganapathy et al., 1995); by RT-PCR and

Western blot analysis (Sun et al., 2002). HPT1 by monoclonal antibody inhibition, identification and subsequent cloning of the gene (Dantzig et al., 1994). MCT1 (SLC16A1) by RT-PCR (Seithel et al., 2004). MCT1 protein and RNA expression profile (Lambert et al., 2002). PAT1 (SLC36A1) by immunofluorescence (Chen et al., 2003).

Supporting list of references for Figure 19. Name within parentheses or single name corresponds to the Homo sapiens Official Gene Symbol, according to the HUGO Gene Nomenclature Committee (http://www.gene.ucl.ac.uk/nomenclature/).

Page 69: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

69

Table VIII (continues): Transporters that have been identified in Caco-2 cells.___________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

Glucose transporters and ion-exchangers: GLUT1 (SLC2A1) by immunoelectron microscopy and by immunofluorescence (Harris et al., 1992); by Nothern blot analysis (Mahraoui et

al., 1994). GLUT2 (SLC2A2) by Nothern blot analysis (Mahraoui et al., 1994). GLUT3 (SLC2A3) by immunoelectron microscopy and by immunofluorescence (Harris et al., 1992); by Nothern blot analysis (Mahraoui et

al., 1994). GLUT5 (SLC2A5) by immunoelectron microscopy and by immunofluorescence (Harris et al., 1992); by Nothern blot analysis (Mahraoui et

al., 1994). SGLT1 (SLC5A1) by Nothern blot analysis (Mahraoui et al., 1994). NHE1 (SLC9A1) by RT-PCR (Thwaites et al., 1999); by Western blot analysis and immunofluorescence (Janecki et al., 1999).NHE2 (SLC9A2) by RT-PCR (Thwaites et al., 1999). NHE3 (SLC9A3) by RT-PCR on poly(A)+RNA using primers specific for NHERF1 and NHERF2 (Thwaites et al., 2002); by Western

blot and immunofluorescence (Janecki et al., 1999). ____________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________________

Other important proteins identified in Caco-2 cells (a), or possibly available (b): (a) CYP1A1 (Boulenc et al., 1992); CYP3A (Carriere et al., 1994; Engman et al., 2001); villin (Chantret et al., 1988);(b) OCTL1 (SLC22A13) (Nishiwaki et al., 1998), OCTL2 (SLC22A14) (Nishiwaki et al., 1998), OST / (Dawson et al., 2005), OATP-E (Tamai et al., 2000); OATP-D (Tamai et al., 2000)

Page 70: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

70

8. ACKNOWLEDGEMENTS The studies in this thesis were performed at DMPK and Bioanalytical Chemistry, AstraZeneca R&D Mölndal and at the Department of Pharmacy, Faculty of Phar-macy, Uppsala University. I am grateful to the generous support and kind help pro-vided to me by people at these research centres.

I would especially like to express my sincere gratitude to:

Professor Per Artursson, my supervisor, for kindly sharing your research skills, for always aiming at high quality and for boldly trying to improve my writing skills.

Associate Professor Anna-Lena Ungell, my supervisor, for having faith in my ability, for sharing your knowledge and ideas, for stimulating discussions… (Minimalistic: For being who you are!)

Prof. Göran Alderborn for providing good working facilities and enjoying the Aul-ton with me.

ISMA (Dr. Ismael Zamora). I need more than this page to sum up the reasons.

Carola Wassvik, for keeping up with my papers and me. You are a wonderful person to share office with and you will handle even the TITANIC. ;o)

Alex Avdeef, Lucia Lazorová, Staffan Tavelin, Johan Gråsjö, co-authors on the PAMPA-paper, for being such inspiring and happy people. I enjoyed working with you!

Eva Ragnarsson, for helping to format the manuscripts in a stile acceptable in size for this small thesis!

All other colleges and friends, past and present, at the department of Pharmacy, for keeping up the good atmosphere.

Past and present members of the Abs. group in Mölndal. I really appreciate all your help and support; especially the feeding of my Caco’s, when I have been away.

Kurt-Jürgen Hoffmann, Göran Nilsson and Roger Simonsson, for supplying and cleaning up all the cold and hot material; thus, I was able to do so many (nice) ex-periments.

‘Calle’ Regårdh and Prof. Peter Langguth for allowing me to come to AstraHässle and Ulf Bredberg and Anders Tunek for allowing me to stay at AstraZeneca during the last years.

My friends and family, for encouragement and support at the right time. I know I have neglected you lately but I seriously hope that there will be an end to that now.

All the people who were willing to water my plants, when I was away! :o) /Sibylle *

The financial support of AstraZeneca R&D Mölndal is gratefully acknowledged. This work was also supported by the IF foundation for Pharmaceutical Research, ULLA, the Swedish

Animal Welfare Agency and the Swedish Medical Research Council (grant 9478).

Page 71: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

71

9. REFERENCES

AAPS-Workshop, 2003, Drug transport: From the bench to the bedside, in: Abstract book Peachtree City, GA, February 10-12).

Adson, A., P.S. Burton, T.J. Raub, C.L. Barsuhn, K.L. Audus and N.F.H. Ho, 1995, Passive diffusion of weak organic electrolytes across Caco-2 cell monolayers: Uncoupling the contributions of hydrodynamic, transcellular, and paracellular barriers, J Pharm Sci 84, 1197.

Adson, A., T.J. Raub, P.S. Burton, C.L. Barsuhn, A.R. Hilgers, K.L. Audus and N.F.H. Ho, 1994, Quantitative approaches to delineate paracellular diffusion in cultured epithelial cell monolayers, J Pharm Sci 83, 1529.

Altenberg, G.A., G. Young, J.K. Horton, D. Glass, J.A. Belli and L. Reuss, 1993, Changes in intra- or extracellular pH do not mediate P-glycoprotein-dependent multidrug resistance, Proc Nati Acad Sci U S A 90, 9735.

Amidon, G.L., H. Lennernäs, V.P. Shah and J.R. Crison, 1995, A Theroretical Basis for a Biopharmaceutic Drug Classification: The Correlation of in Vitro Drug Product Dissolution and in Vivo Bioavailability, Pharm Res 12, 413.

Anderberg, E.K. and P. Artursson, 1993, Epithelial transport of drugs in cell culture. VIII: Effects of sodium dodecyl sulfate on cell membrane and tight junction per-meability in human intestinal epithelial (Caco-2) cells, J Pharm Sci 82, 392.

Anderberg, E.K., C. Nystrom and P. Artursson, 1992, Epithelial transport of drugs in cell culture. VII: Effects of pharmaceutical surfactant excipients and bile acids on transepithelial permeability in monolayers of human intestinal epithelial (Caco-2) cells, J Pharm Sci 81, 879.

Anderle, P., E. Niederer, W. Rubas, C. Hilgendorf, H. Spahn-Langguth, H. Wunderli-Allenspach, H.P. Merkle and P. Langguth, 1998, P-Glycoprotein (P-gp) mediated efflux in Caco-2 cell monolayers: the influence of culturing condi-tions and drug exposure on P-gp expression levels, J Pharm Sci 87, 757.

Artursson, P., 1990, Epithelial transport of drugs in cell culture. I: A model for studying the passive diffusion of drugs over intestinal absorptive (Caco-2) cells, J Pharm Sci 79, 476.

Artursson, P., 1991, Cell cultures as models for drug absorption across the intestinal mucosa, Critical Reviews in Therapeutic Drug Carrier Systems 8, 305.

Artursson, P. and J. Karlsson, 1991, Correlation between oral drug absorption in humans and apparent drug permeability coefficients in human intestinal epithelial (Caco-2) cells, Biochem Biophys Res Commun 175, 880.

Page 72: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

72

Artursson, P. and C. Magnusson, 1990, Epithelial transport of drugs in cell culture. II: Effect of extracellular calcium concentration on the paracellular transport of drugs of different lipophilicities across monolayers of intestinal epithelial (Caco-2) cells, J Pharm Sci 79, 595.

Artursson, P., K. Palm and K. Luthman, 1996, Caco-2 monolayers in experimental and theoretical predictions of drug transport., Adv Drug Deliv Rev 22, 67.

Artursson, P., A.L. Ungell and J.E. Lofroth, 1993, Selective paracellular permeabil-ity in two models of intestinal absorption: cultured monolayers of human intesti-nal epithelial cells and rat intestinal segments, Pharm Res 10, 1123.

Aungst, B.J., 2000, Intestinal Permeation Enhancers, J Pharm Sci 89, 429. Aungst, B.J., N.H. Nguyen, J.P. Bulgarelli and K. Oates-Lenz, 2000, The influence

of donor and reservoir additives on Caco-2 permeability and secretory transport of HIV protease inhibitors and other lipophilic compounds, Pharm Res 17, 1175.

Avdeef, A., P. Artursson, S. Neuhoff, L. Lazorova, J. Grasjo and S. Tavelin, 2005, Caco-2 permeability of weakly basic drugs predicted with the Double-Sink PAMPA pKa

flux method, Eur J Pharm Sci 24, 333. Balimane, P.V. and S. Chong, 2005, Cell culture-based models for intestinal perme-

ability: a critique, DDT 10, 335. BASF and AG, July 1997, Cremophor® EL, in: technical leaflet ME 074 e

(888)Ludwigshafen, Germany) p. 7. Bender, H.M. and J. Dasenbrock, 1998, Brain concentration of asimadoline in mice:

the influence of coadministration of various P-glycoprotein substrates, Int J Clin Pharmacol Ther 36, 76.

Benet, L.Z., C.L. Cummins and C.Y. Wu, 2003, Transporter-enzyme interactions: implications for predicting drug-drug interactions from in vitro data, Curr Drug Metab 4, 393.

Benet, L.Z., S. Oie and J.B. Schwartz, 1996, Goodman & Gilman's: The Pharmacol-ogical Basis of Therapeutics. (McGraw-Hill, New York) p. 1707.

Bertz, R.J. and G.R. Granneman, 1997, Use of in vitro and in vivo data to estimate the likelihood of metabolic pharmacokinetic interactions, Clin Pharmacokinet 32, 210.

Bissonnette, P., H. Gagne, M.J. Coady, K. Benabdallah, J.Y. Lapointe and A. Berte-loot, 1996, Kinetic separation and characterization of three sugar transport modes in Caco-2 cells, Am J Physiol 270, G833.

Bogman, K., Y. Zysset, L. Degen, G. Hopfgartner, H. Gutmann, J. Alsenz and J. Drewe, 2005, P-glycoprotein and surfactants: effect on intestinal talinolol absorp-tion, Clin Pharmacol Ther 77, 24.

Boulenc, X., M. Bourrie, I. Fabre, C. Roque, H. Joyeux, Y. Berger and G. Fabre, 1992, Regulation of cytochrome P450IA1 gene expression in a human intestinal cell line, Caco-2, J Pharmacol Exp Ther 263, 1471.

Bravo, S.A., C.U. Nielsen, J. Amstrup, S. Frokjaer and B. Brodin, 2004, In-depth evaluation of Gly-Sar transport parameters as a function of culture time in the Caco-2 cell model, Eur J Pharm Sci 21, 77.

Bree, F. and J.P. Tillement, 1986, Propranolol binding to human serum proteins studied by high-performance liquid chromatography, J Chromatography 375, 416.

Carlstedt, I., J.K. Sheehan, A.P. Corfield and J.T. Gallagher, 1985, Mucous glyco-proteins: a gel of a problem, Essays Biochem 20, 40.

Page 73: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

73

Carriere, V., T. Lesuffleur, A. Barbat, M. Rousset, E. Dussaulx, P. Costet, I. de Waziers, P. Beaune and A. Zweibaum, 1994, Expression of cytochrome P-450 3A in HT29-MTX cells and Caco-2 clone TC7.[erratum appears in FEBS Lett 1995 Mar 27;362(1):99], FEBS Letters 355, 247.

Cavet, M.E., S. Akhter, R. Murtazina, F. Sanchez de Medina, C.M. Tse and M. Donowitz, 2001, Half-lives of plasma membrane Na(+)/H(+) exchangers NHE1-3: plasma membrane NHE2 has a rapid rate of degradation, Am J Physiol Cell Physiol 281, C2039.

Chang, E.B. and M.C. Rao, 1994, Intestinal water and electrolyte transport. Mecha-nism of physiological and adaptive responses, in: Physiology of the gastrointesti-nal tract, Vol. 62, ed. R. Johnson (Raven Press, New York) p. 2027.

Chantret, I., A. Barbat, E. Dussaulx, M.G. Brattain and A. Zweibaum, 1988, Epithe-lial polarity, villin expression, and enterocytic differentiation of cultured human colon carcinoma cells: a survey of twenty cell lines, Cancer Res 48, 1936.

Chen, Z., Y.J. Fei, C.M. Anderson, K.A. Wake, S. Miyauchi, W. Huang, D.T. Thwaites and V. Ganapathy, 2003, Structure, function and immunolocalization of a proton-coupled amino acid transporter (hPAT1) in the human intestinal cell line Caco-2, J Physiol 546, 349.

Colmenarejo, G., 2003, In silico prediction of drug-binding strengths to human se-rum albumin, Med Res Rev 23, 275.

Craig, P.N., 1990, Drug compendium., in: Cumulative subject index & drug com-pendium, Vol. 6, ed. C.J. Drayton (Pergamon press, Oxford) p. 237.

Cuff, M.A., D.W. Lambert and S.P. Shirazi-Beechey, 2002, Substrate-induced regu-lation of the human colonic monocarboxylate transporter, MCT1, J Physiol 539, 361.

Daniel, H., B. Neugebauer, A. Kratz and G. Rehner, 1985, Localization of acid mi-croclimate along intestinal villi of rat jejunum, Am J Physiol 248, G293.

Dantzig, A.H. and L. Bergin, 1990, Uptake of the cephalosporin, cephalexin, by a dipeptide transport carrier in the human intestinal cell line, Caco-2, Biochim Bio-phys Acta 1027, 211.

Dantzig, A.H., J.A. Hoskins, L.B. Tabas, S. Bright, R.L. Shepard, I.L. Jenkins, D.C. Duckworth, J.R. Sportsman, D. Mackensen and P.R. Rosteck, Jr., 1994, Associa-tion of intestinal peptide transport with a protein related to the cadherin super-family, Science 264, 430.

Davis, S.S., J.G. Hardy and J.W. Fara, 1986, Transit of pharmaceutical dosage forms through the small intestine, Gut 27, 886.

Dawson, P.A., M.L. Hubbert, J.H. Haywood, A.L. Craddock, N. Zerangue, W.V. Christian and N. Ballatori, 2005, The heteromeric organic solute transporter a-b, Osta-Ostb, is a ileal basolateral bile acid transporter, J Biol Chem 280, 6960.

Deferme, S. and P. Augustijns, 2003, The effect of food components on the absorp-tion of P-gp substrates: a review, J Pharm Pharmacol 55, 153.

Dey, S., P. Hafkemeyer, I. Pastan and M.M. Gottesman, 1999, A single amino acid residue contributes to distinct mechanisms of inhibition of the human multidrug transporter by stereoisomers of the dopamine receptor antagonist flupentixol, Biochem 38, 6630.

Dressman, J.B., R.R. Berardi, L.C. Dermentzoglou, T.L. Russell, S.P. Schmaltz, J.L. Barnett and K.M. Jarvenpaa, 1990, Upper gastrointestinal (GI) pH in young, healthy men and women, Pharm Res 7, 756.

Page 74: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

74

Dudeja, P.K., K.M. Anderson and J.S. Harris, 1995, Reversal of Multidrug Resis-tance Phenotype by Surfactants: Relationship to Membrane Lipid Fluidity, Arch Biochem Biophys 319, 309.

Edgar, B., C. Regardh, G. Johnsson, L. Johansson, P. Jundborg, I. Löfberg and O. Rönn, 1985, Felodipine kinetics in healthy men, Clin Pharmacol Ther 38, 205.

Elimrani, I., K. Lahjouji, E. Seidman, M.J. Roy, G.A. Mitchell and I. Qureshi, 2003, Expression and localization of organic cation/carnitine transporter OCTN2 in Caco-2 cells, Am J Physiol Gastrointest Liver Physiol 284, G863.

Emi, Y., D. Tsunashima, K.-I. Ogawara, K. Higaki and T. Kimura, 1998, Role of P-glycopotein as a secretory mechanism in quinidine absorption from rat small in-testine., J Pharm Sci 87, 295.

Englund, G., P. Hallberg, P. Artursson, K. Michaelsson and H. Melhus, 2004, Asso-ciation between the number of coadministered P-glycoprotein inhibitors and se-rum digoxin levels in patients on therapeutic drug monitoring, BMC Med 2:8.

Engman, H., C. Tannergren, P. Artursson and H. Lennernas, 2003, Enantioselective transport and CYP3A4-mediated metabolism of R/S-verapamil in Caco-2 cell monolayers, Eur J Pharm Sci 19, 57.

Engman, H.A., H. Lennernas, J. Taipalensuu, C. Otter, B. Leidvik and P. Artursson, 2001, CYP3A4, CYP3A5, and MDR1 in human small and large intestinal cell lines suitable for drug transport studies, J Pharm Sci 90, 1736.

Evans, D.F., G. Pye, R. Bramley, A.G. Clark, T.J. Dyson and J.D. Hardcastle, 1988, Measurement of gastrointestinal pH profiles in normal ambulant human subjects, Gut 29, 1035.

Fallingborg, J., L.A. Christensen, M. Ingelman-Nielsen, B.A. Jacobsen, K. Abildgaard and H.H. Rasmussen, 1989, pH-Profile and regional transit times of the normal gut measured by radiotelemetry device, Aliment Pharmacol Ther 3, 605.

Feliciano, N.R., A.A. Bouvet, E. Redalieu, J. Castellana, R.C. Luders, D.J. Schwartz, L. Shum, S. Zak, J.D. Arnold and P.T. Leese, 1990, Pharmacokinetic and pharmacodynamic comparison of an osmotic release oral metoprolol tablet and the metoprolol conventional tablet, Am Heart J 120, 483.

Fisher, J.M., S.A. Wrighton, J.C. Calamia, D.D. Shen, K.L. Kunze and K.E. Thum-mel, 1999, Midazolam metabolism by modified Caco-2 monolayers: effects of extracellular protein binding, J Pharmacol Exp Ther 289, 1143.

Flanagan, S.D., L.H. Takahashi, X. Liu and L.Z. Benet, 2002, Contributions of saturable active secretion, passive transcellular, and paracellular diffusion to the overall transport of furosemide across adenocarcinoma (Caco-2) cells, J Pharm Sci 91, 1169.

Florence, A.T., 1981, Surfactant interactions with biomembranes and drug absorp-tion, Pure Appl Chem 53, 2057.

Fogh, J., J.M. Fogh and T. Orfeo, 1977, One hundred and twenty-seven cultured human tumor cell lines producing tumors in nude mice, J Nati Cancer Inst 59, 221.

Food and Drug Administration, U.S. Department of Health and Human Services, Center for Drug Evaluation and Research, 2000, Guidance for industry: Waiver of in vivo bioavailability and bioequivalence studies for immediate-release solid oral dosage forms based on a biopharmaceutics classification system, http://www.fda.gov/cder/guidance/3618fnl.pdf.

Page 75: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

75

Fromm, M. and K. Hierholzer, 2000, Epithelien, in: Physiology des Menschen, Vol. 34, eds. R.F. Schmidt, G. Thews and F. Lang (Springer, Berlin, Heidelberg, New York) p. 719.

Galia, E., E. Nicolaides, D. Horter, R. Lobenberg, C. Reppas and J.B. Dressman, 1998, Evaluation of various dissolution media for predicting in vivo performance of class I and II drugs, Pharm Res 15, 698.

Gan, L.-S.L., S. Yanni and D.R. Thakker, 1998, Modulation of the Tight Junction of the Caco-2 Cell Monolayers by H2-antagonists, Pharm Res 15, 53.

Ganapathy, M.E., M. Brandsch, P.D. Prasad, V. Ganapathy and F.H. Leibach, 1995, Differential recognition of beta -lactam antibiotics by intestinal and renal peptide transporters, PEPT 1 and PEPT 2, J Biol Chem 270, 25672.

Goda, K., L. Balkay, T. Marian, L. Tron, A. Aszalos and G. Szabo, Jr., 1996, Intra-cellular pH does not affect drug extrusion by P-glycoprotein, J Photochem Photobiol. B - Biology 34, 177.

Goodwin, J.T., B. Mao, T.J. Vidmar, R.A. Conradi and P.S. Burton, 1999, Strategies toward predicting peptide cellular permeability from computed molecular de-scriptors, J Pept Res 53, 355.

Griffin, W.C., 1950, Classification of surface-active agents by 'HLB', J Soc Cosmet Chem 1, 311.

Grundemann, D., S. Harlfinger, S. Golz, A. Greets, A. Lazar, R. Berkels, N. Jung, A. Rubbert and E. Schomig, 2005, Discovery of the ergothioneine transporter, Proc Nati Am Soci U S A 102, 5256.

Gullberg, E., 2005, Particle transcytosis across the human intestinal epithelium. Model development and target identification for improved drug delivery, in: De-partment of Pharmacy (Uppsala University, Uppsala) p. 62.

Gulyaev, A.E., S.E. Gelperina, I.N. Skidan, A.S. Antropov, G.Y. Kivman and J. Kreuter, 1999, Significant transport of doxorubicin into the brain with polysor-bate 80-coated nanoparticles, Pharm Res 16, 1564.

Halestrap, A.P. and N.T. Price, 1999, The proton-linked monocarboxylate trans-porter (MCT) family: structure, function and regulation, Biochem J 343 Pt 2, 281.

Hansch, C., J.E. Quinlan and G.L. Lawrence, 1968, The linear free-energy relation-ship between partition coefficient and the aqueous solubility of organic liquids, J Org Chem 33, 347.

Harris, D.S., J.W. Slot, H.J. Geuze and D.E. James, 1992, Polarized distribution of glucose transporter isoforms in Caco-2 cells, Proc Nati Acad Sci U S A 89, 7556.

Hasselbalch, K.A., 1916, Die Berechnung der Wasserstoffzahl des Blutes aus der freien und gebundenen Kohlensäure desselben und die Sauerstoffbindungen des Blutes als Funktion der Wasserstoffzahl, Biochem Z 78, 112.

Hayer-Zillgen, M., M. Bruss and H. Bonisch, 2002, Expression and pharmacological profile of the human organic cation transporters hOCT1, hOCT2 and hOCT3, Br J Pharmacol 136, 829.

Hayton, W.L., 1980, Rate-limiting barriers to intestinal drug absorption: a review, J Pharmacokinet Biopharm 8, 321.

Herrera-Ruiz, D., Q. Wang, O.S. Gudmundsson, T.J. Cook, R.L. Smith, T.N. Faria and G.T. Knipp, 2001, Spatial expression patterns of peptide transporters in the human and rat gastrointestinal tracts, Caco-2 in vitro cell culture model, and mul-tiple human tissues, AAPS PharmSci 3, E9.

Page 76: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

76

Hidalgo, I.J. and R.T. Borchardt, 1990a, Transport of a large neutral amino acid (phenylalanine) in a human intestinal epithelial cell line: Caco-2, Biochim Bio-phys Acta 1028, 25.

Hidalgo, I.J. and R.T. Borchardt, 1990b, Transport of bile acids in a human intesti-nal epithelial cell line, Caco-2., Biochim Biophys Acta 1035, 97.

Hidalgo, I.J., T.J. Raub and R.T. Borchardt, 1989, Characterization of the Human Colon Carcinoma Cell Line (Caco-2) as a Model System for Intestinal Epithelial Permeability, Gastroenterology 96, 736.

Hilgers, A.R., R.A. Conradi and P.S. Burton, 1990, Caco-2 cell monolayers as a model for drug transport across the intestinal mucosa, Pharm Res 7, 902.

Hochman, J. and P. Artursson, 1994, Mechanisms of absorption enhancement and tight junction regulation, J Control Release 29, 253.

Hogben, C.A.M., D.J. Tocco, B.B. Brodie and L.S. Schanker, 1959, On the mechan-sim of intestinal absorption of drugs, J Pharmacol Exp Ther 125, 275.

Hogerle, M.L. and D. Winne, 1983, Drug absorption by the rat jejunum perfused in situ. Dissociation from the pH-partition theory and role of microclimate-pH and unstirred layer, Naunyn Schmiedebergs Arch Pharmacol 322, 249.

Horter, D. and J.B. Dressman, 2001, Influence of physicochemical properties on dissolution of drugs in the gastrointestinal tract, Adv Drug Deliv Rev 46, 75.

Hosoya, K.-i., K.-J. Kim and V.H.L. Lee, 1996, Age-dependent Expression of P-Glycoprotein gp170 in Caco-2 Cell Monolayers, Pharm Res 13, 885.

Hubatsch, I., L. Lazorova, A. Vahlne and P. Artursson, 2004, Orally active antiviral tripeptide glycyl-prolyl-glycinamide is activated by CD26 (Dipeptidy peptidase IV) before transport across the intestinal epithelium, Antimicrob Agents Chemo-ther 49, 1087.

Hubbert, M.L., A.L. Craddock, J.H. Haywood, N. Ballatori and P.A. Dawson, 2004, The heteromeric organic solute transporter (Ost) alpha and beta complex is an ileal basolateral bile acid transporter, Falk Symposium No. 141, Stockholm, Sweden, abstract No. 52.

Hunter, J., M.A. Jepson, T. Tsuruo, N.L. Simmons and B.H. Hirst, 1993, Functional Expression of P-glycoprotein in Apical Membranes of Human Intestinal Caco-2 Cells, J Biol Chem 268, 14991.

Ingels, F., S. Deferme, E. Destexhe, M. Oth, G. Van den Mooter and P. Augustijns, 2002, Simulated intestinal fluid as transport medium in the Caco-2 cell culture model, Int J Pharm 232, 183.

Ingels, F.M. and P.F. Augustijns, 2003, Biological, pharmaceutical, and analytical considerations with respect to the transport media used in the absorption screen-ing system, Caco-2, J Pharm Sci 92, 1545.

Irvine, J.D., L. Takahashi, K. Lockhart, J. Cheong, J.W. Tolan, H.E. Selick and J.R. Grove, 1999, MDCK (Madin-Darby canine kidney) cells: A tool for membrane permeability screening, J Pharm Sci 88, 28.

Jackson, M.J., 1987, Drug transport across gastrointestinal epithelia, in: Physiology of the Gastrointestinal tract, Vol. 59, ed. L.R. Johnson (Raven Press, New York) p. 1597.

Jackson, M.J., Y.F. Shiau, S. Bane and M. Fox, 1974, Intestinal transport of weak electrolytes. Evidence in favor of a three-compartment system, J Gen Physiol 63, 187.

Page 77: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

77

Janecki, A.J., M.H. Montrose, C.M. Tse, F.S. de Medina, A. Zweibaum and M. Donowitz, 1999, Development of an endogenous epithelial Na(+)/H(+) ex-changer (NHE3) in three clones of caco-2 cells, Am J Physiol 277, G292.

Johnstone, R.W., A.A. Ruefli, K.M. Tainton and M.J. Smyth, 2000, A role for P-glycoprotein in regulating cell death, Leuk Lymphoma 38, 1.

Jonker, J.W., E. Wagenaar, L. van Deemter, R. Gottschlich, H.M. Bender, J. Dasen-brock and A.H. Schinkel, 1999, Role of blood-brain barrier P-glycoprotein in limiting brain accumulation and sedative side-effects of asimadoline, a peripher-ally acting analgaesic drug, Br J Pharmacol 127, 43.

Jumarie, C. and C. Malo, 1991, Caco-2 cells cultured in serum-free medium as a model for the study of enterocytic differentiation in vitro, J Cell Physiol 149, 24.

Kamm, W., J. Hauptmann, I. Behrens, J. Stürzebecher, F. Dullweber, H. Gohlke, M. Stubbs, G. Klebe and T. Kissel, 2001, Transport of Peptidomimetic Thrombin In-hibitors with a 3-Amino-Phenylalanine Structure: Permeability and Efflux Mechanism in Monolayers of a Human Intestinal Cell Line (Caco-2), Pharm Res 18, 1110.

Karali, T.T., 1989, Gastrointestinal absorption of drugs, Crit Rev Ther Drug Carrier Systems 6, 39.

Karlsson, J. and P. Artursson, 1991, A method for the determination of cellular per-meabiliy coefficients and aqueous boundery layer thickness in monolayers of in-testinal epithelial (Caco-2) cells grown in permeable filter chambers, Int J Pharm 71, 55.

Karlsson, J., A.L. Ungell, J. Grasjo and P. Artursson, 1999, Paracellular drug trans-port across intestinal epithelia: Influence of charge and induced water flux, Eur J Pharm Sci 9, 47.

Keppler, D. and J. Konig, 1997, Expression and localization of the conjugate export pump encoded by the MRP2 (cMRP/cMOAT) gene in liver, FASEB Journal 11, 509.

Kerns, E.H., 2001, High throughput physicochemical profiling for drug discovery, J Pharm Sci 90, 1838.

Kobayashi, D., T. Nozawa, K. Imai, J. Nezu, A. Tsuji and I. Tamai, 2003, Involve-ment of human organic anion transporting polypeptide OATP-B (SLC21A9) in pH-dependent transport across intestinal apical membrane, J Pharmacol Exp Ther 306, 703.

Kosa, T., T. Maruyama and M. Otagiri, 1997, Species differences of serum albu-mins: I. Drug binding sites, Pharm Res 14, 1607.

Krishna, G., K.-j. Chen, C.-c. Lin and A.A. Nomeir, 2001, Permeability of lipophilic compounds in drug discovery using in-vitro human absorption model, Caco-2, Int J Pharm 222, 77.

Kulkarni, A., Y. Han and A.J. Hopfinger, 2002, Predicting Caco-2 cell permeation coefficients of organic molecules using membrane-interaction QSAR analysis, J Chem Inf Comput Sci 42, 331.

Kullak-Ublick, G.A., B. Stieger and P.J. Meier, 2004, Enterohepatic bile salt trans-porters in normal physiology and liver disease, Gastroenterology 126, 322.

Lambert, D.W., I.S. Wood, A. Ellis and S.P. Shirazi-Beechey, 2002, Molecular changes in the expression of human colonic nutrient transporters during the tran-sition from normality to malignancy, Br J Cancer 86, 1262.

Lande, M.B., N.A. Priver and M.L. Zeidel, 1994, Determinats of apical membrane permeabilities of barrier epithelia, Am J Physiol 267, C367.

Page 78: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

78

Landwojtowicz, E., P. Nervi and A. Seelig, 2002, Real-time monitoring of P-glycoprotein activation in living cells, Biochem 41, 8050.

Lee, K., C. Ng, K.L. Brouwer and D.R. Thakker, 2002, Secretory transport of raniti-dine and famotidine across Caco-2 cell monolayers, J Pharmacol Exp Ther 303, 574.

Lee, K. and D.R. Thakker, 1999, Saturable transport of H2-antagonists ranitidine and famotidine across Caco-2 cell monolayers, J Pharm Sci 88, 680.

Lennernäs, H., 1998, Human Intestinal Permeability, J Pharm Sci 87, 403. Lentz, K.A., J. Hayashi, L.J. Lucisano and J.E. Polli, 2000, Development of a more

rapid, reduced serum culture system for Caco-2 monolayers and application to the biopharmaceutics classification system, Int J Pharm 200, 41.

Li, A.P., 2001, Screening for human ADME/Tox drug properties in drug discovery., DDT 6, 357.

Liang, E., K. Chessic and M. Yazdanian, 2000, Evaluation of an accelerated Caco-2 cell permeability model, J Pharm Sci 89, 336.

Lin, J.H. and A.Y. Lu, 2001, Interindividual variability in inhibition and induction of cytochrome P450 enzymes, Annu Rev Pharmacol Toxicol 41, 535.

Lin, J.H. and M. Yamazaki, 2003, Role of P-glycoprotein in pharmacokinetics: clinical implications, Clin Pharmacokinet 42, 59.

Lipinski, C.A., 2003, Chris Lipinski discusses life and chemistry after the Rule of Five, DDT 8, 12.

Lipinski, C.A., F. Lombardo, B.W. Dominy and P.J. Feeney, 2001, Experimental and computational approaches to estimate solubility and permeability in drug dis-covery and development settings, Adv Drug Deliv Rev 46, 3.

Lu, S., A.W. Gough, W.F. Bobrowski and B.H. Stewart, 1996, Transport properties are not altered across Caco-2 cells with heightened TEER despite underlying physiological and ultrastructural changes, J Pharm Sci 85, 270.

Lucas, M., 1983, Determination of acid surface pH in vivo in rat proximal jejunum, Gut 24, 734.

Lucas, M.L., B.T. Cooper, F.H. Lei, I.T. Johnson, G.K. Holmes, J.A. Blair and W.T. Cooke, 1978, Acid microclimate in coeliac and Crohn's disease: a model for folate malabsorption, Gut 19, 735.

Lucas, M.L., W. Schneider, F.J. Haberich and J.A. Blair, 1975, Direct measurement by pH-microelectrode of the pH microclimate in rat proximal jejunum, Proc R Soc London B Biol Sci 192, 39.

Lucas, M.L. and R.R. Whitehead, 1994, A re-evaluation of the properties of the three-compartment model of intestinal weak-electrolyte absorption, J Theor Biol 167, 147.

Mackay, M., I. Williamson and J. Hastewell, 1991, (A) Cell biology of epithelia, Adv Drug Deliv Rev 7, 313.

Madara, J.L. and J.S. Trier, 1994, The functional morphology of the mucosa of the small intestine., in: Physiology of the gastrointestinal tract., ed. R. Johnson (Ra-ven Press, New York) p. 1577.

Mahraoui, L., A. Rodolosse, A. Barbat, E. Dussaulx, A. Zweibaum, M. Rousset and E. Brot-Laroche, 1994, Presence and differential expression of SGLT1, GLUT1, GLUT2, GLUT3 and GLUT5 hexose-transporter mRNAs in Caco-2 cell clones in relation to cell growth and glucose consumption, Biochem J 298 Pt 3, 629.

Page 79: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

79

Martel, F., D. Grundemann, C. Calhau and E. Schomig, 2001, Apical uptake of organic cations by human intestinal Caco-2 cells: putative involvement of ASF transporters, Naunyn Schmiedebergs Arch Pharmacol 363, 40.

McNeil, N.I., K.L. Ling and J. Wager, 1987, Mucosal surface pH of the large intes-tine of the rat and of normal and inflamed large intestine in man, Gut 28, 707.

McNeil, N.I. and K.L.E. Ling, 1984, Large intestinal mucosal surface pH in rat and man, in: Intestinal absorption and secretion, Vol. 8, eds. E. Skadhauge and E. Heintze (MTP Press, Lancaster) p. 103.

Mehvar, R., M.E. Gross and R.N. Kreamer, 1990, Pharmacokinetics of atenolol enantiomers in humans and rats, J Pharm Sci 79, 881.

Meylan, W.M., P.H. Howard and R.S. Boethling, 1996, Improved method for esti-mating water solubility from octanol/water partition coefficient, Environ Toxicol Chem 15, 100.

Nagahara, N., S. Tavelin and P. Artursson, 2004, The contribution of the paracellu-lar route to pH dependent permeability of ionizable drugs, J Pharm Sci 93, 2972.

Nellans, H.N., 1991, (B) Mechanisms of peptide and protein absorption. (1) Paracel-lular intestinal transport: modulation of absorption, Adv Drug Deliv Rev 7, 339.

Nerurkar, M.M., P.S. Burton and R.T. Borchardt, 1996, The use of surfactants to enhance the permeability of peptides through caco-2 cells by inhibition of an api-cally polarized efflux system, Pharm Res 13, 528.

Nerurkar, M.M., N.F. Ho, P.S. Burton, T.J. Vidmar and R.T. Borchardt, 1997, Mechanistic roles of neutral surfactants on concurrent polarized and passive membrane transport of a model peptide in Caco-2 cells, J Pharm Sci 86, 813.

Ng, K.Y. and R.T. Borchardt, 1993, Biotin transport in a human intestinal epithelial cell line (Caco-2), Life Sci 53, 1121.

Nishiwaki, T., Y. Daigo, M. Tamari, Y. Fujii and Y. Nakamura, 1998, Molecular cloning, mapping, and characterization of two novel human genes, ORCTL3 and ORCTL4, bearing homology to organic-cation transporters, Cytogenet Cell Genet 83, 251.

Norberg, A., A.W. Jones, R.G. Hahn and J.L. Gabrielsson, 2003, Role of variability in explaining ethanol pharmacokinetics: research and forensic applications, Clin Pharmacokinet 42, 1.

Pade, V. and S. Stavchansky, 1997, Estimation of the relative contribution of the transcellular and paracellular pathway to the transport of passively absorbed drugs in the Caco-2 cell culture model, Pharm Res 14, 1210.

Palm, K., 1998, Experimental and Theoretical Predictions of Intestinal Drug Ab-sorption, in: Department of Pharmacy (Uppsala Univeristy, Uppsala) p. 59.

Palm, K., K. Luthman, J. Ros, J. Grasjo and P. Artursson, 1999, Effect of molecular charge on intestinal epithelial drug transport: pH-dependent transport of cationic drugs, J Pharmacol Exp Ther 291, 435.

Palm, K., K. Luthman, A.L. Ungell, G. Strandlund and P. Artursson, 1996, Correla-tion of drug absorption with molecular surface properties, J Pharm Sci 85, 32.

Peters, T., 1996, All about albumin. Biochemistry, genetics, and medical applica-tions (Academic Press, Inc., San Diego) p. 432.

Petri, N., C. Tannergren, D. Rungstad and H. Lennernas, 2004, Transport character-istics of fexofenadine in the Caco-2 cell model, Pharm Res 21, 1398.

Pinto, M., S. Robine-Leon, M.-D. Appay, M. Kedinger, N. Triadou, E. Dussaulx, B. Lacroix, P. Simon-Assmann, K. Haffen, J. Fogh and A. Zweibaum, 1983, En-

Page 80: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

80

terocyte-like differentiation and polarization of the human colon carcinoma cell line Caco-2 in culture, Biol Cell 47, 323.

Powell, D.W., 1981, Barrier function of epithelia, Am J Physiol 241, G275. Quaroni, A. and J. Hochman, 1996, Development of intestinal cell culture models

for drug transport and metabolism studies, Adv Drug Deliv Rev 22, 3. Rakhit, A., N.H.G. Holford, T.W. Guentert, K. Maloney and S. Riegelman, 1984,

Pharmacokinetics of quinidine and three of its metabolites in Man, J Pharma-cokinet Biopharm 12, 1.

Rau, S.E., J.R. Bend, J.M.O. Arnold, L.T. Tran, J.D. Spence and D.G. Baily, 1997, Grapefruit juice - terfenadine single-dos interaction: Magnitude, mechanism, and relevance, Clin Pharmacol Ther 61, 401.

Raub, T.J., C.L. Barsuhn, L.R. Williams, D.E. Decker, G.A. Sawada and N.F. Ho, 1993, Use of a biophysical-kinetic model to understand the roles of protein bind-ing and membrane partitioning on passive diffusion of highly lipophilic mole-cules across cellular barriers, J Drug Target 1, 269.

Rege, B.D., J.P. Kao and J.E. Polli, 2002, Effects of nonionic surfactants on mem-brane transporters in Caco-2 cell monolayers, Eur J Pharm Sci 16, 237.

Rodrigues, A.D., 2002, Drug-drug interactions, in: Drugs and the pharmaceutical sciences, Vol. 116 (Marcel Dekker, Inc., New York) p. 645.

Rowland, M. and T.N. Tozer, 1989, Clinical pharmacokinetics. Concepts and appli-cations (Lea and Febiger, Philadelphia, London) p. 546.

Rubio-Aliaga, I. and H. Daniel, 2002, Mammalian peptide transporters as targets for drug delivery, Trends Pharmacol Sci 23, 434.

Saha, P. and J.H. Kou, 2000, Effect of solubilizing excipients on permeation of poorly water-soluble compounds across Caco-2 cell monolayers, Eur J Pharm Biopharm 50, 403.

Saha, P. and J.H. Kou, 2002, Effect of bovine serum albumin on drug permeability estimation across Caco-2 monolayers, Eur J Pharm Biopharm 54, 319.

Schmiedlin-Ren, P., K.E. Thummel, J.A. Fisher, M.F. Paine, K.S. Lown and P.B. Watkins, 1997, Expression of Enzymatically Active CYP3A4 by Caco-2 Cells Grown on Extracellular Matrix-Coated Permeable Supports in the Presence of 1 ,25-Dihydroxyvitamin D3, Mol Pharmacol 51, 741.

Scholz, A., B. Abrahamsson, S.M. Diebold, E. Kostewicz, B.I. Polentarutti, A.L. Ungell and J.B. Dressman, 2002, Influence of hydrodynamics and particle size on the absorption of felodipine in labradors, Pharm Res 19, 42.

Seithel, A., J. Karlsson, C. Hilgendorf, A. Björquist and A.L. Ungell, 2004, Vari-ability in gene expression of drug transporters by age and passage in Caco-2 cells, EUFEPS, Copenhagen.

Shen, D.D., A.A. Artru and K.K. Adkison, 2004, Principles and applicability of CSF sampling for the assessment of CNS drug delivery and pharmacodynamics, Adv Drug Deliv Rev 56, 1825.

Shiau, Y.F., P. Fernandez, M.J. Jackson and S. McMonagle, 1985, Mechanisms maintaining a low-pH microclimate in the intestine, Am J Physiol 248, G608.

Shore, P.A., B.B. Brodie and C.A.M. Hogben, 1957, The gastric secretion of drugs: A pH partition hypothesis, J Pharmacol Exp Ther 119, 361.

Stein, J., M. Zores and O. Schroder, 2000, Short-chain fatty acid (SCFA) uptake into Caco-2 cells by a pH-dependent and carrier mediated transport mechanism, Eur J Nutr 39, 121.

Page 81: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

81

Stephens, R.H., C.A. O'Neill, A. Warhurst, G.L. Carlson, M. Rowland and G. War-hurst, 2001, Kinetic profiling of P-glycoprotein-mediated drug efflux in rat and human intestinal epithelia, J Pharmacol Exp Ther 296, 584.

Sun, D., H. Lennernas, L.S. Welage, J.L. Barnett, C.P. Landowski, D. Foster, D. Fleisher, K.D. Lee and G.L. Amidon, 2002, Comparison of human duodenum and Caco-2 gene expression profiles for 12,000 gene sequences tags and correla-tion with permeability of 26 drugs, Pharm Res 19, 1400.

Suzuki, H. and Y. Sugiyama, 2000, Role of metabolic enzymes and efflux transport-ers in the absorption of drugs from the small intestine, Eur J Pharm Sci 12, 3.

Taipalensuu, J., H. Tornblom, G. Lindberg, C. Einarsson, F. Sjoqvist, H. Melhus, P. Garberg, B. Sjostrom, B. Lundgren and P. Artursson, 2001, Correlation of gene expression of ten drug efflux proteins of the ATP-binding cassette transporter family in normal human jejunum and in human intestinal epithelial Caco-2 cell monolayers, J Pharmacol Exp Ther 299, 164.

Takanaga, H., I. Tamai and A. Tsuji, 1994, pH-dependent and carrier-mediated transport of salicylic acid across Caco-2 cells, J Pharm Pharmacol 46, 567.

Tamai, I., J. Nezu, H. Uchino, Y. Sai, A. Oku, M. Shimane and A. Tsuji, 2000, Mo-lecular identification and characterization of novel members of the human or-ganic anion transporter (OATP) family, Biochem Biophys Res Commun 273, 251.

Tamai, I., Y. Sai, A. Ono, Y. Kido, H. Yabuuchi, H. Takanaga, E. Satoh, T. Ogihara, O. Amano, S. Izeki and A. Tsuji, 1999, Immunohistochemical and functional characterization of pH-dependent intestinal absorption of weak organic acids by the monocarboxylic acid transporter MCT1, J Pharm Pharmacol 51, 1113.

Tamai, I., H. Takanaga, H. Maeda, Y. Sai, T. Ogihara, H. Higashida and A. Tsuji, 1995, Participation of a proton-cotransporter, MCT1, in the intestinal transport of monocarboxylic acids, Biochem Biophys Res Commun 214, 482.

Taub, M.E., L. Kristensen and S. Frokjaer, 2002, Optimized conditions for MDCK permeability and turbidimetric solubility studies using compounds representative of BCS classes I-IV, Eur J Pharm Sci 15, 331.

Tavelin, S., 2003, New approaches to studies of paracellular drug transport in intes-tinal epithelial cell monolayers, in: Department of Pharmacy (Uppsala University, Uppsala) p. 66.

Taylor, D.C., R. Pownall and W. Burke, 1985, The absorption of beta-adrenoceptor antagonists in rat in-situ small intestine; the effect of lipophilicity, J Pharm Phar-macol 37, 280.

Thwaites, D.T., D. Ford, M. Glanville and N.L. Simmons, 1999, H(+)/solute-induced intracellular acidification leads to selective activation of apical Na(+)/H(+) exchange in human intestinal epithelial cells, J Clin Invest 104, 629.

Thwaites, D.T., D.J. Kennedy, D. Raldua, C.M. Anderson, M.E. Mendoza, C.L. Bladen and N.L. Simmons, 2002, H+/dipeptide absorption across the human in-testinal epithelium is controlled indirectly via a functional Na+/H+ exchanger, Gastroenterology 122, 1322.

Tsuji, A., E. Miyamoto, N. Hashimoto and T. Yamana, 1978, GI absorption of -lactam antibiotics II: Deviation from pH-partition hypothesis in penicillin absorp-tion through in situ and in vitro lipoidal barriers., J Pharm Sci 67, 1705.

Tsuji, A., H. Takanaga, I. Tamai and T. Terasaki, 1994, Transcellular transport of benzoic acid across Caco-2 cells by a pH-dependent and carrier-mediated trans-port mechanism, Pharm Res 11, 30.

Page 82: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

82

Ungell, A.L., 2004, Caco-2 replace or refine?, DDT 1, 423. Ungell, A.-L., 1997, In vitro absoption studies and their relevance to absorption

from the GI tract, Drug Dev Ind Pharm 23, 879. Ungell, A.L. and J. Karlsson, 2003, Cell cultures in drug discovery: An industrial

perspective, in: Drug bioavailability. Estimation of solubility, permeability, ab-sorption and bioavailability, eds. H. van De Waterbeemd, H. Lennernäs and P. Artursson (Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim) p. 90.

Wacher, J.V., J.A. Silverman, Y. Zhang and L.Z. Benet, 1998, Role of P-Glycoprotein and Cytochrome P450 3A in Limiting Oral Absorption of Peptides and Peptidometics, J Pharm Sci 87, 1322.

Walgren, R.A. and T. Walle, 1999, The influence of plasma binding on absorp-tion/exsorption in the Caco-2 model of human intestinal absorption, J Pharm Pharmacol 51, 1037.

van Hoogdalem, E.J., A.G. de Boer and D.D. Breimer, 1989, Intestinal drug absorp-tion enhancement: an overview, Pharmacol Ther 44, 407.

van Meer, G., 1989, Polarity and polarized transport of membrane lipids in a cul-tured epithelium, in: Functional epithelial cells in culture, eds. K.S. Matlin and J.D. Valentish (Allan R. Liss, New York) p. 43.

van Osdol, W. and T. Watanabe, 1999-07-01, Novel formulations for the transder-mal administration of asimadoline, Patent: WO9932096, application number WO 1998US26652 19981215.

Westphal, K., A. Weinbrenner, T. Giessmann, M. Stuhr, G. Franke, M. Zschiesche, R. Oertel, B. Terhaag, H.K. Kroemer and W. Siegmund, 2000, Oral bioavailabil-ity of digoxin is enhanced by talinolol: evidence for involvement of intestinal P-glycoprotein, Clin Pharmacol Ther 68, 6.

Vidanovic, D., J. Milic Askrabic, M. Stankovic and V. Poprzen, 2003, Effects of nonionic surfactants on the physical stability of immunoglobulin G in aqueous solution during mechanical agitation, Pharmazie 58, 399.

Wils, P., V. Phung-Ba, A. Warnery, D. Lechardeur, S. Raeissi, I.J. Hidalgo and D. Scherman, 1994, Polarized transport of docetaxel and vinblastine mediated by P-glycoprotein in human intestinal epithelial cell monolayers, Biochem Pharmacol 48, 1528.

Wilson, G., I.F. Hassan, C.J. Dix, I. Williamson, R. Shah, M. Mackay and P. Arturs-son, 1990, Transport and permeability properties of human caco-2 cells: an in vi-tro model of the intestinal epithelial cell barrier, J Controlled Release 11, 25.

Volpe, D.A., 2004, Permeability classification of representative fluoroquinolones by a cell culture method, AAPS PharmSci 6, Article 13.

Woodcock, D.M., M.E. Linsenmeyer, G. Chojnowski, A.B. Kriegler, V. Nink, L.K. Webster and W.H. Sawyer, 1992, Reversal of multidrug resistance by surfactants, Br J Cancer 66, 62.

Wu, C.-Y. and L.Z. Benet, 2005, Predicting drug disposition via application of BCS: Transport/absorption/elimination interplay and development of a biopharmaceu-tical drug disposition classification system, Pharm Res 22, 11.

Wu, C.-Y., L.Z. Benet, M.F. Hebert, S.K. Gupta, M. Rowland, D.Y. Gomez and V.J. Wacher, 1995, Differentiation of absorption and first-pass gut and hepatic me-tabolism in humans: Studies with cyclosporine, Clin Pharmacol Ther 58, 492.

Yabuuchi, H., I. Tamai, J. Nezu, K. Sakamoto, A. Oku, M. Shimane, Y. Sai and A. Tsuji, 1999, Novel membrane transporter OCTN1 mediates multispecific, bidi-

Page 83: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery

83

rectional, and pH-dependent transport of organic cations, J Pharmacol Exp Ther 289, 768.

Yalkowsky, S.H. and S.C. Valvani, 1980, Solubility and partitioning I: Solubility of nonelectrolytes in water, J Pharm Sci 69, 912.

Yamashita, S., T. Furubayashi, M. Kataoka, T. Sakane, H. Sezaki and H. Tokuda, 2000, Optimized conditions for prediction of intestinal drug permeability using Caco-2 cells, Eur J Pharm Sci 10, 195.

Yamashita, S., K. Konishi, Y. Yamazaki, Y. Taki, T. Sakane, H. Sezaki and Y. Fu-ruyama, 2002, New and better protocols for a short-term Caco-2 cell culture sys-tem, J Pharm Sci 91, 669.

Yazdanian, M., K. Briggs, C. Jankovsky and A. Hawi, 2004, The "high solubility" definition of the current FDA Guidance on Biopharmaceutical Classification Sys-tem may be too strict for acidic drugs, Pharm Res 21, 293.

Yee, S., 1997, In vitro permeability across Caco-2 cells (colonic) can predict in vivo (small intestine) absorption in man - Fact or myth, Pharm Res 14, 763.

Yu, L., A. Bridgers, J. Polli, A. Vickers, S. Long, A. Roy, R. Winnike and M. Cof-fin, 1999, Vitamin E-TPGS increases absorption flux of an HIV protease inhibi-tor by enhancing its solubility and permeability, Pharm Res 16, 1812.

Zhang, L., W. Gorset, M.J. Dresser and K.M. Giacomini, 1999, The interaction of n-tetraalkylammonium compounds with a human organic cation transporter, hOCT1, J Pharmacol Exp Ther 288, 1182.

Zhang, X., J.K. Jackson and H.M. Burt, 1996, Determination of surfactant critical micelle concentration by a novel fluorescence depolarization technique, J Bio-chem Biophys Methods 31, 145.

Zhou, S.Y., N. Piyapolrungroj, L. Pao, C. Li, G. Liu, E. Zimmermann and D. Flei-sher, 1999, Regulation of paracellular absorption of cimetidine and 5-aminosalicylate in rat intestine, Pharm Res 16, 1781.

Zimmermann, C., H. Gutmann, P. Hruz, J.-P. Gutzwiller, C. Beglinger and J. Drewe, 2005, Mapping of multidrug resistance gene 1 and multidrug resistance-associated protein isoform 1 to 5 mRNA expression along the human intestinal tract, Drug Metab Dispos 33, 219.

Zschiesche, M., G.L. Lemma, K.J. Klebingat, G. Franke, B. Terhaag, A. Hoffmann, T. Gramatte, H.K. Kroemer and W. Siegmund, 2002, Stereoselective disposition of talinolol in man, J Pharm Sci 91, 303.

Zweibaum, A., H.P. Hauri, E. Sterchi, I. Chantret, K. Haffen, J. Bamat and B. Sor-dat, 1984, Immunohistological evidence, obtained with monoclonal antibodies, of small intestinal brush border hydrolases in human colon cancers and foetal co-lons, Int J Cancer 34, 591.

Ca

co

Ca

co

Page 84: in vitro s for Druguu.diva-portal.org/smash/get/diva2:166491/FULLTEXT01.pdf · bioavailability Oral administration is currently considered the most convenient and safest drug delivery