Sunteți pe pagina 1din 18

PersPecTives

estimation of the accuracy of important


OpiniOn
pharmacokinetic parameters, such as liver
and renal clearance, can be improved.
Coexistence of passive and In the following sections, we will first
discuss the basics of membrane permeability
carrier-mediated processes and the experimental methods applied in
the measurement of passive transport and
in drug transport carrier­mediated transport. We will then give
an overview on the role of joint contribution
of passive transport and carrier­mediated
Kiyohiko Sugano, Manfred Kansy, Per Artursson, Alex Avdeef, Stefanie Bendels, transport to the total permeation of a drug,
Li Di, Gerhard F. Ecker, Bernard Faller, Holger Fischer, Grégori Gerebtzoff, focusing on in vitro and in vivo results.
Finally, some implications for drug discovery
Hans Lennernaes and Frank Senner
and development will be discussed.
Abstract | The permeability of biological membranes is one of the most important
determinants of the pharmacokinetic processes of a drug. Although it is often Basics of membrane permeation
Passive transcellular transport. Biological
accepted that many drug substances are transported across biological membranes
cellular membranes are fluidic patches com­
by passive transcellular diffusion, a recent hypothesis speculated that carrier- prising various bilayer­forming amphiphilic
mediated mechanisms might account for the majority of membrane drug transport phospholipids, cholesterol and membrane­
processes in biological systems. Based on evidence of the physicochemical anchored proteins — including transporters
characteristics and of in vitro and in vivo findings for marketed drugs, as well as — that form thin (~5 nm) hydrophobic
barriers separating aqueous environments.
results from real-life discovery and development projects, we present the view that
Consequently, lipid bilayer and carrier
both passive transcellular processes and carrier-mediated processes coexist and proteins coexist in a cellular membrane
contribute to drug transport activities across biological membranes. when the transporter is expressed in the
cells. In this article, the passive transcellular
The permeability of biological membranes permeation for many (but not all) drugs transport (for uptake, efflux, absorptive and
is a key determinant of drug pharmaco­ in many (but not all) cell types, at least in excretive directions (BOX 1)) is defined as a
kinetics. Historically, passive diffusion intestinal membrane permeation. Carrier­ concentration gradient­driven mass transport
through the lipid bilayer portion of a mediated transport is probably important of a compound (called the permeant, which
biological membrane was suggested as for drugs with low passive permeation, includes drug and drug­like molecules) from
the dominant route. However, in the past especially for brain and liver distribution, one side of the cellular membrane to the other
couple of decades, many carrier proteins and in renal excretion. However, even in through the lipid bilayer portion (FIG. 1).
(transporters) have been discovered to those three cases, passive diffusion coexists. As the fluidic lipid bilayer portion of a
transport various drugs. Consequently, Drug absorption across the intestinal membrane does not have specific binding
it would be interesting to discuss whether epithelium will be used as an example in sites, passive transcellular transport is usually
carrier-mediated transport is the ubiquitous this article because abundant drug transport not saturable, not subject to inhibition
and dominant mechanism of biological studies have been done in this tissue. and less sensitive (or not sensitive at all)
membrane permeation for most drugs in The theory for intestinal transport is to the stereospecific structure of a drug
most organs1,2 or whether it is a specialized qualitatively valid for other tissues in vivo (BOX 1). As the centre of the lipid bilayer is
mechanism for a certain drug in a certain (the quantitative contribution of each highly hydrophobic, a compound diffuses
cell type. process differs case by case, by tissues and across the lipid bilayer portion mainly as an
Based on a literature review described drugs, in accordance with the varying uncharged and largely desolvated species,
in this article, we conclude that passive levels of transporters expressed in different depending on its molecular size and affinity
transport and carrier­mediated transport tissues, and the drug concentration at the to the centre of the lipid bilayer (known as
coexist in biological membrane permeation biological membrane and the binding sites). its lipophilicity). Therefore, passive trans­
to various extents. The weight of contribu­ We are aware that there is also large interest cellular transport generally depends on the
tion from each route to the total membrane in understanding carrier­mediated trans­ uncharged fraction of a permeant (that is,
permeation should be carefully considered port in other organs, especially in the liver, it depends on its pH and pKa) and the whole
on a case by case basis (for each drug and the kidney and the brain3. Once carrier­ molecule physicochemical property of a
each cell membrane). Passive transport is mediated transport mechanisms can be permeant such as its octanol–water partition
probably the primary route of membrane properly quantified in these organs, an coefficient, its total hydrogen bond strength

nATure revIeWs | Drug Discovery vOlume 9 | AugusT 2010 | 597

© 2010 Macmillan Publishers Limited. All rights reserved


PersPectives

(ABC) and the solute carrier (slC) families.


Box 1 | Terminology of transport routes and direction of transport
About two dozen of these are of clinical rele­
A schematic representation is shown of how compounds can permeate the intestinal membrane. vance with regard to drug disposition and/or
Transcellular permeation is the route passing through the epithelial cells (that consists of passive side effects3.
transcellular and carrier-mediated permeation). Paracellular permeation is the route passing If the transport process directly or
between the cells. The aqueous boundary layer (ABL) is adjacent to a membrane and could be a indirectly consumes energy (ATP), it is
permeation barrier. Uptake refers to transport into a cell and efflux refers to transport out of a cell.
said to be active and may not require a
Apical (A) side Intestinal membrane Basolateral (B) side concentration gradient of a permeant.
The ABC transporter superfamily requires
binding and hydrolysis of ATP to function
and to mediate the primary active transport
Paracellular permeation (passive) process. many uptake transporters of
the slC superfamily 13,14 are driven by ion
ABL permeation (passive)
gradients created by ATP­dependent
Passive transcellular Passive transcellular primary transporters, such as the na+/
transport transport K+­ATPase. These co­transporters are said
to mediate secondary active transport.
Carrier-mediated Carrier-mediated
When carrier­mediated transport is not
uptake efflux energy driven, it is defined as a facilitated
Carrier-mediated Carrier-mediated transport process and relies on a concentra­
efflux uptake tion gradient of a substrate, as well as a
transporter protein. unlike passive trans­
cellular transport, the carrier­mediated
process is expected to be saturable (BOXEs 2,3,
Tight junction see supplementary information s1 (table)
for examples). This occurs when the total
number of permeant molecules exceeds
the number of carrier protein binding sites
Concentration gradient available for transport (that is, when the
Absorptive direction (A to B) concentration of a substrate is higher
Excretive direction (B to A) than the michaelis constant (Km) (BOX 3)).
Carrier­mediated transport is also subject to
inhibition. As carrier proteins have a more
and its molecular size (TABLE 1), whereas Natureall
is common to almost Reviews Drug Discovery
cells. |Therefore, rigid structure than fluidic lipid bilayers and
stereochemically­specific interactions basic passive transcellular transport occurs as carrier proteins are made of chiral amino
determine the carrier­mediated transport. regardless of cell type (for example between acids, carrier­mediated transport is stereo­
This is called the pH partition theory4,5. in vivo organs and in vitro cells). The extent specific and in many cases enantioselective,
When pKa is neglected, the correlation of passive transcellular transport could be whereas the molecular properties determining
between the lipophilicity of a compound dependent on the lipid composition of the passive transcellular membrane transport are
and the biological membrane permeation membrane, but is usually of comparable generally identical between enantiomers.
can become unclear for compounds that magnitude between different cell types. Therefore, even though the physical
undergo ionization6,7. In addition, because For instance, Caco-2 cells (derived from the forces (for example, a hydrogen bond)
passive transcellular transport is a diffusion human colon) and madin–Darby canine governing both inter­molecular interactions
process, after normalizing for lipophilicity, kidney (mDCK) cells show similar mag­ (that is, drug–carrier protein and drug–lipid
a correlation with the molecular size of a nitudes of passive transcellular transport 12. bilayer) are the same, the spatial (stereo)
compound (the molecular size correlates with Therefore, in general, a lipophilic drug patterns are different between passive and
the diffusion coefficient) can be observed8,9. that displays a high passive transcellular carrier­mediated transport. Hence, for
This dependency on molecular size that is transport across the intestinal epithelial cell passive transport, molecular chemical
observed in passive transcellular permeation membrane may also display a high passive descriptors (such as total hydrogen bond
should not be confused with that in paracell­ transcellular transport across, for example, strength, lipophilicity and dipole/polariz­
ular or aqueous­pore permeation. However, the sinusoidal and canalicular cell membranes ability) are often used to investigate the
the molecular size dependency of passive of the hepatocyte. quantitative structure–permeation relation­
transcellular transport is usually less sig­ ship (QsPr)15. By contrast, for carrier­
nificant than lipophilicity dependency, and Carrier-mediated transport. A membrane mediated transport, stereostructure specific
often is difficult to detect for loosely packed transport process involving a protein in the QsPr (three­dimensional QsPr) and/
membranes such as the intestinal membrane. transcellular permeation process is called or structure­based drug design approaches
But it could be observed for the more densely carrier­mediated in this article (the protein tend to be used16,17. In the literature on
packed blood–brain barrier 10,11. is called the transporter). There are more carrier­mediated transport, dependence on
Although there are variations in the than 400 membrane transporters, and these lipophilicity is not well documented, but can
lipid composition of membranes between belong to two major superfamilies of mem­ be identified in a narrow range of chemical
different cell types, the lipid bilayer portion brane transporters: the ATP­binding cassette structural classes. most importantly,

598 | AugusT 2010 | vOlume 9 www.nature.com/reviews/drugdisc

© 2010 Macmillan Publishers Limited. All rights reserved


PersPectives

Octanol–water Artificial
partitioning membranes Cell culture systems

a b c d e f
Carrier-
Transcellular Paracellular mediated Cell suspension Membrane
or adherent cell vesicles

Cell
monolayer Transport
protein

Passive transmembrane Passive transcellular and paracellular transport Passive transmembrane


transport transport

Active/carrier-mediated transport
Figure 1 | The potential of a drug to pass biological membranes by epithelium or the blood–brain barrier endothelium. The monolayer-forming
passive and carrier-mediated processes can be assessed by multiple cells provide a paracellular transport routeNature
between the cells.
Reviews | Drug The para-
Discovery
techniques. Dotted and solid arrows and lines indicate passive and carrier- cellular route is important for the permeation of small hydrophilic drugs in
mediated transport routes, respectively. a | Organic solvent–water par- leaky barriers, such as the upper small intestine, whereas it lacks significance
titioning systems (for example, octanol, hexadecane) are often used to in tighter barriers, such as the colon or the blood–brain barrier. d | A mono-
model the distribution of drugs into the cell membrane, rather than their layer-forming cell line with both passive drug transport and carrier-mediated
permeation across the cell membrane. b | Artificial membranes are available drug transport routes. A cell monolayer which expresses a carrier protein
in different formats and compositions, such as black lipid membranes231, can be used to investigate both passive transport and carrier-mediated
unilamellar vesicle liposomes232,233 or parallel artificial membrane per- transport processes. To examine the contribution of passive transport, mock
meation assays. Artificial membranes allow transport studies across the lipid transfected cell lines are usually used as controls. e | A suspension or an
bilayer through a passive transmembrane route in isolation. Membrane adherent cell line that overexpresses either an uptake or an efflux transport
composition can influence compound distribution and transport234 and protein, for example, HeK293 cells236. cell lines that do not form monolayers
transmembrane pH gradients are important for the distributions of acids are also used for in vitro studies of transport processes. However, these
and bases, thus supporting the importance of compound partitioning for studies are limited to uptake into the cells or efflux out of the cells, rather
membrane passage234,235. c | A monolayer-forming cell line with passive than transport across monolayers. Their most common application is in
transcellular and passive paracellular pathways, and negligible carrier- studies of carrier-mediated drug transport. f | inverted membrane vesicles
mediated transport. some cell lines originating from epithelial and from a cell line expressing an efflux transport protein. As they are inverted,
endothelial membranes differentiate into cell monolayers that form barriers the vesicles allow efflux protein-mediated accumulation of a substrate
resembling the physiological barriers of, for example, the intestinal in the vesicles, which simplifies the experiments237.

the carrier­mediated transport of a drug a compound. However, such models are not and Xenopus oocytes (see supplementary
occurs through the specific cells that express available. A good, but not perfect, choice for information s1 (table)). usually when trans­
the transporter. These characteristics for such a reference membrane is the black lipid fected cells are used, non­transfected cells
carrier­mediated transport (BOX 2) are membrane model20–23 and unilamellar vesicles (mock) are simultaneously used as a control
different from those of passive transcellular (liposomes)8,9. The black lipid membrane experiment to evaluate the contribution of
transport. Detailed definitions and char­ model and liposomes consist of a single passive transcellular membrane transport.
acterization of carrier­mediated transport bilayer formed from predominantly zwitter­ Ex vivo and in situ rodent models such as
processes are found elsewhere18,19. ionic phospholipids (for example, egg lecithin, the Ussing chamber, everted sac and intestinal
There are some exceptions to these rules, which is a mixture of phosphatidylcholines). perfusion models4,26 can also be used to inves­
and therefore these may be classed as gen­ Another way of investigating passive tigate carrier­mediated transport. Although
eral rules rather than being solely sufficient transport is to use a monolayer­forming cell extremely costly to obtain, permeation data
evidence. For example, temperature and pH line that does not express functional drug from human in vivo intestinal membrane
dependency are now not considered criteria transporting proteins, such as 2/4/A1 cells24,25. would be the most direct and the most rele­
for judging carrier­mediated transport, as vant evidence of the extent of carrier­mediated
they are also observed in passive transcellular Methods used to investigate carrier-mediated transport of a drug in the human intestine33–35.
membrane transport (discussed later, and transport. monolayer­forming cell lines, Comparison of the plasma concentration–
excluded from BOX 1). such as the Caco­2 (REFs 26–30) and mDCK time profile and tissue distribution data
cell lines31,32, in which one, two or even four between normal and knockout animals36
investigating permeation mechanisms transport proteins can be simultaneously provides direct evidence of the involvement of
Methods used to investigate passive expressed are commonly used intestinal a transporter in pharmacokinetics. methods to
transport. A membrane with exactly the same barrier models. Other adherent or suspension deconvolute various components of transport
lipid composition and lipid bilayer structure cell lines that can overexpress transport pro­ in the Caco­2 model have also been studied37.
as a biological membrane, but without any teins by gene transfection are also commonly In addition, the saturation and the inhibi­
membrane proteins, would be ideal to inves­ used for studies of carrier­mediated transport tion of transport of a compound by a specific
tigate the passive transcellular transport of (FIG. 1); for example, Hela cells, HeK293 cells inhibitor also suggests that the compound

nATure revIeWs | Drug Discovery vOlume 9 | AugusT 2010 | 599

© 2010 Macmillan Publishers Limited. All rights reserved


PersPectives

Table1 | References describing relationships between biological membrane transport and passive transport indicators
Parameter names¶ Biological membrane permeation¶ Number of compounds**,‡‡
Passive transport indicator: physicochemical properties*
Log Poct, pKa, Mr Human oral Fa 92 drugs7, 258 drugs80, 567 drugs85
Log Poct, pKa, polar surface area, hydrogen bond number Human oral Fa 170 drugs76
Polar surface area Human oral Fa 92 drugs7, 20 drugs238
Hydrogen bond strength, polarizability, dipole moment, Human oral Fa 178 drugs90
molecular volume
Polar surface area, Mr, log Poct, pKa Human Peff 22 drugs84
Log Poct, pKa, polar surface area, hydrogen bond number Human oral absorption rate 22 drugs76
Log Poct, pKa, polar surface area, hydrogen bond number rat bioavailability 1,117 in-house compounds83
Log Poct, pKa rat intestinal absorption rate 75 drugs78
Log Poct, Mr caco-2 16,227 in-house compounds75
Polar surface area caco-2 77 drugs88
Log Poct, pKa, Mr caco-2 35 drugs74
Hydrogen bond strength, polarizability, dipole moment, rat blood–brain barrier permeation 30 drugs87
molecular volume
surface activity cNs permeation 42 drugs86
Log Poct, pKa Human tissue distribution 31 drugs81
Log Poct, pKa Human volume of distribution 64 drugs77, 670 drugs79
Passive transport indicator: artificial membrane partition ‡

Liposome binding Human oral Fa 27 drugs109, 20 drugs110


immobilized artificial membrane chromatography rat oral bioavailability 12 drugs111
immobilized liposome chromatography Human oral Fa 13 drugs108

is a substrate of carrier­mediated transport. drugs that are reported to cross membranes variability, partly because of missing
However, being an inhibitor is not necessarily by carrier­mediated transport have been standardization practices53. second, carrier­
equal to being a substrate, and many inhibi­ studied in these cell lines, although they mediated and paracellular transport are
tors are nonspecific, as they also interact with express many of the carrier­mediated trans­ absent in standard PAmPA assays40,54.
several transporters and drug metabolizing port proteins found in the human small In cases in which these aspects are neglected
enzymes38. For instance, P-glycoprotein (P­gp) intestine, at levels only partly comparable to (for example, compiling data from different
and CYP3A4 have a significant overlap in those found in vivo45,46. The permeation of laboratories or including paracellular route
substrate specificity 39. a multitude of compound series have been permeants in the data set), it could result in a
Analyses of concentration dependency ranked in Caco­2 cells to gain an apprecia­ superficially insignificant correlation among
data using kinetic models, such as the tion of the relative oral absorption potential these assays55,56, possibly leading to a conclu­
michaelis–menten equation (BOX 3), are also of lead analogues47–49. sion that passive permeation cannot occur
used to differentiate carrier­mediated trans­ However, Caco­2 cells are more labour in the permeation of biological membranes.
port from passive transcellular membrane intensive and have a lower throughput than
transport. According to the michaelis– artificial membranes, and thus are used as Separating different forms of transport
menten equation, the fraction of a permeant a secondary lower throughput screen when All rules have exceptions. The criteria
that crosses the membrane by passive tran­ results from a more physiological model are outlined in BOX 2 are general characteristics
scellular membrane transport can become desired. Other specific models of membrane of passive transcellular transport and
more significant when the concentration transport such as genetically transfected carrier­mediated transport and are not strict
of the permeant is higher than the Km. mDCK cells that express P­gp (mDCK­ criteria. For example, in some high capacity
MDR1 cells) can also be used to determine transporters such as amino acid and glucose
practical methods in drug discovery how a compound may be transported across transporters57, it may often be difficult to
As permeation measurements using black a membrane50,51. obtain saturation, without which it may be
lipid membrane and liposomes are difficult, There are two aspects that should be erroneously concluded that the mechanism
the parallel artificial membrane permeation carefully considered when correlating of permeation is passive diffusion. A strictly
assay (PAmPA)40 is an alternative used by PAmPA, Caco­2 and mDCK data. linear concentration dependency is only
many pharmaceutical companies, and has First, these experiments can give variable observed for electrically neutral molecules
been utilized in the study of many thousands results when not performed under stand­ or charged molecules at low concentrations.
of compounds41–43. In addition, Caco­2 cells ardized conditions52 and there is large many drugs are charged at the physiological
and mDCK cells are routinely used44. many inter­laboratory and intra­laboratory pH range and under these conditions

600 | AugusT 2010 | vOlume 9 www.nature.com/reviews/drugdisc

© 2010 Macmillan Publishers Limited. All rights reserved


PersPectives

Table1 (cont.) | References describing relationships between biological membrane transport and passive transport indicators
Parameter names¶ Biological membrane permeation¶ Number of compounds**,‡‡
Passive transport indicator: artificial membrane permeation§
Octanol membrane caco-2 16 drugs74
egg lecithin PAMPA# Human oral Fa 92 drugs7, 25 drugs40
Hexadecane PAMPA Human oral Fa 32 drugs106
Biomimetic PAMPA Human oral Fa 80 drugs54,101,104
Phospholipid–octanol membrane Human oral Fa 21 drugs6
Liposome PAMPA Human oral Fa 21 drugs96
Trilayer PAMPA Human oral Fa 35 drugs239
Biomimetic PAMPA Human Peff 18 drugs105
Double-sink PAMPA Human Peff 8 drugs91
Double-sink PAMPA rat Peff 17 drugs26
Double-sink PAMPA caco-2 18 drugs37
egg lecithin PAMPA caco-2 92 drugs7
Biomimetic PAMPA caco-2 20 drugs103
Trilayer PAMPA caco-2 35 drugs239
PAMPA MDcK-MDR1 40 drugs and 72 in-house compounds97
Blood–brain barrier PAMPA Brain to plasma ratio 30 drugs and 14 in-house compounds95
Blood–brain barrier PAMPA Mouse blood–brain barrier permeation 130 drugs93
silicon PAMPA Human skin Kp 19 drugs102
Double-sink PAMPA Ki/ic90 34 in-house compounds99
silicon PAMPA Uptake in fish 20 chemicals98
Passive transport indicator: cells without functioning transporters||
2/4/A1 cells Human oral Fa 30 drugs100
Fa, fraction of a dose absorbed; ic90, concentration that produces 90% inhibition in a cell-based assay; Kp, human skin permeability coefficient value; Ki, inhibition
dissociation constant or binding affinity of the inhibitor; log Poct, octanol–water partition coefficient of a compound; Mr, molecular mass; Peff, effective intestinal
membrane permeation and is the sum of passive transport and carrier-mediated transport; pKa, dissociation constant. *Partition coefficients in different solvent
systems and additional physicochemical properties can be used to describe biological membrane permeation and can suggest that passive transport is dominant
for these drugs and compounds. ‡Partition into artificial membranes can be used to describe biological membrane permeation and can suggest that passive
transport is dominant for these drugs and compounds. §Permeations derived by passive permeation measurements using artificial membranes are predictive for
biological membrane permeation and can suggest that passive transport is dominant for these drugs and compounds. ||Permeations derived by measurements
using cells that do not express a functioning transporter are predictive for biological membrane permeation and can suggest that passive transport is dominant for
these drugs and compounds. ¶For details of each parameter please refer to the references. #There are various parallel artificial membrane permeation assay
(PAMPA) technologies to improve predictability for biological membrane permeation. For details of each PAMPA please refer to the references. **Drugs refer to a
compound that has been used clinically. in-house compounds refers to a compound in the drug discovery process. Both are cited as these two could have different
drug-like properties. ‡‡Typical drugs used for these investigations (the number in the parentheses is oral Fa in humans): acebutolol (80–90), acetaminophen (80),
acetylsalicylic acid (84–100), acyclovir (20–23), allopurinol (90), alprenolol (93–96), amiloride (50), ampicillin (62), antipyrine (97), atenolol (50–54), aztreonam (1),
barbital (90), bromocriptine (28), bupropion (87), caffeine (100), carbamazepine (70–100), ceftriaxone (1), cefuroxime (5), cephalexin (0), chloramphenicol (90),
chlorothiazide (13), chlorpromazine (100), cidofovir (3), cimetidine (64–95), ciprofloxacin (69), cloxacillin (37–60), clozapine (100), corticosterone (100), coumarin
(100), creatinine (80), cymarin (47), cytarabine (20), desferrioxamine (2), desipramine (100), dexamethasone (80–100), diazepam (100), diclofenac (100), dicloxacillin
(35–76), dilthiazem (80), diltiazem (92), dipyridamole (66), doxycycline (90–100), enalapril (66), erythritol (90), ethambutol (80), ethionamide (80), etoposide (50),
famotidine (38–45), fenoterol (60), flecainide (81), flucytosin (75–90), foscarnet (17), furosemide (50–61), ganciclovir (3), gentamycin (0), guanabenz (75–80),
HBeD (5), hydrocortisone (55–91), imipramine (99–100), indomethacin (100), isoniazid (80), ketoprofen (100), labetalol (90), lactulose (0.6), lansoprazole (85),
lincomycin (28), mannitol (16–26), metaproterenol (44), metformin (86), methotrexate (20), methylprednisolone (82), metolazone (64), metoprolol (95),
nadolol (35–57), naltrexone (96), naproxen (99–100), nicotinic acid (88), norfloxacin (71), olsalazine (2), oxacillin (30–35), oxprenolol (97), oxybutynin (6),
oxytetracycline (60), phenobarbital (100), phenytoin (90), pindolol (87–92), piroxicam (100), practolol (100), pravastatin (34), prazosin (77–95), prednisolone
(99), procainamide (75–95), progesterone (91), propranolol (90–100), propylthiouracil (76), quinidine (81), quinine (90), raffinose (0.3), ranitidine (50–64),
ribavirin (33), salicylic (acid) (100), sotalol (60), streptomycin (1), sulphasalazine (12–59), sulindac (90), sulpiride (35–44), sumatriptan (57), terbutaline (62–73),
testosterone (98–100), tetracycline (75–80), theophylline (98–100), tiacrilast (99), timolol (72–95), tolbutamide (85), tranexamic (acid) (55), valsartan (55),
verapamil (95–100), warfarin (93–98), zidovudine (100).

membrane binding and passive diffusion components or by electric potential differ­ membranes (or octanol as a surrogate) can be
could be nonlinear (owing to a change in ences across membranes), saturation and regarded as the precondition for membrane
the surface potential of the membrane inhibition of passive permeation were also passage, partition processes are not sufficient
on drug partitioning 10) and could closely observed58,59. several efflux transporters show to describe membrane permeation for all
resemble carrier­mediated transport (BOX 3). a high degree of promiscuity in their sub­ compound classes correctly.
In some cases of passive transcellular strate recognition, but the promiscuity is still For instance, highly charged or polar
transport (such as transport of ionized narrower than passive transport. Although amphiphilic compounds have the potential to
molecules by ion pairing with lipid partitioning processes of compounds into accumulate in the polar membrane region,

nATure revIeWs | Drug Discovery vOlume 9 | AugusT 2010 | 601

© 2010 Macmillan Publishers Limited. All rights reserved


PersPectives

Box 2 | General features of passive and carrier-mediated transports rapidly permeated, whereas compounds
with low lipophilicity slowly permeated
carrier-mediated transport (for example, glycerol (log Doct = –1.76) and
• Concentration dependent (saturable)
urea (log Doct = –1.66)72). These studies indi­
• Subject to inhibition cate that many drug­like compounds can
• More structure specific than passive transport, but dependence on lipophilicity could be pass through the lipid bilayer in proportion
identified in a narrow chemical series to their lipophilicity 73. Correlation between
• Cell type specific; requires expression of the transporter indicators of biological membrane permea­
Passive transport
tion and passive permeation (the whole
• Not concentration dependent (non-saturable) molecule physicochemical property and
artificial membrane permeation) for struc­
• Not subject to inhibition
turally diverse compounds also suggests that
• Less structure specific than carrier-mediated transport; there is a general dependence on
passive transcellular membrane transport of
lipophilicity for structurally diverse compounds
drugs exists (TABLE 1).
• Less cell type specific than carrier mediated transport
Rationale for the existence of carrier-
mediated transport. given the difficulty in
identifying carrier­mediated transport as
mimicking compounds that have a high many cephalosporins have oral discussed above, we revisited 55 publica­
permeation potential but without passing bioavailability higher than expected from tions (published later than 1996) that have
this barrier 60. lipophilic basic amidine­ their lipophilicity properties, which is due in been cited to provide evidence for a more
containing inhibitors of the coagulation part to peptide transporter 1 (PePT1; also prominent role of carrier­mediated trans­
pathway can belong to this class of com­ known as sCl15A1)­mediated transport 68. port1 (total 100 drug–carrier protein com­
pounds61,62. The influence of charge As PePT1­mediated transport is proton­ binations (see supplementary information
distribution on the permeation of cationic dependent, pH dependency was used to s1 (table)). These reports tended to use a
amphiphilic compounds has been described, support the involvement of PePT1­mediated thorough investigation of the criteria listed
showing that permanently charged lipophilic transport. However, several cephalosporins in BOX 2, especially uptake into transfected
molecules can pass membranes provided have an acidic functional group69, which cells overexpressing the transporter of
that the charge can be spread over several could explain the apparent pH gradient interest (for example, PePT1, PePT2 (also
aromatic ring systems or neutralized by effect as being solely due to the consequence known as sCl15A2), P­gp, breast cancer
making an ion pair complex with a counter of the pH partition theory of passive trans­ resistance protein (BCrP; also known as
anion63,64. Therefore, a thorough study of cellular permeation. As both passive ABCg2), multidrug resistance­associated
carrier­mediated transport should include transport and PePT1­mediated transport protein 1 (mrP1; also known as ABCC1),
multiple indicators of carrier­mediated are pH dependent, it is difficult to quantify organic anion transporting polypeptides
transport 65. the contribution from each mechanism (OATPs), organic cation transporters
by the pH dependency of the permeation. (OCTs)), as demonstrating the involvement
Why not temperature and pH dependency? The pH­dependent in situ intestinal mem­ of carrier­mediated transport. It is clear
Two experimental observations that were brane permeation of benzoic acid in the rat that carrier­mediated transport exists for
used to support carrier­mediated transport is a good example showing the importance drugs permeating biological membranes,
— temperature and pH dependency — can of thorough consideration of the effect of but it is interesting that in most of these
be commonly observed for passive diffusion. pH4 (FIG. 3), which shows that the acid micro­ publications, passive transport was also
energy consumption and active trans­ climate is the most important factor that reported, usually in control experiments in
port processes are minimal at 0–4 °C, determines the effective intestinal membrane non­transfected cells.
whereas energy consumption is normal and permeation (Peff) of benzoic acid70. Clear indications of the involvement of
active transport processes are functioning passive transport were presented in 81 cases.
at 37 °C. Consequently, it has often been Theoretical aspects In 46 cases, passive transcellular transport
assumed that when a compound is taken up In the following sections we first discuss contributed to more than 30% of total
at 37 °C but not at 0–4 °C, the transport is the existence or non­existence of both pas­ uptake and/or permeance. This suggests
carrier­mediated, but when the compound sive transcellular membrane transport and that even among these publications focusing
is taken up at both temperatures then carrier­mediated transport, and then discuss on carrier­mediated transport, passive
transport is thought to be passive. However, the relative contribution of these transport transcellular transport coexisted. Therefore,
passive transport is also highly temperature processes on the pharmacokinetics of drugs. it is common practice to subtract the base­
dependent, because partition/distribution line passive transcellular transport from the
coefficients and passive transcellular Rationale for the existence of passive trans- total permeation in order to separate the
(and paracellular) permeation can be cellular membrane transport. molecules contribution of carrier­mediated transport.
strongly influenced by temperature66,67, diffuse across the membrane in proportion However, these experiments cannot suggest
as exemplified in FIG. 2. For example, to their concentration gradient across the the quantitative contribution of passive trans­
the activation energy values for passive membrane. In black lipid membranes and cellular transport and of carrier­mediated
transport and active transport across cell liposomal membranes, numerous reports transport in the total (net) permeation
monolayers were indistinguishable for suggest that compounds with mid to high in vivo, as there are differences between
indomethacin and salicylic acid27. lipophilicity (for example, a log Doct > 0)23,71 in vitro, in vivo and clinical conditions.

602 | AugusT 2010 | vOlume 9 www.nature.com/reviews/drugdisc

© 2010 Macmillan Publishers Limited. All rights reserved


PersPectives

were obtained when high quality data


Box 3 | Michaelis–Menten equation for membrane permeation
were used (however, again with some
The rate of total mass transport across a cellular membrane (dM/dt) is the sum of passive transport outliers)7,26,37,43,95,99,118–121 (FIG. 5a). In a recent
(PT) and carrier-mediated (CM) transport, and can be described by using the Michaelis–Menton study that compared a Caco­2 cell­based
equation (see equation and figure below). Passive transcellular transport is in proportion to the assay with PAmPA, 85% concordance
concentration gradient across the membrane (in this equation the concentration on the right side between both assays was obtained for struc­
of the membrane is C and that in the right side is assumed to be zero).
turally diverse compounds122, suggesting
F/ #×
2  2 ×%#×2 ×% 8OCZ×% that these compounds permeated across
26 %/ 26
FV -O % Caco­2 and mDCK cells mainly by passive
Where A is the surface area of a membrane (length2); P is the permeability (length/time); C is
transcellular transport.
the concentration of a permeant (amount/length3); Km is the Michaelis constant (amount/length3); The permeation coefficients of 30 drugs
Vmaxis the maximum carrier-mediated transport (amount/time) in 2/4/A1 cells showed a good correla­
tion with the human Fa of these drugs123.
In these studies, the relative importance
of passive transport and carrier­mediated
Total (net) transport transport was investigated using drugs
(carrier-mediated
+ passive) that are more hydrophilic (therefore have
Carrier- Carrier- lower passive transcellular transport) than
mediated
lipophilic (which would have high passive
dM/dt

mediated
Vmax transport transport
transcellular transport). Therefore,
Passive
transcellular
carrier­mediated transport should be more
transport readily observed than passive transcellular
Passive transport. These drugs are usually incom­
transport
pletely absorbed in the human intestine100
(Biopharmaceutics classification system (BCs)
Km Concentration class III). The data set was selected so that
half of the drugs were known to be at least
partly transported by different carrier
proteins of the human intestine, whereas
so we conclude that both passive trans­ PAmPA6,7,26,37,40,43,54,91–107
Nature Reviews and other
| Drugartificial
Discovery for the remaining part of the data set, no
cellular transport and carrier­mediated membrane systems83,108–111 showed good clear indications of carrier­mediated trans­
transport exist in the biological membrane correlations with clinical Fa for more than port had been published. The permeation
permeation of drugs. 100 structurally diverse drugs (TABLE 1) of these compounds was tested in three
albeit with some outliers40,97,112. most regis­ in vitro models: artificial membranes,
practical aspects tered oral drugs (for example, propranolol, Caco­2 cells and 2/4/A1 cells. The best
In this section, we discuss the relative con­ desipramine and naproxen), as well as correlation between the human Fa and
tribution of passive transcellular transport typical drug candidates, have mid to high in vitro permeance was obtained for 2/4/A1
and carrier­mediated transport in vivo, PAmPA permeations43, suggesting that cells, which do not express any functioning
using intestinal membrane permeation as these compounds are absorbed mainly by carrier proteins. Importantly, 3 out of the
an example. As there are more than 1,000 passive transcellular transport. A bilinear approximately 15 actively transported com­
marketed drugs and much larger numbers relationship between PAmPA permeations pounds were outliers. Indeed, two of these,
of drug­like compounds in discovery, it is and log Poct/Doct (FIG. 4) have been described a model substrate for PePT1 (gly­sar) and
difficult to have an exact percentage but we for structurally diverse drugs113,114. Those methotrexate (a substrate for the reduced
attempt to draw a conclusion. results supported the principles described hydrofolate transporter) are known to be
by Kubinyi115,116, which state that the rate highly dependent on active transport for
Contribution of passive transport in constants of transport of a drug from an intestinal permeation.
intestinal membrane permeation. Although aqueous phase into an organic phase (k1) and Correlations between indicators of
there are large differences between octanol in the reverse direction (k2) can be described passive permeation (that is, log Doct, artificial
and lipid bilayers, octanol–water partition by nonlinear relationships. As lipophilicity membrane permeabilities and 2/4/A1 cell
coefficients or distribution coefficients increases, k1 increases whereas k2 decreases, permeation) and in vivo intestinal mem­
(log Poct, log Doct) can still be regarded as the resulting in the bilinear relationship. brane permeations (oral absorption rates,
gold standard lipophilicity scale that is used Interestingly, the most frequent lipophilicity Peff and Fa) for structurally diverse drugs and
to describe distribution processes of drugs range for structurally diverse registered oral compounds were calculated. These correla­
into membranes. In a study of more than 550 drugs (70% of oral drugs have log Doct values tions provided the evidence supporting the
structurally diverse drugs, log Doct correlated of +0.5 to +3) overlapped with the range that premise that passive transcellular membrane
with the oral absorption rate in rats, the frac­ gives a high membrane permeance, which transport is the primary transport route for
tion of a dose absorbed (Fa) in humans and depends on the method applied83,117 (FIG. 4) . most drugs in in vivo intestinal membrane
Caco­2 membrane permeation7,74–85 (TABLE 1). Despite the restrictions described permeation (TABLE 1). At the same time, the
In addition, various other physicochemical earlier in the direct comparison of Caco­2, existence of outliers suggested a possibility
properties also correlated with Fa and mDCK and PAmPA measurements, good that carrier­mediated transport could be
human Peff (REFs 7,86–90) (TABLE 1). correlations for structurally diverse drugs present for these drugs.

nATure revIeWs | Drug Discovery vOlume 9 | AugusT 2010 | 603

© 2010 Macmillan Publishers Limited. All rights reserved


PersPectives

20 concentration is larger than Km, passive


4 °C
transcellular transport contributes signifi­
18 cantly to the total permeance, or is even
37 °C
the dominating mechanism.
16
This aspect was recently proposed for
angiotensin­converting enzyme (ACe)
inhibitors132,133, some of which are his­
14 torically claimed to be a substrate for the
PAMPA permeation (1 × 10–6 cm s–1)

proton­dependent oligopeptide transporters


12 PePT1 and PePT2. Of the 14 investigated
ACe inhibitors, the majority displayed weak
10 or no carrier­mediated transport. low trans­
port currents were observed for only four of
the inhibitors, including enalapril and lisi­
8
nopril, which have previously been cited as
drugs that are transported by PePT1. When
6 the authors considered the low affinities,
low transport activities, relatively moderate
4 to high lipophilicity (log Doct > 0 in 10 out
of 14 drugs) and daily dose, they thought it
2 highly probable that passive diffusion mainly
controls the in vivo intestinal absorption of
many of the ACe inhibitors. Although this
0
controversial conclusion is supported by
rbu e
e
e

n
Pir ran

asa lol

Va ne

rap n
il
ofe ole

Pit icam
Pro tatin
ole ntip e

oth atin
Fur azide

er fen
Me llow
ide
kin e

Na rolo
Da 8

am
in
din
De pson
pin

Na roxe
n

a
in-

other publications118,134,135, further in vivo

t
o

i
Dip rami
yri

laz
tal
gat
em
Fex dam

r
Lu opro
t

Su pran

lsa
ye
na
ze

top

s
ch vas

ox
i

ava
p
psa
os

studies are needed to confirm the findings.


ma

sip

Ve
sto

yri

dro Flu

lph

Te
Ke
A

cif
lor
rba

cy

In addition, many drugs that were identi­


Ca

fied as a P­gp substrate in an in vitro study


Ch

Hy

were later shown to be completely absorbed


Figure 2 | Temperature dependency (4 °c and 37 °c) of passive transport through an artificial
membrane system of various drugs. Permeation experiments for numerous drugs were in vivo136,137, for example, verapamil. One
Nature Reviews | Drugperformed
Discovery
under conditions as previously described but at two different temperatures63,65,114, 4 °c and 37 °c. reason suggested for this discrepancy
PAMPA, parallel artificial membrane permeation assay. was the difference of the concentrations
between in vitro assays and in vivo intestinal
fluid137,138. usually, in vitro cellular assays
use a low concentration to identify a P­gp
Contribution of carrier-mediated transport On the other hand, the contribution of substrate, whereas a higher concentration
to intestinal membrane permeation — results PePT1 to in vivo oral absorption is ~60% for can be achieved in vivo and clinically (for
from knockout mice experiments. Knockout cefixime (comparing wild­type and knock­ example, when 100 mg dose of a compound
mice124 and transgenic mice with organ­ out mice)129 and is ~50% for cephalexin of molecular mass 400 daltons is dosed
specific expression of human transporters125 (from rat in situ permeation studies)130. in humans, the concentration could be
have been used to evaluate the contribution In addition, several clinical studies show 1,000 μm (intestinal fluid volume is assumed
of carrier­mediated transport in vivo. The that genetic variance of transporters affected to be 250 ml))139. The apparent Km of P­gp is
effect of P­gp on the area under the plasma the AuC after oral administration of several <200 μm for most cases140. Therefore, results
concentration–time curve (AuC) (or plasma drugs (reviewed in REF. 131). This evidence obtained solely from cell culture studies
concentration at a specific time point such as suggests that, at least for several propor­ using a low concentration might have led to
tmax) was determined by oral or intravenous tions of drugs, carrier­mediated transport an overestimation of the importance of P­gp
administration of well­established P­gp sub­ determines more than 50% of membrane transport on the understanding of the main
strates in P­gp­deficient mice (P­gp knockout) permeation in a certain organ. intestinal absorption process.
and wild­type animals36. even though Another reason suggested for the discrep­
bioavailability also includes the effect of gut wall Cases when carrier-mediated transport ancy between in vitro and in vivo findings is
and liver metabolism, and excretion (so that does or does not display a significant effect that passive transcellular transport could be
there is overestimation of the role of P­gp in on oral absorption. For several substrates rapid enough to overcome the efflux carrier­
intestinal membrane permeation126–128), these of P­gp, oral absorption was not affected mediated transport 141–144. Direct comparison
data contain one of the most relevant pieces by P­gp knockout (TABLE 2). As shown in of mDCK­MDR1 cells and PAmPA permea­
of information for the contribution of carrier­ BOX 3, being a transporter substrate does tions highlighted the importance of the
mediated transport on intestinal membrane not necessarily mean that carrier­mediated balance of passive and carrier­mediated
permeation36,126,128. For several compounds, transport dominates the total (net) permea­ efflux transport in total permeance145 (FIG. 5b).
the bioavailability ratio of drugs after oral tion of the substrate in vivo. According to As shown using a large set of compounds,
administration was >2. However, as shown in the michaelis–menten equation for mem­ active efflux from the basolateral to the
TABLE 2, P­gp affects brain distribution more brane permeation (BOX 3), when passive apical side of the cells can mainly be expected
significantly than it affects oral absorption. transcellular permeation is high and/or drug for those compounds that exhibit low passive

604 | AugusT 2010 | vOlume 9 www.nature.com/reviews/drugdisc

© 2010 Macmillan Publishers Limited. All rights reserved


PersPectives

Luminal pH permeation144. similar findings are described


2 3 4 5 7 9 10 11 for bidirectional transport in Caco­2 cells146.
a usually the fraction of compounds with
a large efflux ratio (efr >5) increases with
decreasing permeation. This experimental
–2 observation was further confirmed in
pKa (3.98) Caco­2 cells147 (FIG. 6). These data suggest
pH partition that in order for the efflux P­gp transport
theory to be significant, it has to surmount passive
O
transcellular transport in the absorptive
Air-segmented perfusion OH direction148,149. For instance, digoxin,
–3
hABL 103 µm (estimated) a typical P­gp substrate150, had absorption of
>80%151–153. verapamil is another example154.
Log Peff (cm s–1)

Benzoic acid
For intestinal membrane permeation,
Normal perfusion (0.5 ml min–1) (5.35) the contribution of the paracellular pathway
hABL 531 µm can also be significant54,155. For example, in
–4 (6.05)
the case of ampicillin (a PePT1 substrate),
a significant contribution of the paracellular
route on the total permeance was suggested
Ionic in rats156 in addition to PePT1. moreover,
the total permeance of a paracellular pathway
compound, for example, ranitidine, is less
3 4 5 6 7 8
affected by P­gp efflux 143. As cellular uptake
Microclimate pH assays cannot assess the paracellular route
and as in vitro monolayer cells (for example,
Caco­2) can have a tighter paracellular route,
b pH 5.0 pH 6.5 pH 7.4 these assays would underestimate the contri­
bution of the paracellular pathway. notably,
2/4/A1 cells have paracellular permeance
most comparable to that of the human small
intestine and show good predictability for Fa
of both passive transcellular and paracellular
transport157.
Therefore, carrier­mediated transport
processes in the intestine can be expected to
be more important for low dose or low solu­
bility compounds (hence low concentration
in the intestinal fluid) with low passive
Passive (neutral) Passive (ion) Aqueous boundary layer Paracellular
permeance (in both passive transcellular
membrane transport and paracellular trans­
Figure 3 | pH-dependent permation of benzoic acid in the rat intestine. a | rat in situ intes-
tinal membrane perfusion of benzoic acid as a function of the microclimate pH on the surface of
port). Indeed, this aspect has been repeat­
enterocytes based on published absorption–pH rate data4, which were transformed edly cited in the literature3,136,142–144,154,158–162.
Nature Reviews | into
Drugpermea-
Discovery
tion values. The pH partition theory curve without the interference by the aqueous boundary layer
(ABL) is indicated by the green solid line. The blue dashed curve corresponds to the theoretical Human intestinal membrane permeability.
curve under the air-segmented flow perfusion conditions (efficient stirring in the perfusate, less even though it can be expensive to obtain,
ABL effect) and the red dashed curve corresponds to that in the normal flow perfusion conditions the human Peff is the most authentic method
(large ABL effect). The red circles represent experimental data in normal flow perfusion conditions to investigate intestinal membrane permeation
and the green circles represent experimental data in air-segmented flow perfusion conditions. The relevant to clinical pharmacokinetics. Among
numbers in parentheses are apparent pKa values under air-segmented flow (5.35) and normal flow more than 40 drugs investigated for human
(6.05) conditions, indicating the extent of the ABL-induced ‘artefact shift’ in absorption curves Peff (REFs 34,35), amoxicillin, cephalexin, enal­
from the intrinsic pKa value (3.98). The departure of red and blue dashed curves from the solid green april, lisinopril and valacyclovir are PePT1
curve for pH > 6.5 indicates the minor contribution from the transport of the ionic form of benzoic
substrates, and cyclosporine, cimetidine, fex­
acid. At pH 6, under normal flow perfusion conditions, 44% of the transport follows the original
ofenadine and verapamil are P­gp substrates.
prediction of the pH partition theory that is passive diffusion. Under air-segmented flow condi-
tions, the number increases to 78% of the transport, as a result of the reduction of the ABL resist- Fexofenadine and l­3,4­dihydroxyphenyla­
ance (54% to 18%). When the luminal pH was used as it is (without converting to the microclimate lanine (l­DOPA) are absorbed through an
pH), the pH–permeation curve did not follow the pH partition theory. Therefore, it could be mis- OATP and an amino acid transporter, respec­
leading to conclude that the pH partition theory is inapplicable. The most important factor for tively. In addition, cimetidine is an OCT1 and
rationalizing the observed curve shape is the acid microclimate. b | estimated factors controlling an OCT2 substrate.
transport in the rat in situ intestinal membrane perfusion experiment described in part a, at micro- Amoxicillin (log Doct = –1.7), cephalexin
climates of pH 5.0, pH 6.5 and pH 7.4. Part a is reproduced, with permission from, REF. 41 © (2008) (log Doct= –1.1), valacyclovir (log Doct= –1.4)
elsevier science. and l­DOPA (log Doct < –2) showed Peff values

nATure revIeWs | Drug Discovery vOlume 9 | AugusT 2010 | 605

© 2010 Macmillan Publishers Limited. All rights reserved


PersPectives

a b 100 membrane permeation, although, as we


Partition Hydrocortisone Corticosterone discussed above, a sole experimental find­
90 Naproxen
Aqueous phase Coumarin Testosterone ing should not be used as the main evidence
Cocaine

Flux: mass tranfer via artificial


80
C1 to support this transport mechanism. As
P = C1 /C2
70 Papaverine the substrate concentration in the cytosol

membrane (%)
C2 60 is lower than that on the apical side169, the
Proquazone
Lipid phase 50
apparent Km value could be higher than that
Pindolol
of the intrinsic Km value, depending on the
40
Permeation Metergoline intrinsic passive permeation and the mem­
30 Procainamide brane surface area170 of both apical and baso­
k1 Tetracycline
20 Chlorprothixen lateral membranes. The apparent Km of P­gp
10
Acyclovir value in the apical to the basolateral direction
was seen to be higher than that in the baso­
0 Saccharin
k2 –1 0 1 2 3 4
lateral to the apical direction and can reach >
Log D pH 7.4
1,000 μm172,173 in specific cases. In addition,
the clinical interaction between fexofenadine
Figure 4 | Partitioning versus permeation. a | Principal difference between the partitioning and and nagrignin suggests that about 50–70%
distribution, and membrane permeation of drug-like compounds. b | relationship between permea-
Nature Reviews | Drug Discovery of absorptive permeation in the intestine
tion through a parallel artificial membrane permeation assay (PAMPA) and octanol–water partitioning
might be mediated by passive transcellular
(expressed as the percentage flux: mass transfer via artificial membrane) at physiological pH (log D pH
7.4). A rough nonlinear relationship between lipophilicity and permeation is observed. Bilinear rela- membrane transport and 30–50% mediated
tionships in compound mass transfer as described by Kubinyi115,116, depend on the rate constants of by OATP174. Therefore, the possibility of a
transfer of the drug through aqueous and organic compartments. in simple in vitro systems the rate cancelling out effect between P­gp and OATP
constant k1 of transport of a drug from an aqueous phase into an organic phase and the rate constant that results in a linear increase of exposure
k2 of the reverse process can be described as functions of the partition coefficient P, where c cannot be entirely excluded.
represents concentration. Data in panel b plotted from REF. 114. most reported clinical drug–drug inter­
actions of fexofenadine, including those for
verapamil and ketoconazole, which partly
that are higher than that expected from their to apical direction being 28­fold to 85­fold result in an increase of plasma levels (AuC)
lipophilicity. This is in good agreement with higher than the Papp (apical to basolateral) of fexofenadine, were explained by the inhi­
the notion that carrier­mediated transport in the concentration range 10–1,000 μm163. bition of intestinal efflux by P­gp. The clini­
has a significant role in oral absorption. Indeed, Papp from the basolateral to the apical cal fexofenadine and verapamil interaction
enalapril (log Doct = 0.08) and lisinopril (log direction decreased with increasing con­ suggests that verapamil increased the AuC
Doct = –1.32) were already discussed above. centration (Vmax = 5.21 nmol cm per second of fexofenadine by up to 2–3­fold in total,
Cyclosporine and verapamil showed a good and Km = 150 μm), suggesting the saturation although it was difficult to identify which
permeation (1.6 × 10–4 and 6.80 × 10–4 cm of an apical efflux transporter by this drug. transporter is responsible for this increase175.
per second, respectively), although they are In addition, data obtained using Caco­2 However, there was little or no acute effect
P­gp substrates, suggesting that high passive cells suggest that the in vitro permeation by either of these P­gp inhibitors, verapamil
transcellular transport overcame P­gp efflux. was increased in the apical to basolateral and ketoconazole, on the Peff of fexofenadine
Other drugs (more than 30) were thought to direction by approximately 2–3 fold in the in humans and rats165,176. An in vivo per­
be absorbed by passive diffusion35,84. presence of various P­gp inhibitors, such fusion study with simultaneous assessment
To further investigate the joint contribu­ as verapamil, ketoconazole and gF 120918 of intestinal transport and plasma pharma­
tion of passive transcellular transport and (REFs 163,164,166). However, the Papp (apical cokinetics suggests that liver uptake of fexo­
carrier­mediated transport, fexofenadine to basolateral) was independent of the con­ fenadine is mediated by OATP1B1 and/or
and cimetidine are discussed in detail below centration applied, suggesting that the effect OATP1B3, which could also be inhibited by
as two examples of BCs class III drugs, of carrier­mediated transport in the apical verapamil and ketoconazole177. In addition,
which are probably influenced by carrier­ to the basolateral direction is minimal or the by using double transfected cells expressing
mediated transport and because in vitro, apparent Km in the apical to the basolateral OATP1B1/mrP2 or OATP1B3/mrP2, it
in vivo and clinical permeation data were direction is higher than 1 mm. was shown that OATP1B1 and OATP1B3 are
available in the literature. In a clinical investigation, the plasma involved in the hepatic uptake of fexofena­
exposure of fexofenadine was discovered to dine178. Therefore, this evidence suggests
Fexofenadine. Fexofenadine is a well­ be linear over a wide dose range of 40–800 that the drug–drug interaction could occur
established substrate for P­gp (efflux mg 167. In addition fexofenadine had similar in the liver uptake process rather than at the
direction basolateral to apical) and OATP bioavailability when given orally as a micro­ intestinal membrane permeation process,
(absorptive direction apical to basola­ dose (<100 μg) and at a higher dose of which is also supported by a physiologically
teral)163,164. The Peff in humans was low 120 mg (41% versus 30%, respectively)168. based pharmacokinetic model179.
(0.1–0.2 × 10–4 cm per second) and variable, This is in agreement with the linear in vitro In conclusion, for fexofenadine, passive
which classifies it as a compound with low permeance in the apical to basolateral direc­ transcellular transport is suggested as one
membrane permeation (BCs class III)165. tion as described above in the Caco­2 model. of the main intestinal membrane transport
Fexofenadine displays polarized transport This linear increase of exposure after oral mechanisms in humans, but several other
in Caco­2 cells, with the apparent perme­ administration suggests that passive diffu­ transporters may also be involved, especially
ability coefficient (Papp) from the basolateral sion would be the main route in intestinal in the liver and the kidney 180. The intrinsic

606 | AugusT 2010 | vOlume 9 www.nature.com/reviews/drugdisc

© 2010 Macmillan Publishers Limited. All rights reserved


PersPectives

a –3 b 30 inhibitors used in these studies (for example,


Correlation = 0.84
cyclosporine) could be both inhibitors of the
Log Papp MDCK (1 × 10–6 cm s–1)

–4 25 transporter P­gp and the enzyme CYP3A4

RGF: PappGF120918a–b/Pappa–b
(REF. 39). In addition, it is suggested that P­gp
20
–5 and CYP3A4 are synergistically operating to
15
avoid absorption of xenobiotics in the intes­
–6
tine196,197. P­gp can reduce the permeance of
10 the drug, thereby increasing the chance of
the drug to be metabolized by CYP3A4 in
–7
5 the gut wall.
An interesting example of this phenom­
–8 0
–8 -7 -6 -5 -4 -3 0.01 0.1 1 10 100
enon is paclitaxel, which is reported to have
Log Papp HDM-PAMPA (1 × 10–6 cm s–1) Papp HDM-PAMPA (1 × 10–6 cm s–1) a low bioavailability following oral admin­
istration. The reason for its low bioavail­
Figure 5 | competition between passive and active transport. a | correlation between the
Nature Reviews | Drug Discovery
ability was first thought to be mainly due to
apparent passive permeation (Papp) for 37 drugs in MDcK cells and in hexadecane membrane (HDM)-
parallel artificial membrane permeation assay (PAMPA). compounds studied were acyclovir, alprenolol, intestinal efflux 193,198,199 but it could be due
amiloride, antipyrine, atenolol, ceftriaxon, chloramphenicol, chloroquine, cimetidine, cyclosporine, to a combination of its low solubility and its
desipramine, digoxin, doxorubicin, fluvastatin, furosemide, guanabenz, imipramine, lansoprazole, extensive first pass metabolism198,200. Paclitaxel
methotrexate, metoprolol, midazolam, mitoxantrone, naproxen, omeprazole, pantoprazole, prazosine, has a relatively high in vitro permeance in
propranolol, pumafentrine, quinidine, ranitidine, ritonavir, sulphasalazine, sulpiride, terbutaline, Caco­2 cells (4.4 × 10–6 cm per second, apical
testosterone, tolafentrine, topotecan and verapamil. A good correlation (r = 0.84) was observed to basolateral)201, which suggests that the
between the PAMPA and MDcK cell permeations of the drugs, suggesting that passive transcellular drug will be effectively absorbed in the intes­
transport is dominant in the permeation of these compounds in MDcK cells. b | correlation between tinal tract and that presystemic metabolism
P-glycoprotein (P-gp) (efflux) function determined as the ratio of apparent permeation across caco-2 is the major reason for incomplete bioavaila­
cells in the absence and presence of the P-gp inhibitor GF120918 (REF. 145) and passive permeation
bility. Increased solubility and dissolution rate
determined by HDM-PAMPA. All compounds analysed are known to show P-gp-ATPase activity. The
dashed lines indicate the threshold ratios of 0.5 and 1.25, which discriminate between compounds provided higher luminal concentration that
that are actively transported (above the lines). A compound between the threshold ratios of 0.5 and could more efficiently saturate the intestinal
1.25 indicates the marginal involvement of P-gp. Only compounds with low passive permeation show efflux and/or metabolism for paclitaxel200,202.
significant active transport expressed by the rGF ratio (apical (a) to basolateral (b) transport rate ratio It is worth noting that extensive investi­
(inhibited/non inhibited)). These data confirm findings that transporter-related efflux has to surmount gations are currently being undertaken to
passive permeation in order to be significant. Data plotted from REF. 145. identify a specific inhibitor for each trans­
porter 38, which will allow us to quantify
the contribution of each transport process
in vivo and in clinical pharmacokinetics.
low passive membrane permeance of fexo­ permeation mechanism in intestinal mem­
fenadine, together with it being a transporter brane permeation of cimetidine in humans. intestinal membrane transport: summary
substrate, determines the low permeance If the organic cation transporters, OCT1 For many drugs that are substrates of
in the small intestine and the liability of the and OCT2, were the primary mechanism it carrier­mediated transport (for example,
AuC for fexofenadine to be largely affected by would be expected that the human intestinal ACe inhibitors, cimetidine, fexofenadine
hepatic uptake and/or other transporters. in vivo permeance would have been signifi­ and verapamil), passive transport con­
cantly higher than the Caco­2 permeation, tributes significantly to the total intesti­
Cimetidine. Cimetidine is a relatively as the expression of OCT1 and OCT2 is nal membrane permeation, especially at
hydrophilic compound (log Poct = 0.48). higher in the human small intestine than it clinical doses.
The rate and the extent of intestinal absorp­ is in the Caco­2 model34,185,187. Therefore, the Conversely, some antibiotics (for example,
tion of cimetidine has been extensively main determinant of the intestinal membrane cefixime)129, folic acid derivatives (for
investigated, with the Fa after oral admin­ permeation of cimetidine is considered to be example methotrexate)203,204 and amino
istration estimated to be 75%181–183 (BCs passive diffusion188. acid derivatives (for example, l­DOPA)205,
class III). Cimetidine is a substrate for P­gp the structures of which are closely related
and/or OCT1 and OCT2 (REFs 184–187). interplay of mechanisms and properties to those of nutritional compounds206, are
The human Peff (in vivo) is 0.22 ± 0.13 × In the above discussion, only passive trans­ mainly absorbed in the intestine by carrier­
10–4 cm per second and 0.32 ± 0.18 × 10–4 cellular transport and carrier­mediated mediated transport. These are hydrophilic
cm per second at luminal concentrations of transport were considered. However, the drugs with little expected passive transport
1.0 mm and 2.7 mm (that is, clinical dose situation surrounding the bioavailability but good oral absorption. In addition, from
200–400 mg (3–6 mm, 250 ml intestinal of a drug may be more complex. the pharmacokinetic data of P­gp knockout
fluid volume)), respectively 34. several clinical studies have claimed that animals and clinical genetic variants of P­gp,
The lack of difference between perme­ an inhibition of intestinal efflux (especially oral absorption of several P­gp substrates
ance at the two concentrations, together with by inhibiting P­gp) is the major cause behind have been proven to be limited by P­gp
the observation that the human intestinal increased bioavailability when certain in vivo36,126,128 (TABLE 2).
membrane permeance quantitatively cor­ drugs are coadministered189–195. However, These results collectively suggest that pas­
relates well to that of the Caco­2 model, in many of these reports, involvement of sive transport and carrier­mediated transport
suggests that passive transport is the primary CYP3A4 cannot be excluded51,138,184–187, as the coexist in intestinal membrane permeation.

nATure revIeWs | Drug Discovery vOlume 9 | AugusT 2010 | 607

© 2010 Macmillan Publishers Limited. All rights reserved


PersPectives

Table 2 | Data from knockout mice for bioavailability and brain to plasma ratio*
Drug Plasma Auc or concentration ratio Bioavailability ratio Brain to plasma ratio
intravenous orally
Amprenavir NA 1.3 <1.3 21
Asimadoline 1.0 1.1 1.1 9.1
Benzo(a)pyrene 0.8 0.8 1.0 1.6
cyclosporine 1.1 0.6–0.9 0.5–0.8, 1.6 11–29
Digoxin NA 2.4 2 4–28
Dihydroergocryptine 1.8 1.8 1.0 1.1
erythromycin 1.5 3.4 2.3 1.2
Fexofenadine 1.0, 4.6 4.6, 6.5 1.0, 6.5 1.9
Fluconazone 1.2 1.2 1.0 0.9
indinavir 0.7 2.0 2.9 3–10
ivermectin NA 1.9–3.7 <1.9–3.7 17–27
Loperamide 2.0 2.0 1.0 6.7
Nelfinavir 1.3 4.8 3.7 31
Paclitaxel 1.1–2 5.0–6.0 2.5–5.5, 3.18–6.71 7.9
resperine 1.2 1.2 1.0 2.4
retinoic acid 1.0 1.1 1.1 1.0
rifampicin NA 3.5 NA NA
ritonavir 1.0 1.0 1.0 6.9
s 09788 NA 2.4 3.4 NA
salinomycin NA NA 1.5 NA
saquinavir 0.7–1.1 6.5 1.55, 5.9–9.3 2.2–6.8
Tacrolimus 2.3 8.2 3.5, 3.6 6
Talinolol NA 2.9 NA NA
Topotecan NA 2.3 <2.3 2.0
UK-224671 1.1 >40 >36 NA
verapamil NA 1 1 8.3
vinorelbine NA NA 1.5 NA
AUc, area under the plasma concentration–time curve; NA, not available. For details see reFs 36, 125, 126, 128. *Brain to plasma ratio refers to the concentration
ratio of a drug in the brain compared with that in the plasma.

For most drugs, the intestinal membrane transcellular transport and carrier­mediated concentration of a drug that is not bound to
permeation is primarily determined by pas­ transport should be carefully discussed in plasma proteins is lower than the concentra­
sive transport207, and for a proportion of a case by case manner. tion in the intestinal fluid. In addition, the
drugs it is affected by carrier­mediated trans­ transporters in the excretive direction, such
port84,208. As exemplified by fexofenadine, the Other drug disposition processes as P­gp, tend to have wide substrate specifi­
contribution of carrier­mediated transport We have already highlighted that there is city 213,214. The in vivo and clinical effect of
could become more significant in post­ evidence showing that carrier­mediated carrier­mediated transport in the excretive
absorption processes, for example, for liver transport has a more important role in direction has been shown for many drugs by
clearance, than in the intestinal membrane blood–brain barrier permeation126,209 using knockout animals and by investigating
permeation. even when carrier­mediated (TABLE 2), renal210 and hepatic211 secretions humans with genetic variants124.
transport is involved, the contribution of pas­ than it does in the intestine. recent progress
sive diffusion to the overall permeance could in the understanding of carrier­mediated Distribution: blood–brain barrier. The
become significant when the concentration transport showed that this could be more blood–brain barrier expresses various trans­
at the permeation site is high and/or passive important than usually thought1,3. For exam­ porters such as P­gp, mrPs and OATPs. P­gp
transcellular permeation is rapid. In vitro ple, 7 out of 50 compounds that had been affects brain distribution more significantly
permeation studies at a low concentra­ progressed to clinical studies were catego­ compared to oral absorption (TABLE 2). At
tion could overestimate the contribution of rized as being eliminated by carrier­mediated the same time, passive transcellular trans­
carrier­mediated transport in vivo. Therefore, transport (biliary and/or renal excretion)212. port also has an important role93–95,159,215.
the weight of contribution of passive This is probably due to the fact that the The effect of the efflux transporter on the

608 | AugusT 2010 | vOlume 9 www.nature.com/reviews/drugdisc

© 2010 Macmillan Publishers Limited. All rights reserved


PersPectives

brain distribution of a drug (Kpp,b, the brain


concentration of a drug in the presence and
the absence of an efflux transporter) can be
expressed as Kpp,b = 1 + Clcarrier­mediated, efflux/
Clpassive transcellular membrane (REF. 216), showing that
the balance of passive transcellular transport 12
and carrier­mediated transport affect the > 10
distribution of a drug in the brain. < 10
<8
10 <6
Distribution to other organs. Passive trans­ <4
cellular transport is essential in describ­ <2
ing in vivo distribution properties of drug
8
candidates (TABLE 1). The tissue distribution
coefficients were successfully predicted by
log Poct and pKa in physiologically based
6
pharmacokinetic modelling. For an overview
Efflux ratio

of this technique, see REFs 82,217.


At the same time, for some specific drugs
4
in certain organs, carrier­mediated transport
could be crucial. For example, genetically
modified mice showed increased sensitivity
2
to a neurotoxic drug and a carcinostatic 0.2
drug218, suggesting that efflux carrier­mediated

n –1
)
transport significantly contributes to the 0.4

mi
1.0
permeation of a drug into these tissues.

d,ml
0.8 0.6

(CL
Excretion and metabolism: hepatic uptake

on
0.6

usi
and biliary excretion. About a dozen drug 0.8

iff
0.4

d
transporters are known to be expressed in Apical efflux by

ve
0.2

ssi
hepatocytes211,219. It is established that OATPs transporters

Pa
(CLint,sec, ml min–1) 1.0
have a significant role in the hepatic uptake 0.0

of some hydrophilic statins (for example, Figure 6 | influence of passive permeation and carrier-mediated efflux on the efflux ratio in
pravastatin, rosuvastatin and cerivasta­ caco-2 cells, described by a theoretical model. The efflux ratio depends on a saturable apical
tin)3,220. At the same time, passive transcell­ efflux (cLint,sec) as well as on passive diffusion (cLd = A × Ppassive (BOX 3)).Nature Reviews
The efflux ratio| Drug Discovery
approaches unity
with high passive permeation. Data plotted from REF. 147.
ular transport also has an important role in
hepatic uptake221. It has been suggested that
carrier­mediated transport becomes impor­
tant when passive transcellular transport is per second, the renal clearance often exceeds example, antibiotics) and non­oral delivery
low, whereas passive transcellular transport the glomerular filtration rate159,223, suggesting routes, about 80% (based on experience at
has a dominant role when passive transcell­ a significant contribution of carrier­mediated F. Hoffmann­la roche) of compounds cur­
ular transport is high210,211 (for example, Papp, transport in proximal tubular secretion. rently in lead identification or lead optimiza­
passive transcellular membrane > 5 × 10–6 tion and preclinical or clinical development
cm per second)159,222. Excretion and metabolism: drug–drug have medium to high passive permeance
interactions. Drug–drug interactions that (m.K. and F.s., unpublished observations).
Excretion and metabolism: renal excretion. occur through carrier­mediated transport usually such compounds with high passive
renal excretion is determined by the balance processes can be a safety concern3, for permeance have adequate in vivo exposure
of glomerular filtration, tubular secretion example drug–drug interactions between when not modified by metabolic degrada­
and reabsorption. The glomerular filtration statins and other drugs by OATP transport tion or restricted by low intestinal solubility.
is the ultrafiltration of unbound drugs and in hepatic uptake220. even when mem­ PAmPA experiments can be used to identify
no carrier­mediated transport is involved. In brane permeation of a drug is determined high or low passive transcellular transport
addition, many drugs are reabsorbed by pas­ by passive diffusion, the drug could be an compounds. In addition, low passive tran­
sive transcellular transport depending on the inhibitor for carrier­mediated transport of a scellular transport compounds are further
lipophilicity of the drug. However, carrier­ co­administered drug. characterized in cell­based systems to inves­
mediated transport has a significant role in tigate carrier­mediated transport 225.
tubular secretion210. OCTs, OATPs and mrPs implications for practical drug discovery Being a substrate of carrier­mediated
are expressed in the renal epithelium. When most drug candidates in development are transport can add value to a drug. For exam­
Papp, passive transcellular membrane is > 10 × 10–6 cm per in all probability highly permeable and ple, an acyclovir prodrug (valacyclovir)
second, the passive transcellular membrane low solubility compounds (BCs class II)224. was designed to be transported by PePT1
reabsorption becomes predominant and little This is as a result of a large number of per­ (REFs 226,227). Being a P­gp substrate would
renal excretion is anticipated. By contrast, meation and solubility screening processes. also decrease the central nervous system side
when Papp, passive transcellular membrane is < 1 × 10–6 cm excluding drugs for specific indications (for effects of a drug, for example, ivermectin228.

nATure revIeWs | Drug Discovery vOlume 9 | AugusT 2010 | 609

© 2010 Macmillan Publishers Limited. All rights reserved


PersPectives

glossary
Active transport CYP3A4 metabolism can occur in the gut wall and in Papp is the sum of passive transport and carrier-
An energy dependent, carrier protein-mediated transport the liver. mediated transport when the effect of the aqueous
process that can be against a concentration gradient. boundary layer and the paracellular pathway is
Efflux ratio neglected.
Area under the plasma concentration–time curve (EfR). The ratio of the apparent permeation of a
(AUC). A measure of how much of an administered drug compound in the absorptive (apical to basolateral) Passive transport
reaches the bloodstream in a defined period of time. direction to that in the secretory (basolateral to apical) This refers to the movement of a permeant across a
direction, as determined in cell-based experiments. membrane from a region of high concentration to
Bioavailability The EfR is used to determine possible active transport. that of low concentration by a process of diffusion.
In this article, bioavailability means absolute The rate of passive transport is proportional to the
bioavailability. Bioavailability describes the fraction of an Fa concentration gradient of the permeant across the
administered dose (unchanged drug) that reaches the The fraction of a dose of drug that is absorbed after oral membrane.
systemic circulation. Intravenously administered drugs administration. Fa depends on membrane permeance,
by definition have a bioavailability of 100%. Other routes solubility and the dissolution rate of a compound. It is Peff
of administration may lead to lower bioavailability, expressed as a percentage. (Effective intestinal membrane permeation). Peff is
usually expressed by a ratio or percentage of the measured by an in vivo experiment, for example, the
maximum value. First pass metabolism in situ perfusion method. Peff is the sum of passive
After intestinal absorption, drug molecules first pass transport and carrier-mediated transport.
Biopharmaceutics classification system through the liver before entering the systemic circulation.
(BCs). A classification system for drug molecules that During this process, the drug can be metabolized in the P-glycoprotein
considers solubility and permeation. There are four liver and the systemic bioavailability can be reduced. (P-gp). Also called ABCB1, P-gp is a protein involved in
classes: BCs class I (molecules have high permeation/high active efflux transport processes.
solubility), BCs class II (molecules have high permeation/ Lipophilicity
low solubility), BCs class III (molecules have low The affinity of a compound for a lipid environment, for pH partition theory
permeation/high solubility) and BCs class IV (molecules example, the hydrocarbon core of a phospholipid bilayer. A theory that describes passive transport based on the
have low permeation/low solubility). secondary factors It can also be described as the inter-molecular interaction pH of the aqueous phase, the dissociation constant (pKa)
considered in BCs include the rate of dissolution and between a compound and the solvent environment, such and the lipophilicity of a permeant. It states that an
the pH. as the total hydrogen bond strength and dipole/ ionizable drug moves across a membrane by passive
polarizability. diffusion as its uncharged form, depending on its
Black lipid membrane lipophilicity.
A single bilayer phospholipid membrane and optically Log Doct
black in appearance. It slowly forms when a small quantity The octanol–water apparent partition coefficient at a pKa
of an egg lecithin dissolved in n-decane is placed over a certain pH. Both dissociated and undissociated The dissociation constant pKa for an ionizable molecule is
small hole in a thin sheet of Teflon suspended in an (uncharged) molecular species are taken into account. In the pH at which it would be ionized by 50%. The degree
aqueous buffer solution. such membranes have been this article, when log Doct is less than zero, it is referred to (%) of ionization is pH dependent. For an acid it
viewed as useful models of the more complex natural as having low lipophilicity, when log Doct is greater than decreases by lowering pH. For a base it increases by
membranes. zero but less than 2, it is referred to as having moderate lowering pH.
lipophilicity and when log Doct is greater than 2 it is
Caco-2 cells referred to as having high lipophilicity. Tissue distribution coefficient
An immortalized line of heterogeneous human epithelial This represents the degree of drug molecule distributing
colorectal adenocarcinoma cells used as a drug transport Log Poct into a tissue. This value depends on both passive
model for assessing intestinal absorption. Transport The octanol–water partition coefficient of a compound transport and carrier-mediated transport across the
measurements can be performed in two directions: (neutral form). This parameter is most widely used as a cellular membrane of the tissue.
apical to basolateral or basolateral to apical. whole molecule lipophilicity parameter.
Total permeation
Carrier-mediated transport Oral bioavailability This refers to the combination of passive transport and
This refers to active or facilitated transport that has a The fraction or percentage of a drug that reaches the carrier-mediated transport processes.
limited capacity and thus is saturable and subject to systemic circulation following oral administration. It is
inhibition. determined by the fraction absorbed and the fraction not Ussing chamber
metabolized in the gut wall and in the liver. An instrument used to measure transport (for example,
CYP3A4 of drugs) across epithelial barriers. A sheet of epithelia
(Cytochrome P450 3A4). The most important member Papp (for example, intestinal mucosa) is clamped between
of the cytochrome P450 mixed-function oxidase (Permeability coefficient). The apparent permeation two chambers and drug transport across the epithelia is
system. It is involved in the metabolism of xenobiotics. for example, in Caco-2 cell-based permeation assays. measured.

However, currently, it is difficult to inten­ At the same time, carrier­mediated transport (and often primary) factor of membrane
tionally design a compound that is a dual could be a reason for drug–drug inter­ permeation of drug­like molecules. At the
substrate for both carrier­mediated transport actions and other side effects. Therefore, it same time, a significant proportion of drugs
(especially for transport in the absorptive is important to survey the involvement of were proved to be carrier­mediated trans­
direction) and a pharmacological target. carrier­mediated transport for individual port substrates in specific organs, potentially
Inevitably, when designing an orally available drugs, especially when passive transcellular leading to significant drug–drug interactions
drug, it is the passive transport across mem­ transport of the drug is low to moderate. and side effects. Therefore, further studies of
branes in the intestine that offers the highest the balance of how passive transcellular and
probability of success. For instance, most of Concluding thoughts carrier­mediated transport influences tissue
the prodrug approaches to increase the intes­ We conclude that passive transport and and organ distribution, as well as safety rele­
tinal membrane permeation were designed carrier­mediated transport coexists, and that vant aspects, are highly desirable. Direct cor­
to increase passive transcellular transport229. passive permeation remains a significant relation between in vitro transport and the

610 | AugusT 2010 | vOlume 9 www.nature.com/reviews/drugdisc

© 2010 Macmillan Publishers Limited. All rights reserved


PersPectives

accurate in vivo pharmacokinetic variables, 11. Seelig, A. The role of size and charge for blood–brain 35. Lennernas, H. Modeling gastrointestinal drug
barrier permeation of drugs and fatty acids. J. Mol. absorption requires more in vivo biopharmaceutical
and development of in vitro–in vivo extrapo­ Neurosci. 33, 32–41 (2007). data: experience from in vivo dissolution and
lation models are crucial for further progress 12. Irvine, J. D. et al. MDCK (Madin–Darby canine kidney) permeability studies in humans. Curr. Drug Metab.
cells: a tool for membrane permeability screening. 8, 645–657 (2007).
in the field230. Therefore, efforts should be J. Pharm. Sci. 88, 28–33 (1999). 36. Collett, A., Tanianis-Hughes, J., Hallifax, D. &
increased to improve our understanding of 13. Fredriksson, R., Nordstrom, K. J., Stephansson, O., Warhurst, G. Predicting P-glycoprotein effects on oral
Hagglund, M. G. & Schioth, H. B. The solute carrier absorption: correlation of transport in Caco-2 with
the interplay between passive transport and (SLC) complement of the human genome: phylogenetic drug pharmacokinetics in wild-type and mdr1a–/– mice
carrier­mediated transport processes, and to classification reveals four major families. FEBS Lett. in vivo. Pharm. Res. 21, 819–826 (2004).
582, 3811–3816 (2008). 37. Avdeef, A. et al. Caco-2 permeability of weakly
apply the gained knowledge in the discovery 14. He, L., Vasiliou, K. & Nebert, D. W. Analysis and basic drugs predicted with the double-sink PAMPA
and development of successful drugs. update of the human solute carrier (SLC) gene pKaflux method. Eur. J. Pharm. Sci. 24, 333–349
superfamily. Hum. Genomics 3, 195–206 (2009). (2005).
Kiyohiko Sugano is at Pfizer, Research Formulation, 15. Van de Waterbeemd, H. in Comprehensive Medicinal 38. Matsson, P., Pedersen, J. M., Norinder, U.,
Chemistry II Vol. 5 (eds Testa, B. & Van de Bergstroem, C. A. & Artursson, P. Identification of
Sandwich Laboratories, Ramsgate Road,
Waterbeemd, H.) 669–697, (Elsevier Science, novel specific and general inhibitors of the three major
Sandwich, Kent CT13 9NJ, UK. New York, 2006). human ATP-binding cassette transporters P-gp, BCRP
16. Chang, C., Ray, A. & Swaan, P. In silico strategies and MRP2 among registered drugs. Pharm. Res. 26,
Manfred Kansy, Stefanie Bendels, Holger Fischer,
for modeling membrane transporter function. 1816–1831 (2009).
Grégori Gerebtzoff and Frank Senner are at Drug Discov. Today 10, 663–671 (2005). 39. Kim, R. B. et al. Interrelationship between substrates
The Non-Clinical Safety Department, F. Hoffmann-La 17. Wiese, M. & Pajeva, I. K. in Comprehensive and inhibitors of human CYP3A and P-glycoprotein.
Roche, CH-4070 Basel, Switzerland. Medicinal Chemistry II Vol. 5 (eds Testa, B. & Pharm. Res. 16, 408–414 (1999).
Van de Waterbeemd, H.) 767–794 (Elsevier Science, 40. Kansy, M., Senner, F. & Gubernator, K.
Per Artursson and Hans Lennernaes are at the New York, 2006). Physicochemical high throughput screening:
Department of Pharmacy, Biomedical Centre, Uppsala 18. Banerjee, A., Johnston, J. S. & Swaan, P. W. in parallel artificial membrane permeation assay
Cellular Drug Delivery: Principles and Practice. Ch. 7 in the description of passive absorption processes.
University, S-752 63 Uppsala, BOX 580, Sweden.
(eds Lu, D. R. & Oie, S.) 107–128 (Humana Press, J. Med. Chem. 41, 1007–1010 (1998).
Alex Avdeef is at pION, 5 Constitution Way, New Jersey, 2004). 41. Avdeef, A., Kansy, M., Bendels, S. & Tsinman, K.
19. Lee, V. H. Membrane transporters. Eur. J. Pharm. Sci. Absorption–excipient–pH classification gradient
Woburn, Massachussetts 01801, USA.
11, S41–S50 (2000). maps: sparingly soluble drugs and the pH partition
Li Di is at the Pharmacokinetics, 20. Mueller, P., Rudin, D. O., Tien, H. T. & Wescott, W. C. hypothesis. Eur. J. Pharm. Sci. 33, 29–41 (2008).
Reconstitution of cell membrane structure in vitro and 42. Nielsen, P. E. & Avdeef, A. PAMPA — a drug
Dynamics and Metabolism Department, its transformation into an excitable system. Nature absorption in vitro model 8. Apparent filter porosity
Pfizer, Groton, Connecticut 06340, USA. 194, 979–980 (1962). and the unstirred water layer. Eur. J. Pharm. Sci. 22,
21. Saparov, S. M., Antonenko, Y. N. & Pohl, P. A new 33–41 (2004).
Gerhard F. Ecker is at the Department model of weak acid permeation through membranes 43. Sugano, K. in Comprehensive Medicinal Chemistry II
of Medicinal Chemistry, University of Vienna, revisited: does Overton still rule? Biophys. J. 90, Vol. 5 (eds Testa, B. & Van de Waterbeemd, H.)
Althanstrasse 141090 Wien, Austria. L86–L88 (2006). 453–487 (Elsevier, Oxford, 2007).
22. Walter, A. & Gutknecht, J. Monocarboxylic acid 44. Artursson, P. & Borchardt, R. T. Intestinal drug
Bernard Faller is at the Novartis Institutes permeation through lipid bilayer membranes. absorption and metabolism in cell cultures: Caco-2
for Biomedical Research, WSJ-350.3.04, J. Membr. Biol. 77, 255–264 (1984). and beyond. Pharm. Res. 14, 1655–1658 (1997).
CH-4002 Basel, Switzerland. 23. Walter, A. & Gutknecht, J. Permeability of small 45. Englund, G. et al. Regional levels of drug transporters
nonelectrolytes through lipid bilayer membranes. along the human intestinal tract: co-expression of ABC
Correspondence to M.K. and K.S. J. Membr. Biol. 90, 207–217 (1986). and SLC transporters and comparison with Caco-2
24. Paul, E. C., Hochman, J. & Quaroni, A. Conditionally cells. Eur. J. Pharm. Sci. 29, 269–277 (2006).
e-mails: manfred.kansy@roche.com;
immortalized intestinal epithelial cells: novel approach 46. Hilgendorf, C. et al. Expression of thirty-six drug
kiyohiko.sugano@pfizer.com for study of differentiated enterocytes. Am. J. Physiol. transporter genes in human intestine, liver, kidney,
doi:10.1038/nrd3187 265, C266–C278 (1993). and organotypic cell lines. Drug Metab. Dispos. 35,
25. Tavelin, S. et al. An improved cell culture model based 1333–1340 (2007).
1. Dobson, P. D. & Kell, D. B. Carrier-mediated cellular on 2/4/A1 cell monolayers for studies of intestinal 47. Hubatsch, I., Ragnarsson, E. G. & Artursson, P.
uptake of pharmaceutical drugs: an exception or drug transport: characterization of transport routes. Determination of drug permeability and prediction of
the rule? Nature Rev. Drug Discov. 7, 205–220 Pharm. Res. 20, 373–381 (2003). drug absorption in Caco-2 monolayers. Nature Protoc.
(2008). 26. Bermejo, M. et al. PAMPA — a drug absorption 2, 2111–2119 (2007).
2. Dobson, P. D., Lanthaler, K., Oliver, S. G. & Kell, D. B. in vitro model 7. Comparing rat in situ, Caco-2, 48. Neuhoff, S., Ungell, A.-L., Zamora, I. & Artursson, P.
Implications of the dominant role of transporters and PAMPA permeability of fluoroquinolones. pH-dependent bidirectional transport of weakly basic
in drug uptake by cells. Curr. Top. Med. Chem. 9, Eur. J. Pharm. Sci. 21, 429–441 (2004). drugs across Caco-2 monolayers: implications for
163–181 (2009). 27. Neuhoff, S., Ungell, A. L., Zamora, I. & Artursson, P. drug-drug interactions. Pharm. Res. 20, 1141–1148
3. Giacomini, K. M. et al. Membrane transporters in drug pH-dependent passive and active transport of acidic (2003).
development. Nature Rev. Drug Discov. 9, 215–236 drugs across Caco-2 cell monolayers. Eur. J. Pharm. 49. Pickett, S. D., McLay, I. M. & Clark, D. E. Enhancing
(2010). Sci. 25, 211–220 (2005). the hit-to-lead properties of lead optimization
4. Hoegerle, M. L. & Winne, D. Drug absorption by 28. Pade, V. & Stavchansky, S. Estimation of the relative libraries. J. Chem. Inf. Comput. Sci. 40, 263–272
the rat jejunum perfused in situ. Dissociation from contribution of the transcellular and paracellular (2000).
the pH-partition theory and role of microclimate-pH pathway to the transport of passively absorbed drugs 50. Feng, B. et al. In vitro P-glycoprotein assays to predict
and unstirred layer. Naunyn Schmiedebergs in the Caco-2 cell culture model. Pharm. Res. 14, the in vivo interactions of P-glycoprotein with drugs in
Arch. Pharmacol. 322, 249–255 (1983). 1210–1215 (1997). the central nervous system. Drug Metab. Dispos. 36,
5. Hogben, C. A., Tocco, D. J., Brodie, B. B. & 29. Takanaga, H., Tamai, I. & Tsuji, A. pH-dependent and 268–275 (2008).
Schanker, L. S. On the mechanism of intestinal carrier-mediated transport of salicylic acid across 51. Thiel-Demby, V. E. et al. Biopharmaceutics
absorption of drugs. J. Pharmacol. Exp. Ther. Caco-2 cells. J. Pharm. Pharmacol. 46, 567–570 classification system: validation and learnings of an
125, 275–282 (1959). (1994). in vitro permeability assay. Mol. Pharm. 6, 11–18
6. Corti, G., Maestrelli, F., Cirri, M., Zerrouk, N. & 30. Tsuji, A., Simanjuntak, M. T., Tamai, I. & Terasaki, T. (2009).
Mura, P. Development and evaluation of an in vitro pH-dependent intestinal transport of monocarboxylic 52. Volpe, D. A. Variability in Caco-2 and MDCK cell-based
method for prediction of human drug absorption. acids: carrier-mediated and H+-cotransport intestinal permeability assays. J. Pharm. Sci. 97,
Eur. J. Pharm. Sci. 27, 354–362 (2006). mechanism versus pH-partition hypothesis. J. Pharm. 712–725 (2008).
7. Zhu, C., Jiang, L., Chen, T. M. & Hwang, K. K. Sci. 79, 1123–1124 (1990). 53. Hayeshi, R. et al. Comparison of drug transporter
A comparative study of artificial membrane 31. Kopplow, K., Letschert, K., Konig, J., Walter, B. & gene expression and functionality in Caco-2 cells from
permeability assay for high throughput profiling of Keppler, D. Human hepatobiliary transport of organic 10 different laboratories. Eur. J. Pharm. Sci. 35,
drug absorption potential. Eur. J. Med. Chem. 37, anions analyzed by quadruple-transfected cells. 383–396 (2008).
399–407 (2002). Mol. Pharmacol. 68, 1031–1038 (2005). 54. Sugano, K., Takata, N., Machida, M., Saitoh, K. &
8. Cohen, B. E. & Bangham, A. D. Diffusion of small 32. Sasaki, M., Suzuki, H., Ito, K., Abe, T. & Sugiyama, Y. Terada, K. Prediction of passive intestinal absorption
nonelectrolytes across liposome membranes. Transcellular transport of organic anions across a using bio-mimetic artificial membrane permeation
Nature 236, 173–174 (1972). double-transfected Madin–Darby canine kidney II cell assay and the paracellular pathway model.
9. Xiang, T. X. & Anderson, B. D. Influence of chain monolayer expressing both human organic anion- Int. J. Pharm. 241, 241–251 (2002).
ordering on the selectivity of dipalmitoylphosphatidyl- transporting polypeptide (OATP2/SLC21A6) and 55. Avdeef, A. et al. Parallel artificial membrane
choline bilayer membranes for permeant size multidrug resistance-associated protein 2 (MRP2/ permeability assay (PAMPA)-critical factors for
and shape. Biophys. J. 75, 2658–2671 (1998). ABCC2). J. Biol. Chem. 277, 6497–6503 (2002). better predictions of absorption. J. Pharm. Sci.
10. Meier, M., Blatter, X. L., Seelig, A. & Seelig, J. 33. Lennernas, H. Human intestinal permeability. 96, 2893–2909 (2007).
Interaction of verapamil with lipid membranes and J. Pharm. Sci. 87, 403–410 (1998). 56. Galinis-Luciani, D., Nguyen, L. & Yazdanian, M. Is
P–glycoprotein: connecting thermodynamics 34. Lennernas, H. Intestinal permeability and its parallel artificial membrane permeability assay
and membrane structure with functional activity. relevance for absorption and elimination. a useful tool for discovery? J. Pharm. Sci. 96,
Biophys. J. 91, 2943–2955 (2006). Xenobiotica 37, 1015–1051 (2007). 2886–2892 (2007).

nATure revIeWs | Drug Discovery vOlume 9 | AugusT 2010 | 611

© 2010 Macmillan Publishers Limited. All rights reserved


PersPectives

57. Rautio, J., Laine, K. & Gynther, M. Enhanced brain 79. Obach, R. S., Lombardo, F. & Waters, N. J. Trend 102. Ottaviani, G., Martel, S. & Carrupt, P.-A. Parallel
drug delivery and targeting. Pharm. Technol. Eur. analysis of a database of intravenous pharmacokinetic artificial membrane permeability assay: a new
20, 27–29, 32–33 (2008). parameters in humans for 670 drug compounds. membrane for the fast prediction of passive human
58. Sugano, K., Nabuchi, Y., Machida, M. & Asoh, Y. Drug Metab. Dispos. 36, 1385–1405 (2008). skin permeability. J. Med. Chem. 49, 3948–3954
Permeation characteristics of a hydrophilic basic 80. Obata, K. et al. Prediction of oral drug absorption (2006).
compound across a bio-mimetic artificial membrane. in humans by theoretical passive absorption model. 103. Saitoh, R. et al. Correction of permeability with
Int. J. Pharm. 275, 271–278 (2004). Int. J. Pharm. 293, 183–192 (2005). pore radius of tight junctions in Caco-2 monolayers
59. Takagi, M. et al. A new interpretation of salicylic acid 81. Rodgers, T. & Rowland, M. Physiologically based improves the prediction of the dose fraction of
transport across the lipid bilayer: implications of pharmacokinetic modelling 2: predicting the tissue hydrophilic drugs absorbed by humans. Pharm. Res.
pH-dependent but not carrier-mediated absorption distribution of acids, very weak bases, neutrals and 21, 749–755 (2004).
from the gastrointestinal tract. J. Pharmacol. zwitterions. J. Pharm. Sci. 95, 1238–1257 (2006). 104. Sugano, K., Hamada, H., Machida, M. & Ushio, H.
Exp. Ther. 285, 1175–1180 (1998). 82. Rodgers, T. & Rowland, M. Mechanistic approaches to High throuput prediction of oral absorption:
60. Valko, K. & Reynolds, D. P. High-throughput volume of distribution predictions: understanding the improvement of the composition of the lipid solution
physicochemical and in vitro ADMET screening: processes. Pharm. Res. 24, 918–933 (2007). used in parallel artificial membrane permeation assay.
a role in pharmaceutical profiling. Am. J. Drug Deliv. 83. Veber, D. F. et al. Molecular properties that influence J. Biomol. Screen. 6, 189–196 (2001).
3, 83–100 (2005). the oral bioavailability of drug candidates. J. Med. 105. Sugano, K., Nabuchi, Y., Machida, M. & Aso, Y.
61. Gustafsson, D. et al. The direct thrombin inhibitor Chem. 45, 2615–2623 (2002). Prediction of human intestinal permeability using
melagatran and its oral prodrug H 376/95: 84. Winiwarter, S. et al. Correlation of human jejunal artificial membrane permeability. Int. J. Pharm.
intestinal absorption properties, biochemical and permeability (in vivo) of drugs with experimentally and 257, 245–251 (2003).
pharmacodynamic effects. Thromb. Res. 101, theoretically derived parameters. A multivariate data 106. Wohnsland, F. & Faller, B. High-throughput
171–181 (2001). analysis approach. J. Med. Chem. 41, 4939–4949 permeability pH profile and high-throughput alkane/
62. Sugano, K. et al. Quantitative structure–intestinal (1998). water log P with artificial membranes. J. Med. Chem.
permeability relationship of benzamidine analogue 85. Reynolds, D. P., Lanevskij, K., Japertas, P., 44, 923–930 (2001).
thrombin inhibitor. Bioorg. Med. Chem. Lett. 10, Didziapetris, R. & Petrauskas, A. Ionization-specific 107. Avdeef, A. The rise of PAMPA. Expert Opin. Drug
1939–1942 (2000). analysis of human intestinal absorption. J. Pharm. Sci. Metab. Toxicol. 1, 325–342 (2005).
63. Fischer, H., Kansy, M., Avdeef, A. & Senner, F. 98, 4039–4054 (2009). 108. Beigi, F. et al. Immobilized liposome and
Permeation of permanently positive charged 86. Fischer, H., Gottschlich, R. & Seelig, A. Blood–brain biomembrane partitioning chromatography of
molecules through artificial membranes — influence barrier permeation: molecular parameters governing drugs for prediction of drug transport. Int. J. Pharm.
of physico-chemical properties. Eur. J. Pharm. Sci. passive diffusion. J. Membr. Biol. 165, 201–211 164, 129–137 (1998).
31, 32–42 (2007). (1998). 109. Danelian, E. et al. SPR biosensor studies of the direct
64. Ruifrok, P. G. & Meijer, D. K. Transport of organic 87. Gratton, J. A., Abraham, M. H., Bradbury, M. W. & interaction between 27 drugs and a liposome surface:
ions through lipid bilayers. Naunyn Schmiedebergs Chadha, H. S. Molecular factors influencing drug correlation with fraction absorbed in humans. J. Med.
Arch. Pharmacol. 316, 266–272 (1981). transfer across the blood–brain barrier. J. Pharm. Chem. 43, 2083–2086 (2000).
65. Poirier, A. et al. Design, data analysis and simulation Pharmacol. 49, 1211–1216 (1997). 110. Loidl-Stahlhofen, A., Eckert, A., Hartmann, T. &
of in vitro drug transport kinetic experiments using a 88. Hou, T. J., Zhang, W., Xia, K., Qiao, X. B. & Xu, X. J. Schottner, M. Solid-supported lipid membranes as
mechanistic in vitro model. Drug Metab. Dispos. 36, ADME evaluation in drug discovery. 5. Correlation of a tool for determination of membrane affinity:
2434–2444 (2008). Caco-2 permeation with simple molecular properties. high-throughput screening of a physicochemical
66. Lei, Y. D., Wania, F., Shiu, W. Y. & Boocock, D. G. B. J. Chem. Inf. Comput. Sci. 44, 1585–1600 (2004). parameter. J. Pharm. Sci. 90, 599–606 (2001).
HPLC-based method for estimating the temperature 89. Palm, K. et al. Evaluation of dynamic polar molecular 111. Pidgeon, C. et al. IAM chromatography: an in vitro
dependence of n-Octanol–water partition coefficients. surface area as predictor of drug absorption: screen for predicting drug membrane permeability.
J. Chem. Eng. Data 45, 738–742 (2000). comparison with other computational and experimental J. Med. Chem. 38, 590–594 (1995).
67. Tanaka, M., Fukuda, H. & Nagai, T. Permeation predictors. J. Med. Chem. 41, 5382–5392 (1998). 112. Ano, R. et al. Relationships between structure and
of a drug through a model membrane consisting of 90. Zhao, Y. et al. Evaluation of human intestinal high-throughput screening permeability of peptide
millipore filter with oil. Chem. Pharm. Bull. 26, 9–13 absorption data and subsequent derivation of a derivatives and related compounds with artificial
(1978). quantitative structure–activity relationship (QSAR) membranes: application to prediction of Caco-2 cell
68. Raeissi, S. D., Li, J. & Hidalgo, I. J. The role of an with the Abraham descriptors. J. Pharm. Sci. 90, permeability. Bioorg. Med. Chem. 12, 257–264
α-amino group on H+-dependent transepithelial 749–784 (2001). (2004).
transport of cephalosporins in Caco-2 cells. J. Pharm. 91. Avdeef, A. in Absorption and Drug Development: 113. Fujikawa, M., Nakao, K., Shimizu, R. & Akamatsu, M.
Pharmacol. 51, 35–40 (1999). Solubility, Permeability and Charge State 1–312 QSAR study on permeability of hydrophobic
69. Takacs-Novak, K., Box, K. J. & Avdeef, A. (Wiley-Interscience, Hoboken, New Jersey, 2003). compounds with artificial membranes. Bioorg.
Potentiometric pKa determination of water-insoluble 92. Camenisch, G., Folkers, G. & van de Waterbeemd, H. Med. Chem. 15, 3756–3767 (2007).
compounds: validation study in methanol/water Comparison of passive drug transport through Caco-2 114. Kansy, M. et al. in Pharmacokinetic Optimization in
mixtures. Int. J. Pharm. 151, 235–248 (1997). cells and artificial membranes. Int. J. Pharm. 147, Drug Research: Biological, Physicochemical and
70. Said, H. M., Blair, J. A., Lucas, M. L. & Hilburn, M. E. 61–70 (1997). Computational Strategies (eds Testa, B., Van de
Intestinal surface acid microclimate in vitro and 93. Dagenais, C., Avdeef, A., Tsinman, O., Dudley, A. & Waterbeemd, H., Folkers, G. & Guy, R.) 447–464
in vivo in the rat. J. Lab. Clin. Med. 107, 420–424 Beliveau, R. P-glycoprotein deficient mouse in situ (Wiley-VCH, Zurich, 2001).
(1986). blood–brain barrier permeability and its prediction 115. Kubinyi, H. Drug partitioning: relationships between
71. Xiang, T. X. & Anderson, B. D. Substituent using an in combo PAMPA model. Eur. J. Pharm. Sci. forward and reverse rate constants and partition
contributions to the transport of substituted p-toluic 38, 121–137 (2009). coefficient. J. Pharm. Sci. 67, 262–263 (1978).
acids across lipid bilayer membranes. J. Pharm. Sci. 94. Di, L., Kerns, E. H., Bezar, I. F., Petusky, S. L. & 116. Kubinyi, H. Lipophilicity and drug activity. Prog. Drug
83, 1511–1518 (1994). Huang, Y. Comparison of blood–brain barrier Res. 23, 97–198 (1979).
72. Cohen, B. E. Permeability of liposomes to permeability assays: in situ brain perfusion, 117. Wenlock, M. C., Austin, R. P., Barton, P., Davis, A. M.
nonelectrolytes. II. Effect of nystatin and gramicidin A. MDR1–MDCKII and PAMPA–BBB. J. Pharm. Sci. & Leeson, P. D. A comparison of physiochemical
J. Membr. Biol. 20, 235–268 (1975). 98, 1980–1991 (2009). property profiles of development and marketed oral
73. Cao, Y., Xiang, T.-X. & Anderson, B. D. Development of 95. Di, L., Kerns, E. H., Fan, K., McConnell, O. J. & drugs. J. Med. Chem. 46, 1250–1256 (2003).
structure–lipid bilayer permeability relationships for Carter, G. T. High throughput artificial membrane 118. Balimane, P. V. et al. Peptide transporter substrate
peptide-like small organic molecules. Mol. Pharm. 5, permeability assay for blood–brain barrier. identification during permeability screening in drug
371–388 (2008). Eur. J. Med. Chem. 38, 223–232 (2003). discovery: comparison of transfected MDCK-hPepT1
74. Camenisch, G., Alsenz, J., van de Waterbeemd, H. 96. Flaten, G. E., Dhanikula, A. B., Luthman, K. & cells to Caco-2 cells. Arch. Pharm. Res. 30, 507–518
& Folkers, G. Estimation of permeability by passive Brandl, M. Drug permeability across a phospholipid (2007).
diffusion through Caco-2 cell monolayers using vesicle based barrier: a novel approach for studying 119. Balimane, P. V. et al. A novel high-throughput
the drugs’ lipophilicity and molecular weight. passive diffusion. Eur. J. Pharm. Sci. 27, 80–90 (2006). automated chip-based nanoelectrospray tandem mass
Eur. J. Pharm. Sci. 6, 313–319 (1998). 97. Kerns, E. H. et al. Combined application of parallel spectrometric method for PAMPA sample analysis.
75. Johnson, T. W., Dress, K. R. & Edwards, M. artificial membrane permeability assay and Caco-2 J. Pharm. Biomed. Anal. 39, 8–16 (2005).
Using the Golden Triangle to optimize clearance permeability assays in drug discovery. J. Pharm. Sci. 120. Di, L. & Kerns, E. H. Profiling drug-like properties
and oral absorption. Bioorg. Med. Chem. Lett. 19, 93, 1440–1453 (2004). in discovery research. Curr. Opin. Chem. Biol. 7,
5560–5564 (2009). 98. Kwon, J.-H. & Escher, B. I. A modified parallel artificial 402–408 (2003).
76. Linnankoski, J., Ranta, V.-P., Yliperttula, M. & Urtti, A. membrane permeability assay for evaluating the 121. Sugano, K., Obata, K., Saitoh, R., Higashida, A. &
Passive oral drug absorption can be predicted more bioconcentration of highly hydrophobic chemicals in Hamada, H. in Pharmacokinetic Profiling in Drug
reliably by experimental than computational models fish. Environ. Sci. Technol. 42, 1787–1793 (2008). Research (eds Testa, B., Krämer, S., Wunderli-
— fact or myth. Eur. J. Pharm. Sci. 34, 129–139 99. Li, C. et al. Correlation between PAMPA permeability Allenspach, H. & Folkers, G.) 441–458 (Wiley-VCH,
(2008). and cellular activities of hepatitis C virus protease Zurich, 2006).
77. Lombardo, F., Obach, R. S., Shalaeva, M. Y. & Gao, F. inhibitors. Biochem. Pharmacol. 75, 1186–1197 122. Masungi, C. et al. Parallel artificial membrane
Prediction of volume of distribution values in humans (2008). permeability assay (PAMPA) combined with a 10-day
for neutral and basic drugs using physicochemical 100. Matsson, P. et al. Exploring the role of different drug multiscreen Caco-2 cell culture as a tool for assessing
measurements and plasma protein binding data. transport routes in permeability screening. J. Med. new drug candidates. Pharmazie 63, 194–199
J. Med. Chem. 45, 2867–2876 (2002). Chem. 48, 604–613 (2005). (2008).
78. Martin, Y. C. A practitioner’s perspective of the role 101. Miret, S., Abrahamse, L. & de Groene, E. M. 123. Tavelin, S. et al. Prediction of the oral absorption of
of quantitative structure–activity analysis in Comparison of in vitro models for the prediction of low-permeability drugs using small intestine-like
medicinal chemistry. J. Med. Chem. 24, 229–237 compound absorption across the human intestinal 2/4/A1 cell. monolayers. Pharm. Res. 20, 397–405
(1981). mucosa. J. Biomol. Screen. 9, 598–606 (2004). (2003).

612 | AugusT 2010 | vOlume 9 www.nature.com/reviews/drugdisc

© 2010 Macmillan Publishers Limited. All rights reserved


PersPectives

124. Lagas, J. S., Vlaming, M. L. & Schinkel, A. H. 145. von Richter, O. et al. A novel screening strategy to 169. Korjamo, T., Kemilainen, H., Heikkinen, A. T. &
Pharmacokinetic assessment of multiple ATP-binding identify ABCB1 substrates and inhibitors. Naunyn Monkkonen, J. Decrease in intracellular concentration
cassette transporters: the power of combination Schmiedebergs Arch. Pharmacol. 379, 11–26 (2009). causes the shift in Km value of efflux pump substrates.
knockout mice. Mol. Interv. 9, 136–145 (2009). 146. Faller, B. in Hit and Lead Profiling. Identification and Drug Metab. Dispos. 35, 1574–1579 (2007).
125. van de Steeg, E. et al. Methotrexate pharmacokinetics Optimization of Drug-like Molecules. Methods and 170. Tachibana, T. et al. Model analysis of the
in transgenic mice with liver-specific expression of Principles in Medicinal Chemistry Vol. 43 (eds Faller, concentration-dependent permeability of P-gp
human organic anion-transporting polypeptide 1B1 B. & Laszlo, U.) (Wiley VCH, Weinheim 2009). substrates. Pharm. Res. 27, 442–446 (2010).
(SLCO1B1). Drug Metab. Dispos. 37, 277–281 147. Sun, H. & Pang, K. S. Permeability, transport, and 171. Trotter, P. J. & Storch, J. Fatty acid uptake and
(2009). metabolism of solutes in Caco-2 cell monolayers: metabolism in a human intestinal cell line (Caco-2):
126. Chen, C., Liu, X. & Smith, B. J. Utility of Mdr1-gene a theoretical study. Drug Metab. Dispos. 36, comparison of apical and basolateral incubation.
deficient mice in assessing the impact of P-glycoprotein 102–123 (2008). J. Lipid Res. 32, 293–304 (1991).
on pharmacokinetics and pharmacodynamics in drug 148. Eytan, G. D. Mechanism of multidrug resistance in 172. Troutman, M. D. & Thakker, D. R. Novel experimental
discovery and development. Curr. Drug Metab. 4, relation to passive membrane permeation. Biomed. parameters to quantify the modulation of absorptive
272–291 (2003). Pharmacother. 59, 90–97 (2005). and secretory transport of compounds by
127. Chiou, W. L., Chung, S. M. & Wu, T. C. Apparent lack 149. Wu, C.-Y. & Benet, L. Z. Predicting drug disposition via P-glycoprotein in cell culture models of intestinal
of effect of P-glycoprotein on the gastrointestinal application of BCS: transport/absorption/ elimination epithelium. Pharm. Res. 20, 1210–1224 (2003).
absorption of a substrate, tacrolimus, in normal mice. interplay and development of a biopharmaceutics 173. Troutman, M. D. & Thakker, D. R. Efflux ratio
Pharm. Res. 17, 205–208 (2000). drug disposition classification system. Pharm. Res. cannot assess P-glycoprotein-mediated attenuation
128. del Amo, E. M., Heikkinen, A. T. & Moenkkoenen, J. 22, 11–23 (2005). of absorptive transport: asymmetric effect of
In vitro–in vivo correlation in P-glycoprotein mediated 150. Fenner, K. S. et al. Drug–drug interactions mediated P-glycoprotein on absorptive and secretory transport
transport in intestinal absorption. Eur. J. Pharm. Sci. through P-glycoprotein: clinical relevance and across Caco-2 cell monolayers. Pharm. Res. 20,
36, 200–211 (2009). in vitro–in vivo correlation using digoxin as a probe 1200–1209 (2003).
129. Kato, Y. et al. Investigation of the role of oligopeptide drug. Clin. Pharmacol. Ther. 85, 173–181 (2009). 174. Bailey, D. G., Dresser, G. K., Leake, B. F. & Kim, R. B.
transporter PEPT1 and sodium/glucose cotransporter 151. Oh, D.-M., Curl, R. L., Yong, C.-S. & Amidon, G. L. Naringin is a major and selective clinical inhibitor
SGLT1 in intestinal absorption of their substrates Effect of micronization on the extent of drug of organic anion-transporting polypeptide 1A2
using small GTP-binding protein Rab8-null mice. absorption from suspensions in humans. Arch. Pharm. (OATP1A2) in grapefruit juice. Clin. Pharmacol. Ther.
Drug Metab. Dispos. 37, 602–607 (2009). Res. 18, 427–433 (1995). 81, 495–502 (2007).
130. Hironaka, T., Itokawa, S., Ogawara, K., Higaki, K. 152. Shaw, T. R. & Carless, J. E. Effect of particle size on 175. Yasui-Furukori, N., Uno, T., Sugawara, K. & Tateishi, T.
& Kimura, T. Quantitative evaluation of PEPT1 the absorption of digoxin. Eur. J. Clin. Pharmacol. Different effects of three transporting inhibitors,
contribution to oral absorption of cephalexin in rats. 7, 269–273 (1974). verapamil, cimetidine, and probenecid, on
Pharm. Res. 26, 40–50 (2009). 153. Yu, L. X. An integrated model for determining causes fexofenadine pharmacokinetics. Clin. Pharmacol. Ther.
131. Johnson, W. W. P-glycoprotein-mediated efflux as of poor oral drug absorption. Pharm. Res. 16, 77, 17–23 (2005).
a major factor in the variance of absorption and 1883–1887 (1999). 176. Kikuchi, A., Nozawa, T., Wakasawa, T., Maeda, T. &
distribution of drugs: modulation of chemotherapy 154. Ogihara, T. et al. What kinds of substrates show Tamai, I. Transporter-mediated intestinal absorption of
resistance. Methods Find. Exp. Clin. Pharmacol. 24, P-glycoprotein-dependent intestinal absorption? fexofenadine in rats. Drug Metab. Pharmacokinet. 21,
501–514 (2002). Comparison of verapamil with vinblastine. 308–314 (2006).
132. Brandsch, M., Knutter, I. & Bosse-Doenecke, E. Drug Metab. Pharmacokinet. 21, 238–244 (2006). 177. Tannergren, C. et al. Multiple transport mechanisms
Pharmaceutical and pharmacological importance 155. Nagahara, N., Tavelin, S. & Artursson, P. Contribution involved in the intestinal absorption and first-pass
of peptide transporters. J. Pharm. Pharmacol. 60, of the paracellular route to the pH-dependent extraction of fexofenadine. Clin. Pharmacol. Ther. 74,
543–585 (2008). epithelial permeability to cationic drugs. J. Pharm. Sci. 423–436 (2003).
133. Knutter, I. et al. Transport of angiotensin-converting 93, 2972–2984 (2004). 178. Matsushima, S., Maeda, K., Ishiguro, N., Igarashi, T.
enzyme inhibitors by H+/peptide transporters 156. Lafforgue, G. et al. Oral absorption of ampicillin: & Sugiyama, Y. Investigation of the inhibitory effects
revisited. J. Pharmacol. Exp. Ther. 327, 432–441 role of paracellular route vs. PepT1 transporter. of various drugs on the hepatic uptake of fexofenadine
(2008). Fund. Clin. Pharmacol. 22, 189–201 (2008). in humans. Drug Metab. Dispos. 36, 663–669
134. Brandsch, M. Transport of drugs by proton-coupled 157. Linnankoski, J. et al. Paracellular porosity and pore (2008).
peptide transporters: pearls and pitfalls. Expert Opin. size of the human intestinal epithelium in tissue and 179. Swift, B., Tian, X. & Brouwer, K. L. Integration of
Drug Metab. Toxicol. 5, 887–905 (2009). cell culture models. J. Pharm. Sci. 99 (Suppl. 4), preclinical and clinical data with pharmacokinetic
135. Schoenmakers, R. G., Stehouwer, M. C. & Tukker, J. J. 2166–2175 (2010). modeling and simulation to evaluate fexofenadine as
Structure–transport relationship for the intestinal 158. Eytan, G. D., Regev, R., Oren, G. & Assaraf, Y. G. a probe for hepatobiliary transport function. Pharm.
small-peptide carrier: is the carbonyl group of the The role of passive transbilayer drug movement in Res. 26, 1942–1951 (2009).
peptide bond relevant for transport? Pharm. Res. multidrug resistance and its modulation. J. Biol. Chem. 180. Matsushima, S. et al. The inhibition of human
16, 62–68 (1999). 271, 12897–12902 (1996). multidrug and toxin extrusion 1 is involved in
136. Cao, X. et al. Why is it challenging to predict intestinal 159. Fagerholm, U. The role of permeability in drug the drug–drug interaction caused by cimetidine.
drug absorption and oral bioavailability in human ADME/PK, interactions and toxicity — presentation Drug Metab. Dispos. 37, 555–559 (2009).
using rat model. Pharm. Res. 23, 1675–1686 (2006). of a permeability-based classification system (PCS) 181. Grahnen, A. Cimetidine bioavailability and variable
137. Chiou, W. L., Chung, S. M., Wu, T. C. & Ma, C. for prediction of ADME/PK in humans. Pharm. Res. renal clearance. Eur. J. Clin. Pharmacol. 27, 623–624
A comprehensive account on the role of efflux 25, 625–638 (2008). (1984).
transporters in the gastrointestinal absorption of 13 160. Kohl, C. Transporters — the view from industry. 182. Grahnen, A., Jameson, S., Loof, L., Tyllstrom, J. &
commonly used substrate drugs in humans. Int. J. Clin. Chem. Biodivers. 6, 1988–1999 (2009). Lindstrom, B. Pharmacokinetics of cimetidine in
Pharmacol. Ther. 39, 93–101 (2001). 161. Scherrmann, J. M. Transporters in absorption, advanced cirrhosis. Eur. J. Clin. Pharmacol. 26,
138. Kwon, H., Lionberger, R. A. & Yu, L. X. Impact of distribution, and elimination. Chem. Biodivers. 6, 347–355 (1984).
P-glycoprotein-mediated intestinal efflux kinetics 1933–1942 (2009). 183. Grahnen, A., Von Bahr, C., Lindstroem, B. & Rosen, A.
on oral bioavailability of P-glycoprotein substrates. 162. Shugarts, S. & Benet, L. Z. The role of transporters Bioavailability and pharmacokinetics of cimetidine.
Mol. Pharm. 1, 455–465 (2004). in the pharmacokinetics of orally administered drugs. Eur. J. Clin. Pharmacol. 16, 335–340 (1979).
139. Amidon, G. L., Lennernas, H., Shah, V. P. & Pharm. Res. 26, 2039–2054 (2009). 184. Collett, A., Higgs, N. B., Sims, E., Rowland, M. &
Crison, J. R. A theoretical basis for a biopharmaceutic 163. Petri, N., Tannergren, C., Rungstad, D. & Warhurst, G. Modulation of the permeability of H2
drug classification: the correlation of in vitro drug Lennernaes, H. Transport characteristics of receptor antagonists cimetidine and ranitidine by
product dissolution and in vivo bioavailability. fexofenadine in the Caco-2 cell model. Pharm. Res. P-glycoprotein in rat intestine and the human colonic
Pharm. Res. 12, 413–420 (1995). 21, 1398–1404 (2004). cell line Caco-2. J. Pharmacol. Exp. Ther. 288,
140. Lin, J. H. & Yamazaki, M. Role of P-glycoprotein 164. Cvetkovic, M., Leake, B., Fromm, M. F., Wilkinson, G. R. 171–178 (1999).
in pharmacokinetics: clinical implications. & Kim, R. B. OATP and P-glycoprotein transporters 185. Sun, D. et al. Comparison of human duodenum and
Clin. Pharmacokinet. 42, 59–98 (2003). mediate the cellular uptake and excretion of Caco-2 gene expression profiles for 12,000 gene
141. Cao, X. et al. Permeability dominates in vivo intestinal fexofenadine. Drug Metab. Dispos. 27, 866–871 sequences tags and correlation with permeability
absorption of P-gp substrate with high solubility (1999). of 26 drugs. Pharm. Res. 19, 1400–1416 (2002).
and high permeability. Mol. Pharm. 2, 329–340 165. Tannergren, C., Knutson, T., Knutson, L. & 186. Sun, W., Wu, R. R., van Poelje, P. D. & Erion, M. D.
(2005). Lennernas, H. The effect of ketoconazole on the Isolation of a family of organic anion transporters
142. Doppenschmitt, S., Spahn-Langguth, H., Regardh, C. G. in vivo intestinal permeability of fexofenadine using a from human liver and kidney. Biochem. Biophys. Res.
& Langguth, P. Role of P-glycoprotein-mediated regional perfusion technique. Br. J. Clin. Pharmacol. Commun. 283, 417–422 (2001).
secretion in absorptive drug permeability: 55, 182–190 (2003). 187. Takamatsu, N. et al. Human jejunal permeability
an approach using passive membrane permeability 166. Glaeser, H. et al. Intestinal drug transporter of two polar drugs: cimetidine and ranitidine.
and affinity to P-glycoprotein. J. Pharm. Sci. 88, expression and the impact of grapefruit juice in Pharm. Res. 18, 742–744 (2001).
1067–1072 (1999). humans. Clin. Pharmacol. Ther. 81, 362–370 (2007). 188. Jantratid, E. et al. Biowaiver monographs for
143. Lentz, K. A., Polli, J. W., Wring, S. A., Humphreys, J. E. 167. Russel, T., Stoltz, M. & Weir, S. Pharmacokinetics, immediate release solid oral dosage forms: cimetidine.
& Polli, J. E. Influence of passive permeability on pharmacodynamics, and tolerance of single- and J. Pharm. Sci. 95, 974–984 (2006).
apparent P-glycoprotein kinetics. Pharm. Res. 17, multiple-dose fexofenadine hydrochloride in healthy 189. Christians, U., Jacobsen, W., Benet, L. Z. & Lampen, A.
1456–1460 (2000). male volunteers. Clin. Pharmacol. Ther. 64, 612–621 Mechanisms of clinically relevant drug interactions
144. Varma, M. V., Sateesh, K. & Panchagnula, R. (1998). associated with tacrolimus. Clin. Pharmacokinet. 41,
Functional role of P-glycoprotein in limiting intestinal 168. Lappin, G. et al. Pharmacokinetics of fexofenadine: 813–851 (2002).
absorption of drugs: contribution of passive evaluation of a microdose and assessment of absolute 190. Floren, L. C. et al. Tacrolimus oral bioavailability
permeability to P-glycoprotein mediated efflux oral bioavailability. Eur. J. Pharm. Sci. 40, 125–131 doubles with coadministration of ketoconazole.
transport. Mol. Pharm. 2, 12–21 (2005). (2010). Clin. Pharmacol. Ther. 62, 41–49 (1997).

nATure revIeWs | Drug Discovery vOlume 9 | AugusT 2010 | 613

© 2010 Macmillan Publishers Limited. All rights reserved


PersPectives

191. Fricker, G., Drewe, J., Huwyler, J., Gutmann, H. & 209. Adachi, Y., Suzuki, H. & Sugiyama, Y. Comparative 227. Han, H. et al. 5′-amino acid esters of antiviral
Beglinger, C. Relevance of P-glycoprotein for the enteral studies on in vitro methods for evaluating in vivo nucleosides, acyclovir, and AZT are absorbed by the
absorption of cyclosporin A: in vitro–in vivo correlation. function of MDR1 P-glycoprotein. Pharm. Res. 18, intestinal PEPT1 peptide transporter. Pharm. Res.
Br. J. Pharmacol. 118, 1841–1847 (1996). 1660–1668 (2001). 15, 1154–1159 (1998).
192. Glavinas, H., Krajcsi, P., Cserepes, J. & Sarkadi, B. 210. Shitara, Y., Sato, H. & Sugiyama, Y. Evaluation of 228. Mealey, K. L., Bentjen, S. A., Gay, J. M. &
The role of ABC transporters in drug resistance, drug–drug interaction in the hepatobiliary and renal Cantor, G. H. Ivermectin sensitivity in collies is
metabolism, and toxicity. Curr. Drug Deliv. 1, 27–42 transport of drugs. Annu. Rev. Pharmacol. Toxicol. associated with a deletion mutation of the mdr1
(2004). 45, 689–723 (2005). gene. Pharmacogenetics 11, 727–733 (2001).
193. Kruijtzer, C. M., Beijnen, J. H. & Schellens, J. H. 211. Maeda, K., Suzuki, H. & Sugiyama, Y. Hepatic 229. Testa, B. in Comprehensive Medicinal Chemistry II
Improvement of oral drug treatment by temporary transport. Methods Princ. Med. Chem. 40, 277–332 Vol. 5 (eds. Testa, B. & van de Waterbeemd, H.)
inhibition of drug transporters and/or cytochrome (2009). 1009–1041 (Elsevier Science, New York, 2006).
P450 in the gastrointestinal tract and liver: 212. Hosea, N. A. et al. Prediction of human 230. Watanabe, T., Kusuhara, H., Maeda, K., Shitara, Y.
an overview. Oncologist 7, 516–530 (2002). pharmacokinetics from preclinical information: & Sugiyama, Y. Physiologically based
194. Kunta, J. R. & Sinko, P. J. Intestinal drug transporters: comparative accuracy of quantitative prediction pharmacokinetic modeling to predict transporter-
in vivo function and clinical importance. Curr. Drug approaches. J. Clin. Pharmacol. 49, 513–533 (2009). mediated clearance and distribution of pravastatin in
Metab. 5, 109–124 (2004). 213. Seelig, A. How does P-glycoprotein recognize its humans. J. Pharmacol. Exp. Ther. 328, 652–662
195. Lown, K. S. et al. Role of intestinal P-glycoprotein substrates? Int. J. Clin. Pharmacol. Ther. 36, 50–54 (2009).
(mdr1) in interpatient variation in the oral (1998). 231. Winterhalter, M. Black lipid membranes. Curr. Opin.
bioavailability of cyclosporine. Clin. Pharmacol. Ther. 214. Seelig, A. & Landwojtowicz, E. Structure–activity Colloid Interface Sci. 5, 250–255 (2000).
62, 248–260 (1997). relationship of P-glycoprotein substrates and 232. Balon, K., Riebesehl, B. U. & Muller, B. W. Drug
196. Benet, L. Z., Cummins, C. L. & Wu, C. Y. Unmasking modifiers. Eur. J. Pharm. Sci. 12, 31–40 (2000). liposome partitioning as a tool for the prediction
the dynamic interplay between efflux transporters and 215. Maher Doan, K. M. et al. Passive permeability and of human passive intestinal absorption. Pharm. Res.
metabolic enzymes. Int. J. Pharm. 277, 3–9 (2004). P-glycoprotein-mediated efflux differentiate central 16, 882–888 (1999).
197. Ito, K., Kusuhara, H. & Sugiyama, Y. Effects of nervous system (CNS) and non-CNS marketed drugs. 233. Harrigan, P. R., Wong, K. F., Redelmeier, T. E.,
intestinal CYP3A4 and P-glycoprotein on oral drug J. Pharmacol. Exp. Ther. 303, 1029–1037 (2002). Wheeler, J. J. & Cullis, P. R. Accumulation of
absorption — theoretical approach. Pharm. Res. 16, 216. Adachi, Y., Suzuki, H. & Sugiyama, Y. Quantitative doxorubicin and other lipophilic amines into large
225–231 (1999). evaluation of the function of small intestinal unilamellar vesicles in response to transmembrane pH
198. Bardelmeijer, H. A. et al. Entrapment by Cremophor P-glycoprotein: comparative studies between in situ gradients. Biochim. Biophys. Acta 1149, 329–338
EL decreases the absorption of paclitaxel from the and in vitro. Pharm. Res. 20, 1163–1169 (2003). (1993).
gut. Cancer Chemother. Pharmacol. 49, 119–125 217. Poulin, P. & Theil, F. P. A priori prediction of 234. Yan, E. C. & Eisenthal, K. B. Effect of cholesterol
(2002). tissue:plasma partition coefficients of drugs on molecular transport of organic cations across
199. Stephens, R. H. et al. Resolution of P-glycoprotein and to facilitate the use of physiologically-based liposome bilayers probed by second harmonic
non-P-glycoprotein effects on drug permeability using pharmacokinetic models in drug discovery. generation. Biophys. J. 79, 898–903 (2000).
intestinal tissues from mdr1a-/- mice. Br. J. Pharmacol. J. Pharm. Sci. 89, 16–35 (2000). 235. Xiang, T. X. & Anderson, B. D. The relationship
135, 2038–2046 (2002). 218. Schinkel, A. H. et al. Disruption of the mouse mdr1a between permeant size and permeability in lipid
200. Peltier, S., Oger, J. M., Lagarce, F., Couet, W. & P-glycoprotein gene leads to a deficiency in the bilayer membranes. J. Membr. Biol. 140, 111–122
Benoit, J. P. Enhanced oral paclitaxel bioavailability blood–brain barrier and to increased sensitivity to (1994).
after administration of paclitaxel-loaded lipid drugs. Cell 77, 491–502 (1994). 236. Ahlin, G. et al. Structural requirements for drug
nanocapsules. Pharm. Res. 23, 1243–1250 (2006). 219. Shitara, Y., Horie, T. & Sugiyama, Y. Transporters as a inhibition of the liver specific human organic cation
201. Walle, U. K. & Walle, T. Taxol transport by human determinant of drug clearance and tissue distribution. transport protein. J. Med. Chem. 51, 5932–5942
intestinal epithelial Caco-2 cells. Drug Metab. Dispos. Eur. J. Pharm. Sci. 27, 425–446 (2006). (2008).
26, 343–346 (1998). 220. Ieiri, I., Higuchi, S. & Sugiyama, Y. Genetic 237. Pedersen, J. M. et al. Prediction and identification
202. Sandstrom, M., Simonsen, L. E., Freijs, A. & polymorphisms of uptake (OATP1B1, 1B3) and efflux of drug interactions with the human ATP-binding
Karlsson, M. O. The pharmacokinetics of epirubicin (MRP2, BCRP) transporters: implications for inter- cassette transporter multidrug-resistance associated
and docetaxel in combination in rats. Cancer individual differences in the pharmacokinetics and protein 2 (MRP2; ABCC2). J. Med. Chem. 51,
Chemother. Pharmacol. 44, 469–474 (1999). pharmacodynamics of statins and other clinically 3275–3287 (2008).
203. Strum, W. B. A pH-dependent, carrier-mediated relevant drugs. Expert Opin. Drug Metab. Toxicol. 5, 238. Palm, K., Stenberg, P., Luthman, K. & Artursson, P.
transport system for the folate analog, amethopterin, 703–729 (2009). Polar molecular surface properties predict the
in rat jejunum. J. Pharmacol. Exp. Ther. 203, 640–645 221. Chou, C., McLachlan, A. J. & Rowland, M. intestinal absorption of drugs in humans. Pharm. Res.
(1977). Membrane permeability and lipophilicity in the 14, 568–571 (1997).
204. Yokooji, T., Mori, N. & Murakami, T. Site-specific isolated perfused rat liver: 5-ethyl barbituric acid and 239. Chen, X., Murawski, A., Patel, K., Crespi, C. L. &
contribution of proton-coupled folate transporter/ other compounds. J. Pharmacol. Exp. Ther. 275, Balimane, P. V. A novel design of artificial membrane
haem carrier protein 1 in the intestinal absorption 933–940 (1995). for improving the parallel artificial membrane
of methotrexate in rats. J. Pharm. Pharmacol. 61, 222. Huang, L. et al. Relationship between passive permeability assay model. Pharm. Res. 25,
911–918 (2009). permeability, efflux, and predictability of clearance 1511–1520 (2008).
205. Lennernas, H. et al. The effect of l-leucine on the from in vitro metabolic intrinsic clearance.
absorption of levodopa, studied by regional jejunal Drug Metab. Dispos. 38, 223–231 (2009).
Acknowledgements
perfusion in man. Br. J. Clin. Pharmacol. 35, 223. Fagerholm, U. Prediction of human pharmacokinetics
This paper is the joint effort of several colleagues from
243–250 (1993). — renal metabolic and excretion clearance. J. Pharm.
academia and the pharmaceutical industry working on a
206. Biegel, A. et al. Three-dimensional quantitative Pharmacol. 59, 1463–1471 (2007).
topic of common interest. We have to thank all involved
structure–activity relationship analyses of b-lactam 224. Abrahamsson, B. & Lennernaes, H. Application of the
colleagues for rapid feedback to support this task. We have
antibiotics and tripeptides as substrates of the biopharmaceutics classification system now and in
to thank H. Kubinyi for comments at the initial phase of this
mammalian H+/peptide cotransporter PEPT1. the future. Methods Princ. Med. Chem. 40, 523–558
activity.
J. Med. Chem. 48, 4410–4419 (2005). (2009).
207. Tannergren, C., Bergendal, A., Lennernas, H. & 225. Balimane, P. V., Han, Y. H. & Chong, S. Current
Competing interests statement
Abrahamsson, B. Toward an increased understanding industrial practices of assessing permeability and
The authors declare no competing financial interests.
of the barriers to colonic drug absorption in humans: P-glycoprotein interaction. AAPS J. 8, E1–13
implications for early controlled release candidate (2006).
assessment. Mol. Pharm. 6, 60–73 (2009). 226. Ganapathy, M. E., Huang, W., Wang, H., Ganapathy, V.
SUppLEMEnTARY inFORMATiOn
208. Skold, C. et al. Presentation of a structurally diverse & Leibach, F. H. Valacyclovir: a substrate for the
see online article: s1 (table)
and commercially available drug data set for intestinal and renal peptide transporters PEPT1
correlation and benchmarking studies. J. Med. Chem. and PEPT2. Biochem. Biophys. Res. Commun. 246, All liNks Are AcTive iN THe oNliNe PDF
49, 6660–6671 (2006). 470–475 (1998).

614 | AugusT 2010 | vOlume 9 www.nature.com/reviews/drugdisc

© 2010 Macmillan Publishers Limited. All rights reserved

S-ar putea să vă placă și