Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Advanced Drug Delivery Reviews 65 (2013) 152–168

Contents lists available at SciVerse ScienceDirect

Advanced Drug Delivery Reviews


journal homepage: www.elsevier.com/locate/addr

Modeling the human skin barrier — Towards a better understanding of


dermal absorption☆
Owen G. Jepps a, b, Yuri Dancik e, Yuri G. Anissimov a, b, Michael S. Roberts c, d,⁎
a
Griffith University, School of Biomolecular and Physical Sciences, Brisbane, QLD, Australia
b
Griffith University, Queensland Micro- and Nanotechnology Centre, Brisbane, QLD, Australia
c
The University of Queensland, School of Medicine, Therapeutics Research Centre, Brisbane, QLD, Australia
d
The University of South Australia, Therapeutics Research Centre, School of Pharmacy and Medical Science, Adelaide, SA, Australia
e
Proctor & Gamble Eurocor, Temselaan 100, Strombeek-Bever, Belgium

a r t i c l e i n f o a b s t r a c t

Article history: Many drugs are presently delivered through the skin from products developed for topical and transdermal
Accepted 9 April 2012 applications. Underpinning these technologies are the interactions between the drug, product and skin that
Available online 14 April 2012 define drug penetration, distribution, and elimination in and through the skin. Most work has been focused
on modeling transport of drugs through the stratum corneum, the outermost skin layer widely recognized as
Keywords:
presenting the rate-determining step for the penetration of most compounds. However, a growing body of
Modeling
Human skin
literature is dedicated to considering the influence of the rest of the skin on drug penetration and distribu-
Barrier tion. In this article we review how our understanding of skin physiology and the experimentally observed
Dermal absorption mechanisms of transdermal drug transport inform the current models of drug penetration and distribution
Transdermal drug transport in the skin. Our focus is on models that have been developed to describe particular phenomena observed
Transdermal penetration at particular sites of the skin, reflecting the most recent directions of investigation.
Distribution © 2012 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
2. Overview of skin transport modeling. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
3. Contributions to drug penetration and distribution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
3.1. Stratum corneum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
3.1.1. Morphology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
3.1.2. Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
3.1.3. Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
3.2. Viable epidermis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
3.2.1. Morphology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
3.2.2. Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
3.2.3. Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
3.3. Dermis and deeper layers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
3.3.1. Morphology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
3.3.2. Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
3.3.3. Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
3.4. Appendages . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
3.4.1. Morphology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
3.4.2. Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160
3.4.3. Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160
4. Global perspective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160

☆ This review is part of the Advanced Drug Delivery Reviews theme issue on “Modeling the human skin barrier — Towards a better understanding of dermal absorption”.
⁎ Corresponding author at: Therapeutics Research Centre, School of Medicine, University of Queensland, Princess Alexandra, Hospital, Brisbane, QLD 4120, Australia. Fax: + 61 7
3240 5806.
E-mail address: m.roberts@uq.edu.au (M.S. Roberts).

0169-409X/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
doi:10.1016/j.addr.2012.04.003
O.G. Jepps et al. / Advanced Drug Delivery Reviews 65 (2013) 152–168 153

4.1. Overall distribution models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160


4.2. Prodrugs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
4.3. In vivo vs. in vitro results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164
4.4. Clinical response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164
5. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
Acknowledgment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
Appendix A. Summary of equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
165

1. Introduction Drug transport through the skin is primarily considered to be by


diffusion — the ‘random walk’ of drug molecules through the various
Transdermal drug delivery is an important delivery mode. Over structures within the skin. This walk can be mediated by active trans-
one third of drugs under clinical evaluation involve delivery into or port (whereby protein transporters facilitate drug transport in certain
via the skin [1]: this figure includes both drugs whose sites of action environments) and convective transport (whereby molecules are
are in the skin, and those meant to be taken up by the systemic circu- driven by local currents in the lymphatic or vascular systems or inter-
lation. The advantages of transdermal delivery include the avoidance stitial spaces). Uptake into the lymphatic or vascular systems impacts
of the pain, possible infection and compliance issues related to injec- upon drug distribution, as well as facilitating drug clearance. Apart
tions; the possibility to control drug release over long periods; and from clearance, drug may also be effectively removed through metab-
avoiding the first-pass metabolism via the oral route that can lead olism in the skin.
to large loss of drug [2,3]. The challenges include the significant vari- When modeling skin transport processes quantitatively, we typi-
ability in penetration between people [4], and between sites on an cally consider the various physiological regions of interest (potential-
individual [4]; and the considerable effectiveness of the skin barrier ly including skin and deeper tissue layers, appendages, the circulation
(particularly its uppermost layer, the stratum corneum) in limiting and other tissues) as compartments. Drug levels within the compart-
drug penetration. Predicting the permeability of a given compound ment can be modeled as a single time-dependent value c(t) (the
through the skin is in general quite difficult, due to the highly com- traditional compartment model) or as function c(x,t) of both position
plex nature of the structures and mechanisms that constitute the de- within that compartment (usually skin depth) as well as time. For the
livery pathway. latter choice, drug transport within the compartment may be mod-
In this article we will review the current understanding of the eled using partial differential equations that describe the effects of
structures and mechanisms that most affect transdermal drug deliv- drug diffusion, convection, elimination and metabolism. Drug trans-
ery and the distribution of drugs within the skin. Our focus will be port between compartments, as well as drug elimination or metabo-
on the transport and distribution of the solute drugs themselves: lism, can be described using rate constants. For a comprehensive
we will not consider the delivery of nanoparticles or vesicles, which review of mathematical models related to transdermal drug delivery,
has been reviewed elsewhere [5,6]. We begin with a brief overview we refer the reader to Ref. [9].
of the skin transport processes (that we will consider in more detail One of the principles behind such models is that the parameters
in Section 3 from a physiological perspective), and consider the describing the various transport processes – diffusion, convection,
main quantities and mathematical representations. We then investi- elimination, etc. – are expected to depend on the drug under investi-
gate in greater detail the various contributions to this overall picture. gation, but to remain independent of the dose regimen. From a
Given the importance of skin physiology in determining drug transport mathematical viewpoint it is easiest to establish model parameters
and distribution, we group these contributions according to skin mor- for drug delivery processes that have reached a (pseudo) steady
phology (the stratum corneum; viable epidermis; dermis and deeper state — that is, a state where drug concentrations in the various com-
layers; and appendages). In each section we will consider the various partments remain essentially constant (until, for example, the deple-
transport mechanisms that arise, how their effects are modeled, and tion of the administered dose of drug becomes significant). Once
their overall contribution to percutaneous penetration. We then consid- these model parameters have been determined for a dose regimen
er how these contributions are combined to form comprehensive pen- that produces a steady state, the same model parameters can then
etration and distribution models, as well as other important topics, be used to explore the drug transport and distribution behavior for
before concluding. other doses.
Quantities that can be established experimentally, and that relate
directly to the drug transport and distribution in skin, include the
2. Overview of skin transport modeling steady-state flux of drug JSS into the skin (and related quantities
such as the permeability coefficient kp, which represents the constant
Drugs applied to the skin surface enter the skin through either the of proportionality between the steady-state flux and drug concentra-
uppermost layer of the epidermis – the Stratum Corneum (SC) – or tion C in the applied dose, i.e. JSS = kpC; and the maximum steady-
one of the appendages (including hair follicles and ducts leading to state flux Jmax, obtained when C is maximal, and thus equal to the
eccrine sweat glands) [7,8] (Fig. 1). Passage through the SC leads to a solubility S of the drug in the applied vehicle, i.e. Jmax = kpS); the lag
further series of skin layers — the viable epidermis, the dermis and the time between drug application and attainment of the steady state;
hypodermis/subcutaneous tissue. Permeation into the vascularized the clearance of drug through excretion; and drug concentrations in
dermis (which also houses the lymphatic pathways) and deeper tissues the various skin layers, circulation or other tissues. We note that
offers the possibility of systemic drug distribution. The heavily vascular- this final set has traditionally only been ascertained invasively, al-
ized nature of the hair follicle and appendageal glands also suggests the though modern microscopy techniques now allow the possibility of
possibility of systemic drug distribution via those routes. While the SC in situ concentration estimation [10–12].
barrier, which serves as the major rate-limiting component to skin pen- The influence of specific molecular-scale mechanisms (such as ac-
etration for most drugs, has been the focus of most penetration models, tive transport) on the overall transdermal transport of a drug is often
much less is known about transport beyond the SC, and about transport contained within these compartment-scale model parameters. As un-
via the appendageal routes. derstanding of these processes improves, the sophistication of models
154 O.G. Jepps et al. / Advanced Drug Delivery Reviews 65 (2013) 152–168

Fig. 1. Schematic overview of the skin layers, appendages, blood and lymphatic vessels.

used to describe drug penetration and distribution increases, as does involucrin, and filagrin. The intercellular lipid matrix chiefly com-
our understanding of the dependence of model parameters on drug prises a mixture of ceramides, cholesterol, triglycerides and fatty
properties. This in turn facilitates better prediction of transdermal acids, arranged into lipid lamellae. The relative proportions of these
drug transport, which is the ultimate aim of such work. compounds and their precise physical structure are matters of ongo-
ing research (cf, for example, [16,17]).
3. Contributions to drug penetration and distribution The lower epidermal layers are thus usually referred to as the via-
ble epidermis, in contradistinction to the cells of the SC that are enu-
Given the long-established importance of the SC as the main pen- cleated during differentiation [18,19]. This ‘bricks and mortar’ [20]
etration barrier for many drugs [13], it is important to recognize that structure is primarily responsible for the SC's effectiveness as the
the skin layers below the SC can also play important roles in the pen- main physicochemical barrier, both against the permeation of exoge-
etration and distribution of topically administered drugs. Lower skin nous compound into the skin, as well as endogenous compounds out
layers are important, for example, when considering the delivery of of the skin. It is worth noting that imperfections in the SC structure
lipophilic drugs, or drugs targeting the dermis, deeper tissues, or (such as within the cell structure, or lacunae of water found within
the circulation. In this section, we consider the various layers of the the lipid layers [21]) are likely to impact upon transport properties,
skin and the appendages, outlining the local transport effects that im- although it is not clear to what extent.
pact upon drug penetration and distribution, and the way in which
they can be modeled. 3.1.2. Mechanisms
Transport across the SC is essentially by passive diffusion, whether
3.1. Stratum corneum it be via an intercellular (i.e. restricted to the lipid matrix) or transcel-
lular (i.e. through both corneocytes and lipid matrix) pathway. Perme-
3.1.1. Morphology ation along such pathways is thus contingent upon a compound's
The SC is the outermost layer of the epidermis. Cells migrate from affinity with lipid environments, with the internal environment of
the dermal–epidermal junction at the base of the epidermis to the SC the corneocyte, and upon the capacity of a molecule to permeate the
over a period of approximately 2 weeks during a process known as corneocyte cell wall. The relative importance of these pathways has re-
desquamation [14]. The final stages of desquamation involve the dif- cently become a renewed focus of investigation, both experimentally
ferentiation of epidermal keratinocytes into flat, closely packed inter- and from theoretical considerations. Despite initial support for the
digitated corneocytes, embedded in a highly organized, dense lipid transcellular route [22], early experimental studies indicated that the
matrix [15,16] that is shed from the differentiating keratinocytes. intercellular lipid pathway was the dominant for SC permeation
The corneocytes comprise a dense scaffolding network of keratin fila- [23–26]. More recent microscopy evidence supports the view that the
ments believed to support the corneocyte shape; a range of hygro- transcellular route is an important one [27], as do results from recent
scopic compounds collectively termed Natural Moisturizing Factor SC transport models (Section 3.1.3). Scheuplein suggested different
(NMF) that help maintain SC hydration; and water-insoluble cell pathways for molecules, depending on their polarity [28], a view sup-
walls made up of highly cross-linked proteins including loricrin, ported by early experimental evidence [29,30].
O.G. Jepps et al. / Advanced Drug Delivery Reviews 65 (2013) 152–168 155

3.1.3. Modeling 3.2. Viable epidermis


One of the simplest mathematical models of drug transport
through the SC can be formulated as a very large well-stirred donor 3.2.1. Morphology
phase applied to the top of SC (that is initially devoid of drug), with The viable epidermis (VE) is an avascular environment, consisting
rapid clearance at the lower SC boundary (for example, due to the mainly of keratinocytes and comprised of approximately 40% protein,
phase beneath the SC being maintained at an effective zero drug con- 40% water and 15–20% lipids [19,42,43]. The undulating epidermal–
centration). If the SC is represented as a homogeneous membrane, dermal junction consists of papilla that project into the dermis [19] —
transport can be described by the diffusion equation (Fick's 2nd law) cells in the basal layer of the epidermis form the most important struc-
tural and functional link to the dermis below [43]. We recall that
∂C ðx; t Þ ∂2 C ðx; t Þ the epidermal cells undergo continual migration from the epidermal–
¼ DSC ð1:1Þ dermal junction up toward the SC during desquamation.
∂t ∂x2

together with the initial and boundary conditions: 3.2.2. Mechanisms


Drug diffusion in the VE can be considered as diffusion through an
C ðx; 0Þ ¼ 0; C ð0; t Þ ¼ K SC C d ; C ðhSC ; t Þ ¼ 0 ð1:2Þ aqueous medium that is hindered by the significant presence of pro-
teins and punctuated by cell membrane crossings. The high fraction of
where C(x, t) is the concentration at depth x and time t, D is the diffu- water present in the VE makes this region a more effective barrier
sion coefficient, KSC is the partitioning coefficient, Cd is the concentra- against lipophilic compounds, which have greater affinity for non-
tion in the donor phase and hSC is the thickness of the SC. The solution polar environments [13]. Wenkers and Lippold suggested that the
to Eq. (1.1) with initial and boundary conditions given by Eq. (1.2) is VE provides a decisive barrier to the penetration of NSAIDs from a
most commonly expressed as the amount of solute penetrating SC (Q) lipophilic vehicle, based on correlations between skin permeability
vs. time (t) [31]: and octanol–vehicle and PBS–vehicle partition coefficients [44].
Other VE processes that impact upon the penetration and distribu-
!! tion of drug within skin include drug binding and sequestration within
DSC 1 2X ∞
ð−1Þn D 2 2
Q ðt Þ ¼ K SC AC d hSC t− − exp −t 2SC π n ð1:3Þ the VE, drug metabolism and active transport. The detailed influence of
2
hSC 6 π n¼1 n2 hSC drug binding, sequestration and metabolism on drug penetration and
distribution is not always well established, although it is generally
where A is the area of solute application to SC. The summation term in understood that they can limit the extent to which drug is available to
2
this equation can be ignored at long times (practically at t > hSC /(2D)) permeate into lower regions. Bhatt et al. observed higher than expected
yielding the equation that is most commonly applied to experimental amounts of the lipophilic pesticide tecnazene remaining in skin tissue
data in the transdermal penetration literature: after a 48 h in vitro skin penetration experiment, suggesting that the
! binding of solute molecules to soluble proteins may have sequestered
K D AC h2   them within the tissue [45].
Q ðt Þ ¼ SC SC d t− SC ¼ kp AC d t−t lag ð1:4Þ This effect can be represented in terms of a drug's bioavailability:
hSC 6DSC
that fraction of drug that is available at a particular site for a particular
process (such as permeation into a neighboring region). Experimental
where
evidence suggests that epidermal metabolism occurs in the basal
2
layer [46,47], and that metabolism can play an important role in the
K SC DSC h sequestration of drugs in the VE [41]. It has been recently suggested
kp ¼ and t lag ¼ SC : ð1:5Þ
hSC 6DSC that the presence of keratin in the VE may affect diffusion into the
VE and/or partitioning from the SC into the VE [48].
While it has been formulated here as the solution to diffusion There is a wealth of information regarding carrier-mediated trans-
through a homogeneous membrane with simple boundary condi- porters in human skin [49], and studies of active transport in the VE
tions, Eq. (1.4) is widely applied because its form often remains center on those transport proteins expressed in epidermal keratino-
unchanged for more sophisticated SC models, with only equations cytes. However, studies of the kinetics associated with the active
for kp and tlag modified [32–34]. Indeed, the difficulties in determining transport of drugs – either for enhancing permeation or for serving
the actual path taken by different drugs through the SC are largely re- as a barrier – have yet to appear in the literature.
lated to this observation. These other more complex models are
reviewed in [9]. Solutions to various application scenarios (finite ve- 3.2.3. Modeling
hicle volume, finite time exposure with wash off and evaporation The barrier effects of the VE can be incorporated into skin perme-
from the vehicle) are also presented in Ref. [35]. ation models by treating the VE as a homogeneous unstirred aqueous
Models that explicitly include the structure of the SC facilitate the layer immediately below the SC. Thus the SC and VE form two barriers
comparison of model steady-state and transient lag times with exper- in series: the total epidermal permeability is calculated as
imental values, providing valuable insight into the importance of the
intercellular [36] and transcellular [37,38] routes. A similar approach 1 1 1
¼ þ ð1:6Þ
has provided evidence that corneocyte uptake and/or slow binding in kp;epi kp;SC kp;VE
the SC influence permeant transport behavior, even for simple mole-
cules such as water [34,39,40]. in accordance with solutions to the diffusion equation [32,50,51] —
An important phenomenon of drug delivery through the SC is the see Ref. [9] for more details.
so-called ‘reservoir effect’, where lipophilic compounds are effective- Cleek and Bunge compared such a model to single-membrane
ly sequestered by the SC. Such compounds partition preferentially models (finite or semi-infinite) representing the SC only, to further in-
into the lipophilic SC rather than the aqueous viable tissues — the vestigate the role of the VE [52]. By defining the ratio B= kp,SC/kp,VE,
SC–VE boundary provides a significant energy barrier for such mole- they could consider the overall contribution of the viable epidermal bar-
cules. However, a compound must be sufficiently diffusive within rier in terms of its effectiveness relative to the SC barrier, obtained
the SC for an effective reservoir to be formed. For a recent review of through formal solutions to the diffusion equation for each choice of bar-
the reservoir effect, the reader is referred to Ref. [41]. rier. The quantity B can be related to drug physicochemical properties
156 O.G. Jepps et al. / Advanced Drug Delivery Reviews 65 (2013) 152–168

such as lipophilicity [53], which are expected to impact upon the relative Despite the importance of and recent interest in VE protein trans-
permeability of drug in the lipophilic SC and drug in the aqueous VE. At porters, most studies have focussed on the nature of such transporters
penetration times that are sensitive to the existence of both barriers, rather than on characterizing their effect on drug permeability. Li et al.
the two-membrane model shows that penetration into the skin is depen- [65] showed that the penetration of the NSAIDs flurbiprofen and indo-
dent on B and therefore on drug lipophilicity: for highly hydrophilic methacin is governed by pH- and ATP-dependent saturable kinetics.
drugs (Bb 0.01) penetration is limited primarily by the SC, while for in- Passive diffusion and reversible kinetics alone cannot account for this:
creasingly lipophilic drugs the effects of the aqueous layer become the authors propose active transport as a potential explanation, possibly
more significant, “choking” the flux of such drugs into the skin in agree- by MRP, MCT, OATP or OCTN xenobiotic transporters [65].
ment with experimental observations [44]. Such models can also be used
to correct lag time predictions [52], and various methods have been 3.3. Dermis and deeper layers
proposed for estimating the parameter B [53], from which one can gen-
erate corrections to the SC permeability predictive model of Potts & 3.3.1. Morphology
Guy that take into account epidermal resistance [52,53]. More recently Immediately beneath the epidermis, the dermis is the thickest
this model has been extended to account for ionization and vehicle component of the skin, up to 4 mm in depth. Its upper layer, the
properties, applied to model the penetration of lipophilic (N-nitroso-N- 100–200 μm thick papillary dermis, consists of thin collagen bundles,
methyldodecylamine, NDOMA) and hydrophilic (hydroquinone) per- elastin fibers, fibrocytes and ground substance comprising mainly
meants through skin [54] and incorporated into models of skin diffusion water, electrolytes, plasma proteins and polysaccharides–polypeptide
which include the effect of desquamation [55,56]. complexes [43]. Below this layer is the reticular dermis, made up pre-
Such two-layer models have been used to estimate the permeabil- dominantly of thick collagen bundles and coarse elastic fibers [43].
ity of glucose in the viable epidermis from tape stripping data [57]. Protected by the epidermal barrier, the dermis houses the blood ves-
Glucose permeability was calculated to be 38% of that in dermis, com- sels, lymphatics and nervous system within the skin, as well as the
pared with a value of 30% of the dermal permeability that was various skin appendages. Below the reticular dermis lies the hypoder-
obtained from a side-by-side diffusion cell experiment on human epi- mis (subcutaneous fat tissue), which may have a thickness of up to
dermal membrane. several millimeters. Comprising fat microlobules and fibrous collagen,
Drug binding and sequestration in the viable epidermis affect the it also houses the blood vessels, lymphatics and nerves, and serves to
availability of drug for permeation into the dermis and beyond. The store energy, insulate the skin from below, and to connect the skin to
steady-state concentration of drug cSS at a site can be determined by underlying structures such as muscle and bone while allowing move-
balancing the input flux of drug JSSFavailableA (for input flux JSS through ment over them [42,43,66].
area A, where fraction Favailable of drug is available at the site) with the The structure and organization of the dermal blood vessels have
clearance flux cSSCl (for clearance rate Cl): been described by Braverman [67,68], with models of vessel distribu-
tion with depth subsequently developed by Cevc and Vierl [69]. Most
kp;SC cV Favailable A J SS Favailable A of the microvasculature is concentrated in the papillary dermis: the
cSS ¼ ¼ : ð1:7Þ
Cl Cl number of blood vessels decreases some 20-fold in the first millime-
ter of dermis; blood vessel area increases with depth from the epider-
However, drug binding or sequestration reduces the flux incident mal junction, peaking toward the lower end of the papillary dermis
at lower levels. Siddiqui et al. proposed that epidermal sequestration before decreasing quasi-exponentially. The orientation of vessels
of lipophilic solutes with steroid accumulation in the VE reduces the also has a particular structure, with a dense plexus of vessels running
steady-state flux to JSS = kp, SC(cV − cSS) [27], whence along the epidermal–dermal junction, another less dense plexus in
the reticular dermis, and vessels between the two plexuses that also
kp;SC cV Favailable A connect to the skin appendages (Fig. 1). The system of lymphatic ves-
cSS ¼ ð1:8Þ
Cl þ kp;SC Favailable A sels in the dermis runs roughly parallel to the dermal vasculature
structure [43,70].
leading to a decrease of up to 30% in the predicted steady-state der-
mal steroid concentration. 3.3.2. Mechanisms
The relationship between drug binding in the VE and dermis and Similar to the VE, drug diffusion through the dermis is primarily a
drug physicochemical properties is not generally understood: while hindered diffusion through an aqueous medium. However, there are
correlations with drug lipophilicity have been established [58], such important differences: the vascularization of the dermis contributes
relations are not universally predictive across families of compounds significantly to drug transport and distribution in the skin, as does
[59]. Recent studies show that once a lipophilic compound has per- the lymphatic system; the dermis is predominantly acellular in na-
meated beyond the SC, it can bind with soluble proteins and undergo ture; and the dermis provides many opportunities for the binding
bound transport [60]; and correlations have been recently established and sequestration of drug.
between lipophilicity and preferential distribution in the VE [61], The co-administration of vasoconstricting or vasodilating com-
although not necessarily in the dermis as well [62]. pounds (which alter drug uptake in the vasculature) during penetration
The effects of metabolism are generally incorporated into trans- studies has helped shed further light on the role of the vasculature in
port models by including terms to represent the fraction of available drug penetration and distribution [71–81]. In theory, vasodilation
drug that remains unmetabolized. A more sophisticated approach should allow greater clearance of drug into the vasculature, thereby
was adopted by Sugibayashi et al. in their study of the metabolism lowering steady-state concentrations: vasoconstriction should have
of ethyl nicotinate (EN) into nicotinic acid (NA) by the enzyme ester- the opposite effect. Such behavior has been observed experimentally
ase [63]. Models consistent with their experimental data indicate that for lidocaine delivery together with tolazoline (a vasodilator) or norepi-
enzyme distribution within the skin had little effect on the overall nephrine (a vasoconstrictor) [71,72], for salicylic acid and lidocaine
flux of EN and NA, but had a significant influence on the concentra- with phenylephrine (a vasoconstrictor), for flurbiprofen with epineph-
tion gradients. The relatively poor permeation of drugs whose tissue rine (a vasoconstrictor) [76] and for antipyrine with phenylephrine
residence time is long compared with the metabolic half-life (indicat- [79,81]. A number of microdialysis studies have also confirmed these
ing that a large amount of drug is metabolized in the time taken to observations [74,77,78,80]. Increased drug distribution at contra-
cross the tissue) has been observed experimentally and modeled by lateral sites due to vasodilation [75], and reduced distribution at contra-
Boderke et al. [64]. lateral sites due to vasoconstriction [81] have also been observed.
O.G. Jepps et al. / Advanced Drug Delivery Reviews 65 (2013) 152–168 157

Vasoconstriction has also been observed to augment deeper tissue


(muscle) penetration of drugs inversely to their fraction unbound in
viable skin (which reflects their distribution in muscle) [79].
There is also evidence to suggest that the orientation of blood ves-
sels plays a role in producing a “convective force” for drug transport
into deep tissue, beyond levels that can be explained by systemic de-
livery alone [82]. Such results have been observed for piroxicam de-
livery into deep tissues both at the site of application [83] and away
from the site of application [84], as well as for the diazepram applied
dermally to perfused rat hindlimb [85].
Lymphatic clearance plays an important role in the removal of
macromolecules [70,86–90], which suggests its potential relevance
to drug transport where drug binding occurs.
While transporter proteins are known to be expressed in the dermis
[91], few studies have been undertaken to date on the role of cutaneous
carrier proteins in drug distribution in dermis. Published studies show,
however, that such transport mechanisms may occur and could be
harnessed for the purposes of drug delivery to the deeper tissues or
the systemic circulation. P-Glycoprotein is a transporter protein belong-
ing to the ABC family and mainly expressed in the dermal components
of human skin (including sweat ducts) and muscle tissue [92]. Ito et al.
showed that P-Glycoprotein can transport itraconazole within dermal
tissue [93]. Another study by the same group showed that the Organic
Anion Transporter 2 (OAT 2) was involved in the transport of topically
applied flurbiprofen in an in vitro experiment [94]. Lipophilicity may
play a role in such transport mechanisms. While the transport of the
lipophilic flurbiprofen was vectorial (greater from the epidermal to
the hypodermal side then in the opposite direction) and saturable, no
carrier-mediated transport of mannitol, a hydrophilic permeant, was
observed.

3.3.3. Modeling
In the same way that the SC and VE can be treated as “diffusion re-
sistances in series” (where the diffusive resistance is given by the in-
verse permeability), Eq. (1.6) can be extended to include the dermis
as a third diffusive resistance. Using such a model, Scheuplein and
Blank found that while the permeability of a homologous family of al-
cohols decreased significantly with increasing number of carbon
atoms n, it remained approximately constant for the dermis [13]
(for n > 6, when the main barrier is provided by the SC, the dermal
and epidermal permeabilities are similar in value).
Much of dermal transport modeling concerns the quantification of
Fig. 2. The two-compartment model in Ref. [90]. (a) Schematic for Qp and Qt represent-
properties related to the vasculature and lymphatic clearance. Typi-
ing the perfusate flow rate into and blood flow rate out of tissue respectively; ka and ke
cally, this is achieved via compartment models where a single value represent drug absorption and elimination rates, (b) observed data for diclofenac
represents the mean local concentration of the drug. For example, remaining in dermal cell with time for a dextran and albumin perfusate, (c) observed
Roberts and Cross [90,95] solve a two-compartment model (Fig. 2) data for diclofenac appearing in perfusate from tissue with time after delivery via a der-
analytically to establish the time-dependent cumulative amount of mal cell for a dextran and albumin perfusate.

drug having reached the dermis. By considering the fractions of


drug unbound in tissue and in the vasculature, one may estimate [73]. Higaki et al.'s model (Fig. 4) applied to the delivery of 6 drugs
the effective drug volume of distribution, which then permits an esti- led to similar conclusions at early times, but more significant contri-
mate of the elimination rate constant. From such an approach, impor- butions from systemic circulation at longer times [79], where the
tant clinical parameters such as fractions of sequestered drug and the transport rate from viable skin to muscle was found to be significantly
solute retention half-life can be estimated. This approach has been correlated to the fraction unbound in viable skin. The removal of VE in
validated with diclofenac absorption data [95]. Singh and Roberts' study may explain to some degree the long-time
More complex compartment models that explicitly include the discrepancy between these studies.
systemic circulation and distribution into deeper tissues have also Despite offering valuable insights, a potential drawback of such
been developed by Singh and Roberts (Fig. 3) [96,97] and by Higaki compartment models is the implicit homogeneity, which belies the
et al. [79]. By fitting physiologically based parameters in their model expected variation in drug concentration and other effects. This can
to match experimental data of the distribution of salicylic acid after be overcome by explicitly building such variation, where known,
application to dermis in anesthetized rats, the relative importance into spatially-sensitive models. For example, the variation in vascu-
of the tissue and vasculature transport to the distribution of drug in larization through the papillary and reticular dermis can be incorpo-
muscles could be examined. In particular, the importance of systemic rated using the partial differential equation [69]
circulation (through measuring delivery at contralateral sites) could
be quantified, revealing the dominance of tissue transport for deep
drug delivery at the site of application [96]. This model also describes ∂cðd; t Þ ∂2 cðd; t Þ
¼ DðdÞ −kd P ðdÞcðd; t Þ ð1:9Þ
the concentration–depth profiles observed during vasoconstriction ∂t ∂2 t
158 O.G. Jepps et al. / Advanced Drug Delivery Reviews 65 (2013) 152–168

Fig. 3. Drug distribution in the skin layers and circulation described by the model of Singh et al. [96]. (a) Physiological pharmacokinetic model: distribution is measured as concen-
trations c; model parameters include clearances Cl, blood flow rates Q and volumes V; subscripts refer to relevant site. (b) Experimental (data points) and predicted (solid line)
dermal concentration–time profiles in underlying tissue after dermal application of salicylic acid. The first peak is due to direct penetration from topical application. (c) Experimental
(data points) and predicted (solid line) concentration–time profiles in subcutaneous and lower tissues below the applied site. The second peak at about 10 h is attributed to systemic
distribution as well as some direct penetration of salicylate.

where the diffusion coefficient D and elimination kpP are functions dispersion coefficient for the liver model was later related to the
of depth. Such depth-dependence can then be estimated by fitting liver morphology and physiologically related parameters [101] — a
against concentration–depth profiles [69]. similar relationship remains to be established for the dermal trans-
It is clear from experimental data with vasodilators and vasocon- port model.
strictors that convective capillary transport, while responsible for Kretsos et al. analyzed experimental data on the in vivo penetra-
clearance of the drug, also contributes to transport of solutes to dee- tion of salicylic acid into de-epidermized rat skin using a distributed
per tissues. Recent models [98,99] suggest that transport to lower diffusion-clearance model describing transient distribution in whole
layers occurs mostly by blood/lymphatic flow rather than molecular skin [102]. Their main goal was to characterize dermal transport of
diffusion alone, as assumed previously. This blood/lymphatic trans- salicylic acid (SA) by deriving values for the effective diffusion and
port can be modeled by replacing the dermal diffusion coefficient partition coefficients (with respect to an aqueous solution at a given
with an effective dispersion coefficient (similar to modeling of liver pH) in dermis, and for a volumetric first-order rate constant repre-
clearance by dispersion model [100]). As in the liver model, the dis- senting systemic clearance. The study shows that owing to a balance
persion coefficient is approximated to be a constant parameter, between diffusion within and volumetric clearance from the dermis,
independent of other physiological parameters of the dermis. The the SA concentration–depth profile exhibits an exponential decay
O.G. Jepps et al. / Advanced Drug Delivery Reviews 65 (2013) 152–168 159

Fig. 4. (a) Compartment model of Higaki et al. [79] describing kinetics of topically applied drugs. Distribution is measured as concentrations C; model parameters include clearances
Cl and volumes V; subscripts refer to relevant site. (b) Solid lines describe predicted intradermal kinetics of diclofenac after topical application on the stripped skin in vivo with the
symbols representing donor cell (△), viable skin (▲), muscle (◊), plasma (♦) and contralateral muscle (●). (c) Predicted diclofenac concentrations in muscle — the total concen-
tration (solid line), the concentration due to direct penetration (dashed line), and the concentration due to systemic circulation (dotted line).

profile. While capillary clearance significantly impacts drug distribu- KP the permeant's octanol–water partition coefficient. A detailed de-
tion in dermis, the in vivo dermal diffusion coefficient of SA was scription of Kasting and coworker's correlations (as well as their use
found to be approximately equal to the molecular diffusion coefficient in a general skin penetration model) is presented in Ref. [104]. In gener-
of SA in dermis, implying that the dermal vasculature does not signif- al, protein binding in the skin is often approximated by binding to albu-
icantly increase transport by convection. min since this is the most common protein found in the interstitial fluid.
Kretsos et al. analyzed in vitro dermal penetration data and devel- However, other lipoprotein, α1-glycoprotein [105] and keratin should
oped the most detailed empirical expressions to date for the partition also be considered. Evidence of permeation rate differing as a result of
and diffusion coefficients in dermis [103]. The models incorporate per- differential protein binding is given by Weiss et al. [106].
meant ionization, volume exclusion due to the presence of a fiber phase, Questions that remain to be answered in the study of deep tissue
a lipid phase and binding to albumin present in a portion of the dermis' penetration include [107]: is direct penetration or systemic circula-
aqueous phase. A “binding factor” is derived and serves to modify the tion more important for the delivery of certain drugs to the deep
values of solute concentration, and partition and diffusion coefficients skin tissue?; and why is the increase in the fraction unbound in viable
of unbound solute based on these considerations. This factor is [103] skin not important for systemic absorption of certain drugs?

ϕa 3.4. Appendages
Binding factor ¼ ð1−ϕa Þ þ þ φlip fnon K P ð1:10Þ
fu
3.4.1. Morphology
where ϕa and ϕlip denote the albumin and lipid phase volume fractions The conical entrance into the hair follicle from the skin surface
in dermis, fu and fnon fractions of unbound and non-ionized solute, and (known as the infundibulum) extends approximately 500 μm deep
160 O.G. Jepps et al. / Advanced Drug Delivery Reviews 65 (2013) 152–168

(i.e. well beyond the typical depth of the epidermal–dermis junction) network. Measurements of the effect of the vasodilator benzyl nicotinate
to the sebaceous duct that connects the follicle with the sebaceous (BN) after topical application to the human forehead (rich in follicles),
gland. Sebaceous glands are also found independently of hair follicles, forearm and calf (poor in follicles) suggest a significant role played by
mainly on the face [108]. This gland produces sebum, a lipophilic sub- the heavily vascularized of the hair follicles [127].
stance comprising triglycerides, wax esters, squalene, cholesterol and Studies comparing the penetration of a range of drugs through
its esters [109] that fills the infundibulum and serves various roles human scalp and through abdominal skin demonstrate the influence
(including protection against bacteria, skin over-wetting, heat loss, of greater follicle density [128]. Microscopy techniques suggest that
transporting antioxidants to the skin surface, and maintaining skin the infundibulum wall remains a significant barrier, albeit weaker
integrity) [43,110]. Sebum takes approximately 8 h to pass from the than the SC [66,122]. Visual evidence also suggests that permeation
sebaceous duct to the skin surface [111]. The epithelial wall of the in- beyond the infundibulum is favorable through the junction separat-
fundibulum is a continuation of the SC whose thickness diminishes ing the outer and inner root sheaths [128,129]. Lademann et al.
down the follicle [13,66,112]. Below the sebaceous duct, the hair conclude that massage induces a hair-shaft “pumping” mechanism
follicle comprises a series of tightly packed cell layers – an acellular that improves penetration by pushing nanoparticles deeper into folli-
membrane, the outer root sheath (a continuation of the viable cles [130]. The same group has also found a positive correlation
epidermis), and the inner root sheath – that surround the hair shaft between follicular penetration and the ‘activity’ of hair follicles
[43,66,113,114]. At the base of the follicle lies the hair bulb, a region (measured in terms of sebum production and follicle growth) [131].
that contains the dermal papilla and is instrumental for follicular These results suggest that sebum flow does not impede permeation
growth. The follicle is heavily vascularized: the bulb is surrounded through the follicular pathway [130], although the effects on drug
by interconnected blood vessels running parallel to the length of the permeation are not clear. The influence of binding on permeation
follicle; vessels run parallel to the follicle above the bulb, connecting has also been observed, particularly binding of lipophilic compounds
with the vessels surrounding the sebaceous gland (whose acini are to the hair shaft cuticle [122,123]. Cone suggests that protein and
also highly vascularized) [115]. melanin are primarily responsible for hair-shaft binding and seques-
Eccrine sweat glands produce sweat to help cool the body by tration [132]. Reservoir effects within the cuticle [122,123] and sebaceous
evaporation, improve grip, and sensitize skin [42,43,116]. The glands gland [126,133] have been suggested, although further confirmation is
are housed in the lower dermis or hypodermis, and connected to the required.
surface by a duct approximately 100 μm in diameter that coils about Permeation through sweat gland epithelium appears to depend
the gland before rising straight through the dermis and spiralling up significantly on lipophilicity, as is commonly observed in other epi-
the epidermis [117]. Water, salts and electrolytes can be reabsorbed thelia [133]. Jackson and Davis developed a 1D diffusion model for
through the duct walls to maintain homeostasis as required [116]. permeation through the sweat duct, approximating it as a uniform
The skin also contains apocrine and apoeccrine glands that resemble tube [134], and demonstrating that the sweat duct may act as a reser-
eccrine glands. In adults, apocrine ducts connect to the infundibulum, voir under occlusion and non-sweating conditions.
and their precise product (and its function) has not been isolated Lademann and co-workers have also visualized permeation into
[43,66]. Apoeccrine glands appear during adolescence, and share fea- sweat glands [135,136], with similar notions of ‘activity’ (albeit with
tures of both eccrine and apoeccrine glands [118,119]. These eccrine less well established criteria). The novel potential for delivery from
glands and ducts are also highly vascularized, as are the apocrine the eccrine sweat gland to the skin surface in human patients has
glands [115,120]. also been observed for intravenously (IV) administered compounds
[137,138]. Tight-junction-associated occludin, claudin 1 and claudin
3.4.2. Mechanisms 4 are absent from those parts of the apocrine and apoeccrine ducts
The skin appendages present an attractive alternative pathway for positioned within the epidermis, offering a potentially more perme-
drug delivery into the skin, raising the possibility of avoiding the chal- able route into the epidermis than the SC [119,139,140]. The lack of
lenges of delivering drug through the SC. Lipophilic sebum provides a cornified cell layer further suggests the possibility of drug transport
a natural pathway for the diffusion of lipophilic drugs into the hair via these glands, but this application remains to be investigated.
follicles and sebaceous glands, but presents a barrier for hydrophilic
drugs [66]. While this behavior has been observed experimentally, 4. Global perspective
the potential for drugs to pass from the infundibulum into deeper
parts of the follicle/sebaceous gland, and ultimately to permeate into 4.1. Overall distribution models
the vasculature or the skin, remains to be understood. In a similar
fashion, the potential for drug penetration via the eccrine or apocrine The aforementioned models, based on the mechanisms deduced
ducts has yet to be understood, although recent work indicates this pos- from experimental data and related skin physiology, describe penetra-
sibility [119]. tion and distribution processes of a specific nature and at a specific
A clear understanding of the mechanisms promoting or retarding site. Many are developed to explain experimental data that is intention-
drug transport via the hair follicles and glands remains to be estab- ally of a similarly specific nature, in order to improve understanding of
lished, despite the wealth of experimental studies. a particular phenomenon that would not be possible were other poten-
tially confounding factors not avoided. As a consequence of this
3.4.3. Modeling reductionist approach, global penetration and distribution models do
Scheuplein and Blank identified the importance of the appenda- not generally comprise a combination of these state-of-the-art models.
geal pathway during the initial stages after topical drug application Rather, information garnered from these models is typically fed into
[13,121]. This work has been much more recently verified using mi- global models of a relatively simple nature, typically involving linear
croscopy techniques [122,123]. kinetics (i.e. compartment models). This raises the prospect that the
It appears that lipophilic vehicles such as liposomal formulations or behavior exhibited by the more sophisticated model might not be avail-
nanoemulsions facilitate drug penetration into deeper tissues by improv- able in the approximate global version: however, if the more sophisti-
ing permeation through the follicular pathway, in hairy rats [124], hairy cated model is linear, the underlying approximation can be improved
mouse [125], and hairless rat [126]. This last study, which compared in through the addition of further compartments.
vitro and in vivo data, also highlighted the potential role played by the At its simplest, one can then consider models such as those originally
vasculature — however, no studies have been able to assess the perme- proposed by Scheuplein and Blank [13] that represent the various
ation of drugs between hair follicles and their surrounding vascular permeation pathways as a passage of diffusive resistances, connected
O.G. Jepps et al. / Advanced Drug Delivery Reviews 65 (2013) 152–168 161

Fig. 5. (a) Schematic of drug distribution for topical (left) and intravenous (IV) (right) administrations: distribution is measured as local amounts A; also represented are the frac-
tions of administered drug available at the site of action (fa), and lost to metabolism in the skin (fm, skin) and in other organs (fm, body). (b) Total amount of butylparaben (●) recov-
ered as intact drug (♦) and as metabolized p-hydroxybenzoic acid (×) after application to excised rat skin as a 20% (w/v) drug suspension in phosphate-buffered saline [174].
Panel a: adapted from Ref. [141].
162 O.G. Jepps et al. / Advanced Drug Delivery Reviews 65 (2013) 152–168

in series (e.g. consecutive skin layers) or in parallel (e.g. SC route vs.


appendageal route), leading to expressions such as

!−1
1 1 1
kp;skin ¼ þ þ þ kp;appendages ð1:11Þ
kp;SC kp;VE kp;dermis

where the various local permeabilities are determined by factors out-


lined in the previous section.
Clearance from the skin also plays an important part in these
models [32,141]. For systemic delivery, the steady-state concentra-
tion of drug in plasma cSS can be approximated by Eq. (1.7), which
results from a mass balance of drug entering the skin (and therefore
the circulation, for a steady state) with that being cleared from the
circulation. Comparison with IV administration can provide informa-
tion about first-pass drug availability from topical application, once
the steady-state flux (or equivalent) has been established. This can
be achieved either by comparing the total AUC for topical and IV ad-
Fig. 6. Demonstration of ‘flip–flop’ kinetics, comparing excretion rates following oral
ministrations [142]: (open circles) and topic (closed circles) norephedrine HCl administrations.
Adapted from [175], cited in [3].
AUC topical DoseIV
Favailable ¼ ð1:12Þ
Dosetopical AUC IV
The kinetics of topical delivery into the systemic circulation can be
or by comparing excretion data for topical and IV administrations distinctive (in comparison with IV administration) due to the signifi-
[143] cant differences in absorption rates. Absorption is effectively instanta-
neous for IV administration, and therefore much faster than
Atopical DoseIV elimination (i.e. clearance): for topical delivery, absorption can be
Favailable ¼ ð1:13Þ
Dosetopical AIV much slower than elimination, leading to so-called ‘flip–flop kinetics’
where the usual rate-dependence on clearance observed from IV ad-
where A represents the amount excreted (from topical or IV adminis- ministration is ‘flipped’, and the absorption rate becomes the rate de-
tration). One can see the origin of these equations by considering the termining step (Fig. 6) [151,152].
various contributions to the first-pass skin availability, which takes into
account the combined effect of the fraction of topically applied drug fa
absorbed into the skin, as well as the fraction of drug 1 − fm, skin that re-
mains unaffected by skin metabolism. The amount excreted from the
urine is a fraction 1 − fm, skin of drug that remains unaffected by metab-
olism in the circulation, for both the IV and topical administrations,
which gives us (Fig. 5)
  
  fa 1−f m;skin 1−f m;body Atopical DoseIV
Favailable ¼ fa 1−f m;skin ¼   ¼ :
1−f m;body Dosetopical AIV

ð1:14Þ

Eq. (1.12) has been used to determine a skin first-pass availability


for topically administered fentanyl of 0.92 ± 0.33 [142]. Eq. (1.13)
has been used to determine the skin first-pass availability for
various commercial formulations of methyl salicylate, which varied
from 0.12 to 0.2 [143]; estradiol (Favailable = 0.075) [144]; diclofe-
nac (Favailable = 0.066) [145]; and nitroglycerin (Favailable = 0.72 ±
0.04) [144]. The various factors that affect first-pass availability, in-
cluding skin vascularization, metabolism and the availability of ap-
pendages [90], vary from site to site on the body, and experimental
data obtained for methyl salicylate [143] and nicotine [146]
demonstrate that first-pass availability is indeed site-dependent.
While our understanding of those enzymes responsible for transder-
mal drug metabolism continues to improve [147–149], the influence
of these enzymes on the kinetics of metabolized drugs is difficult to
determine. Specific enzymes that affect transdermally delivered drugs
include esterases (found in the viable skin) that play a role in the
first-pass metabolism of drugs such as methyl salicylate and nitroglyc-
erin [148]. The effect of esterases has been further investigated by Fig. 7. Schematic overview relating the effective clearance Cleff* from the site of action
during topical administration to the clearance from skin into the circulation Clskin, re-
skin pretreatment with an esterase inhibitor, and modeled with a turn clearance into the skin from the circulation Clr, and clearance from the circulation
two-layer diffusion model that predicted the decreasing lag time with ClIV.
increasing enzyme metabolic rate [150]. Adapted from Ref. [141].
O.G. Jepps et al. / Advanced Drug Delivery Reviews 65 (2013) 152–168 163

Fig. 8. Prodrug, soft drug and pro-soft drug mechanisms of action. (a) Metabolism may strongly reduce the internal flux of the active compound, Jact,int, compared to the active com-
pound's external flux Jact,ext. (b) The prodrug approach consists in attaching a promoiety to the active drug which renders it inactive but increases its bioavailability, or external flux
Jact,ext. Cutaneous metabolism is harnessed to render the drug active. The resulting internal flux to the site of action, Jact,int, is greater than in the case of an unchanged drug (Fskin
designates the fraction of unchanged drug in skin). (c) A soft drug is the opposite of a prodrug. Skin metabolism is utilized to render an active drug inactive, which is eliminated
upon pharmacological action. (d) In the pro-soft drug approach, an inactive drug, with external flux Jinact,ext, is metabolized to an active compound which diffuses to the site of action
with internal flux Jact,int (Fskin designates the fraction of un-metabolized inactive drug). Following pharmacological action, the active drug is further metabolized to an inactive com-
pound and eliminated.
Figure adapted from Ref [141].

For cutaneous delivery, the clearance is now elimination (and poten- and skin, respectively. Binding may have an important influence on the
tially metabolism) from the site of action, but the potential for drug to re- local therapeutically active concentration, as well as on clearance from
turn to the site of action after uptake into the circulation must be taken the site (Fig. 5). From Eq. (1.15) we identify two limiting cases: when
into account (Fig. 7): mass balance gives us cbClIV =cskinClskin −cbClr, clearance from the circulation is much greater than the return clearance,
which combined with the definition of the effective skin clearance Cleff * the effective skin clearance is effectively just given by Clskin; however,
(Fig. 7) yields when the return clearance dominates, the effective clearance is much
* b b Clskin. In this second case, the drug exchange between
smaller, Cleff
 cb Cl Cl skin and circulation is much greater than delivery from the skin surface.
Cleff ¼ ClIV ¼ IV skin ð1:15Þ
cskin ClIV þ Clr
4.2. Prodrugs
where the Cl represent clearance from skin into the circulation (Clskin), re-
turn clearance from circulation into the skin (Clr), or clearance from the While drug metabolism can reduce drug availability at the site of ac-
circulation (ClIV), and cb and cskin represent concentrations in the blood tion, metabolism can sometimes play a useful role in the improvement

Fig. 9. Comparison of in vitro vs. in vivo skin penetration data obtained from experiments with full-thickness human skin. The chemicals investigated are (a) nitrobenzene and (b)
2,4-dinitrochlorobenzene from [161]; (c) caffeine in water gel, (d) caffeine in petrolatum, (e) caffeine in ethylene glycol gel, (f) caffeine in benzoic acid and (g) testosterone from
[162]; (h) isofenphos [163]; (i) chloroform [164]; (j) lindane [165]; (k) chlorpyrifos [166,167]; (l) ortho-phenylphenol [168]; (m) methylene bis-benzotriazoyl tetramethylbutyl-
phenol (MBBT) and (n) titanium dioxide from [169].
164 O.G. Jepps et al. / Advanced Drug Delivery Reviews 65 (2013) 152–168

of topical drug delivery, or in the reduction of undesirable effects of a systemic effect is the potency of a drug. Davis [170] has pointed
drug beyond the site of action. Some of the difficulties in delivering out that potency and skin penetration are equally important in deter-
drug to the site of action, including low drug permeability through the mining the efficacy of topical products. He comments that direct skin
skin, fast drug elimination, or difficulties in formulation [153] can be penetration of non-steroidal anti-inflammatory drugs is associated
obviated through the addition of a promoiety to the original drug, form- with local efficacy without systemic adverse effects. He goes on to
ing a prodrug. If the prodrug is metabolized to produce the desired drug point out that the topical application of other drugs, such as corticoste-
at the site of action, this approach can be used to deliver drugs more ef- roids, retinoids and antihistamines is associated with systemic toxicity,
fectively. Metabolism can also play an important role in physiologically especially in severe skin disease.
deactivating drugs to non-toxic by-products once the desired pharma- In practice, therefore, the appropriate dosing of a topical product
cological effect has been achieved at the site of action: drugs where must first begin with drug selection and, here, efficacy is a function
this behavior is observed are called soft drugs. It is also possible to of both how much has penetrated and how potent the compound is.
find a soft drug behavior in prodrugs: such compounds are known as One of the first authors to combine these concepts in a formal manner
pro-soft drugs. The mechanisms of these approaches are outlined in was Lippold [171,172]. He defined the efficacy coefficient as the
Fig. 8. ratio of the maximum flux divided by the drug potency dose. Thus,
One of the key advantages of prodrugs for topical drug delivery is in greater efficacy is provided by either a higher flux or a more potent
circumventing the skin barrier: a naltrexone prodrug improves skin (lower dose for the same effect) drug. Fig. 10 shows an illustration
permeability 8-fold [154], while mouse and rat models demonstrate of this approach using anti-inflammatory data from Cordero et al.
improvements of 9-fold for 5-fluorouracil [155], 12-fold for ketorolac [173]. Here, it is evident that optimal skin penetration occurs at a
[156] and almost 40-fold for nalbuphine [157]. The topical application log octanol–water partition coefficient (log P) of 2 to 3 and that the
of the retinoid prodrug tazarotene may produce superior delivery to
orally administered alternatives due to esterase metabolism in the skin
that generates the more water-soluble active metabolite tazarotenic
acid [145,158].

4.3. In vivo vs. in vitro results

Much of the available drug penetration data has been obtained


through in vitro experiments on excised skin, so the correlation be-
tween the values obtained in these results and the values that
would be expected in vivo is an issue of some significance. There is
an obvious importance in maintaining the skin viability through var-
ious means [159,160], but beyond this issue lies the question of how
well the essential conditions responsible for in vivo penetration are
reproduced in vitro. A number of studies [161–169] have performed
both in vitro and in vivo penetration studies, for the purposes of ex-
plicit comparison, by comparing the percentage of applied dose
absorbed in the receptor of the in vitro experiment to the percentage
of applied dose absorbed and excreted in the urine for the in vivo
experiment.
The overall picture is of a good correlation between in vitro and in
vivo data (Fig. 9), although a number of results from these studies
[161,162,168] show that the in vitro studies under-predict the in
vivo penetration. Such results might be explained by the lower tem-
perature of exposed skin during in vitro experiments than of skin
under clothing in vivo [162], or that a lag in excretion
underestimates the actual amount of penetrated drug in vivo [168].
Indeed, Cnubben et al. found a similar result in viable full-thickness
rat skin [168] — the more significant discrepancy for the same com-
pound (ortho-phenylphenol) in these rat studies was attributed to
greater compound metabolism in viable rat skin. Metabolism may also
play a role in the discrepancy for 2,4-dinitrochlorobenzene [161] and
testosterone [162].
One important exception to this observation, leading to the over-
prediction of the in vivo penetration from in vitro studies, is in the
case where the SC reservoir likely has an important influence in
vivo [165–167]. Important factors that influence in vivo penetration
for such drugs, including the effects of occlusion [165], or significant
partitioning into subcutaneous fatty tissue [166,167], are not neces-
sarily captured by the in vitro experiment. Denaturing of the skin
from the polar receptor fluid may also have influenced these results
[167].

4.4. Clinical response


Fig. 10. Using (a) maximal skin flux and (b) in vitro anti-inflammatory activity to esti-
mate (c) likely clinical efficacy of various non-steroidal anti-inflammatory drugs with
Percutaneous penetration does not necessarily guarantee response. varying values for logarithm octanol–water partition coefficient (log P).
Indeed, a key determinant in the choice of a drug for either local or Data from [173] and Scifinder data base.
O.G. Jepps et al. / Advanced Drug Delivery Reviews 65 (2013) 152–168 165

more lipophilic drugs had greater in-vitro-based indices of anti- Appendix A. Summary of equations
inflammatory activity. As a consequence, diclofenac with log P > 4 has
the highest efficacy. As Davis [170] points out, the validity of this ap- Please refer to the text surrounding each equation for a full expla-
proach is evident from recent clinical studies showing the efficacy of nation of the relevant variables.
topical diclofenac and ketoprofen.
(1.1) SC homogeneous membrane model for solute concentration
(Fick's 2nd law):
5. Conclusions
∂C ðx; t Þ ∂2 C ðx; t Þ
¼ DSC :
∂t ∂x2
The complex structure of the skin reflects the myriad functions
that it performs, often (but by no means solely) related to its role as
(1.2) Usual boundary conditions for transdermal solute application:
a barrier separating us from the world outside. The variety of these
structures and functions makes it exceedingly difficult to give com- C ðx; 0Þ ¼ 0; C ð0; t Þ ¼ K SC C d ; C ðhSC ; t Þ ¼ 0 :
prehensive models that describe the penetration and distribution of
drugs passing through the skin. After many decades of investigation, (1.3) Amount of solute permeating the SC as a function of time:
and despite much progress on many of the fundamental questions,
!!
many details of transdermal drug transport remain elusive. However, DSC 1 2X ∞
ð−1Þn D 2 2
as experimental techniques – non-invasive imaging techniques in Q ðt Þ¼ K SC AC d hSC t− − exp −t 2SC π n :
2
hSC 6 π n¼1 n2 hSC
particular – continue to provide greater access to the details for trans-
dermal transport, our capacity to refine our models in light of these
(1.4) Long-time limit amount of solute permeating the SC as a func-
features continues to improve.
tion of time:
The overall picture of transdermal drug transport is one where the
SC provides the main penetration barrier for most compounds (for !
K D AC h2  
those drugs that readily enter the SC, the aqueous viable skin beneath Q ðt Þ ¼ SC SC d t− SC ¼ kp AC d t−t lag :
hSC 6DSC
provides the main resistance). Permeation through the SC is largely by
passive diffusion, although the nature of that process (transcellular vs
intercellular, the degree of corneocyte uptake) remains to be fully (1.5) Relationships defining permeability coefficient and lag time:
understood. The viable layers beneath the SC provide a less physically
constrictive environment for passive diffusion; however, there are var- K SC DSC h2
kp ¼ and t lag ¼ SC :
ious other processes that become relevant to the overall transport. hSC 6DSC
These include drug binding and sequestration, active transport, metab-
olism and (below the epidermal–dermal junction) transport mediated (1.6) Combining rule for permeabilities in series
by the vasculature and lymphatics. This last contribution can manifest
1 1 1
itself as a dispersive influence on the transport, as well as facilitating ¼ þ :
drug clearance. The appendages also provide an alternative pathway kp;epi kp;SC kp;VE
for topically administered drug — however, the capacity for drug to con-
tinue beyond the appendages into the surrounding skin or systemic cir-
culation is not entirely understood. (1.7) Steady-state solute concentration in viable epidermis, balan-
Because of the many different influences on overall drug transport cing steady-state delivery and clearance:
through skin, studies typically focus on improving our understanding
of one influence, restricting as much as possible the potential for con- kp;SC cV Favailable A J SS Favailable A
cSS ¼ ¼ :
founding effects. In order to provide a more global model of transdermal Cl Cl
transport, the information gained from these individually studies is
usually incorporated into standard pharmacokinetic models (usually (1.8) Steady-state solute concentration in viable epidermis, balancing
physiologically-based compartment models) that allow a representa- steady-state delivery and clearance, incorporating epidermal
tion of the drug distribution in relevant tissues. As our understanding drug binding/sequestration:
of skin transport improves to the extent that we may confidently juxta-
pose the more sophisticated local models, these pharmacokinetic kp;SC cV Favailable A
cSS ¼
models will also be able to provide a more sophisticated global picture Cl þ kp;SC Favailable A
of drug kinetics following topical administration.
An allied challenge is that of predicting transport properties. One mo- (1.9) Dermal diffusion–elimination model (elimination varying with
tivation for physiologically-based models is that the model parameters depth):
have a physiological interpretation. This in turn allows the possibility
that these parameters could be predicted from a fundamental under-
∂cðd; t Þ ∂2 cðd; t Þ
standing of the processes and interactions involved. Identifying those ¼ DðdÞ −kd P ðdÞcðd; t Þ:
∂t ∂2 x
physicochemical properties responsible for the various transport mecha-
nisms, and isolating their dependencies qualitatively and quantitatively, (1.10) Krestos et al. binding factor (accounting for dermal heterogeneity):
remain an on-going challenge. However, such predictive power is of fun-
damental importance in the continuing development of applications for ϕa
therapeutic topical drug administration. Binding factor ¼ ð1−ϕa Þ þ þ φlip fnon K P :
fu

(1.11) Combining rule for permeabilities in series and in parallel:


Acknowledgment
!−1
We are grateful to Ms. Klintean Wunnapuk for the preparation of 1 1 1
kp;skin ¼ þ þ þ kp;appendages :
Figs. 2, 5 and 7. kp;SC kp;VE kp;dermis
166 O.G. Jepps et al. / Advanced Drug Delivery Reviews 65 (2013) 152–168

(1.12) Drug availability obtained comparing AUC for topical and IV [18] K.C. Madison, Barrier function of the skin: “la raison d'etre” of the epidermis,
J. Invest. Dermatol. 121 (2003) 231–241.
administrations: [19] J.A. McGrath, R.A.J. Eady, F.M. Pope, Anatomy and Organization of Human Skin,
Rook's Textbook of Dermatology, 2004.
AUC topical DoseIV [20] A.C. Michaels, S.K. Chandrasekaran, J.E. Shaw, Drug permeation through human
Favailable ¼ : skin: theory and in vitro experimental measurement, AICHE J. 21 (1975) 985–996.
Dosetopical AUC IV
[21] J.A. Bouwstra, A. de Graaff, G.S. Gooris, J. Nijsse, J.W. Wiechers, A.C. van Aelst,
(1.13) Drug availability obtained comparing excretion from topical Water distribution and related morphology in human stratum corneum at dif-
ferent hydration levels, J. Invest. Dermatol. 120 (2003) 750–758.
and IV administrations:
[22] T. Yotsuyanagi, W.I. Higuchi, A two phase series model for the transport of steroids
across the fully hydrated stratum corneum, J. Pharm. Pharmacol. 24 (1972)
Atopical DoseIV 934–941.
Favailable ¼ : [23] C.A. Squier, The permeability of keratinized and nonkeratinized oral epithelium
Dosetopical AIV
to horseradish peroxidase, J. Ultrastruct. Res. 43 (1973) 160–177.
[24] P.M. Elias, D.S. Friend, The permeability barrier in mammalian epidermis, J. Cell
(1.14) Drug availability, related to fraction of drug absorbed into the Biol. 65 (1975) 180–191.
[25] W.J. Albery, J. Hadgraft, Percutaneous absorption: theoretical description, J. Pharm.
skin and fraction of drug undergoing metabolism in skin
Pharmacol. 31 (1979) 129–139.
   [26] H.E. Bodde, I. van den Brink, H.K. Koerten, F.H.N. de Haan, Visualization of in
  f a 1−f m;skin 1−f m;body vitro percutaneous penetration of mercuric chloride; transport through intercel-
Favailable ¼ f a 1−f m;skin ¼   lular space versus cellular uptake through desmosomes, J. Control. Release 15
1−f m;body (1997) 227–236.
[27] O. Siddiqui, M.S. Roberts, A.E. Polack, Percutaneous absorption of steroids: relative
Atopical DoseIV contributions of epidermal penetration and dermal clearance, J. Pharmacokinet.
¼ : Biopharm. 17 (1989) 405–424.
Dosetopical AIV
[28] R.J. Scheuplein, Mechanism of percutaneous adsorption. I. Routes of penetration
and the influence of solubility, J. Invest. Dermatol. 45 (1965) 334–346.
(1.15) Effective clearance, accounting for potential return of drug to [29] M.S. Roberts, R.A. Anderson, J. Swarbrick, Permeability of human epidermis to
phenolic compounds, J. Pharm. Pharmacol. 29 (1977) 677–683.
the site of action:
[30] K.D. Peck, A.H. Ghanem, W.I. Higuchi, Hindered diffusion of polar molecules
through and effective pore radii estimates of intact and ethanol treated human
 cb Cl Cl epidermal membrane, Pharm. Res. 11 (1994) 1306–1314.
Cleff ¼ ClIV ¼ IV skin :
cskin ClIV þ Clr [31] J. Crank, The Mathematics of Diffusion, Clarendon Press, Oxford, UK, 1975.
[32] Y.G. Anissimov, M.S. Roberts, Diffusion modeling of percutaneous absorption ki-
netics. 1. Effects of flow rate, receptor sampling rate, and viable epidermal resis-
References tance for a constant donor concentration, J. Pharm. Sci. 88 (1999) 1201–1209.
[33] Y.G. Anissimov, M.S. Roberts, Diffusion modeling of percutaneous absorption kinet-
[1] S.E. Cross, M.S. Roberts, Physical enhancement of transdermal drug application: ics: 3. Variable diffusion and partition coefficients, consequences for stratum cor-
is delivery technology keeping up with pharmaceutical development? Curr. neum depth profiles and desorption kinetics, J. Pharm. Sci. 93 (2004) 470–487.
Drug Deliv. 1 (2004) 81–92. [34] Y.G. Anissimov, M.S. Roberts, Diffusion modelling of percutaneous absorption
[2] A. Joshi, J. Raje, Sonicated transdermal drug transport, J. Control. Release 83 kinetics: 4. Effects of a slow equilibration process within stratum corneum on
(2002) 13–22. absorption and desorption kinetics, J. Pharm. Sci. 98 (2009) 772–781.
[3] M.S. Roberts, S.E. Cross, M.A. Pellett, Skin transport, in: K.A. Walters (Ed.), Derma- [35] Y.G. Anissimov, Mathematical models for different exposure conditions, in: M.S.
tological and Transdermal Formulations, Marcel Dekker, New York, 2002. Roberts, K.A. Walters (Eds.), Dermal Absorption and Toxicity Assessment,
[4] H. Schaefer, T.E. Redelmeier, J. Lademann, Skin penetration, in: J.D. Johansen, P.J. Informa Healthcare, New York, 2008, pp. 271–286.
Frosch, J.-P. Lepoittevin (Eds.), Contact Dermatitis, Springer-Verlag, Berlin, 2011, [36] H.F. Frasch, A.M. Barbero, Steady-state flux and lag time in the stratum corneum
pp. 215–227. lipid pathway: results from finite element models, J. Pharm. Sci. 92 (2003)
[5] M. Schneider, F. Stracke, S. Hansen, U.F. Schaefer, Nanoparticles and their inter- 2196–2207.
actions with the dermal barrier, Dermatoendocrinol 1 (2009) 197–206. [37] A.M. Barbero, H.F. Frasch, Transcellular route of diffusion through stratum cor-
[6] R.V. Contri, L.A. Fiel, A.R. Pohlmann, S.S. Guterres, R.C.R. Beck, Transport of sub- neum: results from finite element models, J. Pharm. Sci. 95 (2006) 2186–2194.
stances and nanoparticles across the skin and in vitro models to evaluate skin [38] T.F. Wang, G.B. Kasting, J.M. Nitsche, A multiphase microscopic diffusion model
permeation and/or penetration, in: R.C.R. Beck, S.S. Guterres, A.R. Pohlmann for stratum corneum permeability. II. Estimation of physicochemical parame-
(Eds.), Nanocosmetics and Nanomedicines: New Approaches for Skin Care, ters, and application to a large permeability database, J. Pharm. Sci. 96 (2007)
Springer-Verlag, Berlin, 2011, pp. 3–36. 3024–3051.
[7] B.W. Barry, Novel mechanisms and devices to enable successful transdermal [39] P.V. Raykar, M.C. Fung, B.D. Anderson, The role of protein and lipid domains in the
drug delivery, Eur. J. Pharm. Sci. 14 (2001) 101–114. uptake of solutes by human stratum corneum, Pharm. Res. 5 (1988) 140–150.
[8] J. Hadgraft, Skin, the final frontier, Int. J. Pharm. 224 (2001) 1–18. [40] J.M. Nitsche, T.F. Wang, G.B. Kasting, A two-phase analysis of solute partitioning
[9] Y.G. Anissimov, O.G. Jepps, Y. Dancik, M.S. Roberts, Mathematical and pharmaco- into the stratum corneum, J. Pharm. Sci. 95 (2006) 649–666.
kinetic modelling of epidermal and dermal transport processes, Adv. Drug Del. [41] M.S. Roberts, S.E. Cross, Y.G. Anissimov, The skin reservoir for topically applied
Rev. 65 (2013) 169–190. solutes, in: R.L. Bronaugh, H.I. Maibach (Eds.), Percutaneous Absorption:
[10] K. Samuelsson, C. Simonsson, C.A. Jonsson, G. Westman, M.B. Ericson, A.T. Karlberg, Drugs, Cosmetics, Mechanisms, Methodology, Taylor & Francis, Boca Raton,
Accumulation of FITC near stratum corneum-visualizing epidermal distribution of a 2005, pp. 213–234.
strong sensitizer using two-photon microscopy, Contact Dermatitis 61 (2009) [42] W.A. Ritschel, A.S. Hussain, The principles of permeation of substances across
91–100. the skin, Methods Find. Exp. Clin. Pharmacol. 10 (1988) 39–56.
[11] K. Konig, A.P. Raphael, L. Lin, J.E. Grice, H.P. Soyer, H.G. Breunig, M.S. Roberts, [43] H.R. Jakubovic, A.B. Ackerman, Structure and function of skin: development,
T.W. Prow, Applications of multiphoton tomographs and femtosecond laser morphology, and physiology, in: S.L. Moschella, H.J. Hurley (Eds.), Dermatology,
nanoprocessing microscopes in drug delivery research, Adv. Drug Deliv. Rev. W.B. Saunders, Philadelphia, 1992, pp. 3–87.
63 (2011) 388–404. [44] B.P. Wenkers, B.C. Lippold, Skin penetration of nonsteroidal antiinflammatory
[12] M.S. Roberts, Y. Dancik, T.W. Prow, C.A. Thorling, L.L. Lin, J.E. Grice, T.A. drugs out of a lipophilic vehicle: influence of the viable epidermis, J. Pharm.
Robertson, K. Konig, W. Becker, Non-invasive imaging of skin physiology and Sci. 88 (1999) 1326–1331.
percutaneous penetration using fluorescence spectral and lifetime imaging [45] V.D. Bhatt, R.S. Soman, M.A. Miller, G.B. Kasting, Permeation of tecnazene
with multiphoton and confocal microscopy, Eur. J. Pharm. Biopharm. 77 (2011) through human skin in vitro as assessed by HS-SPME and GC–MS, Environ. Sci.
469–488. Technol. 42 (2008) 6587–6592.
[13] R.J. Scheuplein, I.H. Blank, Permeability of the skin, Physiol. Rev. 51 (1971) [46] P. Liu, W.I. Higuchi, W.Q. Song, T. Kurihara-Bergstrom, W.R. Good, Quantitative
702–747. evaluation of ethanol effects on diffusion and metabolism of beta-estradiol in
[14] K.A. Walters, M.S. Roberts, The structure and function of skin, in: K.A. Walters hairless mouse skin, Pharm. Res. 8 (1991) 865–872.
(Ed.), Dermatological and Transdermal Formulations, Marcel Dekker, New [47] T. Hikima, K. Yamada, T. Kimura, H.I. Maibach, K. Tojo, Comparison of skin distri-
York, 2002, pp. 1–41. bution of hydrolytic activity for bioconversion of beta-estradiol 17-acetate be-
[15] J.A. Bouwstra, J. Thewalt, G.S. Gooris, N. Kitson, A model membrane approach to tween man and several animals in vitro, Eur. J. Pharm. Biopharm. 54 (2002)
the epidermal permeability barrier: an X-ray diffraction study, Biochemistry 36 155–160.
(1997) 7717–7725. [48] C.M. Heard, B.V. Monk, A.J. Modley, Binding of primaquine to epidermal mem-
[16] L. Norlen, The physical structure of the skin barrier, in: M.S. Roberts, K.A. Walters branes and keratin, Int. J. Pharm. 257 (2003) 237–244.
(Eds.), Dermal Absorption and Toxicity Assessment, Informa Healthcare, New [49] K. Bleasby, J.C. Castle, C.J. Roberts, C. Cheng, W.J. Bailey, J.F. Sina, A.V. Kulkarni,
York, 2008, pp. 37–68. M.J. Hafey, R. Evers, J.M. Johnson, R.G. Ulrich, J.G. Slatter, Expression profiles of
[17] P.M. Elias, Epidermal lipids, barrier function, and desquamation, J. Invest. Dermatol. 50 xenobiotic transporter genes in humans and pre-clinical species: a resource
80 (1983) 44s–49s. for investigations into drug disposition, Xenobiotica 36 (2006) 963–988.
O.G. Jepps et al. / Advanced Drug Delivery Reviews 65 (2013) 152–168 167

[50] G.L. Flynn, S.H. Yalkowsky, Correlation and prediction of mass transport across [82] S.E. Cross, M.S. Roberts, Targeting local tissues by transdermal application: un-
membranes. I. Influence of alkyl chain length on flux-determining properties derstanding drug physicochemical properties that best exploit protein binding
of barrier and diffusant, J. Pharm. Sci. 61 (1972) 838–852. and blood flow effects, Drug Dev. Res. 46 (1999) 309–315.
[51] M.S. Roberts, R.A. Anderson, J. Swarbrick, D.E. Moore, The percutaneous absorp- [83] S.C. McNeill, R.O. Potts, M.L. Francoeur, Local enhanced topical delivery (LETD)
tion of phenolic compounds: the mechanism of diffusion across the stratum cor- of drugs: does it truly exist? Pharm. Res. 9 (1992) 1422–1427.
neum, J. Pharm. Pharmacol. 30 (1978) 486–490. [84] N.A. Monteiro-Riviere, A.O. Inman, J.E. Riviere, S.C. McNeill, M.L. Francoeur, Topical
[52] R.L. Cleek, A.L. Bunge, A new method for estimating dermal absorption from penetration of piroxicam is dependent on the distribution of the local cutaneous
chemical exposure. 1. General approach, Pharm. Res. 10 (1993) 497–506. vasculature, Pharm. Res. 10 (1993) 1326–1331.
[53] A.L. Bunge, R.L. Cleek, B.E. Vecchia, A new method for estimating dermal absorp- [85] S.E. Cross, Z. Wu, M.S. Roberts, The effect of protein binding on the deep tissue
tion from chemical exposure. 3. Compared with steady-state methods for pre- penetration and efflux of dermally applied salicylic acid, lidocaine and diazepam
diction and data analysis, Pharm. Res. 12 (1995) 972–982. in the perfused rat hindlimb, J. Pharmacol. Exp. Ther. 277 (1996) 366–374.
[54] S. Gregoire, C. Ribaud, F. Benech, J.R. Meunier, A. Garrigues-Mazert, R.H. Guy, [86] V. Bocci, G.P. Pessina, L. Paulesu, C. Nicoletti, The lymphatic route. VI. Distribu-
Prediction of chemical absorption into and through the skin from cosmetic tion of recombinant interferon-alpha 2 in rabbit and pig plasma and lymph,
and dermatological formulations, Br. J. Dermatol. 160 (2009) 80–91. J. Biol. Response Mod. 7 (1988) 390–400.
[55] M.B. Reddy, R.H. Guy, A.L. Bunge, Does epidermal turnover reduce percutaneous [87] A. Supersaxo, W.R. Hein, H. Steffen, Effect of molecular weight on the lymphatic
penetration? Pharm. Res. 17 (2000) 1414–1419. absorption of water-soluble compounds following subcutaneous administra-
[56] L. Simon, A. Goyal, Dynamics and control of percutaneous drug absorption in the tion, Pharm. Res. 7 (1990) 167–169.
presence of epidermal turnover, J. Pharm. Sci. 98 (2009) 187–204. [88] H. Yoshikawa, K. Takada, S. Muranishi, Molecular weight-dependent lymphatic
[57] E. Khalil, K. Kretsos, G.B. Kasting, Glucose partition coefficient and diffusivity in transfer of exogenous macromolecules from large intestine of renal insufficiency
the lower skin layers, Pharm. Res. 23 (2006) 1227–1234. rats, Pharm. Res. 9 (1992) 1195–1198.
[58] K. Walter, H. Kurz, Binding of drugs to human skin: influencing factors and the [89] S.E. Cross, M.S. Roberts, Subcutaneous absorption kinetics and local tissue
role of tissue lipids, J. Pharm. Pharmacol. 40 (1988) 689–693. distribution of interferon and other solutes, J. Pharm. Pharmacol. 45 (1993)
[59] S. Yagi, K. Nakayama, Y. Kurosaki, K. Higaki, T. Kimura, Factors determining drug 606–609.
residence in skin during transdermal absorption: studies on beta-blocking [90] S.E. Cross, M.S. Roberts, Dermal blood flow, lymphatics, and binding as determi-
agents, Biol. Pharm. Bull. 21 (1998) 1195–1201. nants of topical absorption, clearance, and distribution, in: J.E. Riviere (Ed.), Dermal
[60] R. Ibrahim, G.B. Kasting, Improved method for determining partition and diffu- Absorption Models in Toxicology and Pharmacology, Taylor & Francis, Boca Raton,
sion coefficients in human dermis, J. Pharm. Sci. 99 (2010) 4928–4939. 2006.
[61] S.E. Cross, B.M. Magnusson, G. Winckle, Y. Anissimov, M.S. Roberts, Determination [91] M. Schmuth, A.M. Ortegon, M. Mao-Qiang, P.M. Elias, K.R. Feingold, A. Stahl, Dif-
of the effect of lipophilicity on the in vitro permeability and tissue reservoir charac- ferential expression of fatty acid transport proteins in epidermis and skin ap-
teristics of topically applied solutes in human skin layers, J. Invest. Dermatol. 120 pendages, J. Invest. Dermatol. 125 (2005) 1174–1181.
(2003) 759–764. [92] C. Skazik, J. Wenzel, Y. Marquardt, A. Kim, H.F. Merk, D.R. Bickers, J.M. Baron,
[62] B.M. Magnusson, S.E. Cross, G. Winckle, M.S. Roberts, Percutaneous absorption P-glycoprotein (ABCB1) expression in human skin is mainly restricted to dermal
of steroids: determination of in vitro permeability and tissue reservoir charac- components, Exp. Dermatol. 20 (2011) 450–452.
teristics in human skin layers, Skin Pharmacol. Physiol. 19 (2006) 336–342. [93] K. Ito, H.T. Nguyen, Y. Kato, T. Wakayama, Y. Kubo, S. Iseki, A. Tsuji, P-glycopro-
[63] K. Sugibayashi, T. Hayashi, Y. Morimoto, Simultaneous transport and metabo- tein (Abcb1) is involved in absorptive drug transport in skin, J. Control. Release
lism of ethyl nicotinate in hairless rat skin after its topical application: the effect 131 (2008) 198–204.
of enzyme distribution in skin, J. Control. Release 62 (1999) 201–208. [94] K. Ito, Y. Kato, H. Tsuji, H.T. Nguyen, Y. Kubo, A. Tsuji, Involvement of organic
[64] P. Boderke, K. Schittkowski, M. Wolf, H.P. Merkle, Modeling of diffusion and con- anion transport system in transdermal absorption of flurbiprofen, J. Control.
current metabolism in cutaneous tissue, J. Theor. Biol. 204 (2000) 393–407. Release 124 (2007) 60–68.
[65] Q. Li, H. Tsuji, Y. Kato, Y. Sai, Y. Kubo, A. Tsuji, Characterization of the transder- [95] M.S. Roberts, S.E. Cross, A physiological pharmacokinetic model for solute dispo-
mal transport of flurbiprofen and indomethacin, J. Control. Release 110 (2006) sition in tissues below a topical below a topical application site, Pharm. Res. 16
542–556. (1999) 1392–1398.
[66] Y.Y. Grams, J.A. Bouwstra, Penetration and distribution in human skin focusing [96] P. Singh, M.S. Roberts, Dermal and underlying tissue pharmacokinetics of salicylic
on the hair follicle, in: R.L. Bronaugh, H.I. Maibach (Eds.), Percutaneous Absorp- acid after topical application, J. Pharmacokinet. Biopharm. 21 (1993) 337–373.
tion: Drugs–Cosmetics–Mechanisms–Methods, Taylor & Francis, Boca Raton, [97] P. Singh, H.I. Maibach, M.S. Roberts, Site of effects, in: M.S. Roberts, K.A. Walters
2005. (Eds.), Dermal Absorption and Toxicity Assessment, Marcel Dekker, New York,
[67] I.M. Braverman, A. Keh, D. Goldminz, Correlation of laser Doppler wave patterns 1998, pp. 353–370.
with underlying microvascular anatomy, J. Invest. Dermatol. 95 (1990) 283–286. [98] Y. Dancik, Y.G. Anissimov, O.G. Jepps, M.S. Roberts, Convective transport of high-
[68] I.M. Braverman, The cutaneous microcirculation, J. Investig. Dermatol. Symp. ly plasma protein bound drugs facilitates direct penetration into deep tissues
Proc. 5 (2000) 3–9. after topical application, Br J Clin Dermatol 73 (2012) 841–855.
[69] G. Cevc, U. Vierl, Spatial distribution of cutaneous microvasculature and local drug [99] Y.G. Anissimov, M.S. Roberts, Modelling dermal drug distribution after topical
clearance after drug application on the skin, J. Control. Release 118 (2007) 18–26. application in human, Pharm. Res. 28 (2011) 2119–2129.
[70] C.M. O'Driscoll, Anatomy and physiology of the lymphatics, in: W.N. Charman, [100] M.S. Roberts, M. Rowland, A dispersion model of hepatic elimination: 1. Formu-
V.J. Stella (Eds.), Lymphatics Transport of Drugs, CRC Press, Boca Raton, 1992. lation of the model and bolus considerations, J. Pharmacokinet. Biopharm. 14
[71] J.E. Riviere, B. Sage, P.L. Williams, Effects of vasoactive drugs on transdermal li- (1986) 227–260.
docaine iontophoresis, J. Pharm. Sci. 80 (1991) 615–620. [101] Y.G. Anissimov, A.J. Bracken, M.S. Roberts, Interconnected-tubes model of hepatic
[72] J.E. Riviere, N.A. Monteiro-Riviere, A.O. Inman, Determination of lidocaine con- elimination, J. Theor. Biol. 188 (1997) 89–101.
centrations in skin after transdermal iontophoresis: effects of vasoactive drugs, [102] K. Kretsos, G.B. Kasting, J.M. Nitsche, Distributed diffusion-clearance model
Pharm. Res. 9 (1992) 211–214. for transient drug distribution within the skin, J. Pharm. Sci. 93 (2004)
[73] P. Singh, M.S. Roberts, Effects of vasoconstriction on dermal pharmacokinetics 2820–2835.
and local tissue distribution of compounds, J. Pharm. Sci. 83 (1994) 783–791. [103] K. Kretsos, M.A. Miller, G. Zamora-Estrada, G.B. Kasting, Partitioning, diffusivity
[74] N. Borg, E. Gotharson, E. Benfeldt, L. Groth, L. Stahle, Distribution to the skin of and clearance of skin permeants in mammalian dermis, Int. J. Pharm. 346
penciclovir after oral famciclovir administration in healthy volunteers: compar- (2008) 64–79.
ison of the suction blister technique and cutaneous microdialysis, Acta Derm. [104] Y. Dancik, M.A. Miller, J. Jaworska, G.B. Kasting, Design and performance of a
Venereol. 79 (1999) 274–277. spreadsheet-based model for estimating bioavailability of chemicals from der-
[75] S.E. Cross, S.A. Megwa, H.A. Benson, M.S. Roberts, Self promotion of deep tissue mal exposure, Adv. Drug Del. Rev. 65 (2013) 221–236.
penetration and distribution of methylsalicylate after topical application, Pharm. [105] T. Peyret, P. Poulin, K. Krishnan, A unified algorithm for predicting partition co-
Res. 16 (1999) 427–433. efficients for PBPK modeling of drugs and environmental chemicals, Toxicol.
[76] K. Sugibayashi, G. Yanagimoto, T. Hayashi, T. Seki, K. Juni, Y. Morimoto, Analysis Appl. Pharmacol. 249 (2010) 197–207.
of skin disposition of flurbiprofen after topical application in hairless rats, J. Control. [106] H.M. Weiss, M. Fresneau, T. Moenius, A. Stuetz, A. Billich, Binding of pimecrolimus
Release 62 (1999) 193–200. and tacrolimus to skin and plasma proteins: implications for systemic exposure
[77] P. Boutsiouki, J.P. Thompson, G.F. Clough, Effects of local blood flow on the per- after topical application, Drug Metab. Dispos. 36 (2008) 1812–1818.
cutaneous absorption of the organophosphorus compound malathion: a micro- [107] C.M. Lee, H.I. Maibach, Deep percutaneous penetration into muscles and joints,
dialysis study in man, Arch. Toxicol. 75 (2001) 321–328. J. Pharm. Sci. 95 (2006) 1405–1413.
[78] G.F. Clough, P. Boutsiouki, M.K. Church, C.C. Michel, Effects of blood flow on the [108] R. Agarwal, O.P. Katare, S.P. Vyas, The pilosebaceous unit: a pivotal route for top-
in vivo recovery of a small diffusible molecule by microdialysis in human skin, ical drug delivery, Methods Find. Exp. Clin. Pharmacol. 22 (2000) 129–133.
J. Pharmacol. Exp. Ther. 302 (2002) 681–686. [109] S. Valiveti, J. Wesley, G.W. Lu, Investigation of drug partition property in artificial
[79] K. Higaki, M. Asai, T. Suyama, K. Nakayama, K. Ogawara, T. Kimura, Estimation of sebum, Int. J. Pharm. 346 (2008) 10–16.
intradermal disposition kinetics of drugs: II. Factors determining penetration of [110] C.C. Zouboulis, Sebaceous gland in human skin—the fantastic future of a skin
drugs from viable skin to muscular layer, Int. J. Pharm. 239 (2002) 129–141. appendage, J. Invest. Dermatol. 120 (2003) xiv–xv.
[80] C.J. Morgan, A.G. Renwick, P.S. Friedmann, The role of stratum corneum and der- [111] V.M. Meidan, M.C. Bonner, B.B. Michniak, Transfollicular drug delivery—is it a
mal microvascular perfusion in penetration and tissue levels of water-soluble reality? Int. J. Pharm. 306 (2005) 1–14.
drugs investigated by microdialysis, Br. J. Dermatol. 148 (2003) 434–443. [112] H. Schaefer, J. Lademann, The role of follicular penetration. A differential view,
[81] K. Higaki, K. Nakayama, T. Suyama, C. Amnuaikit, K. Ogawara, T. Kimura, En- Skin Pharmacol. Appl. Skin Physiol. 14 (Suppl. 1) (2001) 23–27.
hancement of topical delivery of drugs via direct penetration by reducing [113] G.F. Odland, Structure of the skin, in: L.A. Goldsmith (Ed.), Physiology, Morphology
blood flow rate in skin, Int. J. Pharm. 288 (2005) 227–233. and Molecular Biology of the Skin, Oxford University Press, New York, 1991, p. 110.
168 O.G. Jepps et al. / Advanced Drug Delivery Reviews 65 (2013) 152–168

[114] A.C. Lauer, Percutaneous drug delivery to the hair follicle, in: R.L. Bronaugh, H.I. [146] S. Sobue, K. Sekiguchi, H. Kikkawa, S. Irie, Effect of application sites and multiple
Maibach (Eds.), Percutaneous Absorption: Drugs–Cosmetics–Mechanisms– doses on nicotine pharmacokinetics in healthy male Japanese smokers following
Methods, Taylor & Francis, Boca Raton, 2005. application of the transdermal nicotine patch, J. Clin. Pharmacol. 45 (2005)
[115] R.A. Ellis, Vascular patterns in the skin, in: W. Montagna, R.A. Ellis (Eds.), Ad- 1391–1399.
vances in Biology of the Skin, Pergamon Press, New York, 1961, pp. 20–36. [147] S.A.M. Hotchkiss, Dermal metabolism, in: M.S. Roberts, K.A. Walters (Eds.), Dermal
[116] W. Montagna, A.M. Kligman, K.S. Carslile, Atlas of Normal Human Skin, Spring- Absorption and Toxicity Assessment, Marcel Dekker, New York, 1998, pp. 43–101,
er-Verlag, New York, 1992. New York.
[117] S.E. Cross, C. Anderson, M.S. Roberts, Topical penetration of commercial salicy- [148] Y. Dancik, C.A. Thorling, G. Krishnan, M.S. Roberts, Cutaneous metabolism and
late esters and salts using human isolated skin and clinical microdialysis studies, active transport in transdermal drug delivery, in: M.-R. N. (Ed.), Toxicology of
Br. J. Clin. Pharmacol. 46 (1998) 29–35. the Skin, Taylor & Francis, Boca Raton, 2010, pp. 69–82.
[118] K. Sato, R. Leidal, F. Sato, Morphology and development of an apoeccrine sweat [149] Y.-T. Wu, Y.G. Anissimov, M.S. Roberts, Introduction to dermatokinetics, in: N.
gland in human axillae, Am. J. Physiol. 252 (1987) R166–R180. Murthy (Ed.), Dermatokinetics of Therapeutic Agents, CRC Press Taylor & Francis,
[119] K. Wilke, R. Wepf, F.J. Keil, K.P. Wittern, H. Wenck, S.S. Biel, Are sweat glands an Boca Raton, 2011.
alternate penetration pathway? Understanding the morphological complexity [150] N. Seko, H. Bando, C.W. Lim, F. Yamashita, M. Hashida, Theoretical analysis of the
of the axillary sweat gland apparatus, Skin Pharmacol. Physiol. 19 (2006) 38–49. effect of cutaneous metabolism on skin permeation of parabens based on a
[120] H.J. Hurley, The eccrine sweat glands: structure and function, in: R.K. Freinkel, D.T. two-layer skin diffusion/metabolism model, Biol. Pharm. Bull. 22 (1999) 281–287.
Woodley (Eds.), The Biology of the Skin, Parthenon, New York, 2001, pp. 47–76. [151] R. Oliyai, V.J. Stella, Prodrugs of peptides and proteins for improved formulation
[121] R.J. Scheuplein, Mechanism of percutaneous absorption. II. Transient diffusion and delivery, Annu. Rev. Pharmacol. Toxicol. 33 (1993) 521–544.
and the relative importance of various routes of skin penetration, J. Invest. [152] G. Lefevre, G. Sedek, H.L. Huang, M. Saltzman, M. Rosenberg, B. Kiese, P.
Dermatol. 48 (1967) 79–88. Fordham, Pharmacokinetics of a rivastigmine transdermal patch formulation
[122] Y.Y. Grams, L. Whitehead, P. Cornwell, J.A. Bouwstra, Time and depth resolved in healthy volunteers: relative effects of body site application, J. Clin. Pharmacol.
visualisation of the diffusion of a lipophilic dye into the hair follicle of fresh 47 (2007) 471–478.
unfixed human scalp skin, J. Control. Release 98 (2004) 367–378. [153] N. Bodor, P. Buchwald, Soft drug design: general principles and recent applica-
[123] Y.Y. Grams, L. Whitehead, P. Cornwell, J.A. Bouwstra, On-line visualization of dye tions, Med. Res. Rev. 20 (2000) 58–101.
diffusion in fresh unfixed human skin, Pharm. Res. 21 (2004) 851–859. [154] A.L. Stinchcomb, P.W. Swaan, O. Ekabo, K.K. Harris, J. Browe, D.C. Hammell, T.A.
[124] A.C. Lauer, J.T. Elder, N.D. Weiner, Evaluation of the hairless rat as a model for in Cooperman, M. Pearsall, Straight-chain naltrexone ester prodrugs: diffusion and
vivo percutaneous absorption, J. Pharm. Sci. 86 (1997) 13–18. concurrent esterase biotransformation in human skin, J. Pharm. Sci. 91 (2002)
[125] H. Wu, C. Ramachandran, N.D. Weiner, B.J. Roessler, Topical transport of hydrophilic 2571–2578.
compounds using water-in-oil nanoemulsions, Int. J. Pharm. 220 (2001) 63–75. [155] H.D. Beall, K.B. Sloan, Topical delivery of 5-fluorouracil (5-FU) by 1,3-bisalkylcarbonyl-
[126] E. Bernard, J.L. Dubois, J. Wepierre, Importance of sebaceous glands in cutaneous 5-FU prodrugs, Int. J. Pharm. 231 (2002) 43–49.
penetration of an antiandrogen: target effect of liposomes, J. Pharm. Sci. 86 [156] H.J. Doh, W.J. Cho, C.S. Yong, H.G. Choi, J.S. Kim, C.H. Lee, D.D. Kim, Synthesis and
(1997) 573–578. evaluation of Ketorolac ester prodrugs for transdermal delivery, J. Pharm. Sci. 92
[127] U. Jacobi, M. Kaiser, W. Sterry, J. Lademann, Kinetics of blood flow after topical (2003) 1008–1017.
application of benzyl nicotinate on different anatomic sites, Arch. Dermatol. [157] K.C. Sung, J.Y. Fang, O.Y. Hu, Delivery of nalbuphine and its prodrugs across skin
Res. 298 (2006) 291–300. by passive diffusion and iontophoresis, J. Control. Release 67 (2000) 1–8.
[128] T. Ogiso, T. Shiraki, K. Okajima, T. Tanino, M. Iwaki, T. Wada, Transfollicular drug [158] R. Marks, Pharmacokinetics and safety review of tazarotene, J. Am. Acad. Dermatol.
delivery: penetration of drugs through human scalp skin and comparison of 39 (1998) S134–S138.
penetration between scalp and abdominal skins in vitro, J. Drug Target. 10 [159] P. Liu, W.I. Higuchi, A.H. Ghanem, W.R. Good, Transport of beta-estradiol in
(2002) 369–378. freshly excised human skin in vitro: diffusion and metabolism in each skin
[129] L.M. Lieb, A.P. Liimatta, R.N. Bryan, B.D. Brown, G.G. Krueger, Description of the layer, Pharm. Res. 11 (1994) 1777–1784.
intrafollicular delivery of large molecular weight molecules to follicles of human [160] A. Mavon, V. Raufast, D. Redoules, Skin absorption and metabolism of a new vi-
scalp skin in vitro, J. Pharm. Sci. 86 (1997) 1022–1029. tamin E prodrug, delta-tocopherol-glucoside: in vitro evaluation in human skin
[130] J. Lademann, H. Richter, U.F. Schaefer, U. Blume-Peytavi, A. Teichmann, N. models, J. Control. Release 100 (2004) 221–231.
Otberg, W. Sterry, Hair follicles—a long-term reservoir for drug delivery, Skin [161] R.L. Bronaugh, H.I. Maibach, Percutaneous absorption of nitroaromatic compounds:
Pharmacol. Physiol. 19 (2006) 232–236. in vivo and in vitro studies in the human and monkey, J. Invest. Dermatol. 84
[131] E.J. Cone, Mechanisms of drug incorporation into hair, Ther. Drug Monit. 18 (1985) 180–183.
(1996) 438–443. [162] R.L. Bronaugh, T.J. Franz, Vehicle effects on percutaneous absorption: in vivo and
[132] A. Kelch, S. Wessel, T. Will, U. Hintze, R. Wepf, R. Wiesendanger, Penetration in vitro comparisons with human skin, Br. J. Dermatol. 115 (1986) 1–11.
pathways of fluorescent dyes in human hair fibres investigated by scanning [163] R.C. Wester, H.I. Maibach, J. Melendres, L. Sedik, J. Knaak, R. Wang, In vivo and in
near-field optical microscopy, J. Microsc. 200 (2000) 179–186. vitro percutaneous absorption and skin evaporation of isofenphos in man, Fundam.
[133] H.L. Johnson, H.I. Maibach, Drug excretion in human eccrine sweat, J. Invest. Appl. Toxicol. 19 (1992) 521–526.
Dermatol. 56 (1971) 182–188. [164] D. Dick, K.M. Ng, D.N. Sauder, I. Chu, In vitro and in vivo percutaneous absorp-
[134] R.J. Jackson, W.B. Davis, A mathematical approach to the prediction of the rate of tion of 14C-chloroform in humans, Hum. Exp. Toxicol. 14 (1995) 260–265.
penetration of ions in the sweat duct, J. Theor. Biol. 97 (1982) 481–489. [165] I.P. Dick, P.G. Blain, F.M. Williams, The percutaneous absorption and skin distri-
[135] J. Lademann, H. Richter, N. Otberg, F. Lawrenz, U. Blume-Peytavi, W. Sterry, Appli- bution of lindane in man. II. In vitro studies, Hum. Exp. Toxicol. 16 (1997)
cation of a dermatological laser scanning confocal microscope for investigation in 652–657.
skin physiology, Laser Phys. 13 (2003) 756–760. [166] P. Griffin, H. Mason, K. Heywood, J. Cocker, Oral and dermal absorption of chlor-
[136] J. Lademann, N. Otberg, H. Richter, L. Meyer, H. Audring, A. Teichmann, S. Thomas, pyrifos: a human volunteer study, Occup. Environ. Med. 56 (1999) 10–13.
A. Knuttel, W. Sterry, Application of optical non-invasive methods in skin physiol- [167] P. Griffin, M. Payne, H. Mason, E. Freedlander, A.D. Curran, J. Cocker, The in vitro
ogy: a comparison of laser scanning microscopy and optical coherent tomography percutaneous penetration of chlorpyrifos, Hum. Exp. Toxicol. 19 (2000)
with histological analysis, Skin Res. Technol. 13 (2007) 119–132. 104–107.
[137] U. Jacobi, E. Waibler, P. Schulze, J. Sehouli, G. Oskay-Ozcelik, T. Schmook, W. [168] N.H. Cnubben, G.R. Elliott, B.C. Hakkert, W.J. Meuling, J.J. van de Sandt, Comparative
Sterry, J. Lademann, Release of doxorubicin in sweat: first step to induce the in vitro–in vivo percutaneous penetration of the fungicide ortho-phenylphenol,
palmar-plantar erythrodysesthesia syndrome? Ann. Oncol. 16 (2005) 1210–1211. Regul. Toxicol. Pharmacol. 35 (2002) 198–208.
[138] T. Schmook, U. Jacobi, J. Lademann, M. Worm, E. Stockfleth, Detection of doxorubicin [169] A. Mavon, C. Miquel, O. Lejeune, B. Payre, P. Moretto, In vitro percutaneous ab-
in the horny layer in a patient suffering from palmar-plantar erythrodysaesthesia, sorption and in vivo stratum corneum distribution of an organic and a mineral
Dermatology 210 (2005) 237–238. sunscreen, Skin Pharmacol. Physiol. 20 (2007) 10–20.
[139] A.S. Zelickson, Electron microscopic study of epidermal sweat duct, Arch. Dermatol. [170] A.F. Davis, Getting the dose right in dermatological therapy, in: K.A. Walter, M.S.
83 (1961) 106–111. Roberts (Eds.), Dermatologic, Cosmeceutic, and Cosmetic Development: Therapeu-
[140] K. Hashimoto, B.G. Gross, W.F. Lever, The ultrastructure of the skin of human tic and Novel Approaches, Informa Healthcare, New York, 2008, pp. 197–213.
embryos. I. The intraepidermal eccrine sweat duct, J. Invest. Dermatol. 45 [171] D. Mertin, B.C. Lippold, In-vitro permeability of the human nail and of a keratin
(1965) 139–151. membrane from bovine hooves: prediction of the penetration rate of antimyco-
[141] Y. Dancik, O.G. Jepps, M.S. Roberts, Physiologically based pharmacokinetics and tics through the nail plate and their efficacy, J. Pharm. Pharmacol. 49 (1997)
pharmacodynamics of skin, in: M.S. Roberts, K.A. Walters (Eds.), Dermal Absorp- 866–872.
tion and Toxicity Assessment, Informa Healthcare, New York, 2008, pp. 179–207. [172] B.P. Wenkers, B.C. Lippold, Prediction of the efficacy of cutaneously applied non-
[142] J.R. Varvel, S.L. Shafer, S.S. Hwang, P.A. Coen, D.R. Stanski, Absorption character- steroidal anti-inflammatory drugs from a lipophilic vehicle, Arzneimittel-
istics of transdermally administered fentanyl, Anesthesiology 70 (1989) forschung 50 (2000) 275–280.
928–934. [173] J.A. Cordero, M. Camacho, R. Obach, J. Domenech, L. Vila, In vitro based index of
[143] M.S. Roberts, W.A. Favretto, A. Meyer, M. Reckmann, T. Wongseelashote, Topical topical anti-inflammatory activity to compare a series of NSAIDs, Eur. J. Pharm.
bioavailability of methyl salicylate, Aust. N. Z. J. Med. 12 (1982) 303–305. Biopharm. 51 (2001) 135–142.
[144] E. Nakashima, P.K. Noonan, L.Z. Benet, Transdermal bioavailability and first-pass [174] H. Bando, S. Mohri, F. Yamashita, Y. Takakura, M. Hashida, Effects of skin metab-
skin metabolism: a preliminary evaluation with nitroglycerin, J. Pharmacokinet. olism on percutaneous penetration of lipophilic drugs, J. Pharm. Sci. 86 (1997)
Biopharm. 15 (1987) 423–437. 759–761.
[145] D.D. Tang-Liu, R.M. Matsumoto, J.I. Usansky, Clinical pharmacokinetics and drug [175] A.H. Beckett, J.W. Gorrod, D.C. Taylor, Comparison of oral and percutaneous routes
metabolism of tazarotene: a novel topical treatment for acne and psoriasis, Clin. in man for the systemic administration of ‘ephedrines’, J. Pharm. Pharmacol.
Pharmacokinet. 37 (1999) 273–287. 24 (1972) 65P–70P (Suppl.).

You might also like