1-s2.0-s0142961205005867-main

Upload: ymiyazy

Post on 13-Apr-2018

213 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/27/2019 1-s2.0-S0142961205005867-main

    1/11

    Biomaterials 27 (2006) 785795

    Mechanical properties of human stratum corneum: Effects of

    temperature, hydration, and chemical treatment

    Kenneth S. Wua, William W. van Osdolb, Reinhold H. Dauskardtc,

    aDepartment of Mechanical Engineering, Stanford University, Stanford, CA 94305, USAbALZA Corporation, Mountain View, CA 94039, USA

    cDepartment of Materials Science and Engineering, Stanford University, Stanford, 416 Escondido Mall, Bldg 550, Rm 550G, CA 94305-2205, USA

    Received 4 March 2005; accepted 24 June 2005

    Available online 10 August 2005

    Abstract

    An in vitro mechanics approach to quantify the intercellular delamination energy and mechanical behavior of isolated human

    stratum corneum (SC) in a direction perpendicular to the skin surface is presented. The effects of temperature, hydration, and a

    chloroformmethanol treatment to remove intercellular lipids were explored. The delamination energy for debonding of cells within

    the SC layer was found to be sensitive to the moisture content of the tissue and to the test temperature. Delamination energies for

    untreated stratum corneum were measured in the range of 18J/m2 depending on test temperature. Fully hydrated specimen

    energies decreased with increasing temperature, while room-humidity-hydrated specimens exhibited more constant values of 24 J/

    m2. Lipid-extracted specimens exhibited higher delamination energies of12J/m2, with values decreasing to 4J/m2 with increasingtest temperature. The peak separation stress decreased with increasing temperature and hydration, but lipid-extracted specimens

    exhibited higher peak stresses than untreated controls. The delaminated surfaces revealed an intercellular failure path with no

    evidence of tearing or fracture of cells. The highly anisotropic mechanical behavior of the SC is discussed in relation to the

    underlying SC structure.

    r 2005 Elsevier Ltd. All rights reserved.

    Keywords: Mechanical properties; Fracture toughness; Epithelial cell; Stratum corneum; Tissue treatment

    1. Introduction

    The layered construction of skin has components that

    possess mechanical properties needed to accommodate

    intrinsic and imposed mechanical stresses, abrasions and

    penetration of foreign objects under variable ambient

    moisture and temperature conditions. It must also haveother physical properties that resist the presence of toxic

    environmental chemicals, pathogens, and radiation [1].

    More severe environmental conditions may result from

    occlusion of the skin via application of adhesive

    dressings or transdermal drug delivery patches which

    can locally elevate moisture content and affect mechan-

    ical behavior. The detailed cellular and intercellular

    structures of the outermost layer of the epidermis, the

    stratum corneum (SC), have been widely studied [25].However, relatively few studies have examined the

    mechanical and fracture properties of SC to determine

    their dependence on tissue treatment and environmental

    conditions[611].In particular, surprisingly few studies

    have reported on the mechanical and delamination

    properties of the SC in the direction normal to the skin

    surface[1215].

    The SC consists of layered anucleated cells that

    mature and subsequently detach in the natural renewing

    ARTICLE IN PRESS

    www.elsevier.com/locate/biomaterials

    0142-9612/$- see front matter r 2005 Elsevier Ltd. All rights reserved.

    doi:10.1016/j.biomaterials.2005.06.019

    Abbreviations: SC, stratum corneum; RH, relative humidity; FH,

    fully hydrated; RHH, room-humidity hydrated; SS, stress separation;

    DCB, double-cantilever beam; CMT, chloroformmethanol treated;

    SEM, standard error of the meanCorresponding author. Tel.: +1 650725 0679;

    fax: +1 650725 4034.

    E-mail address: [email protected] (R.H. Dauskardt).

    http://www.elsevier.com/locate/biomaterialshttp://www.elsevier.com/locate/biomaterials
  • 7/27/2019 1-s2.0-S0142961205005867-main

    2/11

    process of desquamation. The disk-shaped SC cells, or

    corneocytes, composed largely of aligned keratin fila-

    ments, have been likened to bricks bound together by a

    lipid-rich mortar [16]. Intercellular lipids have been

    identified as the primary pathway for chemical diffusion

    and as the barrier to water permeability through the SC

    layer [1720]. In addition to the intercellular lipids,degraded desmosomal protein junctions, or corneo-

    somes, at the cell boundaries are central to SC cohesion

    and renewal, and their degradation is necessary for

    desquamation[2126]. Ultrastructural studies of the SC

    have suggested that increased moisture content is

    associated with elevated desmosome degradation and

    disruption of intercellular lipid structures, which affect

    both mechanical properties and permeability [25,27].

    Changes in delamination properties of SC with

    environmental conditioning have been simply demon-

    strated by a cellophane tape-stripping technique. SC

    hydrated for 24 h prior to tape stripping permits cells to

    be removed more easily, requiring fewer applications of

    tape to fully strip the SC from the epidermis [12].

    Another approach involved use of a cohesograph to

    measure SC cohesive strength[13]. Similar devices have

    been used to examine the bond strength of materials

    applied to the skin, such as adhesive dressings [14].

    While these techniques provide relative measures of

    SC delamination strength, several inherent problems

    exist that make the results qualitative and unreliable.

    Most significantly, the delamination of SC adhered to

    underlying tissue leads to a combined measurement of

    both SC and substrate properties. Additionally, the SC

    loading is nonuniform and can result in highly variableresults. In the case of the cohesograph, the force

    necessary to pull the SC apart may be affected

    substantially by the deformation behavior of the under-

    lying tissue substrate. The tape-stripping technique

    effectively peels the cell layers apart but, similar to the

    peel adhesion test for thin films on elastic substrates,

    provides qualitative results that are difficult to quantify

    [28]. There is clearly a need for quantitative test

    techniques to measure cohesion and strength properties

    of SC accurately and reproducibly.

    In the present study, a quantitative in vitro experi-

    mental mechanics approach to examine the SC inter-cellular delamination energy and out-of-plane

    mechanical behavior is presented. The delamination

    energy of human SC was examined as a function of

    selected testing temperatures and moisture precondi-

    tioning treatments and related to the underlying cellular

    structure. For comparison, human SC delipidized with

    a chloroformmethanol treatment was examined in the

    same manner. While some of the treatments and

    temperatures are nonphysiological, they provide insight

    into the microstructural mechanisms of SC cohesion and

    help to isolate the role of individual components of the

    SC such as intercellular lipids. SC delamination energy

    was quantified in terms of the energy required to

    propagate a debond through the SC layer and defined

    in terms of the strain energy release rate G, measured in

    units of J/m2. Stress-separation (SS) tests were per-

    formed to measure the out-of-plane mechanical beha-

    vior of the SC. Resulting failure surface morphologies

    were examined with scanning electron microscopy toprovide an indication of the delamination mechanisms.

    The delamination energy and strength property depen-

    dence on temperature and moisture are rationalized in

    terms of the underlying SC cell structure and inter-

    cellular lipids.

    2. Materials and methods

    2.1. Tissue preparation

    Human cadaver SC tissue was isolated for these experiments

    from three female Caucasian donors, 7688 years of age, from

    the thigh, abdomen or lower back. For each study as detailed

    in the following sections, comparative tests were performed on

    the same donor tissue to avoid variability between donors. The

    SC tissue was separated from the underlying epidermis via a

    trypsin enzymatic digest, then stored at 4 1C in a fully hydrated

    (FH) state on water-moistened filter paper (Grade 595

    General-Purpose Filter Paper, Schleicher & Schuell Micro-

    Science GmbH, Dassel, Germany). For the untreated SC, two

    sets of tissue were prepared. One set consisted of SC that was

    FH at 100% relative humidity (RH) by storing on water-

    moistened filter paper, and the other was room-humidity

    hydrated (RHH) at 45% RH. Both sets were allowed to

    equilibrate for at least 24 h. These initial hydration and testingconditions represent significantly different equilibrium SC

    water contents corresponding to 300400% wt/wt and5%wt/wt SC water content, respectively[29,30].The thickness of

    untreated RHH SC was measured with a micrometer to be

    between 1535mm. The FH SC thickness was observed to

    increase by 30% similar to values reported by others [31].

    Additional tissue,60 60mm2, was delipidized with a 120-min 30 mL chloroform:methanol (2:1 by volume) soak with

    two subsequent 30 min 30 mL water rinses. The SC thickness

    was not observed to change significantly with treatment.

    Chemically treated SC for SS tests was treated similarly, but

    with a shortened 60-min chloroformmethanol soak. The

    different treatment time is not believed to influence the lipid

    extraction significantly. Studies with a similar chloroformmethanol treatment have shown that the majority of lipids are

    removed within the first 15 min[32]. The treated SC was either

    fabricated into FH test specimens immediately after treatment

    or allowed to dehydrate in an ambient 45% RH environmentsimilar to the untreated SC.

    2.2. Delamination energy measurements

    The delamination energy of the SC tissue was examined

    using fracture mechanics techniques developed to measure the

    adhesive properties of highly viscoelastic pressure-sensitive

    adhesives [33]. Similar techniques have been employed to

    ARTICLE IN PRESS

    K.S. Wu et al. / Biomaterials 27 (2006) 785795786

  • 7/27/2019 1-s2.0-S0142961205005867-main

    3/11

    measure the adhesive properties of polymer bone cements and

    polymer layers in elastic substrates [3436]. The technique

    involves sandwiching the SC between two elastic polycarbo-

    nate substrates (Hyzods GP, Sheffield Plastics Inc., Sheffield,

    MA) with a cyanoacrylate adhesive (Instant Krazyglue Gel,

    Elmers Products Inc., Columbus, OH) to form fracture-

    mechanics-based double-cantilever beam (DCB) specimens

    (Fig. 1(a)). Cyanoacrylate adhesive polymerization is readilyinitiated by the presence of small amounts of water on the

    bonding surfaces limiting the adhesive to the SC exterior. The

    transparent polycarbonate beams facilitate optical inspection

    of the inner sandwich structure during specimen preparation

    and testing. To enable the use of linear elastic fracture

    mechanics to determine the strain energy release rates,

    substrate dimensions were chosen to ensure purely elastic

    deformation of the substrates during testing [33,34].

    To fabricate the specimens, a thin layer of cyanoacrylate

    adhesive was applied to one face of a nominally

    40 10 2.88 mm3 polycarbonate substrate, leaving a 710-mm region of the beam end uncoated. The substrate was

    pressed against the SC on the filter paper backing and a scalpelwas used to cut around the substrate to detach the adhered SC

    from surrounding tissue. In the case of RHH SC, the

    freestanding film was prepared similarly with a sheet of paper

    as a backing. To form the final sandwich structure, another

    substrate coated with adhesive in the same manner was pressed

    against the SC face of the complimentary beam with adhesive-

    free ends aligned. Excess adhesive along the sandwich edges

    was removed with a scalpel to ensure that the two halves of the

    sandwich structure were bound together by SC only.

    DCB specimens containing either untreated or chloroform

    methanol-treated (CMT) SC were placed in an environmental

    chamber (Model LH-6, Associated Environmental Systems,

    Ayer, MA) at selected temperature (10, 25, 75 1C) and RH

    (45, 85% RH) conditions and allowed to equilibrate for

    10 min. Tests involving FH specimens were conducted in an85% RH environment with RHH specimens examined at 45%

    RH unless specified otherwise. The specimens were loaded via

    attached loading tabs to propagate a debond within the SC

    layers. The specimens were tested in a custom-built mechanical

    test system with a computer-controlled DC servoelectric

    actuator operated in displacement control. Tests were per-formed at a constant displacement rate of 2 mm/s. Correspond-

    ing loads were measured using a 222 N load cell. The

    delamination length, a, was measured from recorded load

    displacement,PD, and their elastic compliance relationship:

    C DP

    23

    a 0:64h 3E0I

    , (1)

    where I bh3=12, and C is the specimen compliance, Pis theload, D=2 is the corresponding displacement of each beamfrom its original position at the loading point, E0 E=1 n2is the plane strain Youngs modulus for the polycarbonate, n is

    Poissons ratio, I is the area moment of inertia, b is the

    polycarbonate substrate width, and h is the height of eachpolycarbonate beam.

    By measuring the critical load, P, and the delamination

    length, a, at incipient crack extension, the delamination

    resistance, Gc, was determined from critical values of the

    strain energy release rate, G[33,34,37]:

    G 12P2a2

    b2h3E0 1

    ffiffiffi5

    p

    2

    h

    a 1

    2

    h

    a

    2 !. (2)

    Multiple delamination energies, Gc, were measured for each

    DCB specimen by recording the critical loads, Pc, and

    associated delamination lengths, a, during delamination

    extension (Fig. 2). For the present specimens, the values of

    E and n for the polycarbonate were 2.379 GPa and 0.38,respectively. Given the thin film nature of the SC compared to

    the massive polycarbonate substrate, the contribution of the

    ARTICLE IN PRESS

    StratumCorneum

    P

    Delamination

    a

    P

    Polycarbonate

    Polycarbonate

    b

    h

    h

    b

    h

    P

    P

    StratumCorneum

    Polycarbonate

    b

    h

    (a)

    (b)

    Fig. 1. Fracture mechanics specimen geometries. (a) DCB geometry

    illustrating relevant loading parameters (P,D), delamination extension

    as measured from loading axis points (a) and relevant specimen

    dimensions (b 10mm, h 2:88 mm, length 40 mm). (b) SS speci-men configuration showing loading axis and relevant dimensions

    (b

    10mm, h

    2:88mm).

    0 10 200.0

    1.0

    2.0

    3.0

    Delamination

    Energy,

    Gc

    (J/m2)

    Crack Extension, a (mm)

    Fig. 2. Typical delamination energy as a function of delamination

    extension. Illustration of the variation in delamination energy (Gc) as a

    function of crack length (Da) for a RHH (45% RH) SC specimentested in a DCB configuration at 10 1C, 45% RH.

    K.S. Wu et al. / Biomaterials 27 (2006) 785795 787

  • 7/27/2019 1-s2.0-S0142961205005867-main

    4/11

    elastic strain energy in the SC layer can be ignored in the

    analysis[28].

    2.3. Stress-separation measurements

    Structures containing SC sandwiched between polycarbo-

    nate substrates were fabricated in a similar manner to the DCB

    specimens (Fig. 1(b)). The nominal substrate dimensions were

    10 10 2.88 mm3. The specimens were loaded normal to theSC face via attached loading tabs in an environmental

    chamber with controlled temperature and humidity.

    SS measurements were performed under the same condi-

    tions and SC hydrations as those for the DCB experiments.

    Testing was conducted at a constant displacement rate of

    1mm/s, yielding a machine compliance-corrected initial strain

    rate of0.007 s1 in the SC for an SC thickness of 10 mm. Thespecimens were allowed to equilibrate for20 min prior totesting.

    2.4. Scanning electron microscopy

    Both DCB and SS specimens were examined after mechan-

    ical testing using a scanning electron microscope (Hitachi

    S-2500, Hitachi, Tokyo, Japan) to characterize the fracture

    surface morphologies. Selected specimens were allowed to dry

    in ambient conditions (25 1C, 45% RH), gold or gold-palladium coated, then examined in the electron microscope

    operated at 15 kV. Multiple specimens from each testing

    condition were inspected to ensure representative character-

    ization.

    2.5. Statistical analysis

    Delamination energies are presented as mean 71.96 standard error of the mean (SEM) in which the mean values

    reported are expected to fall within these bounds with 95%

    confidence. On average, n 38 for each test condition.Delamination energies were compared using the Wilcoxon

    test for independent samples. Statistical significance was set at

    1% or 5% as specified in the figures. SS measurements are

    presented as mean7standard deviation (SD). Further statis-

    tical analysis was not performed due to the small sample sizes

    for each test condition (n 324).

    3. Results

    3.1. Quantifying delamination energies

    The delamination energy, Gc, measured as a function

    of the delamination length, Da, for a specimen contain-

    ing RHH SC is shown inFig. 2. The data represent a so-

    called delamination resistance curve, and were ob-

    tained from testing of a single DCB specimen. The

    variability in the measured Gcvalues as seen in the curve

    was predominantly related to inhomogeneities in the SC

    along the DCB length. Typically, four cantilever speci-

    mens were tested for each condition, each yielding

    approximately 712 data points. All data points

    (navg 38) for a given test condition were subsequentlyaveraged to produce the delamination resistance values,

    Gc, represented in Figs. 3 and 4. Note that the

    delamination tip strain rates do not vary significantlyas a function ofDa during such tests, as determined by

    modeling and experimental examination of similar tests

    on pressure-sensitive adhesives [38]. This reduces the

    effects of varying strain rate on the measured delamina-

    tion energies.

    3.2. Delamination energy variation with temperature and

    hydration

    Delamination energy values, Gc, for specimens con-

    taining RHH and FH SC measured at selected

    temperatures are presented inFigs. 3(a) and 4(a). Each

    ARTICLE IN PRESS

    00.0

    1.0

    2.0

    3.0

    4.0

    5.0

    6.0

    +

    ++*

    *

    DelaminationEnergy,

    Gc

    (J/m2)

    Temperature (C)

    0.0

    0.2

    0.4

    0.6

    0.8

    1.0

    45% R.H.

    100% R.H.

    PeakStress,

    (

    MPa)

    Temperature (C)

    10 20 30 40 50 60 70 80

    0 10 20 30 40 50 60 70 80

    45% R.H.

    100% R.H.

    (a)

    (b)

    Fig. 3. The variation in delamination energy and peak stress with

    temperature and hydration for untreated SC. (a) Delamination energy

    values from DCB experiments performed at selected temperatures and

    hydrations (Po0:01 except *Po0:05, +P40:05). (b) Peak stressesfrom SS experiments similarly performed at selected temperatures and

    hydrations. Error bars: (A) mean71.96 SEM and (B) mean7SD.

    K.S. Wu et al. / Biomaterials 27 (2006) 785795788

  • 7/27/2019 1-s2.0-S0142961205005867-main

    5/11

    plot represents data from a single SC donor to mitigate

    tissue specimen variations. SS experiments were per-

    formed on the same SC as that in Fig. 3(a). The resultsare shown in Fig. 3(b). Further DCB tests were

    performed on the same SC as that in Fig. 4(a) after

    chloroformmethanol treatment. These results are pre-

    sented inFig. 4(a). With increasing test temperature, the

    untreated FH specimens exhibited markedly lower

    delamination energies. In contrast, the RHH specimen

    energies were not statistically different with increasing

    test temperature, except between 10 and 25 1C for results

    shown in Fig. 4(a). Note that the Gc values for the

    untreated 45% RH specimens in Fig. 3(a) were

    significantly higher than the untreated 45% RH speci-

    mens of Fig. 4(a). The reasons for these differences

    remain unclear. One possible explanation is that SC

    specimens from different donors possess variable

    properties that elicit different delamination energy

    measurements.

    3.3. Comparison to stress-separation results

    The results of SS tests performed on RHH and FH

    specimens from the same SC as that represented in

    Fig. 3(a) and under the same environmental testing

    conditions are presented in Fig. 3(b). Typically, 34

    measurements were taken at each testing condition, with

    each measurement requiring one specimen. SS tests

    resulted in specimen loading up to a peak stress, after

    which the stress decayed rapidly with continued

    displacement. Peak stress values obtained from the SS

    curves are shown in Fig. 3(b). The peak stresses were

    observed to decrease with increasing hydration and

    testing temperature, exhibiting similar behavior to the

    delamination energy trends seen for the DCB specimens.

    3.4. Comparison to delipidized stratum corneum

    Delamination in the CMT SC specimens exhibited

    markedly different behavior compared to the untreated

    specimens. The corresponding data are shown in

    Fig. 4(a). Most notably, at each test temperature the

    measured delamination energies were significantly high-

    er than those of their untreated counterparts and

    proportionally smaller differences were observed be-

    tween the delamination energies of the RHH and FH

    specimens. In contrast to the untreated specimens,statistically significant decreases in Gc values were

    observed for the highest test temperature compared

    to the lower test temperatures regardless of initial

    hydration.

    A comparison between untreated and CMT SS

    specimens reveals similar trends to those observed

    during delamination testing. Delipidized SS specimens

    exhibited generally higher peak stresses compared to

    their untreated counterparts as shown inFig. 4(b). The

    significant scatter in the untreated control specimens in

    Fig. 4(b) was not characteristic of most of the SS tests

    performed. These specimens possessed the same loadingcharacteristics with loading to a peak stress, followed by

    significant load decreases with increasing displacement.

    Unlike the control group, the treated specimen peak

    stresses exhibited the same trend with testing tempera-

    ture regardless of hydration.

    3.5. Examination of delamination surface morphologies

    Representative scanning electron micrographs of the

    fracture surfaces of DCB specimens containing FH and

    RHH hydrated SC tested at 25 1C are shown inFig. 5.

    Similar micrographs were obtained for such specimens

    ARTICLE IN PRESS

    0.0

    4.0

    8.0

    12.0

    16.0

    ****

    ++++

    +

    +*

    *

    DelaminationEnergy,

    Gc

    (J/m2)

    Temperature (C)

    45% R.H.

    100% R.H.

    CMT 45% R.H.

    CMT 100% R.H.

    0.0

    0.5

    1.0

    1.5

    2.0

    45% R.H.

    100% R.H.

    CMT 45% R.H.

    CMT 100% R.H.

    PeakStress,

    (

    MPa)

    Temperature (C)

    0 10 20 30 40 50 60 70 80

    0 10 20 30 40 50 60 70 80

    (a)

    (b)

    Fig. 4. Comparison of untreated and chloroformmethanol-extracted

    SC. (a) Delamination energy measured in DCB tests performed

    on untreated and CMT SC at selected temperature (10, 25, 75 1C)

    and hydration (45, 100% RH) (Po0:01 except *Po0:05,+,++,z,**P40:05). (b) Peak stress data from SS measurementsconducted at the same temperatures and hydrations for both untreated

    and CMT SC. Data offset for clarity. Error bars: (a) mean71.96-

    SEM and (b) mean7SD.

    K.S. Wu et al. / Biomaterials 27 (2006) 785795 789

  • 7/27/2019 1-s2.0-S0142961205005867-main

    6/11

    at each temperature. For the FH specimens, individual

    SC cells are difficult to discern and surface features

    were characterized by tortuous surface undulations

    (Fig. 5(a)). The RHH specimen micrograph reveals a

    different surface morphology compared to FH fracture

    surfaces, showing partial pull-up of individual SC cells.

    This produced fracture surface roughness characterized

    by individual cellular features (Fig. 5(b)). The micro-

    graphs of delipidized specimens revealed the same

    difference in fracture surface appearance between

    RHH and FH specimens. Fracture surfaces of SS

    specimens yielded morphologies similar to those seen

    in the DCB specimens in which the RHH specimens

    exhibited SC cell pull-up, while the FH specimens lacked

    similar cellular features.

    4. Discussion

    4.1. Origin of delamination energies

    As far as fracture properties of materials are

    concerned, the SC delamination energy values are

    comparatively low. Polymer layers bonded weakly to

    elastic substrates may exhibit similar values, but with

    more strongly covalently bonded interfaces, fracture

    energies increase significantly above 10 J/m2 [35,36,39].

    In most of these cases, plastic deformation at the cracktip contributes significantly to the delamination fracture

    energy. During delamination of the SC layer, energy is

    dissipated by separation of intercellular boundaries,

    plastic deformation of the SC layer, and by the work

    done during cell pull-up.

    To determine the extent of possible plastic deforma-

    tion near the crack tip, plastic zone radius estimates, rp,

    were obtained using the well-known plane strain plastic

    zone expression:

    rp 16p

    GcE0

    s2ys

    , (3)

    where the plane strain modulus, E0, and the yieldstrength, sys, are for the SC layer. The values ofE

    0 wereobtained from the SS initial loading slopes and found to

    be in the range 125MPa. We note that there isconsiderable experimental uncertainty in the values ofE0

    estimated. These are due to inherent difficulties in

    determining the through thickness strain in the thin SC

    layer, and values should be treated accordingly. Thesysvalues were taken as the peak stresses in the SS tests as

    no obvious yielding was apparent prior to the peak. The

    resulting plastic zone size estimates were 212mm,

    corresponding to an average of 25% of the original

    SC thickness, suggesting that only a fraction of the SC

    thickness may have undergone plastic deformation. For

    all of the tissue conditions examined, the failure path

    occurred between corneocytes. Given the averagecorneocyte thickness of0.51mm, multiple layers ofcells on either side of the delamination may have

    undergone plastic deformation. However, energy dis-

    sipation by plastic deformation in a plastic zone

    represents only one possible energy dissipation mechan-

    ism. Alternatively, the cohesive properties of the cell

    boundaries themselves may dominate the energy dis-

    sipation with minimal general plastic deformation. This

    has been substantiated by studies of graded delamina-

    tion properties in which delamination energies increased

    towards the lower SC as upper layers were removed[40].

    Plasticity-dominated dissipation would suggest theopposite trend of decreasing delamination energy with

    decreasing SC thickness.

    Examination of the delamination surfaces of the DCB

    and SS specimens supports the idea that changes in the

    intercellular cohesive properties play a central role in

    determining delamination energy. The scanning electron

    micrographs reveal significant differences between the

    RHH and FH specimens (Fig. 5), with both untreated

    and delipidized SC showing similar topographic char-

    acteristics. Notably, cell pull-up was present in the RHH

    specimens (Fig. 5(b)), and similar results were observed

    for the CMT specimens. The FH specimen delamination

    ARTICLE IN PRESS

    Fig. 5. DCB specimen delamination surface morphology. Scanning electron micrographs of DCB specimens containing (a) FH (100% RH), and (b)

    RHH (45% RH) SC. Specimens tested at 25 1C corresponding to delamination energies reported inFig. 4(a). Images representative of DCB specimen

    delamination surfaces regardless of treatment.

    K.S. Wu et al. / Biomaterials 27 (2006) 785795790

  • 7/27/2019 1-s2.0-S0142961205005867-main

    7/11

    surfaces presented no obvious cell pull-up and indivi-

    dual corneocytes were not easily discernible (Fig. 5(a)).

    Similar features were observed for CMT SC.

    Consistent with the small variations inGcvalues with

    temperature for RHH-untreated specimens (Figs. 3(a)

    and 4(a)), corresponding fracture surface morphologies

    exhibited no discernible differences in cell pull-updensity with temperature. However, no obvious mor-

    phological changes were observed for the FH-untreated

    specimens despite large variations in Gc values with

    temperature (Figs. 3(a) and 4(a)). These observations

    suggest that the delamination energy must be more

    intimately tied to the cohesive properties of the

    intercellular boundaries of the SC and not to the failure

    morphology. This notion is supported by the similar

    trends exhibited by the peak stress data. Similar to the

    untreated specimens, the delipidized SC fracture surface

    morphologies remained unchanged as a function of

    temperature for each hydration condition despite

    significantGc variations with temperature.

    4.2. Environmental effects on SC structure

    The marginal changes in delamination energy and

    peak cohesive strength with increasing test temperature

    exhibited by the untreated RHH SC compared to that of

    the FH specimens indicate that SC cohesive strength

    depends strongly on initial hydration. The commonly

    observed failure of SC between corneocytes highlights

    the relevance of the intercellular space to SC fracture

    properties[22]. In a simplistic model, SC cells are viewed

    as highly keratinized bricks surrounded by a lipidmortar (Fig. 6) [16]. This notion led to the proposal

    that SC cohesive behavior is largely influenced by the

    intercellular lipid characteristics[16,41,42,].

    Additional studies probing the effect of hydration and

    chemical treatment provide further insights. The micro-

    structure of FH human SC has been examined using

    freeze-fracture electron microscopy to reveal the swel-

    ling of corneocytes, as well as the presence of water

    pools in the intercellular spaces [30,43]. Other studies

    have indicated that water disrupts the lipid lamellae to

    varying degrees and that intercellular corneosomes

    become degraded with time in the presence of water

    [25,27,30]. Hydration also has been shown to influencelamellar lipid spacing in hairless mouse SC [44].

    Examination of the affect of hydration on the lipid

    orthorhombic to hexagonal phase transition tempera-

    ture near 35 1C has revealed small decreases in transition

    temperature with increasing hydration, suggesting that

    hydration helps to fluidize the SC lipids [30]. However,

    the precise effects of water on lipid ordering within the

    lamellae remain undetermined with current evidence

    suggesting that the lipids remain relatively unperturbed

    by changes in SC hydration[30,45,46,]. The separation

    of corneocyte interfaces by the presence of intercellular

    water, particularly in conjunction with increases in

    temperature, may be the cause of decreases in delamina-

    tion energy and peak stress values for FH untreated SC

    as seen inFigs. 3 and 4.

    While the precise effects of hydration on lipid

    ordering remain unclear, temperature has been observed

    to affect the ordering of intercellular lipids by inducing

    lipid disorder and phase changes. Specifically, heating-

    induced alkyl-chain disordering of porcine SC lipids has

    been correlated with increased permeability of water

    through the SC membrane [47]. In addition, transmis-

    sion electron diffraction has revealed gradual changes in

    human SC lipid organization from an orthorhombic to

    hexagonal to fluid phase when heated from 20 to 90 1C,representing changes toward a more disordered state

    [48,49]. Differential scanning calorimetry of human SC

    reveals major transitions near 70 and 80 1C, which have

    been deduced to reflect intercellular lamellar lipid

    melting then subsequent dissociation of lipidprotein

    complexes formed at corneocyte envelope interfaces,

    respectively [5052]. Despite such transitions, abrupt

    changes in delamination energy with temperature are

    not necessarily expected due to a competition between

    decreased lipid cohesion and increased plastic dissipa-

    tion from greater lipid fluidity. The lack of substantial

    changes in delamination energy for RHH specimens asseen inFigs. 3(a) and 4(a)as well as unpublished results

    examining a larger temperature range from 10 to 90 1C

    further corroborate this notion.

    The lack of drastic changes in delamination energies

    and peak stress values for the RHH SC compared to

    FH SC indicate that lipid disordering is not enough to

    weaken the intercellular structure significantly (Figs. 3

    and 4). Increased hydration in concert with increasing

    temperature seems to play a key role in reducing

    delamination energy and peak stress values. In patho-

    logic skin states including atopic dermatitis and lamellar

    ichthyosis, impaired barrier properties and a prevalence

    ARTICLE IN PRESS

    Fig. 6. Illustration of SC delamination front and microstructural

    components. Schematic of the SC during delamination, illustrating

    intercellular failure and showing cellular structure, including aligned

    keratin filaments as well as lipid intercellular space with corneosomes.

    K.S. Wu et al. / Biomaterials 27 (2006) 785795 791

  • 7/27/2019 1-s2.0-S0142961205005867-main

    8/11

    of hexagonal and fluid lipid phases suggest that these

    lipid regimes may be associated with higher diffusive

    mobility of substances through the SC [49]. At inter-

    mediate temperatures where multiple phases coexist,

    certain substances exhibit increased SC penetration

    attributed to diffusion-enhancing phase boundaries

    [45,49,53]. As temperature is increased, the lipid phasetransitions likely enable additional diffusion paths that

    allow water molecules to penetrate and separate

    interfaces otherwise inaccessible at lower temperatures

    consistent with our observed decreases in delamination

    energy for FH SC and the lack of large changes for the

    RHH SC.

    Given the implied significance of the lipid structures

    on the cohesive behavior of the SC, removal of the

    intercellular lipids remains important to understanding

    their role in SC delamination properties. The inter-

    cellular domain has been noted as a pathway for

    chemical diffusion, as observed in transepidermal water

    loss experiments and other chemical species tests

    [17,20,54,55]. Studies have shown that isolated and fully

    delipidized non-plantar SC cells require lipids for

    aggregation and proper cohesion [42]. Furthermore,

    mechanical tests on individual SC cells have yielded

    substantially higher modulus values compared to

    coherent SC tissue, highlighting the effect of intercel-

    lular constituents on SC mechanical behavior[56]. Still,

    understanding of the contribution of the intercellular

    lipids to the cohesive strength of the SC remains

    incomplete [54,57,58]. The chloroformmethanol lipid

    extractions used in the present study have been noted to

    remove key lipids believed to be responsible forregulation of desquamation, in particular, cholesterol

    sulfate, whose excess is partly responsible for ichthyosis

    [5860]. However, such chemical treatment does not

    lead to cell dissociation as observed here and by others

    [60]. Indeed, the measured delamination energy for the

    treated specimens was substantially higher compared to

    the untreated controls (Fig. 4(a)). SS tests on the treated

    specimens also resulted in higher peak stresses (Fig.

    4(b)). Similar observations have been noted by others

    using less quantitative cohesometry techniques[22].

    The effects of chloroformmethanol lipid extraction

    on the SC reveal marked modifications to the inter-cellular space. Treatment with similar chloroform

    methanol solutions has shown via thin-layer chromato-

    graphy that the majority of the lipid lamellae are leached

    from the specimen[32]. The treatment results in intimate

    contact between unextracted lipids covalently bound to

    the cornified envelopes of adjacent corneocytes and

    seemingly leaves corneosomes unaltered[22,6164]. The

    increased cohesion of delipidized SC has been attributed

    to the interaction between remaining lipids in opposing

    corneocyte envelopes and postulated to result from

    interdigitation of the opposing envelope lipids [64,65].

    Supporting this notion, extraction of the lamellae and

    envelope lipids leads to dissociation of SC tissue into

    individual cells [42]. While lipid extraction significantly

    increased delamination energies compared to those of

    untreated SC, the resulting fracture surfaces exhibited

    similar features dependent on initial tissue hydration

    (Fig. 5). Even with these differences in failure surface

    morphologies, the chemically treated RHH and FHspecimens possessed delamination energy values and

    peak stresses proportionately more similar than their

    untreated counterparts, further highlighting the impor-

    tance of intercellular lipids on SC cohesive properties

    (Fig. 4).

    In addition to intercellular lipid contributions to SC

    integrity, the corneosome protein linkages between cells

    are known to play a critical role in SC cohesion.

    Corneosome degradation is accelerated with increasing

    SC hydration [25]. Chapman et al. have shown that

    adhesion testing on porcine SC reveals a gradient in SC

    intercellular cohesion with progressively weaker bond-

    ing from the interior toward the more superficial SC

    layers[22]. These results were correlated with a gradient

    in corneosome areal density, in which the number of

    corneosomes between adjacent SC layers, or nonper-

    ipheral corneosomes, was observed to decrease progres-

    sively from the inner to outer cell layers[21,22]. Similar

    observations have been made in human SC [26].

    However, the DCB test results reveal more complex

    trends than simply that increased hydration leads to

    decreased cohesion (Figs. 3 and 4). It is likely that the

    delamination energies are not correlated highly with

    expected corneosome cohesive contributions due to the

    location of the delamination in the few outer layers ofthe SC, as determined by examination of the graded

    properties through the thickness of the SC[40].

    4.3. Comparison of out-of-plane to in-plane mechanical

    studies of stratum corneum

    The reported in-plane properties of SC are signifi-

    cantly different from the out-of-plane behavior pre-

    sented in this study. These differences reveal the highly

    anisotropic nature of the SC composite. With respect to

    elastic behavior, in-plane tensile tests on newborn rat SC

    led to measurements of decreasing tensile moduliranging from8800 to 12 MPa with increasing humid-ity, while similar tests on human SC revealed moduli

    decreasing from80 to 20 MPa, which we calculatedassuming a SC thickness of 10 mm since only force, not

    stress, data were reported[6,9]. Our own initial in-plane

    tensile studies have yielded moduli decreasing from

    1000 to 5 MPa with increasing hydration. Similardecreases in modulus for rat SC have been reported with

    increasing temperature [9]. These variations have been

    associated with the in-plane orientation of the inter-

    mediate filaments in the keratinized corneocytes and the

    substitution of existing proteinprotein hydrogen bonds

    ARTICLE IN PRESS

    K.S. Wu et al. / Biomaterials 27 (2006) 785795792

  • 7/27/2019 1-s2.0-S0142961205005867-main

    9/11

    with water-mediated bonding to facilitate greater fiber

    mobility[7,9,66]. Associated changes in the intercellular

    lipids with temperature and hydration may also affect

    in-plane tensile moduli. In FH SC tensile tests,

    corneocytes were observed to slide past each other,

    highlighting the importance of intercellular components.

    Compared to our human SC in-plane modulusmeasurements from1000 to 5 MPa with increasinghydration, initial out-of-plane modulus measurements

    yielded values in the range of 125 MPa, substantiallylower than those of the in-plane values. Interestingly,

    while SC lipid transitions have been observed to be

    largely hydration independent, in the present study

    mechanical properties are sensitive to the presence of

    water to facilitate decreases in structural integrity and

    cohesive strength, albeit for different reasons out-of-

    plane and in-plane[30,45,46,66].

    The out-of-plane SC peak cohesive strengths were

    also found to be substantially reduced compared to in-

    plane values. For human SC, in-plane strength values

    have been reported to decrease from 22 to 5 MPa whentesting in 0100% RH environments with ether-treated

    specimens yielded elevated strengths ranging from25to 19 MPa in the same environments [6]. Our prelimin-

    ary in-plane tensile measurements gave similar results

    with strengths ranging from18 to 2 MPa in testingenvironments of 10100% RH. In contrast, the out-of-

    plane strengths are significantly lower in the range

    0.10.8 MPa (Figs. 3(b) and 4(b)). The CMT SCspecimens exhibited somewhat higher out-of-plane

    strength values of

    0.41.4 MPa, but still significantly

    lower than in-plane values. The lower out-of-planestrength values are not unexpected, given the continual

    renewal of the SC which sheds corneocytes perpendi-

    cular to the skin surface.

    Only limited work to measure in-plane fracture

    properties of human SC as a function of varying

    environmental conditions has been reported. Mean

    fracture energies of 3600 J/m2 for 76% RH-condi-tioned SC specimens have been measured in a tearing

    configuration [10]. Subsequent work suggested that

    delamination energy increases with increasing hydration

    [11]. This trend was not observed in the delamination

    results of the present study. The in-plane fracture energyis expected to be higher than that of the out-of-plane

    direction, where lower cohesive values of18 J/m2 areneeded to facilitate natural desquamation. The magni-

    tude of the difference is, however, surprising. Peripheral

    corneosomes that help to connect the cells in-plane may

    contribute to the high toughness values. Note, however,

    that the in-plane tearing energies reported above were

    calculated assuming linear elastic behavior. Viscoplastic

    behavior of the SC, particularly at higher hydrations,

    and the unconstrained nature of the tearing configura-

    tion suggest that the reported in-plane fracture energies

    should be treated with caution.

    5. Conclusions

    In conclusion, a fracture-mechanics-based approach

    using DCB specimens has been developed to examine

    quantitatively the delamination properties of human SC.

    In conjunction with out-of-plane SS measurements, a

    decrease in delamination energy with increasing testtemperature was observed in FH (100% RH) SC tissue.

    Little effect of testing temperature was observed for

    RHH (45% RH) SC tissue. The observed decrease in

    delamination energy was associated with a reduction in

    SC cohesive strength. In comparison, delamination

    energy values of CMT specimens exhibited little hydra-

    tion dependence and were higher than those of

    untreated controls, as explained by corneocyte envelope

    interactions. Further exploration is necessary to under-

    stand the microstructural changes that occur with

    delipidization, particularly in conjunction with tempera-

    ture variation. Initial measurements of modulus reveal

    that SC stiffness in this mode is much lower than that

    reported in-plane. From comparisons to in-plane

    experiments, the fracture energy and modulus data

    reveal that a simple bricks and mortal model requires

    refinement to explain the highly anisotropic mechanical

    behavior exhibited by SC.

    Acknowledgements

    The authors would like to thank Dr. Jay Audett and

    the ALZA Corporation for funding in support of the

    research, and Ms. Nieves Crisologo and Ms. MrudulaPatel from ALZA Corporation for their assistance in

    providing the stratum corneum. K.W. was supported by

    a Stanford Graduate Fellowship.

    References

    [1] Thody AJ, Friedmann PS. Functions of the skin. In: Thody AJ,

    Friedmann PS, editors. Scientific basis of dermatology: a

    physiological approach. Edinburgh, London: Churchill Living-

    stone Inc.; 1986. p. 15.

    [2] Montagna W, Lobitz Jr. WC, editors. The epidermis. New York:

    Academic Press, 1964.

    [3] Wilkes GL, Brown IA, Wildnauer RH. The biomechanicalproperties of skin. CRC Crit Rev Biomed Eng 1973;1:45395.

    [4] Marks R, Plewig G, editors. Stratum corneum. Berlin: Springer;

    1983.

    [5] Harding CR. The stratum corneum: structure and function in

    health and disease. Dermatol Ther 2004;17:615.

    [6] Wildnauer RH, Bothwell JW, Douglass AB. Stratum corneum

    biomechanical properties. I. Influence of relative humidity on

    normal and extracted human stratum corneum. J Invest Dermatol

    1971;56:728.

    [7] Wilkes GL, Wildnauer RH. Structureproperty relationships of

    the stratum corneum of human and neonatal rat. II. Dynamic

    mechanical studies. Biochim Biophys Acta 1973;304:27689.

    [8] Wolfram MA, Wolejsza NF, Laden K. Biomechanical properties

    of delipidized stratum corneum. J Invest Dermatol 1973;59:4216.

    ARTICLE IN PRESS

    K.S. Wu et al. / Biomaterials 27 (2006) 785795 793

  • 7/27/2019 1-s2.0-S0142961205005867-main

    10/11

    [9] Papir YS, Hsu K-H, Wildnauer RH. The mechanical properties of

    the stratum corneum. I. The effect of water and ambient

    temperature on the tensile properties of newborn rat stratum

    corneum. Biochim Biophys Acta 1975;399:17080.

    [10] Koutroupi KS, Barbenel JC. Mechanical and failure behavior of

    the stratum corneum. J Biomech 1990;23:2817.

    [11] Nicolopoulos CS, Giannoudis PV, Glaros KD, Barbenel JC. In

    vitro study of the failure of skin surface after influence ofhydration and preconditioning. Arch Dermatol Res 1998;290:

    63840.

    [12] Weigand D, Gaylor JR. Removal of stratum corneum in vivo: an

    improvement on the cellophane tape stripping technique. J Invest

    Dermatol 1973;60:847.

    [13] Marks R, Nicholls S, Ritzbeorge D. Measurement of intracorneal

    cohesion in man using in vivo techniques. J Invest Dermatol

    1977;69:299302.

    [14] Dykes PJ, Heggie R, Hill SA. Effects of adhesive dressings on the

    stratum corneum of the skin. J Wound Care 2001;10:710.

    [15] Wu KS, Van Osdol WW, Dauskardt RH. Mechanical and

    microstructural properties of stratum corneum. In: McKittrick J,

    Aizenberg J, Orme C, Vekilov P, editors. Biological and

    biomimetic materialsProperties to function. Materials Research

    Society Proceedings, 2002 April 15, San Francisco, CA, USA.Warrendale, PA: Materials Research Society; 2002. v. 724.

    p. 2733.

    [16] Elias PM, Grayson S, Lampe MA, Williams ML, Brown BE. The

    intercorneocyte space. In: Marks R, Plewig G, editors. Stratum

    corneum. Berlin: Springer-Verlag; 1983. p. 5367.

    [17] Friedmann PS. The skin as a permeability barrier. In: Thody AJ,

    Friedmann PS, editors. Scientific basis of dermatology: a

    physiological approach. Edinburgh, London: Churchill Living-

    stone, Inc.; 1986. p. 2635.

    [18] Simonetti O, Hoogstraate AJ, Bialik W, Kempenaar JA,

    Schrijvers AH, Bodde HE, Ponec M. Visualization of diffusion

    pathways across the stratum corneum of native and in-vitro-

    reconstructed epidermis by confocal laser scanning microscopy.

    Arch Dermatol Res 1995;287:46573.

    [19] Johnson ME, Mitragotri S, Patel A, Blankschtein D, Langer R.

    Synergistic effects of chemical enhancers and therapeutic ultra-

    sound on transdermal drug delivery. J Pharm Sci 1996;85:6709.

    [20] Fartasch M. Epidermal barrier in disorders of the skin. Microsc

    Res Tech 1997;38:36172.

    [21] Chapman SJ, Walsh A. Desmosomes, corneosomes and desqua-

    mation An ultrastructural study of adult pig epidermis. Arch

    Dermatol Res 1990;282:30410.

    [22] Chapman SJ, Walsh A, Jackson SM, Friedmann PS. Lipids,

    proteins and corneocyte adhesion. Arch Dermatol Res

    1991;283:16773.

    [23] Walsh A, Chapman SJ. Sugars protect desmosome and corneo-

    some glycoproteins from proteolysis. Arch Dermatol Res 1991;

    283:1749.

    [24] Lundstro m A, Serre G, Haftek M, Egelrud T. Evidence for a roleof corneodesmosin, a protein which may serve to modify

    desmosomes during cornification in stratum corneum cell cohe-

    sion and desquamation. Arch Dermatol Res 1994;286:36975.

    [25] Rawlings A, Harding C, Watkinson A, Banks J, Ackerman C,

    Sabin R. The effect of glycerol and humidity on desmosome

    degradation in stratum corneum. Arch Dermatol Res 1995;287:

    45764.

    [26] Simon M, Bernard D, Minondo A-M, Camus C, Fiat F, Corcuff

    P, Schmidt R, Serrre G. Persistence of both peripheral and non-

    peripheral corneodesmosomes in the upper stratum corneum of

    winter xerosis skin versus only peripheral in normal skin. J Invest

    Dermatol 2001;116:2330.

    [27] Warner RR, Boissy YL, Lilly NA, Spears MJ, McKillop K,

    Marshall JL, Stone KJ. Water disrupts stratum corneum lipid

    lamellae: damage is similar to surfactants. J Invest Dermatol

    1999;113:9606.

    [28] Dauskardt RH, Lane M, Ma Q, Krishna N. Adhesion and

    debonding of multi-layer thin film structures. Eng Fract Mech

    1998;61:14162.

    [29] Takahashi M, Machida Y. The influence of hydroxy acids on the

    rheological properties of stratum corneum. J Soc Cosmet Chem

    1985;36:17787.[30] Bouwstra JA, de Graff A, Gooris GS, Nijsse J, Wieches JW, van

    Aelst AC. Water distribution and related morphology in human

    stratum corneum at different hydration levels. J Invest Dermatol

    2003;120:7508.

    [31] Norle n L, Emilson A, Forslind B. Stratum corneum swelling.

    Biophysical and computer assisted quantitative assessments. Arch

    Dermatol Res 1997;289:50613.

    [32] Sznitowska M, Janicki S, Williams A, Lau S, Sto"yhwo A. pH-

    induced modifications to stratum corneum lipids investigated

    using thermal, spectroscopic, and chromatographic techniques.

    J Pharm Sci 2003;92:1739.

    [33] Taub MB, Dauskardt RH. Adhesion of pressure sensitive

    adhesives with applications in transdermal drug delivery. In:

    Mallapragada S, Korsmeyer R, Mathiowitz E, Narasimhan B,

    Tracy M, editors. Biomaterials for drug delivery and tissueengineering. Materials Research Society Proceedings,

    2000 November 27December 1, Boston, MA, USA. Warren-

    dale, PA: Materials Research Society; 2001. v. 662.

    p. NN4.9.1NN4.9.6.

    [34] Ohashi KL, Romero AC, McGowan PD, Maloney WJ, Daus-

    kardt RH. Adhesion and reliability of interfaces in cemented total

    joint arthroplasties. J Orthopaed Res 1998;16:70514.

    [35] Hohlfelder RJ, Maidenberg DA, Dauskardt RH, Wei Y,

    Hutchinson JW. Adhesion of benzocyclobutene-passivated silicon

    in epoxy layered structures. J Mater Res 2001;16:24355.

    [36] Snodgrass JM, Pantelidis D, Jenkins ML, Bravman JC, Daus-

    kardt RH. Subcritical debonding of polymer/silica interfaces

    under monotonic and cyclic loading. Acta Mater 2002;50:

    2395411.

    [37] Kanninen MF. An augmented double cantilever beam model for

    studying crack propagation and arrest. Int J Fract 1973;9:8392.

    [38] Taub MB, Dauskardt RH. Unpublished results. Stanford

    University; 2004.

    [39] Jenkins ML, Dauskardt RH, Bravman JC. Important factors for

    silane adhesion promoter efficacy: surface coverage, functionality,

    and chain length. J Adhesion Sci Tech 2004;18:1483588.

    [40] Wu KS, Dauskardt RH. Unpublished results. Stanford Uni-

    versity; 2005.

    [41] Williams ML, Elias PM. Stratum corneum lipids in disorders of

    cornification. Increased cholesterol sulfate content of stratum

    corneum in recessive xlinked ichthyosis. J Clin Invest 1981;

    68:140410.

    [42] Smith WP, Christensen MS, Nacht S, Gans EH. Effect of lipids

    on the aggregation and permeability of human stratum corneum.J Invest Dermatol 1982;78:711.

    [43] van Hal DA, Jeremiasse E, Junginger HE, Spies F, Bouwstra JA.

    Structure of fully hydrated human stratum corneum: a freeze

    fracture electron microscopy study. J Invest Dermatol 1996;

    106:8995.

    [44] Ohta N, Ban S, Tanaka H, Nakata S, Hatta I. Swelling of

    intercellular lipid lamellar structure with short repeat distance in

    hairless mouse stratum corneum as studied by X-ray diffraction.

    Chem Phys Lipids 2003;123:18.

    [45] Bouwstra JA, Gooris GS, Bras W, Downing DT. Lipid

    organization in pig stratum corneum. J Lipid Res 1995;36:68595.

    [46] Mak VH, Potts RO, Guy RH. Does hydration affect intercellular

    lipid organization in the stratum corneum? Pharm Res

    1991;8:10645.

    ARTICLE IN PRESS

    K.S. Wu et al. / Biomaterials 27 (2006) 785795794

  • 7/27/2019 1-s2.0-S0142961205005867-main

    11/11

    [47] Potts RO, Francoeur ML. Lipid biophysics of water loss through

    the skin. Proc Natl Acad Sci USA 1990;87:38713.

    [48] Pilgram GS, Engelsma-van Pelt AM, Bouwstra JA, Koerten HK.

    Electron diffraction provides new information on human stratum

    corneum lipid organization studied in relation to depth and

    temperature. J Invest Dermatol 1999;113:4039.

    [49] Pilgram GS, Vissers DC, van der Meulen H, Pavel S, Lavrijsen

    SP, Bouwstra JA, Koerten HK. Aberrant lipid organization instratum corneum of patients with atopic dermatitis and lamellar

    ichthyosis. J Invest Dermatol 2001;117:7107.

    [50] Van Duzee BF. Thermal analysis of human stratum corneum.

    J Invest Dermatol 1975;65:4048.

    [51] Golden GM, Guzek DB, Harris RR, McKie JE, Potts RO. Lipid

    thermotropic transitions in human stratum corneum. J Invest

    Dermatol 1986;86:2559.

    [52] Gay CL, Guy RH, Golden GM, Mak VHW, Francoeur ML.

    Characterization of low-temperature (i.e., o651C) lipid transi-

    tions in human stratum corneum. J Invest Dermatol 1994;

    103:2339.

    [53] Ogiso T, Hirota T, Iwaki M, Hina T, Tanino T. Effect of

    temperature on percutaneous absorption of terodiline, and

    relationship between penetration and fluidity of the stratum

    corneum lipids. Int J Pharm 1998;176:6372.[54] Madison K. Barrier function of the skin: la raison detre of the

    epidermis. J Invest Dermatol 2003;121:23141.

    [55] Yu B, Kim KH, So PT, Blankschtein D, Langer R. Visualization

    of oleic acid-induced transdermal diffusion pathways using two-

    photon fluorescence microscopy. J Invest Dermatol 2003;120:

    44855.

    [56] Leveque JL, Poelman MC, de Rigal J, Kligman AM. Are

    corneocytes elastic? Dermatologica 1988;176:659.

    [57] Elias PM. Epidermal lipids, barrier function, and desquamation.

    J Invest Dermatol 1983;80:44s9s.

    [58] Egelrud T, Lundstro m A. The dependence of detergent-induced

    cell dissociation in non-palmo-plantar stratum corneum on

    endogenous proteolysis. J Invest Dermatol 1990;95:4569.

    [59] Long SA, Wertz PW, Strauss JS, Downing DT. Human stratum

    corneum polar lipids and desquamation. Arch Dermatol Res

    1985;277:2847.[60] Takahashi M, Aizawa M, Miyazawa K, Machida Y. Effects of

    surface active agents on stratum corneum cell cohesion. J Soc

    Cosmet Chem 1987;38:218.

    [61] Swartzendruber DC, Wertz PW, Madison KC, Downing DT.

    Evidence that the corneocyte has a chemically bound lipid

    envelope. J Invest Dermatol 1987;88:70913.

    [62] Wertz PW, Swartzendruber DC, Madison KC, Downing DT.

    Composition and morphology of epidermal cyst lipids. J Invest

    Dermatol 1987;89:41925.

    [63] Wertz PW, Madison KC, Downing DT. Covalently bound lipid

    of human stratum corneum. J Invest Dermatol 1989;92:10911.

    [64] Wertz PW, Swartzendruber DC, Kitko DJ, Madison KC,

    Downing DT. The role of the corneocyte lipid envelopes in

    cohesion of the stratum corneum. J Invest Dermatol 1989;

    93:16972.[65] Stewart ME, Downing DT. The o-hydrocyceramides of pig

    epidermis are attached to corneocytes solely through o-hydroxyl

    groups. J Lipid Res 2001;42:110510.

    [66] Wilkes GL, Nguyen A-L, Wildnauer R. Structureproperty

    relations of human and neonatal rat stratum corneum. I. Thermal

    stability of the crystalline lipid structure as studied by X-ray

    diffraction and differential thermal analysis. Biochim Biophys

    Acta 1973;304:26775.

    ARTICLE IN PRESS

    K.S. Wu et al. / Biomaterials 27 (2006) 785795 795