Advance Drug Delivery Reviews

Embed Size (px)

Citation preview

  • 7/28/2019 Advance Drug Delivery Reviews

    1/22

    Nanoparticles and microparticles for skin drug delivery

    Tarl W. Prow a, Jeffrey E. Grice a, Lynlee L. Lin a, Rokhaya Faye a, Margaret Butler c, Wolfgang Becker d,Elisabeth M.T. Wurm e, Corinne Yoong e, Thomas A. Robertson a,b, H. Peter Soyer e, Michael S. Roberts a,b,a The University of Queensland, School of Medicine, Therapeutics Research Centre, Brisbane, QLD, Australiab The University of South Australia, Therapeutics Research Centre, School of Pharmacy and Medical Science, Adelaide, SA, Australiac Australian Institute for Bioengineering and Nanotechnology, The University of Queensland, Brisbane, QLD, Australiad Becker & Hickl, GmbH, Berlin, Germanye Dermatology Research Centre, The University of Queensland, School of Medicine, Princess Alexandra Hospital, Brisbane, QLD, Australia

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

    Article history:

    Received 21 September 2010

    Accepted 31 January 2011

    Available online 23 February 2011

    Keywords:

    Nanoparticle

    Drug delivery

    Topical

    Skin

    Percutaneous penetration

    Transdermal delivery

    Skin is a widely used route of delivery for local and systemic drugs and is potentially a route for their delivery

    as nanoparticles. The skin provides a natural physical barrier against particle penetration, but there are

    opportunities to deliver therapeutic nanoparticles, especially in diseased skin and to the openings of hair

    follicles. Whilst nanoparticle drug delivery has been touted as an enabling technology, its potential in treating

    local skin and systemic diseases has yet to be realised. Most drug delivery particle technologies are based on

    lipid carriers, i.e. solid lipid nanoparticles and nanoemulsions of around 300 nm in diameter, which are now

    considered microparticles. Metal nanoparticles are now recognized for seemingly small drug-like

    characteristics, i.e. antimicrobial activity and skin cancer prevention. We present our unpublished clinical

    data on nanoparticle penetration and previously published reports that support the hypothesis that

    nanoparticles N10 nm in diameter are unlikely to penetrate through the stratum corneum into viable human

    skin but will accumulate in the hair follicle openings, especially after massage. However, significant uptake

    does occur after damage and in certain diseased skin. Current chemistry limits both atom by atom

    construction of complex particulates and delineating their molecular interactions within biological systems.

    In this review we discuss the skin as a nanoparticle barrier, recent work in the field of nanoparticle drug

    delivery to the skin, and future directions currently being explored. 2011 Elsevier B.V. All rights reserved.

    Contents

    1. Skin as a site for particle delivery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 471

    1.1. Skin as a natural particle barrier . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 471

    1.2. Barrier properties of the skin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 472

    1.2.1. Skin structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 472

    1.2.2. The stratum corneum barrier . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 473

    1.2.3. Transport routes of exogenous substances across the stratum corneum . . . . . . . . . . . . . . . . . . . . . . . . . . . 473

    1.2.4. Acid mantle maintains the stratum corneum barrier . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 473

    1.2.5. Viable epidermis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 473

    1.2.6. Hair follicles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4741.3. Barrier properties of diseased skin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 474

    1.4. Skin flexing and massage for enhanced delivery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 474

    2. Small drug particle delivery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 475

    2.1. Anti-inflammatory drugs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 475

    2.1.1. Corticosterone . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 475

    2.1.2. Flufenamic acid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 476

    Advanced Drug Delivery Reviews 63 (2011) 470491

    This review is part of the Advanced Drug Delivery Reviews theme issue on Nanodrug Particles and Nanoformulations for Drug Delivery.

    Corresponding author at: The University of Queensland, Therapeutics Research Unit, School of Medicine, Princess Alexandra Hospital, Woolloongabba, QLD, 4102, Australia.

    Tel.: +61 7 3176 2546.

    E-mail address: [email protected] (M.S. Roberts).

    0169-409X/$ see front matter 2011 Elsevier B.V. All rights reserved.

    doi:10.1016/j.addr.2011.01.012

    Contents lists available at ScienceDirect

    Advanced Drug Delivery Reviews

    j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / a d d r

    http://dx.doi.org/10.1016/j.addr.2011.01.012http://dx.doi.org/10.1016/j.addr.2011.01.012http://dx.doi.org/10.1016/j.addr.2011.01.012mailto:[email protected]://dx.doi.org/10.1016/j.addr.2011.01.012http://www.sciencedirect.com/science/journal/0169409Xhttp://www.sciencedirect.com/science/journal/0169409Xhttp://dx.doi.org/10.1016/j.addr.2011.01.012mailto:[email protected]://dx.doi.org/10.1016/j.addr.2011.01.012
  • 7/28/2019 Advance Drug Delivery Reviews

    2/22

    2.2. Anti-photoageing drugs and antioxidants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 478

    2.2.1. Retinoids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 478

    2.2.2. Tocopheryl acetate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 479

    2.2.3. Multiphoton imaging for in vivo lipid nanoparticle delivery studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . 479

    3. Other topical nanoparticle drug delivery applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 479

    3.1. Antimicrobial agents. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 479

    3.1.1. Antifungal drugs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 479

    3.1.2. Silver nanoparticles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 479

    3.2. Anti-proliferative agents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 483

    3.2.1. 5-aminolevulinic acid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4833.2.2. Podophyllotoxin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 484

    3.3. Other topical particulate delivered drugs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 484

    4. Recent advances in particle-based drug delivery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 485

    5. Systemic safety . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 486

    6. Limitations of nanoparticle and nanocarrier chemical methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 486

    7. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 486

    Acknowledgments. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 487

    References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 487

    1. Skin as a site for particle delivery

    The theory and practical aspects of percutaneous penetration of

    drugs, particulate material and contaminants have been covered in a

    number of excellent reference texts. For comprehensive treatment of

    the topics in this section, the reader is referred to chapters by Roberts

    et al. [1], Norlen [2], Monteiro-Riviere and Baroli [3], Mller et al. [4]

    and other relevant chapters in these books.

    Whilst the skin has historically been used for the topical delivery of

    compounds, it is only since the 1970s with the advent of transdermal

    patches that it has widely been used as a route for systemic delivery [5].

    Nanoparticle delivery to the skin is being increasingly used to facilitate

    local therapies. The nanoparticle definition designated by the National

    Nanotechnology Initiative has been adopted by the American National

    Standards Institute as particles with all dimensions between 1 nm and

    100 nm [6,7]. Fig. 1 shows that the potential sites for targeting

    nanoparticles include the surface of the skin, furrows, and hair follicles.

    A recent review by Baroli discusses nanoparticle penetration largely

    from the skin structure perspective.The title of this reviewPenetration

    of Nanoparticles and Nanomaterials in the Skin: Fiction or Reality?

    highlights the ongoing debate of nanoparticle penetration [8]. This

    debate is fuelled by the need for more rigorous, multidisciplinary

    approaches to shed light on mechanisms of particle penetration and

    interactionswith skin. Likewise,Schneider et al. make a compelling case

    in their review on nanoparticle skin interactions for more rigorous

    studies on the effects of hydration and mechanical stress on skin with

    regard to nanoparticleskin interactions [9].

    Particles can interact with skin at a cellular level as adjuvants. This

    nanoparticleskin interaction can be used to enhance immune

    reactivity for topical vaccine applications [10]. Another example of

    nanoparticleskin interactions the topical use of silver nanoparticles

    as over the counter antimicrobial agents [11], where the nanoparticles

    provide a slow release of silver ions that have wound healing and

    antimicrobial properties. The silver ions released from nanoparticle

    can inhibit microbial proliferation, but also accelerate wound healing.

    This controlled release of silver ions while the nanoparticles remainon the skin surface highlights one of the most successful topical

    nanoparticle drug delivery strategies.

    Generally, the promise of nanoparticle-mediated drug delivery into

    the epidermis and dermis without barrier modification has met with

    little success. Where the barrier is compromised, however, such as in

    aged or diseased skin, there may be potential for enhanced particle

    penetration. Ulcerated squamous cell carcinoma is one example. The

    opportunities and obstacles for nanoparticle drug delivery are only just

    beginning to be explored in ongoing clinical trials. For instance,

    capsaicin loaded nanoparticles are being used to treat the pain

    associated with diabetic neuropathy [12]. Advances in particle engi-

    neering, formulation science and an improved understanding of

    nanoparticleskin interactions will undoubtedly lead to important

    clinically relevant improvements in topicaldrug delivery. An overridingconcern is the safety of any applied nanoparticle, recognising the

    possibility that non-biodegradable nanoparticles could be taken up and

    retained by the reticulo-endothelial system [13]. Inaddition, there isthe

    potential for local toxicity, shown by a recent report of nanoparticles

    inducing keratinocyte apoptosis, wherethe subtlerelationship between

    longer and shorter phosphatidylcholine chain lengths makes the

    difference between life and death for keratinocytes [14]. This illustrates

    the need for toxicity monitoring in vitro and especially in vivo.

    1.1. Skin as a natural particle barrier

    It is generally recognised that the potential for nanoparticle and

    microparticle skin penetration is negligible [15]. Most environmental

    Fig. 1. Sites in skin for nanoparticle delivery. Topical nanoparticle drug delivery takes

    place in three major sites: stratum corneum (SC) surface (panel a), furrows

    (dermatoglyphs) (panel b), and openings of hair follicles (infundibulum) (c). The

    nanoparticles are shown in green and the drug in red. Other sites for delivery are the

    viable epidermis (E) and dermis (D).

    471T.W. Prow et al. / Advanced Drug Delivery Reviews 63 (2011) 470 491

  • 7/28/2019 Advance Drug Delivery Reviews

    3/22

    nanoparticles, be they viruses, bacteria, dust, allergens or materials,

    do not penetrate human skin unless the skin barrier is disrupted

    [13,16]. Topical drug delivery with nanoparticles, with the aim of

    targeting the nanoparticles into the deeper layers of skin, therefore

    has to overcome a significant barrier honed over millions of years of

    evolution. Human papilloma virus (HPV) is perhaps the best example

    of a nanoparticle able to penetrate barrier compromised skin to cause

    skin warts [17]. Today, biomimetic nanoparticle engineering is

    applying these principles from nature to develop better nanoparticledelivery systems [18] and progress is being made in the area of

    nanoparticle interactions with skin [9]. HPV is also in the same size

    range [19] as topical nanoparticles for drug delivery in skin [20] and

    both utilize lipids to facilitate payload delivery (Fig. 2).

    The benefits of drug delivery resulting from topical penetration of

    nanoparticles are offset by their potential for toxicity, as discussed

    above. This has led to the topic of environmental and synthetic

    particle penetration across human skin becoming a contentious

    subject that has attracted the attention of regulatory agencies around

    the world.

    The likelihood of nanoparticle penetration across the skin has

    recently been reviewed by the Scientific Committee on Consumer

    Products (SCCP) who conclude that in relation to dermal exposure

    [21]:

    1) There is evidence of some skin penetration into viable tissues

    (mainly into the stratum spinosum in the epidermal layer, but

    eventually also into the dermis) for very small particles (less than

    10 nm), such as functionalised fullerenes and quantum dots.

    2) When using accepted skin penetration protocols (intact skin),

    there is no conclusive evidence for skin penetration into viable

    tissue for particles of about 20 nm and larger primary particle size

    as used in sunscreens with physical UV-filters.

    3) The above statements on skin penetration apply to healthy skin

    (human and porcine). There is an absence of appropriate

    information for skin with impaired barrier function, e.g. atopic

    skin or sunburned skin. A few data are available on psoriatic skin.

    4) There is evidence that some mechanical effects (e.g. flexing) on

    skin may have an effect on nanoparticle penetration.5) There is no information on the transadnexal penetration for

    particles under 20 nm. Nanoparticles of 20 nm and above

    penetrate deeply into hair follicles, but no penetration into viable

    tissue has been observed.

    The statement that nanoparticles N20 nm in diameter do not

    penetrate into viable tissue is controversial. There have been reports

    showing nanoparticles N20 nm penetrating through the stratum

    corneum (SC), to the viableepidermis[22,23]. However, our experience

    with gold, silver and quantum dots illustrates that these nanoparticles

    do consistently penetrate into the SC, but not into the viable epidermis.

    The reasons for this important disparity may be due to differences in

    nanoparticle constituents, models, and methodologies.

    We have also investigated nanoparticle gene, vaccine, siRNA, and

    drug delivery in a variety of systems, including skin [2436]. Our past

    and present focus in the field of topical therapeutic nanoparticles is

    developing a better understanding of nanoparticle disposition and

    toxicity [16,3740]. The most common tool used to evaluate

    nanoparticle penetration, delivery, toxicity or localization is the

    confocal microscope, illustrating that some primary aspects we areinvestigating are spatial and concentration related. In other words,

    where are the nanoparticles and how much is there? The Monteiro-

    Riviere, Lademann, Guy and Roberts groups are currently investigat-

    ing nanoparticle penetration, delivery, toxicity and species differ-

    ences. The Monteiro-Riviere group have made significant

    contributions towards understanding nanoparticle penetration, tox-

    icity and model systems, and hasgenerated an extensivebody of work

    that has largely focused on nanoparticle interactions with skin and

    skin cells. These reports have illuminated molecular mechanisms of

    nanoparticle uptake, toxicity, and bioretention [22,4164]. The

    Lademann group has made significant contributions towards a better

    understanding of the role hair follicles play as nanoparticle reservoirs

    [16,6576]. The Guy group has also contributed to this are and has

    investigated nanoparticle versus model drug localization in skin

    compartments, revealing valuable information about drug and

    nanoparticle distributions in skin [7781]. The Roberts groups is

    using novel fluorescence lifetime imaging techniques to simulta-

    neously evaluate NAD(P)H lifetime changes with nanoparticle

    penetration [40,82,83].

    1.2. Barrier properties of the skin

    Among the multiple, complex functions of mammalian skin, one of

    its major roles is to prevent invasion of the organism by acting as a

    defensive barrier to threats from the external environment. The skin

    has evolved defensive mechanisms which give it physical, immuno-

    logical, metabolic and UV-protective barriers to allow it to inhibit

    attacks by microbes, toxic chemicals, UV radiation and particulate

    matter (including nanoparticles, which may occur in the naturalenvironment)[84]. Ontheotherhand, the skincan beusedas a portof

    entry for therapeutic substances such as drugs and vaccines if the

    mechanisms that confer the barrier properties are understood and

    exploited.

    1.2.1. Skin structure

    Skin consists of two main layers. The underlying dermis contains a

    variety of cell types, nerves, blood vessels and lymphatics embedded

    in a dense network of connective tissue. Above the dermis and

    separated from it by the basement membrane, the epidermis is

    Fig. 2. Electron microscopy comparison of size and morphology of a virus (panel a, adapted from [254]) that infects human skin, human papilloma virus, and nanoparticles (panels b

    and c) for topical drug delivery are in similar size ranges. The human papilloma virus is 40 nm in diameter, the same size range as smaller drug delivery vehicles like poly(q-

    caprolactone)-block-poly(ethyleneglycol) nanoparticles, 40 nm, designed to deliver minoxidil to hair follicles (panel b, adapted from [130]). Solid lipid nanoparticles like these

    developed for penciclovir can be several times larger at an average diameter of 250 nm (panel c, adapted from [255]). The bar indicates 200 nm for all panels.

    472 T.W. Prow et al. / Advanced Drug Delivery Reviews 63 (2011) 470 491

    http://localhost/var/www/apps/conversion/tmp/scratch_6/image%20of%20Fig.%E0%B2%80
  • 7/28/2019 Advance Drug Delivery Reviews

    4/22

    composed mostly of layers of stratified keratinocytes, where the SC

    cells or corneocytes are bathed in a protein-rich envelope with an

    outer lipid envelope, surrounded by an extracellular lipid matrix [85].

    Keratinocytes undergo a process of keratinisation, in which the cell

    differentiates and moves upward from the basal layer (stratum

    basale), through the stratum spinosum and stratum granulosum, to

    the outermost layer, the stratum corneum (SC or horny layer). On

    reaching the SC, cells become anucleated and flattened and are

    eventually sloughed off [86]. Interspersed amongst the keratinocytesin theviable epidermisare cells with roles such as melanin production

    (melanocytes), sensory perception (Merkel cells) and immunological

    function (Langerhans and other cells). In addition to the structured

    cellular components of skin, there are appendages including the

    pilosebaceous units (hair follicles and associated sebaceous glands),

    apocrine and eccrine sweat glands.

    1.2.2. The stratum corneum barrier

    TheSC represents themain physical barrier of theskin,so that fora

    substance permeating across the skin, diffusion through the SC is the

    rate limiting step [1]. Conversely, the SC is also the main barrier for

    diffusionof water out of theskin [87]. The flattened, anuclear, protein-

    rich corneocytes of the SC are densely packed within the extracellular

    lipid matrix which is arranged in bilayers [88]. Thisis often referred to

    as a bricks and mortar arrangement [89]. The corneocytes are held

    together by corneodesmosomes, which help to form a tough outer

    layer by maintaining cellular shape and regular packing. Eventual

    degradation of the corneodesmosomes by proteolytic enzymes leads

    to desquamation [90].

    Transport of substances across the SC occurs mainly by passive

    diffusion and based on the dual-compartment bricks and mortar

    structure of the SC, interrupted by appendages, is considered to occur

    via three possible routes. These are the transcellular, the intercellular

    and the appendageal routes.

    Covering the corneocytes on the SC surface is a thin (0.410 m),

    irregular and discontinuous layer consisting of sebum secreted by the

    sebaceous glands, along with sweat, bacteria and dead skin cells. This

    layer is considered to have a negligible effect as an additional barrier

    to permeation through the SC [91].

    1.2.2.1. Inter-cellular spacing. For most penetrants, the intercellular

    route is favoured. Small molecules are able to move freely within the

    inter-cellular spaces and diffusion rates are governed largely by their

    lipophilicity, but also physicochemical properties such as molecular

    weight or volume, solubility and hydrogen bonding ability [92].

    However, the free movement of macromolecules or particles may be

    physically restricted within the lipid channels, which have been

    estimated by van de Merwe et al. to be 19 nm [93] and by Baroli et al.

    to be 75 nm [23]. This suggests that for such materials, the SC could

    present an additional barrier that is not present for small molecules.

    1.2.2.2. Skin turnover as a moving barrier. The outer layers of the SC

    (the so-called stratum dysjunctum) are subject to desquamation,allowing the SC to be completely turned over in a period of about

    14 days in humans, depending on anatomical site and age [94]. Thus,

    these cells might be regarded as a moving, constantly renewable

    barrier, providing an inherent mechanism for preventing foreign

    bodies from penetrating the skin and gaining a foothold [95]. The

    continuous upward migration and sloughing of corneocytes from the

    surface might assist in eliminating pathogens, cancerous cells or solid

    particulate matter.

    1.2.3. Transport routes of exogenous substances across the stratum

    corneum

    Polar and non-polar solutes were originally thought to permeate

    through the SC via separate routes [96], with polar solutes taking a

    transcellular routeand morelipophilic solutes goingvia the intercellular

    lipids. However a perception of the difficulty of repeated partitioning

    between lipophilic and hydrophilic compartments in the SC led to this

    pathway being regarded as unlikely in most cases. This was supported

    by histochemical [87] and theoretical [97,98] evidence showing that

    diffusion through intercellular lipids was more likely for most solutes.

    Despite a recent reaffirmation of importance of the transcellular route,

    even for lipophilic solutes, by Wang et al. [99] the transcellular route

    remains controversial.

    With most work focussing on the hair follicles, delivery of drugs orparticles via the appendageal route is regarded as a realistic

    alternative to delivery across the SC, despite the relative sparseness

    of these features on the skin surface [100]. The follicles extend deep

    into the skin, the thickness of the SC layer is progressively reduced as

    it extends into the structure and there is a rich capillary blood supply

    available to transport solutes diffusing out of the follicle. There is

    considerable interest in targeted follicular delivery with tailored drug

    formulations [98] or nanoparticle-bound drugs [101]. There have also

    been suggestions that partitioning of drugs into sebum itself may be a

    potential delivery route [100].

    1.2.4. Acid mantle maintains the stratum corneum barrier

    The surface of theskin haslong been recognised to be acidic, with a

    pH of 4.25.6 measured in humans by Blank over 70 years ago [102].

    This acidic skin surface is described as the acid mantle. The surface pH

    is influenced by sex and anatomical site [103], sweat, sebum and

    hydration [104]. There isa sharp gradientacross the SC, with the pH in

    the upper viable layers in the stratum granulosum approaching

    neutral. Using two-photon fluorescence lifetime imaging of a pH

    dependent fluorophore applied to the SC of hairless mice, Hanson et

    al. [103] identified acidic microdomains in the extracellular matrix,

    which became less frequent away from the surface towards the viable

    epidermis. This, they suggested, accounted for the pH gradient,

    although specific biochemical mechanisms have been proposed to

    account for the acidification per se, including generation of trans-

    urocanic acid by histidase-catalysed degradation of histidine [105].

    The acid mantle hasa numberof functions, includingantimicrobial

    defence [106], the maintenance of the permeability barrier by effects

    on extracellular lipid organisation and processing [107], the preser-vation of optimal corneocyte integrity and cohesion, regulated by pH-

    sensitive proteolytic enzymes [108] and restriction of inflammation

    by inhibiting therelease of pro-inflammatory cytokines [109]. There is

    a clear association between elevated skin pH and diseases such as

    atopic dermatitis [102], with significant pH differences between

    affected and unaffected skin in individual patients [110].

    Interestingly, theacidic pH of theskin surface mayalso support the

    skin's barrier to nanoparticle penetration in some cases, as carbox-

    ylated polystyrene nanoparticles were found to aggregate due to

    decreasing electrostatic forces as the solution pH was lowered [111].

    Aggregates would be less likelyto penetratethe SC.On theotherhand,

    by maintaining the integrity and cohesion of the SC, embedded

    particles would be less likely to be sloughed off during desquamation.

    Further, for certain nanoparticles, such as zinc oxide, the acidic pHmay greatly affect the nanoparticle aggregation and dissolution

    kinetics, thereby confounding experiments on skin penetration

    [112,113].

    1.2.5. Viable epidermis

    1.2.5.1. Tight junctions. The existence of functional tight junctions has

    been demonstrated in mammalian stratum granulosum [114,115],

    although many constituent tight junction proteins have been

    identified in other epithelial layers, as well as follicles [116]. Tight

    junctions are regarded as important elements of the epidermal barrier

    system and localization and expression of tight junction proteins have

    been shown to be altered in diseases characterised by a compromised

    skin barrier, such as psoriasis [117,118].

    473T.W. Prow et al. / Advanced Drug Delivery Reviews 63 (2011) 470 491

  • 7/28/2019 Advance Drug Delivery Reviews

    5/22

    1.2.5.2. Skin deactivation of nanoparticles by skin metabolism and other

    mechanisms. As well as acting as a physical barrier, the skin functions

    as a chemicalor metabolic barrier, with enzymes mainlylocated in the

    basal layer of the viable epidermis [119], as well as the extracellular

    spaces of the SC [120] and the appendages in the dermis [119]. A

    number of nanoparticles are biodegradable through hydrolysis,

    enzyme activity and physical forces causing, for instance, liposomes

    to merge with intercellular lipids [86]. Zinc oxide nanoparticles

    hydrolyse at neutral pH and even more so at acidic pHs, leading totheir conversion to zinc ions [37,112,113], which is accelerated on

    exposure to light [112]. This hydrolysis can be, however, complex and

    affected by the local environment. Cross et al. showed that there was

    no significant difference in the levels of zinc ions passing through the

    human epidermis from nanoparticle zinc oxide formulations relative

    to controls [37]. More recently, Gulson and colleagues showed

    increased recovery of 68Zn in blood and urine following application

    of 68Zn-enriched zinc oxide particles to human volunteers and

    concluded that zinc was able to penetrate the skin barrier. They

    were unable, however, to distinguish between the penetration of solid

    or solubilised zinc oxide,or zinc ions [121]. Skin metabolism following

    topical delivery has often been referred to a first-pass effect,

    analogous to that seen in the liver [84]. With nanoparticles, a similar

    phenomenon applies but arises from the destruction of nanoparticles

    not only by enzymatic processes but also by physical and chemical

    processes as discussed above. In the same way that skin metabolism

    may be exploited in a pro-drug or soft-drug approach to delivery

    [122], there is a potential for similar exploitation of nanoparticles in

    drug delivery. Indeed, the effective delivery of drugs to the skin by

    nanometer scale (b100 nm) liposomes, formed from engineered lipid

    nanovesicles, hasbeen suggested to involve the release of drugs in the

    intercellular lipid layer [36].

    There is also a wide range of protein transporters present in skin

    [123] which may assist in active transport of drugs or other

    compounds. Nanoparticle uptake in keratinocytes is, however, more

    likely to occur through other specialised processes. For instance,

    Zhang et al. have shown that 18 nm quantum dots (QD) with a

    carboxylic acid surface coating are taken up in human epidermal

    keratinocytes by endocytosis on recognition by lipid rafts [57].Transport involved internalizing into early endosomes before trans-

    ferred to late endosomes or lysosomes. The uptake of solid lipid

    nanoparticles (b180 nm) into keratinocytes has also been studied

    with it being suggested that they easily traverse the cell membrane,

    distribute throughout the cytosol and localize in the perinuclear

    region without any toxic effects [124].

    Delivery systems based on drugs, pro-drugs, soft-drugs or particle-

    bound drugs may need to be designed to exploit, or to overcome, the

    complex array of enzymes and transporters awaiting them in the skin

    [125,126]. Particle bound drugs must be able to reach a site where

    they can be released and as we have seen, the barriers to particle

    penetration of the skin are considerable. One possibility of doing this

    is to target the follicular route, to be discussed below.

    1.2.6. Hair follicles

    Whereas hair follicles were regarded as insignificant as potential

    routes for drug delivery, covering only 0.1% of the human skin surface

    area, their complex vascularisation and deep invagination with a

    thinning SC has led to a reappraisal of this view [101]. Work has been

    done on assessing the contribution of the follicular route to drug

    penetration [127,128], as well as targeted delivery [98]. Follicular

    penetration of solid particles, including liposomes [129], minoxidil-

    loaded [130], and fluorescent polystyrene [78] nanoparticles has been

    demonstrated. Lademann discussed the finding that 300600 nm

    particles penetrated follicles best on massage as a consequence of the

    distance between the scales on the hairs, and suggested that the

    movement of the hair acted as a geared pump to push the particles

    into the follicle [70]. They viewed follicles as an efficient reservoir for

    nanoparticle-based drug delivery [101], havingpreviously shown that

    titanium dioxide particles did not penetrate beyond the follicle [131]

    and would be eliminated in time by outward sebumflow. It is possible

    also that some follicles may be blocked by a plug of sebum or closed

    leading to particle penetration being impeded. Significant penetration

    of 40 nm nanoparticles beyond the follicles into epidermal cells can

    occur when the hair sheath has been pulled out [75].

    1.3. Barrier properties of diseased skin

    As stated in the SCCP nanotechnology report [21], there is limited

    data on skin penetration of nanoparticles through diseased skin.

    Topical nanoparticle drug delivery for local effects is likely to be used

    on diseased skin. Among the most common adult skin diseases are

    atopicdermatitis (6.9%) and psoriasis(6.6%) [132]. The effects of these

    barrier altering diseases on the penetration of nanoparticles are

    unknown.

    A report by the Australian Government, Department of Health and

    Ageing, Therapeutic Goods Administration (TGA) identifies a single

    unpublished report that assessed systemic zinc levels in psoriatic

    patients and foundno evidence of an increase in systemic zinc [133].

    We have carried out pilot clinical studies with the aim of quantifying

    nanoparticle penetration in psoriatic and atopic dermatitis lesions.

    We and others have investigated the penetration of nanomaterial

    contained in sunscreens, the most commonly applied source of

    nanoparticles, [16,134137], but there is a gap in the clinical literature

    where the penetration of nanoparticles has not been quantified in

    subjects with clinically relevant skin diseases like psoriasis and

    dermatitis. Hypothetically, nanoparticles could penetrate these

    lesions more efficiently due to an altered SC, inflammation and

    increased keratinocyte turnover.

    Our investigation included zinc oxide nanoparticle (35 nm)

    penetration profiling by quantifying second harmonic generation of

    the nanoparticles with non-invasive multi-photon imaging using

    time-correlated single photon counting (Fig. 3). There is a low level of

    background signal that can be seen in untreated skin images, taken en

    face, that may be due to melanin, keratin aggregates, or collagen

    (Fig. 3a). We observed concentrated zinc oxide nanoparticle signals inlesion furrows (Fig. 3b, red, Furrow and in Fig. 3c, F). The

    nanoparticles penetrated laterally from the furrows into the SC, but

    remained outside of the viable epidermis in non-lesional and lesional

    tissue (Fig. 3c). There was an intense zinc oxide nanoparticle signal

    from alltreated skin optical biopsies,but the imagesshow a punctuate

    distribution on the SC. Compared to the signal obtained from the

    untreated non-lesional (0.230.17 mg/ml at 0 m) and lesional sites

    (0.23 0.13 mg/ml at 0 m), no higher signal for zinc oxide

    nanoparticles could be detected within the viable epidermis of

    sunscreen treated non-lesional (0.17 0.10 mg/ml at 0 m) or

    lesional skin (0.220.09 mg/ml at 0 m). These results are consistent

    with previous reports [133] and the lack of nanoparticle penetration

    may be a mechanism that explains why applying zinc oxide

    sunscreens to subjects with psoriasis does not give rise to an increasein systemic zinc levels.

    1.4. Skin flexing and massage for enhanced delivery

    Nanoparticle skin penetration has the potential for highly

    controlled drug delivery in therapeutically relevant concentrations.

    The penetration enhancement by skin flexing and massage could

    impact topical nanoparticle drug delivery and nanotoxicology. Tinkle

    et al. investigated the potential for mechanical stress on skin to

    enhance particle penetration [138]. The confocal microscopy results

    showed that human skin treated with 1 m beads and then flexed for

    60 min resulted in dermal penetration of the 1 m beads. They also

    showed rapid penetration through a tear in the skin and significant

    penetration into the deeper layers of the SC using scanning electron

    474 T.W. Prow et al. / Advanced Drug Delivery Reviews 63 (2011) 470 491

  • 7/28/2019 Advance Drug Delivery Reviews

    6/22

    microscopy of SC tape strips. Later, detailed studies by the Monteiro-

    Riviere group showed that mechanical flexion in porcine skin

    increased the rate of nanoparticle penetration of 3.5 nm modified

    fullerenes [48]. However, we replicated Tinkle'sflexing apparatus and

    conditions, using excised human skin treated with 35 nm zinc oxide

    nanoparticles, and did not find penetration (unpublished work).

    Likewise, Lademann et al. has shown no penetration of nanoparticles

    into the SC after massaging [72]. Studies on non-massaged and

    massagedporcine skinrevealed that massaging 400

    700 nm diameterparticles resulted in the highest levels of follicular penetration (Fig. 4)

    [72]. An alternative explanation has been advanced by Lekki et al. as

    they found 20 nm nanoparticles penetrated as deep as 400 mintothe

    follicle [139], a process inconsistent with the much larger sizes of

    nanoparticles used by Lademann et al. to support a gearing

    mechanism. They suggest that a mechanical process is involved and

    is formulation dependent so that certain formulations may be enabling

    the mechanical pushing of products into follicles or that inhomoge-

    neous formulations are involved. Noting Lademann's other work

    showing that 40 nm particlesdo penetratedeep into follicleswhenthe

    hair shaft is removed [66], we propose that massaging may push the

    hair shaft to one side increasing the available area and depth of the

    opening (infundulum). As a consequence, greater penetration of

    nanoparticles may be enabled in a formulation, skin type and massage

    mechanism dependent manner. Togetherthese data suggest thatthere

    may be substantial differences between the penetration and interac-

    tions of nano- and micro-particles that depend greatly on the skin

    type, nature of applied flexion and massage and formulation used.

    2. Small drug particle delivery

    2.1. Anti-inflammatory drugs

    Anti-inflammatory drugs represent a broad range of molecules,

    many with potential for topical delivery. Reports on nanoparticle-

    delivered drugs with anti-inflammatory properties for topical use

    include: aceclofenac [140], betamethasone-17-valerate [141], cele-

    coxib [142], clobetasol propionate [143,144], corticosterone

    [145,146], flufenamic acid [147,148], flurbiprofen [149,150], glycyr-

    rhetic acid [151], ketoprofen [152,153], naproxen [153155], nime-

    sulide [156], prednicarbate [141,157,158], and triptolide [159]. These

    drugs can be divided into steroids, e.g. corticosterone, and non-

    steroidal anti-inflammatory drugs (NSAIDs), e.g. naproxen. Corticos-

    teroids work to reduce inflammation by binding glucocorticoid

    receptors, whereas NSAIDs work through inhibiting cyclooxygenase.

    2.1.1. Corticosterone

    In 2008, Kuntsche et al. investigated the potential for solid lipid

    nanoparticles (SLN) (tripalmitate), smectic nanoparticles (cholesteryl

    myristate and cholesterylnonanoate),and cubic nanoparticles (glycerol

    monooleate)to deliver corticosterone to the skin[146]. TheSLN,smectic

    particles, and cubic microparticles were an average of 94 nm, 106 nm,and 361 nm, respectively. The nanoparticles were highly characterized

    with electron microscopy, particle sizing, small angle X-ray diffraction,

    Fig. 3. Penetration profile of zinc oxide nanoparticles from skin furrows. Panels a and b

    show untreated lesional (a) and zinc oxide nanoparticle containing sunscreen treated

    lesional sites (b). The inset in panel b is an electron microscopy image of the individual

    zinc oxide nanoparticles with a mean diameter of 35 nm, the bar indicates 100 nm.

    FLIM measurements were used to differentiate skin autofluorescence (green) from zinc

    oxide nanoparticles (red). The intensity profile of the autofluorescence and zinc oxide

    nanoparticles is shown in Panel c and is derived from the white lines in panels a and b.

    The nanoparticles aggregate in the furrow region (F, Furrow) that is bordered by the

    stratum corneum (SC), that protects the viable epidermis (VE). The bar in panel b

    indicates 50 m.

    475T.W. Prow et al. / Advanced Drug Delivery Reviews 63 (2011) 470 491

    http://localhost/var/www/apps/conversion/tmp/scratch_6/image%20of%20Fig.%E0%B3%80
  • 7/28/2019 Advance Drug Delivery Reviews

    7/22

    and differential scanning calorimetry. Nanoparticle skin interactions

    were monitored with fluorescence microscopy. Storage of these

    nanoparticles for N15 months did not dramatically changethe diameter

    or polydispersity of any of the formulations. Permeation studies were

    carried out in excised human skin and a rat epidermal keratinocyte

    organotypic culture model mounted in Franz type diffusion cells.

    Cumulative corticosterone in the receptor was measured over 48 h.

    These results showed that the PBS control had the highest cumulative

    permeationat 1.5%, cubic nanoparticles at 1.0%, smectic nanoparticles at

    0.4%, and SLN at 0.1% in human epidermis. In the rat epithelial

    keratinocyte organotypic culture the cumulative permeation was 0.7%

    for PBS, 0.7% for cubic nanoparticles, 0.6% for smectic nanoparticles, and0.2% forSLN. Thepermeability coefficientswere alsocalculated and only

    the cubic nanoparticles showed a distinct improvement over the PBS

    control, 4.3107 cm/s versus 1.8107 cm/s. Fluorescence micros-

    copy only detected fluorescence with the cubic nanoparticle group. In

    conclusion, exposure of skin to the largest particle (361 nm) resulted in

    the highest level of drug delivery. This highlights the importance of

    interactions of the materials over that of shape and size.

    A recent report by Jensen et al. describes topical corticosterone

    delivery with SLN using distearate [145]. A series of corticosteroids in

    distearate (hydrocortisone, hydrocortisone-12-acetate, hydrocorti-

    sone-17-butyrate, hydrocortisone-17-valerate, and hydrocortisone-

    21-caprylate) with calculated log P from 1.28 to 4.16 were tested. The

    mean diameters of these particles ranged from 180 5 to 220 28 nm.

    In vitro release assays using cellulose membranes showed that

    permeation was inversely related to lipophilicity. At 12 h, the

    cumulative amount of drug released was 23.0% for hydrocortisone,

    27.2% for hydrocortisone-12-acetate, 18.3% for hydrocortisone-17-

    butyrate, 7.5% for hydrocortisone-17-valerate, and 4.9% for hydrocor-

    tisone-21-caprylate.

    Although both Kuntsche et al. and Jensen et al. both used SLN

    encapsulated corticosteroids for topical delivery, each used different

    lipids to construct the nanoparticles. The nanoparticles themselves

    were similar in size and both were capable of releasing corticoster-

    oids. One possibility for improvement in corticosteroid delivery would

    be to explore the glycerol monooleate nanoparticle formulation [146]

    to deliver the hydrocortisone derivatives [145] for optimal controlledanti-inflammatory delivery. The major differences in permeation

    models limit further comparison.

    2.1.2. Flufenamic acid

    Flufenamic acid (FFA) is an anti-inflammatory drug that was

    chosen as a model lipophilic drug for testing topical nanoparticle drug

    delivery by the Schaefer and Lehr group from Saarbrucken, Germany

    [147,148]. This model drug also hasfluorescence properties when in a

    non-polar environment and excited at 420 nm. This enables micro-

    scopic localization studies in addition to traditional pharmacokinetic

    studies. In 2006, Luengo et al. described topical FFA delivery with

    328 nm poly-lactic-co-glycolic acid (PLGA) particles loaded with FFA

    [147]. The particles were characterized with atomic force microscopy

    (Fig. 5). Release experiments showed that after 6 h, both the free and

    Fig. 4. Particle accumulation inhair folliclesaftermassage. Massaged skin is shown treated with particulatedye andnon-particulate dyein panelsa andb, respectively.Panels c andd

    show skin, without massage, that has been treated with particulate dye (panel c) and non-particulate dye (panel d). Panel e and f show the scales on hair follicles for human and

    porcine skin which serves as a basis for a geared pump delivery of nanoparticles deep into the opening on massaging. Adapted from [70,72].

    476 T.W. Prow et al. / Advanced Drug Delivery Reviews 63 (2011) 470 491

    http://localhost/var/www/apps/conversion/tmp/scratch_6/image%20of%20Fig.%E0%B4%80
  • 7/28/2019 Advance Drug Delivery Reviews

    8/22

    gel particulate formulations reached 100% FFA release using human

    skin. Cumulative penetration (area under the penetration curve) was

    determined for free and gel particle formulations. After 24 h the

    particulate group showed significantly more FFA (6.7 g/cm2) in the

    deep skin layers than the non-particulate control (3.8 g/cm2) for a

    1.8 fold increase.

    Santander-Ortega et al. investigated novel starch derivatives, with

    two different substitution groups for nanoparticle encapsulation of FFA

    [148]. The particles were fully characterized in addition to stability and

    pharmacokinetic experiments. Thesize of the particlesranged from 150

    to 182 nm. The size remained at 180 nm after 25 days at room

    temperature. The encapsulation efficiency for both starches was N95%.

    A total of 14.8% of drug was released and 4.4 g/cm2 permeated within

    the first 12 h.

    The FFA loaded PLGA particles had 6.7 g/cm2 permeated after 24 h

    and the starch nanoparticles had 8.3g/cm2 permeated at the same time

    point. These data show a substantial improvement when using PLGA or

    starch particles to deliver FFA to skin over non-particle controls.

    Fig. 5. AFM images from Luengo et al. showing FFA loaded PLGA nanoparticles for topical anti-inflammatory drug delivery [147]. Two formulations are shown, an aqueous

    nanoparticle suspension (Panel a) and a gel formulation (Panel b). Multi-photon microscopy images of human skin treated with drug-free nanoparticle gel (Panel c) and FFA

    containing nanoparticle gel (Panel d) are shown where the drug is present in the furrows (white arrow heads)." to "AFM images from Luengo et al. showing FFA loaded PLGA

    particles for topical anti-inflammatory drug delivery [147]. Two formulations are shown, an aqueous particle suspension (Panel a) and a gel formulation (Panel b). Multi-photon

    microscopy images of human skin treated with drug-free particle gel (Panel c) and FFA containing particle gel (Panel d) are shown where the drug is present in the furrows (white

    arrow heads).

    Table 1

    Particle-delivered anti-photoageing drugs and antioxidants. Not determined (ND).

    Drug(s) Nanoparticle

    material

    Diameter range, potential Dose, vehicle Model Ref.

    Tretinoin SLN 173406 nm, 2347 mV 0.05%, Hydroxylethyl-cellulose (0.8%) Rhino mice for skin irritation [162]

    Tr et inoin SLN 5 2.0 20 .8 t o 6 5.7 28.0 nm, ND 0 .0 5% , C ar bopol Ult rez 1 0 ( 1% , w/w ) Exc ised hairless a bd ominal r at skin [163]

    Tretinoin SLN 344 nm, ND 0.05%, Carbopol ETD 2020 Excised hairless abdominal rat skin [166]

    Tretinoin Silica 166.5261.6 nm,52.035.6 mV 0.05%, Lecithin or oleylamine in medium

    chain triglycerides

    Franz type diffusion cells with cellulose

    acetate membrane and porcine skin

    [160,161]

    Tretinoin CaCO3 ND 0.10.4%, Study subjects [167]

    Tretinoin CaCO3 1316 nm, ND 0.1%, aqueous ddY mouse [171]

    Isotretinoin SLN 31.3 to 50.0 nm,14.1 to17.9 mV 0.06%, Tween 80 and soybean lecithin Excised abdominal rat skin [168]

    Retinol SLN 224 nm, 58.10.6 mV 0.5%, glycerol, water and xanthin gum Cellulose membrane in Franz type

    diffusion cells

    [172,173]

    Vitamin A palmitate SLN 350 nm, ND 0.25%, Carbopol 940 (1%) Cadaver skin using Keshary Chien cells [169]

    477T.W. Prow et al. / Advanced Drug Delivery Reviews 63 (2011) 470 491

    http://localhost/var/www/apps/conversion/tmp/scratch_6/image%20of%20Fig.%E0%B5%80
  • 7/28/2019 Advance Drug Delivery Reviews

    9/22

    Importantly, these well controlled and informative studies correlate well

    despite having been published 4 years apart thus allowing comparison.

    2.2. Anti-photoageing drugs and antioxidants

    2.2.1. Retinoids

    Retinoids arecommonly usedin topicalacne andanti-ageingproducts

    and are one of the most studied drug groups in terms of topical

    nanoparticle and microparticle delivery (Table 1) [160

    173]. Retinoidswork through nuclear hormone receptors (retinoic acid receptors) and

    retinoid X receptors. The molecular responses to receptor binding are

    diverse, but in the context of skin disease, topical retinoid treatment

    initially reduces inflammation, helps to reduce comedones and wrinkles.

    Retinoids are used clinically to treat acne vulgaris [174] and cosmetically

    to treat photoageing [175]. Tretinoin (all-trans retinoic acid) is the first

    generation retinoid. Instability in the presence of light and oxygen is a

    drawback for using Tretinoin and is one of the major focus points for

    improvement by nanoparticle encapsulation. In the late 1990s, third

    generation retinoids, i.e. adapalene and tazarotene, overcame this

    drawback and were approved for clinical use [176], but have not been a

    focus for nanoparticle delivery. Theother major issue for retinoid therapy

    is local erythema, peeling, dryness, and pain. Skin irritation side effects

    have been mitigated by nanoparticle encapsulation, where the mecha-

    nism of action is likely to be controlled release.

    Improved stability and controlled release have been the primary

    focal points of nanoparticles for retinoid delivery. Castro et al. used solid

    lipid nanoparticles to encapsulate Tretinoin withan efficiency of 94% by

    using a lipophilic amine (stearylamine) [162]. They exploited ion

    pairing between Tretinoin and the lipophilic amine to increase the

    lipophilicity of the Tretinoin, thus enabling dramatic improvement in

    encapsulation efficiency. The microparticles varied in diameter from

    2283 to 68226 nm depending on the surfactant to lipid ratio

    (0.40.1). Storage of these microparticles at 25 C for 90 days revealed

    that samples with stearylamine maintained higher encapsulation

    efficiency than SLN without, 975% versus 552%. Skin irritation

    studieswere doneby comparing the SLNwith stearylamine to a placebo

    gel and a marketed Tretinoin cream containing a completely different

    formulation utilizing liposomes. Castro et al. used a rhino mouse modelto test irritation. Although the engineered SLN performed better than

    the commercialformulation,the lack of stearylamine controls precludes

    further interpretation. Likewise, the lack of skin retention and flux data

    hampers comparison of this SLN formulation withother similar studies.

    Others including the Patravale group have used SLN to deliver

    Tretinoin to skin with similar success [163,166]. In both studies, 0.05%

    Tretinoin was formulated with Carbopol to form gels with favourable

    spreadability, viscosity, and flow. Two distinct SLN sizes were tested

    in these studies: b100 nm [163] and N300 nm [166]. Both reports

    utilized excised skin from the abdomen of hairless rats. Both the

    smaller and larger SLN had similar, but not identical skin retention at

    4.2% and 6.4%, respectively. The encapsulation efficiency of the small

    and large SNL was similar at 46% and 49%, respectively. The flux

    differed with the larger SLN being exactly twice that of the smallermicroparticle at 75.6 ng/cm2 h versus 37.7 ng/cm2 h. These were

    compared to a commercial formulation with a flux of 64.5 ng/cm2 h,

    which contained a completely different formulation but had the same

    Tretinoin level (0.05%) as the SNL formulations. It is not possible to

    identify any potential stability benefits of nanoparticle encapsulation

    between the nanoparticles (b100 nm) and microparticles (N300 nm)

    because photostability was only investigated with the N300 nm SLN.

    Inorganic microparticles have also been used as Tretinoin carriers,

    including silica [177] and calcium carbonate [167,171]. Tretinoin coated

    silica microparticles were developed by the Prestidge laboratory and

    found to be 166 to 261 nm in diameter depending on whether the

    microparticles were formulated with lecithin or oleylamine, respec-

    tively. The zeta potential of the microparticles was 50 and +34 mV

    for lecithin and oleylamine, respectively. Neither the diameter nor the

    zeta potential was dramatically altered by oil or water phase

    formulation. The microparticle formulations were compared to lecithin

    or oleylamine formulations and the controlled release properties were

    investigated. The silica microparticle groups showed a significant

    reduction in the steady state flux to 90% for lecithin emulsions and

    50%for oleylamine emulsions. The72 h cumulative release was13.4 and

    7.4 g/cm2 for lecithin and oleylamine emulsions with silica micro-

    particles, respectively. However, the skin integrity after 72 h of

    treatment is questionable. There was a maximum of 30% and 76%reduction in the cumulative release by the silica microparticle groups

    compared to the non-microparticle containing controls. The 12 h skin

    retention with silica microparticle formulated with lecithin in oil phase

    wassimilar to that ofb100 nm SLNat 3.9% and4.2%,recognizingthat the

    silica studies were done with porcine skin [177] and the SLN studies in

    rat skin [163]. Clinical studies [167] have been carried out in Japanwith

    CaCO3 nanoparticles (1315 nm) coated with Tretinoin at high doses

    (0.1%). These studies do not include skin retention or flux measure-

    ments, but do describe favourable murine [171] and clinical outcomes.

    The lack of pharmacokinetic data, among other discontinuities with the

    more recent literature, restricts comparisons with silica and SLN

    delivered Tretinoin, however.

    Tretinoin derivatives, isotretinoin [168], retinol [172,173], and

    vitamin A palmitate [169], have also been delivered to skin with SLN.

    The isotretinoin loaded SLN described by Liu et al. [168]were in the

    same size range as the tretinoin loaded SLN described by Mandaw-

    gade et al. [163], 3050 nm and both studies used excised abdominal

    rat skin to investigate the pharmacokinetics. Isotretinoin loaded

    nanoparticles were capable of encapsulation efficiencies of up to

    99.7% depending on the surfactant concentration (8% soybean lecithin

    and 4.5% Tween 80). Stability wastested over 3 monthsat 28 C with

    no changes in isotretinoin levels noted. The steady state flux for the

    non-SLN control (0.06% isotretinoin) was 0.76 0.3 g/cm2 h, but

    there was no isotretinoin found in the receptor chambers after 8 h.

    However, the skin retention of the control (2.81 g) and maximum

    SLN (3.65 g) formulation were comparable. This was an increase of

    30% from the non-SLN control.

    Retinol was also delivered by drug loaded SLN, but these studies

    utilized particles N220 nm in diameter [172,173]. The skin retentionof retinol was 0.68% (over 6 h) in porcine skin, an improvement of

    0.9 g over a nanoemulsion control group. The in vitro flux from the

    retinol loaded SLN was 150 ng/cm2 over 6 h, or 25 ng/cm2 h.

    Vitamin A palmitateis converted into retinol in theskin andfinally to

    tretinoin. Vitamin A palmitate has been formulated into SLN of

    approximately 350 nm in diameter by Pople and Singh [169]. The

    pharmacokinetics of thisformulationwas exploredusing humancadaver

    skin with Keshary Chien cells. After 24 h the SLN drug release was 67.5%

    and the gel control was 54.4%, implying that the flux was unusually

    higher with the SLN formulation, SLN formulation usually results in

    lower flux with longer term release. Pople and Sing also evaluated the

    effects of treatment on SC thickness in rats. SLN formulation treatment

    resulted in an increasein SC thickness of almost 3 times comparedto the

    conventional gel. The conventional gel control was a formulation thatcontainedthe samecomponentsas theSLNformulations. Thisincrease in

    SC thickness is likely to be due to a hydration effect.

    In summary, a number of SLN and inorganic particles have been

    used to improve retinoid pharmacokinetics and stability. SLN were the

    most common particles tested and two major size ranges were tested,

    b100 nm and N200 nm. This could be interpreted as a broad

    comparison of nanoparticle versus microparticle retinoid delivery,

    given the definition of nanoparticles as b100 nm in diameter. Both

    nanoparticles and microparticles could be formulated with near 100%

    encapsulation efficiency and both showed improved stability and

    pharmacokinetics (depending on controls used). Together, no real

    benefits of nanoparticle(b100 nm) formulation are evident compared

    to microparticle (N100 nm) data, but encapsulation appears to have

    advantages over free drug formulations.

    478 T.W. Prow et al. / Advanced Drug Delivery Reviews 63 (2011) 470 491

  • 7/28/2019 Advance Drug Delivery Reviews

    10/22

    2.2.2. Tocopheryl acetate

    Zhao et al. explored thebenefitsof a hydrofluoroalkane foam to aid

    tocopheryl acetate release from lipid nanoparticles upon contact with

    the skin [178]. The uncharged lipid nanoparticles were formulated

    with an oil-in-water emulsion and the size range of the loaded

    nanoparticles was from 53 1 to 57 0 nm. Maximal encapsulation

    efficiency was 90.13.7% and 100% of the drug was recoverable.

    Delivery to human skin was evaluated by treating skin with 400 g/

    cm

    2

    tocopheryl acetate for 24 h. Saturated silicone oil was comparedto the aqueous nanosuspension and nanoparticle containing foam. No

    tocopheryl acetate was found in the receptor fluid for any group and

    the estimate for skin retention was 1.70.4% for the oil control and

    0.70.4% for the foam with nanoparticles, while no drug was

    delivered to the SC by aqueous nanoparticles. These data suggest no

    major differences between the groups and no major benefit of

    nanoparticle formulations in terms of total delivery. However, studies

    showed improved stability over a 4 week trial with the foam

    formulation.

    Similar studies from the same group used solid lipid nanoparticles

    to deliver tocopheryl acetate with hyalauronic acid gels [179,180],

    with similar results. The nanoparticles in this study were comparable

    in size to those described by Zhao et al., at 50 1 nm. No tocopheryl

    acetate was released over 24 h when permeability tests were carried

    out with synthetic membranes treated with SLN formulations [180].

    Synthetic membrane permeation studies revealed that the saturated

    silicone oil formulation gave a maximal flux of 836.2 137.9 g/

    cm2 h. Penetration studies were carried out with porcine and human

    skin using tape stripping to extract material from the skin surface.

    These experiments showed that porcine skin treated with silicone oil

    saturated with tocopheryl acetate or SLN gel retained 1.320.57%

    and 1.650.90% respectively of the total dose applied. Interestingly,

    in human skin, the saturated silicone oil delivered 1.700.36%, while

    tocopheryl acetate could not be detected after SLN gel treatment.

    These species differences could be due to a number of reasons

    including differences in SC lipid orientation and thickness, supporting

    the hypothesis that human skin is less permeable than pig skin [181].

    Another difference described by Lademann was the shrinking of hair

    follicles in human skin which may prevent penetration of nanopar-ticles [71]. These well controlled and coordinated studies show no

    substantial delivery benefit of nanoparticle formulation delivered

    tocopheryl acetate over saturated silicone oil, regardless of vehicle.

    2.2.3. Multiphoton imaging for in vivo lipid nanoparticle delivery studies

    Photoageing severity is significantly associated with skin cancer

    (pb0.05) [182]. Photoageing can result in collagen damage, DNA

    damage and metabolic changes due to reactive oxygen species and

    inflammation. Epidermal and dermal antioxidant enzymes are

    powered by NAD(P)H. NAD(P)H is an endogenous fluorophore that

    can be investigated with fluorescence lifetime imaging. These data

    have direct implications for the metabolic state of the skin.

    Our hypothesis was that commercial topical products containing

    lipid nanoparticle components and tretinoin, niacinamide, ubiqui-none and folic acid as active agents would alter the NAD(P)H

    fluorescence lifetime. This is especially relevant for niacinamide

    containing products, as niacinamide is a precursor to NAD(P)H. Other

    products include folic acid which is a form of vitamin B that is

    essential for cell growth and metabolism.

    We applied Nivea Visage Q10 Plus Day cream (Product A,

    ubiquinone), Nivea Visage DNage Day cream (Product B, folic acid),

    L'Oreal Revitalift Day cream (Product C, retinol) and Olay Regenerist

    Microsculpting Cream (Product D, niacinamide) at 2 mg/cm2 every

    day for up to 6 days. In vivo NAD(P)H changes pre- and post-

    treatment were investigated with multiphoton microscopy using

    fluorescence lifetime imaging microscopy on days 2 and 6.

    All multiphoton images were collected using DermaInspect system

    (JenLab GmbH, Jena, Germany). Excitation emission of 740 nm was

    used, and theaverage incident optical power was30 mW at therear of

    the objective. The light going to the FLIM detector was filtered with a

    band pass filter transmitting 350450 nm light for NAD(P)H. Images

    were analysed with SPCImage 2.9.4 software. Two component decay

    matrices were calculated and the region of interest of each image was

    selected for analysis. The mean value of each NAD(P)H component

    (i.e. 1, 2, a1%, a2% and a1%/a2%) were analysed to assess any

    metabolic changes. Stratum basale treated with folic acid (Product B)

    for 6 days is shown in Fig. 6.The a1%/a2% ratio is inversely related to the metabolic rate and so

    was used to evaluate any anti-ageing cream induced changes. At

    stratum granulosum and stratum spinosum, all treated sites were

    observed to have decreases in relative amounts of free NAD(P)H over

    the course of the experiment. Interestingly, ubiquinone-, folic acid-

    and retinoic acid-treated sites had increases in free NAD(P)H at the

    stratum basale level, indicating lower metabolic states. The only

    significant increases in NAD(P)H levels were found in the retinol

    treated group on day 2 in the stratum spinosum (Fig. 7, *).

    3. Other topical nanoparticle drug delivery applications

    3.1. Antimicrobial agents

    There are several recent studies thatdescribe topical nanoparticles for

    anti-microbialdelivery applications [11,183190]. These studiesexamine

    two types of nanoparticles, SLN for imidazole anti-fungal drugs

    [183,188190] and silver nanoparticles [11,184187] that have broad

    anti-microbial applications. The most prominent topical antimicrobial in

    consumer productsis nanoparticle formulated silver.Silver nanoparticles

    possesses antimicrobial properties [191,192] and the mechanism by

    which silver functions as a disinfectant is not yet fully understood, but

    may be related to silver ion induced metabolic inhibition. Alternately,

    imidazole antifungal agents inhibit the synthesis of ergosterol, an

    important part of fungal cell membranes [193].

    3.1.1. Antifungal drugs

    Econazole nitrate was loaded into SLN by Sanna et al. and the

    particles characterized and permeation studies done to validate thesystem [189]. The particles had diameters ranging between 14013

    and 154 5 nm and the encapsulation efficiency ranged between 97

    and 102%. The cumulative econazole release from non-particle

    containing gel into porcine skin was 124.7221.6 g/cm2 after 24 h

    and the particle containing gel was 48.460.8 g/cm2. This could be

    explained by the drug leaving theouter particle surface, leaving a high

    concentration of drug in the core [20,194,195].

    Passerini et al. have also investigated SLN made of glycerol

    palmitostearate loaded with econazole nitrate for topical delivery,

    but importantly this study compared SLN to solid lipid microparticles

    using identical formulations [188]. The microparticles hada size range

    of 18.03.17 to 44.75.16 m and the SLN had the same size range

    described above (150 nm). Permeation experiments conducted with

    porcine skin revealed a flux of 2428 g/cm2

    /h1/2

    for microparticlesand 1520 g/cm2/h1/2 for SLN, with no statistically significant

    difference between the groups. The cumulative amount of econazole

    nitrate that permeated after 24 h was 124.20.12 g/cm2 for non-

    particulate econazole compared to 78121 g/cm2 for microparticles

    and 4881 g/cm2 for SNL. These results show that there is no

    significant difference between micro and nano-particle formulation

    and econazole delivery kinetics in porcine skin, suggesting no size

    dependent effects on drug delivery.

    3.1.2. Silver nanoparticles

    Topical silver nanoparticles, having antibacterial and antifungal

    effects, are some of the most studied therapeutic nanoparticles

    [11,184187,192], but important questions remain regarding topical

    use. Silver nanoparticles are similar to solid drug nanoparticles in that

    479T.W. Prow et al. / Advanced Drug Delivery Reviews 63 (2011) 470 491

  • 7/28/2019 Advance Drug Delivery Reviews

    11/22

    the active agent appears to be the breakdown product of the particle.

    Silver nanoparticles exhibit minimal penetration into skin and are

    consequently considered safe. Studies of long term occupational

    exposure to silver ions and silver nanoparticles conclude that they are

    relatively non-toxic. The sub-milligram levels of silver present in burn

    dressings are considered low risk [196].

    Samberg et al. recently described an in vivo study in porcine skin

    [61]. The skin was dosed topically with 2080 nm silver nanoparticles

    for 14 days. The authors explored the effects of washing the

    nanoparticles and carbon coating the nanoparticles prior to applica-

    tion. Their results showed that none of the groups penetrated the SC

    and that the sites of application showed focal inflammation. Finally,

    Fig. 6. Metabolic state of stratum basale treated with folic acid cream for 6 days. In vivo multiphoton images show untreated (left) and folic acid-treated (right, Product B) stratum

    basale. Free NAD(P)H lifetime contributionover protein-bound NAD(P)H contribution ratios (a1%/a2%)are inversely related to the metabolic rate.The 1 component is related to free

    NAD(P)H in the cytosol, while the 2 component changes when NAD(P)H protein-binding changes.

    480 T.W. Prow et al. / Advanced Drug Delivery Reviews 63 (2011) 470 491

    http://localhost/var/www/apps/conversion/tmp/scratch_6/image%20of%20Fig.%E0%B6%80
  • 7/28/2019 Advance Drug Delivery Reviews

    12/22

    the authors showed through in vitro and in vivo systems that washed

    or carbon coated silver nanoparticles were the least toxic form of

    silver nanoparticles.

    Tian et al. investigated the relationship between topical silver

    nanoparticle delivery and wound healing in mouse models. The silvernanoparticles in thewounddressings were 14 10 nm in diameter,as

    measured by transmission electron microscopy. Silver nanoparticle

    impregnated dressings were changed daily, making treatment

    continuous for the course of the experiments. Thermal injury took

    35.41.29 days to heal without intervention and took 9 days less

    with silver nanoparticle treatment (26.50.93 days). Interestingly,

    injury treated with the same concentration of silver, but without

    nanoparticles (silver sulfadiazine) took 37.43.43 days to heal, or

    10 days longer than with the nanoparticles. Tian et al. also saw

    improvements in cosmesis in thermal injury, improved healing time

    in diabetic mice, and protection from microorganisms in wound areas.

    All of these experiments were conducted in mouse models.

    There is a knowledge gap in topical silver nanoparticle treatment in

    study subjects. We investigated silver nanoparticle penetrationin intact

    and tape stripped skin, while simultaneously monitoring NAD(P)H

    metabolic rate. We also used reflectance confocal microscopy to

    visualize silver nanoparticle aggregate bio-retention in intact and tape

    stripped skin. An aqueous commercial product containing silver

    nanoparticles with a size range of 1345 nm based on 100 nanoparticlemeasurements (Fig. 8a) was applied under occlusive conditions for 4 h

    (180 l, containing 0.040.045 mg/ml silver). The silver nanoparticles

    were characterized by transmission electron microscopy, FLIM, and

    reflectance confocal microscopy. Skin barrier integrity was assessed by

    TEWL.

    Our results show that the commercial topical silver nanoparticle

    spray was composed of silver nanoparticles b50 nm and that 740 nm

    excitation could be used to excite the silver nanoparticle and NAD(P)

    H simultaneously (Figs. 8a, b and 9). Skin penetration of silver

    nanoparticles was not dramatically enhanced by tape stripping 20

    times compared to silver nanoparticle treated intact skin (Fig. 8c).

    This is likely to be due to the incomplete removal of the stratum

    corneum. While silver nanoparticle treatment for four hours de-

    creased metabolism, as determined by increased a1%/a2%, no

    Fig. 7. Summary of NAD(P)H effects from antioxidants delivered with lipid nanoparticle formulations in commercial products. The total NAD(P)H fluorescence was measured by

    FLIM photon counting by summing the free and protein-bound NAD(P)H signals weighted by intensity after 2 and 6 days of daily anti-ageing cream treatment formulated with lipid

    particles. Optical sections were taken of stratum granulosum, stratum spinosum andstratumbasalefromthe volar forearms ofthreestudysubjects. Theactive ingredients areshown

    on the x-axis of each graph.Initial increases in NAD(P)H at 2 days were not observed after 6 days of treatment. Two-tailed Students t-test revealed statistical significance (*)for only

    the Retinol treated stratum spinosum after 2 days of treatment.

    481T.W. Prow et al. / Advanced Drug Delivery Reviews 63 (2011) 470 491

    http://localhost/var/www/apps/conversion/tmp/scratch_6/image%20of%20Fig.%E0%B7%80
  • 7/28/2019 Advance Drug Delivery Reviews

    13/22

    statistically significant differences were detected (Fig. 8d.). Differ-

    ences between untreated and silver treated groups decreased with

    depth, suggesting a depth dependent effect. Fig. 9 shows a1% color

    coded images illustrating the difference between the NAD(P)H

    (green/yellow, a1% 4585) and silver nanoparticle second harmonic

    generation (orange/red, a1% 90100).

    Silver nanoparticle aggregates (Fig. 10, arrow heads) were tracked

    in furrows up to 10 days after treatment with reflectance confocal

    microscopy (Fig. 10, Video 1). Aggregate formation could be due to

    protein/nanoparticle binding [197]. After 10 days, no silver nanopar-

    ticle aggregates could be detected in intact skin, but aggregates

    persisted in tape stripped skin. Similar results were observed in hair

    follicles (data not shown). These results show that even relatively

    brief treatment with silver nanoparticles can last longer than 10 days

    on treated skin, agreeing with previous work by Lademann et al. [71].

    This suggests that the treatment benefits of silver nanoparticle

    treatment may last longer in damaged skin and that toxicity issues

    may need to be studied over longer time courses.

    Fig. 8. Characterization of silver nanoparticles and in vivo human skin penetration. Nanoparticles were characterized with TEM (panel a) and MPM (panel b) for size and second

    harmonic generation characteristics, respectively. A study subject with intact (non-Tape) and tape stripped (Tape) skin was exposed to a commercial silver nanoparticle containing

    antimicrobial solution under occlusive conditions for 4 h prior to MPM imaging. FLIM was used to isolate and measure the silver nanoparticle second harmonic signal. Non-furrowcontaining sites were analysed for silver nanoparticle signal (Integrated Density) and depth plots were generated from FLIM z-stacks. Silver nanoparticle signal was simultaneously

    measured with NAD(P)H fluorescence. Fluorescence lifetime analysiswas used to measure the free/protein bound NAD(P)Hratio (a1/a2). This ratio is thought to be inversely related

    to the metabolic rate [256,257]. Application of silver nanoparticles was associated with an depth dependent increase in the free/protein bound NAD(P)H contribution which is

    consistent with previous reports of silver inhibition of metabolism. However, this trend was not statistically significant when compared to the untreated data.

    482 T.W. Prow et al. / Advanced Drug Delivery Reviews 63 (2011) 470 491

    http://localhost/var/www/apps/conversion/tmp/scratch_6/image%20of%20Fig.%E0%B8%80
  • 7/28/2019 Advance Drug Delivery Reviews

    14/22

    3.2. Anti-proliferative agents

    Hyperproliferative skin disease is not limited to cancer, but also

    precancerous lesions. It can also be caused by inappropriate

    inflammatory responses. Several anti-cancer and anti-proliferation

    drugs have been delivered with dendrimers and nanoparticles,

    including 5-aminolevulinic acid (ALA), 5-fluorouracil (5FU), paclitax-

    el, podophyllotoxin, and Realgar [198

    204].

    3.2.1. 5-aminolevulinic acid

    Photodynamic therapy is based on the delivery of photosensitive

    drugs to skin lesions followed by targeted light exposure. ALA and

    methylaminolevulinate (MAL) are commonly used as they both

    induce the production of the photosensitiser protoporphyrin IX in

    skin cells. Photodynamic therapy is now being used for an increasing

    number of skin conditions, including actinic keratosis and non-

    melanoma skin cancers. Topical delivery has not been optimized.

    Fig. 9. In vivo z-stack of FLIM images color coded from a1% 0 to 100%, blue to red, respectively. The images are 214214 m2 and show untreated and silver nanoparticle treated skin

    thatwas intact (No tapestrip) or tape stripped. The yellow/green indicates NAD(P)H autofluorescence and orange/red indicates silver nanoparticle second harmonic generation. Bar

    indicates color code and 50 m.

    483T.W. Prow et al. / Advanced Drug Delivery Reviews 63 (2011) 470 491

    http://localhost/var/www/apps/conversion/tmp/scratch_6/image%20of%20Fig.%E0%B9%80
  • 7/28/2019 Advance Drug Delivery Reviews

    15/22

    There are significant gaps in knowledge regarding the relationship

    between photodynamic response and dose parameters. Both drugshave

    poor skin penetration profiles (0.26% methylaminolevulinate penetra-tion/24 h). MAL is faster acting at 3 h versus ALA at 46 h. Importantly,

    19%of those treated do notrespond [Australian clinical trial,PC T305/99]

    and MAL (Metavix, US$225/g) is nine times more expensive than ALA.

    Therefore, enhanced ALA delivery could improve non-responder rates,

    recurrence rates, and decrease the cost of therapy.

    Battah et al., from the MacRobert group, have used dendrimers

    conjugated to ALA for improved efficacy by mitigating the hydrophilic

    nature of ALA and characterized porphyrin production and photo-

    sensitivity [198,200]. However, these compounds have not been

    reported for topical use, only reports of systemic application are

    available [205]. The practical problem with systemic ALA application

    is that the entire person would become photosensitive and much

    more drug would likely be needed, thus eliminating any benefit over

    methylaminolevulinate. The lack of topical data in this area could bedueto insufficient skin penetration levels of the large dendrimers-ALA

    molecules or premature ALA cleavage by esterases in the skin. This

    issue may be resolved with the development of a peptide-ALA

    conjugatethat hasbeen shown to work well in an excised porcine skin

    model [206]. Treatment efficacy of this prodrug has the potential to be

    enhanced by nanoparticle delivery in the future.

    3.2.2. Podophyllotoxin

    Podophyllotoxin is a topical antiproliferative drug that is extracted

    from the roots and rhizomes of Podophyllum species [207]. This toxin

    is used clinically to treat HPV induced warts in a 0.5% gel called

    Condylox. Typical therapy would include twice daily application for

    three days and then no treatment for four days before treating again.

    No more than 10 cm2

    should be exposed and a maximum of 0.5 ml

    applied per day. Although this drug is a first-line treatment for genital

    warts, there are severe side effects and systemic absorption can be

    dangerous [208]. This toxicity issue led Chen et al. to explore thepotential for SLN-podophyllotoxin delivery to the skin surface [199].

    SLN were prepared from tripalmitin, soybean lecithin, poloxamer 188,

    and polysorbate 80 in an aqueous solution. The control tincture

    contained podophyllotoxin in 75% ethanol and water. The resulting

    SLN had two populations of particles with mean diameters of 44 nm

    and 194 nm and all had negative zeta potentials that ranged from

    48 to 17 mV. Atomic force microscopy, differential scanning

    calorimetry, and X-ray diffraction were used to further characterize

    the formulations. Podophyllotoxin is fluorescent and can be excited at

    290 nm with emission at 633 nm, so the endogenous fluorescence

    was used to visualize podophyllotoxin localization after treating

    excised porcine ear skin (Fig. 11). Podophyllotoxin fluorescence

    accumulation can be seen in a furrow and hair follicle, in a manner

    similar to the zinc oxide nanoparticles in Figs. 3, 9 and Video 1, FFAloaded PLGA microparticles are shown in Fig. 5d and the silver

    nanoparticles in Figs. 9, 10 and Video 1. Encapsulation of podophyllo-

    toxin in SLNresulted in an increase in thecumulative podophyllotoxin

    compared to the tincture, 23.380.55 g and 6.080.31 g, respec-

    tively. However, no podophyllotoxincould be detected in thereceptor

    chamber of SLN treated skin, suggesting that the majority of the drug

    was contained in the SLN on the outer layers of the skin. The

    podophyllotoxin activity was not measured, so the therapeutic value

    of SLN delivery cannot be assessed.

    3.3. Other topical particulate delivered drugs

    The field of nanoparticle drug delivery to the skin is fragmented in

    terms of the drugs investigated. Some examples of molecules and

    Fig. 10. In vivo RCM images of untreated and silver nanoparticle treated skin that was intact (No tape strip) or tape stripped. Silver nanoparticle aggregates (arrow heads) were seen

    2 days and 6 days (data not shown) after application but not at 10 days after application in intact skin. In contrast, silver nanoparticle aggregates were still seen 10 days after

    application, but only in discrete areas. Each image is 0.5 0.5 mm2.

    484 T.W. Prow et al. / Advanced Drug Delivery Reviews 63 (2011) 470 491

    http://localhost/var/www/apps/conversion/tmp/scratch_6/image%20of%20Fig.%E0%B1%B0
  • 7/28/2019 Advance Drug Delivery Reviews

    16/22

    combinations of molecules that have been investigated for topical

    nanoparticle delivery that have not been discussed are: diethyltolua-

    mide and ethylhexyl p-methoxycinnamate [209]; fludrocortisone

    acetate and flumethasone pivalate [210]; hinokitiol, glycyrrhetinic acid

    and6-benzylaminopurine [211]; tolterodine tartrate [212]; JSH18[213];

    acyclovir [214]; amlodipine [215]; ascorbyl palmitate [216]; benzylnicotinate [217]; buprenorphine [218]; calcipotriol with methotrexate

    [219]; ceramide-3 [220]; coenzyme Q10 [221]; cyclosporin A [222,223];

    cyproterone acetate [224]; frusemide [225]; insulin [226,227]; lidocaine

    [228]; methoxycinnamate[77]; minoxidil [229,230]; nitrendipine

    [231,232]; oxybenzone [233]; progesterone [234]; psoralen [235]; RU

    58841-myristate [236]; temoporfin [237]; thrombin [238]; and triam-

    cinolone acetonide acetate [239]. These reports describe the use of

    several different nanoparticle carriers: chitosan [214]; invasomes [237];

    iron oxide [238]; lipid [218220,223,230]; micro- and nano-emulsions

    [210,212,215,217,218,227]; non-ionic surfactantvesicles(proniosomes)

    [225]; polymer [77,211,226,240]; solid drug [221]; and solid lipid

    nanoparticles [209,213,216,222,224,228,229,231236,239]. We recog-

    nize that many reports use more than one type of nanoparticle material

    and some of these carrier categories overlap. This integration ofnanoparticle drug delivery technologies has helped to cross validate

    studies.

    4. Recent advances in particle-based drug delivery

    Materials advances are critical for the advancement of topical

    nanoparticle drug delivery and have the potential to dramatically

    improve drug delivery kinetics, but also reveal new biological

    information on skin function. Poly(amidoamine) or PAMAM dendri-

    mers have been used as effective delivery devices for some time,

    reviewed by Jain et al. and Nishiyama et al. [241,242]. Dendrimers are

    considered to be nanoparticles, but are composed of highly branched

    polymers. Recent advances by Venuganti et al. have shown that

    engineering the surface charge on dendrimers can regulate dendrimer

    skin penetration for enhanced small drug delivery [202,203].

    Venuganti et al. manipulated the surface charge of the dendrimers

    by the addition of multiple amine groups (x16, 64, and 256),

    carboxylic acid (x64), and hydroxyl (x64) groups, with isopropyl

    myristate as the control. The antiproliferative drug 5FU was used as

    the model for skin permeation studies in separated porcine ear skin.Dendrimers were used to pre-treat skin, at 0.1 to 10 mM, with 1 mM

    as the chosen concentration, for 24 h. Interestingly, all dendrimer

    groups had significantly higher cumulative amounts of penetrated

    5FU (ranging from 1952 to 2597 g/48 h) than the control (1421 g/

    48 h). The flux of the dendrimer groups (96236 g/cm/h) also

    increased well above the isopropyl myristate control (67 g/cm/h).

    Pre-treatment with the hydroxyl modified dendrimer resulted in the

    highest 5FU levels in skin (23.7 g/mg skin), more than twice that of

    the isopropyl myristate pre-treated group (11.9 g/mg skin). Togeth-

    er, these data suggest that engineering dendrimer surface chemistry

    to match the drug and desired skin penetration characteristics has the

    potential to improve nanoparticle drug delivery.

    Dendrimer drug delivery is currently being explored on many fronts

    and the majority of the knowledge about nanoparticle drug deliverysystems focuses on solid lipid nanoparticles. Kuchler et al. from the

    Shfer-Korting group in Germany, have reported the use of SLN for a

    v