Applications Nano Particles

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Progress in Retinal and Eye Research 29 (2010) 596e609

Contents lists available at ScienceDirect

Progress in Retinal and Eye Research


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

Applications of nanoparticles in ophthalmology


Yolanda Diebold a, b, *, Margarita Calonge a, b
a Ocular Surface Research Group, Edicio IOBA, Campus Miguel Delibes, Instituto de Oftalmobiologa Aplicada (IOBA), Universidad de Valladolid, Paseo de Beln, 17, E-47011 Valladolid, Spain b CIBER de Bioingeniera, Biomateriales y Nanomedicina (CIBER-BBN), Spain

a b s t r a c t
Keywords: Drug delivery Gene delivery Nanomedicine Nanoparticles Ocular drug administration Toxicity

Nanocarriers, such as nanoparticles, have the capacity to deliver ocular drugs to specic target sites and hold promise to revolutionize the therapy of many eye diseases. Results to date strongly suggest that ocular medicine will benet enormously from the use of this nanometric scale technology. One of the most important handicaps of the eye as a target organ for drugs is the presence of several barriers that impede direct and systemic drug access to the specic site of action. Supercial barriers include the ocular surface epithelium and the tear lm, and internal barriers include the bloodeaqueous and blooderetina barriers. Topical application is the preferred route for most drugs, even when the target tissues are at the back part of the eye where intraocular injections are currently the most common route of administration. Direct administration using any of these two routes faces many problems related to drug bioavailability, including side effects and repeated uncomfortable treatments to achieve therapeutic drug levels. In this regard, the advantages of using nanoparticles include improved topical passage of large, poorly water-soluble molecules such as glucocorticoid drugs or cyclosporine for immune-related, vision-threatening diseases. Other large and unstable molecules, such as nucleic acids, delivered using nanoparticles offer promising results for gene transfer therapy in severe retinal diseases. Also, nanoparticle-mediated drug delivery increases the contact time of the administered drug with its target tissue, such as in the case of brimonidine, one of the standard treatments for glaucoma, or corticosteroids used to treat autoimmune uveitis, a severe intraocular inammatory process. In addition, nanocarriers permit the non-steroidal anti-inammatory drug indomethacin to reach inner eye structures using the transmucosal route. Finally, nanoparticles allow the possibility of targeted delivery to reach specic types of cancer, such as melanoma, leaving normal cells untouched. This review summarizes experimental results from our group and others since the beginnings of nanocarrier technology to deliver drugs to different locations in the eye. Also, it explores the future possibilities of nanoparticles not only as drug delivery systems but also as aides for diagnostic purposes. 2010 Elsevier Ltd. All rights reserved.

Contents 1. 2. 3. 4. 5. 6. Introduction: what is nanomedicine? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Drug delivery systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . The eye as a target organ for drug delivery systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Types of NPs for ocular delivery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . NPs and the anterior segment of the eye . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . NPs and the posterior segment of the eye . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .597 .597 .597 599 600 603

Abbreviations: AS-ODNs, Antisense oligonucleotides; CS, Chitosan; CSO, Chitosan oligomers; CsA, Cyclosporine A; ESF, European Science Foundation; HA, Hyaluronic acid; HA-CS NPs, Hyaluronic acid and chitosan-based nanoparticles; HA-PECL NPs, Hyaluronic acid-coated poly-e-caprolactone nanoparticles; IOP, Intraocular pressure; LCS-NPs, Liposomeechitosan nanoparticle complexes; NIH, National Institutes of Health; NPs, Nanoparticles; OIR, Oxygen-induced retinopathy; PBCA, Poly(butyl-cyanoacrylate); PECL, Poly-e-caprolactone; PEG, Polyethyleneglycol; pGFP, plasmid green uorescent protein; PIBCA, Poly(isobutyl-cyanoacrylate); PLA, Poly-D-lactic acid; PLGA, Poly-Dlactic-co-glycolide; RNAi, RNA interference; rAAV, adeno-associated virus vectors; RPE, Retina pigment epithelium; siRNA, small interfering RNA. * Corresponding author. Ocular Surface Research Group, Edicio IOBA, Campus Miguel Delibes, Instituto de Oftalmobiologa Aplicada (IOBA), Universidad de Valladolid, Paseo de Beln, 17, E-47011 Valladolid, Spain. Tel.: 34 983 184750; fax: 34 983 184762. E-mail address: yol@ioba.med.uva.es (Y. Diebold). 1350-9462/$ e see front matter 2010 Elsevier Ltd. All rights reserved. doi:10.1016/j.preteyeres.2010.08.002

Y. Diebold, M. Calonge / Progress in Retinal and Eye Research 29 (2010) 596e609

597

7. 8. 9. 10. 11.

NPs and gene delivery/therapy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 604 Nanoparticle safety: toxicity and interaction with the immune system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 605 Future directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 607 Summary and conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 607 Author disclosure statement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 607 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 607 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 607

1. Introduction: what is nanomedicine? In 2003, the European Science Foundation (ESF) initiated a project aimed at gathering European experts from academia and industry to prepare what was called ESF Forward Look on Nanomedicine, published in 2004 (http://www.esf.org/publications/forward-looks. html). Several workshops were conducted to (i) dene the eld, (ii) discuss the future impact of nanomedicine on healthcare practice and society, (iii) review the state-of-the-art of research, (iv) identify Europes strengths and weaknesses, and (v) deliver recommendations on research trends and organization. These experts formally dened the eld of nanomedicine as the science and technology of diagnosing, treating and preventing disease and traumatic injury, of relieving pain, and of preserving and improving human health, using molecular tools and molecular knowledge of the human body. It is noteworthy that this concept is based in complex systems of nanometre-scale size, i.e., from one nanometre to hundreds of nanometres, with the ultimate goal of using them to achieve medical benets. It is important to bear in mind that the nanometre scale is the scale at which molecules and compounds operate inside living cells. Soon afterward the ESFs publication, the National Institutes of Health (NIH) in the U.S.A. developed a Nanomedicine Roadmap Initiative (http://nihroadmap.nih.gov/nanomedicine/). As a centrepiece of this initiative, a national network of eight collaborative Nanomedicine Development Centers was established in 2006. That multidisciplinary research initiative was primarily directed towards gathering extensive information about nanoscale intracellular biological structures. That information was to be used in the application of newly developed nanomedical therapies to treat specic diseases. As an example of the interest that this topic has awakened in the vision research community, an education course Nanotechnology and Nanomedicine: Applications for Vision Research, was organized in 2005, co-sponsored by the Association for Research in Vision and Ophthalmology and the NIHs National Eye Institute. As anyone can envision, nanomedicine has a relevant position in the global agenda for future development of medical research in the 21st Century. One of the main topics in nanomedicine research is the pharmaceutical development of drug delivery systems. Its goal is the development of improved nano-sized drug carriers consisting of at least two components, one of which is the active therapeutic ingredient. These drug-loaded carriers can be termed nanopharmaceuticals or nanomedicines in a broad sense. This review will focus on the ocular applications of nanoparticles (NPs), a particular type of these drug delivery systems. 2. Drug delivery systems Among the different approaches that have been taken to develop more efcient treatments to ght against human and animal life-threatening or debilitating diseases, the development of drug delivery systems is noteworthy. The purpose of a drug delivery system is to act as a carrier or vehicle for an entrapped or bound

therapeutic agent to reach precisely and effectively the desired site of action. Here we focus on delivery systems that target ocular structures. This concept is particularly interesting when one takes into account the physicochemical features of the frequently marketed biotechnological macromolecules, such as peptides, protein, antibodies, and nucleic acids (Conti et al., 2000; Degim and Celebi, 2007; Levy-Nissenbaum et al., 2008). However, a drug delivery system is more than a simple (nano) carrier. All of the science and technology behind the design of drug delivery systems intends to achieve solutions for key aspects of modern treatments: (i) to control the release of the active agent so that a therapeutic concentration is maintained over a prolonged period of time, (ii) to develop organ or site-specic or even diseasespecic targeting, and (iii) to provide new or more convenient routes of administration for drugs able to reach those locations in the body that are difcult to access. The ultimate goals are to better manage relevant drug-related parameters, such as pharmacokinetics, pharmacodynamics, non-specic toxicity, immunogenicity, biorecognition, and to improve therapeutic efcacy (Vanderwoot and Ludwig, 2007; Sahoo et al., 2008). There are different kinds of drug delivery technologies that are designed to serve as drug delivery systems (Medina et al., 2007; Sahoo et al., 2008; Gaudana et al., 2009). These include, among others, transdermal patches, implants, nanodevices, and cell encapsulation devices. Among nanoparticulate-based drug delivery systems (or nanosystems) (Table 1) one can nd different polymeric formulations made of non-degradable polymers and biodegradable polymers that are either hydrophilic or hydrophobic. Examples of nanosystems that differ in composition include the following: (i) Nanoparticles (NPs) consist of 1 mm or smaller particles composed of various polymers or materials. These are described in detail below. (ii) Liposomes are lipidic membranes, similar to plasma membranes, and surround an aqueous core. A variant of liposomes are niosomes, consisting of non-ionic surfactants. (iii) Emulsions consist of stabilised oil-in-water or water-in-oil mixtures. Others include (iv) nanosuspensions, (v) dendrimers, (vi) nanoparticleloaded contact lenses, (vii) nanotubes and fullerenes composed of carbon-based nanomaterials, and (viii) quantum dots made of semiconductor materials with uorescent properties and covered with other materials. However, drug delivery systems other than those collectively named as nanoparticles are out of the scope of this review, and therefore we will not comment on them. 3. The eye as a target organ for drug delivery systems There are a plethora of ocular disorders that may be visionthreatening. The responsiveness towards classically developed drugs is limited and most fail to correct the underlying problem. Thus, there is a scarcity of truly curative treatments for most eye diseases. The main reasons for these limitations are biopharmaceutical problems related to the special characteristic of the eye that restricts drug bioavailability. The eye is partially isolated from the remainder of the body by several types of barriers that impede the effective passage of many drugs (Fig. 1), leading to

598

Y. Diebold, M. Calonge / Progress in Retinal and Eye Research 29 (2010) 596e609

Table 1 Nanoparticulate drug delivery systems (or nanosystems) used as carriers for drug administration. Nanosystem Nanoparticles Liposomes Niosomes Emulsions Nanosuspensions Dendrimers Nanoparticle-loaded contact lenses Nanotubes and fullerenes Quantum dots According to Sahoo et al. (2007) and Gaudana et al. (2009). Composition Natural or synthetic polymers, metals, lipids, phospholipids Phospholipids Non-ionic surfactants Oil-in-water and water-in-oil mixtures that require surfactants Inert polymer resins Synthetic polymers Different hydrogel-based lenses with nanoparticulate-based drugs incorporated in the lens matrix Carbon-based nanomaterials Semiconductor materials covered with other materials Potential application in the eye Yes Yes Yes Yes Yes Yes Yes Not tested yet Yes (diagnostic)

a minimal dose absorption (Urtti, 2006). These barriers consist of the (i) muco-aqueous layer of the tear lm that protects the anterior surface of the eye, (ii) corneal epithelium with abundant tight junctions and desmosomes, (iii) iris blood vessels that lack fenestrations, (iv) non-pigmented layer of the ciliary epithelium that constitutes the bloodeaqueous barrier and limits the passage of molecules from the blood to the inner part of the eye, and (v) retinal pigment epithelium (RPE), along with the endothelium of the retinal vessels, constitute the inner and the outer blooderetina barriers, respectively, that limit the passage of molecules from the blood to the retina and vitreous cavity. Additionally, certain physiological processes contribute to the poor efcacy of conventional drug formulations. For instance, blinking and tear drainage through the lachrymal drainage system act to reduce the residence time of topically applied molecules. Eye drops placed onto the ocular surface are washed away in less than

Fig. 1. Schematic presentation of the ocular structure showing a summary of ocular pharmacokinetics. The numbers refer to following processes: 1) transcorneal permeation from the lachrymal uid into the anterior chamber, 2) non-corneal drug permeation across the conjunctiva and sclera into the anterior uvea, 3) drug distribution from the bloodstream via the bloodeaqueous barrier into the anterior chamber, 4) elimination of drug from the anterior chamber by aqueous humour passage into the trabecular meshwork and Sclemms canal, 5) drug elimination from the aqueous humor into the systemic circulation across the bloodeaqueous barrier, 6) drug distribution from the blood into the posterior eye across the blooderetina barrier, 7) intravitreal drug administration, 8) drug elimination from the vitreous via the posterior route across the blooderetina barrier, and 9) drug elimination from the vitreous via the anterior route to the posterior chamber. (Taken from Urtti A., Adv Drug Deliv Revs (2006), with permission of Elsevier)

30 s (Kaur and Kanwar, 2002). The maintenance of corneal transparency is based upon several strategies, one of them being the sealing of the corneal epithelium by means of specialized structures, such as tight junctions and desmosomes. The corneal epithelium is therefore almost impermeable to any substance larger than 500 Da (Hmlinen et al., 1997). Most of the commonly used topical drugs are larger than that and, consequently, do not cross the cornea. Instead, they permeate throughout the conjunctiva and the underlying sclera in what is known as non-productive passage. Altogether, less than 5% of topically administered drugs reach intraocular tissues (Keister et al., 1991). Therefore, the existence of several ocular tissue and cell barriers along with the physiological processes impede the effective passage of many drugs, leading to a minimum dose absorption into the eye. Drug delivery systems hold promise to be an important part of the new therapeutic armamentarium in ophthalmology. Since the early study of Wood et al. (1985) showing the intrinsic capacity of NPs to adhere to the ocular surface and interact with the epithelium, applications of nanotechnologies to solve eye problems have been sought. Knowledge derived from drug delivery systems using non-ocular routes of administration has stimulated researchers to nd applications for them in ophthalmology. What makes them attractive to treat eye diseases is the possibility of the controlled release of drugs, especially poorly water-soluble ones, that surpasses the ocular barriers and effectively reaches the target. Nanoparticulate systems improve the delivery of poorly watersoluble drugs while signicantly reducing toxicity compared to the free drug. Increasing attention has been focused particularly on this aspect due to the clear therapeutic implications. Also, the micro- or nano-size of such drug carriers is appealing. To envision the potential of nanoparticles as drug carriers that can treat ocular disorders, it is necessary to understand the peculiarities of the drug administration routes in the eye. First, local delivery of the drug is preferred over systemic delivery. There are several different modalities for ocular drug administration (Table 2). The most common include liquids topically applied onto the front of the eye in the form of eye-drops, sub-conjunctival or sub-Tenons injection in the conjunctival tissue or below the Tenons capsule, and intravitreal injection. Also, intraocular implants made of either biodegradable or non-biodegradable polymers loaded with different drugs are used to provide longterm drug presence at the implantation site. Topical administration of drugs is used to alleviate the symptoms and signs caused by ocular surface inammatory disorders, such as dry eye syndrome and allergic diseases that affect millions of people worldwide. They are also used to treat infections and complex, vision-threatening diseases, such as glaucoma or intraocular inammation (uveitis). In most cases, the patients themselves

Y. Diebold, M. Calonge / Progress in Retinal and Eye Research 29 (2010) 596e609 Table 2 Modalities for ocular drug administration. Modality Topical eye-drops Anatomical location Ocular surface (onto the tear lm and corneal and conjunctival epithelia) Kind of administered drug Articial tears Anti-infectives Anti-allergics Anti-inammatories Anti-hypertensives Anaesthetics Mydriatics, miotics, cyclopegics

599

Periocular injection Sub-conjunctival Sub-Tenon Transeptal Retro-orbital Intraocular injection Intracameral Intravitreal

Sub-conjunctval space

Anterior orbit Posterior orbit

Anti-infectives Mydriatics Corticoids Anaesthetics Corticoids Anaesthetics Corticoids Anti-infectives Anti-infectives Anti-angiogenics Corticoids

Anterior chamber Vitreous

instil the drops many times a day, and the treatment is not usually curative but just palliative. To summarize the necessity of using a carrier for ocular drugs, the key is to achieve adequate bioavailability. There are many challenges for ocular drug delivery systems. In some cases, the goal is to stop or reverse problems such as degenerations in the retina and neovascularisation in the cornea and in the retina. In other cases, the drugs must contribute to the success of refractive corneal surgery and the healing process, tissue transplantation and growth factor delivery for retinal and corneal regenerative medicine, and gene therapy for hereditary retinal disorders.

4. Types of NPs for ocular delivery There has been an evolution in the design of nanoparticles (NPs), and Snchez and Alonso (2006) extensively reviewed the features of polymers used to prepare the carriers. We analyze in this review the progress that has been made from a practical point of view in terms of therapeutic and diagnostic applications of NPs

in the eld of ophthalmology. The general term nanoparticle will be used in this review for the sake of simplication. However, this term can refer to nanospheres or nanocapsules, all of which are more properly designated as colloidal systems (Fig. 2). Nanospheres are matrical-type nanostructures that entrap or adsorb the biologically active molecule onto the surface. Nanocapsules are reservoir-type nanostructures within a surrounding polymeric wall containing an oil core where the active molecule is dissolved. These nanostructures can be coated with a hydrophilic polymer or even functionalized with antibodies attached to the coating. There are different methods to prepare NPs and load them with therapeutic molecules (for review, please see Pinto-Reis et al., 2006). NPs can be made of a great variety of materials including organic carbon-based biopolymers and inorganic ceramic, metallic, and semiconductor materials. Some of the most commonly used biomaterials include polyacrylates, polyalkylcyanoacrylates, polylactide (PLA), polylactideepolyglycolide (PLGA), polycaprolactones, dextran, albumin, gelatin, alginate, collagen, hyaluronic acid, and chitosan. The possibilities for their design are almost innite. The intended application inuences the kind of material used for their preparation. The different chemical ways in which bioactive molecules can be associated with polymers and give rise to drugloaded NPs include entrapment, encapsulation, adsorption, or attachment to a polymer. Also, NPs can be prepared in different sizes, charge, and other physicochemical features. This confers great versatility upon them. The physicochemical characteristics of NPs not only confer versatility in terms of the kind of drug to be loaded, but they also inuence the cellular uptake and intracellular trafcking. Additionally, the physicochemical characteristics are critical for other properties, such as interaction with plasma proteins, other blood components (Dobrovolskaia et al., 2008), and with immune cells (Dobrovolskaia and McNeil, 2007), all of which are relevant to the organ distribution. For instance, opsonisation of NPs covers the surface by opsonins present in the blood and creates a molecular signature (Aggarwal et al., 2009) that determines the route of NP internalization in phagocytic cells and eventually their fate. However, if the NP surface is covered by a hydrophilic coat of poly(ethylene glycol) (PEG), opsonisation is prevented. This has been termed a stealth effect (Gref et al., 1994), and it reduces the rate of cellular uptake. Consequently, PEG-NPs stay longer in the bloodstream. In addition, the rate of clearance of NPs is an important consideration with regard to potential toxicity caused by their permanence in the target tissue. Therapeutic targets for drugs can be located extra- or intracellularly. With regard to classical drugs with sizes greater than

Fig. 2. Evolution of colloidal systems: The rst generation of colloidal systems consisted of polymerized matrices known as nanoparticles and nanocapsules composed of a polymeric wall containing an oil core. Second generation systems were similar to nanocapsules, but with improved hydrophilicity associated with a coating polymer. Third generation systems are functionalized by the attachment of antibodies or lectins, among other targeting moieties to the surface structure. (Adapted from Snchez A. and Alonso M.J., 2006).

600

Y. Diebold, M. Calonge / Progress in Retinal and Eye Research 29 (2010) 596e609

1 kDa, stability conditions or hydrophilicity makes them unable to cross plasma membrane. NPs, as well as other drug delivery systems, enable the loaded bioactive molecule to cross cell membranes and epithelial barriers by using different internalization pathways. This is one of the most interesting aspects of this technology. Reports showing that submicrometer delivery systems such as NPs can enter cells and become concentrated into nondiffusible molecules appeared in the 1970s (Couvreur et al., 1977). Much research has been done since then and much knowledge has been gained related to both carrier design and mechanisms of biological action. It is now known that the physicochemical characteristics of nanocarriers, such as size, shape, surface charge, surface coating, and surface functionalization with targeting ligands, along with the target cell type, determine the internalization pathway and the intracellular fate. Increasing numbers of studies about cell targeting and internalization pathways are being accomplished, mainly using nanomedicines to treat infectious diseases and cancers. Recent interesting reviews deal with these topics (Bareford and Swaan, 2007; Frenkel, 2008; Hillaireau and Couvreur, 2009). The main internalization pathways for NPs discussed in those reviews are summarized in Table 3. Phagocytosis is the typical pathway for NPs intravenously administered. It implies NP opsonisation as mentioned above and the involvement of specic cells with phagocytic capacity that will ingest the NPs. In the cytoplasm, NPs and their payload will become accessible to lysosomes and experience enzymatic degradation. This step is important to ensure drug release. Endocytosis occurs in all mammalian cells by means of pitted or invaginated membrane regions that are coated by either of two specic proteins. Pitted membranes coated by clathrin can be formed by receptor-mediated and receptor-independent processes with different internalization rates. Invaginated membranes coated by another protein, caveolin, can also conduct receptor-mediated endocytosis. NPs can be targeted to interact with cell receptors of interest to facilitate internalization by either of these receptormediated endocytic mechanisms. Loaded NPs internalized by clathrin-mediated endocytosis are directed to lysosomal degradation. Those internalized by caveolin-mediated endocytosis accumulate in the endosomes (caveosomes) and are delivered to other subcellular compartments different from lysosomes. This last mechanism has been used, for instance, to deliver chemotherapeutics to cancer cells. Also, clathrin- and caveolin-independent endocytosis mechanisms have been described. Finally, macropynocytosis also occurs in all kind of cells. It is directed by actin-driven membrane protrusions that form large (>1 mm) endocytic vesicles (macropinosomes). Eventually, those vesicles fuse with lysosomes. This pathway is the least specic of all mentioned. Notably, several endocytic mechanisms can take place

simultaneously. The above mentioned review articles present abundant examples of all these mechanisms for different kinds of NPs tested in a variety of cells and tissues. Surprisingly, there is a scarcity of studies about internalization pathways and intracellular trafcking of NPs intended for ocular application (see Table 1). Intracellular metabolism of delivered drugs differs according to the internalization pathway. It implies a physical separation from the NP transporter, and their physicochemical characteristics will determine the degradation ratio. These are key aspects for the pharmacological activity of the carried molecule. Different kinetic processes are described to control the release proles of drugs from NPs, including (i) desorption of the surface bound or adsorbed drug, (ii) diffusion through the NP matrix or the polymer wall, (iii) NP wall erosion, and (iv) a combination of erosion and diffusion processes (Soppimath et al., 2001; Harush-Frenkel et al., 2008). For drugs that uniformly distribute or dissolve through the NP matrix, diffusion and biodegradation govern the release process. If drug diffusion is faster than polymeric degradation, drug release mainly occurs through diffusion; if not, then drug release will occurs through degradation of the polymer. Finally, there are other important factors involved in the design of a nanomedicine, including the modality of administration such as injection or topical application, and the features of the entrapped drug itself. This complex picture is completed with the requirement of minimal potential toxicity (see specic chapter) and the continuous need for improved efciency. Different kinds of NPs prepared with some of the materials described above have been studied in the search for alternatives in ophthalmic treatments. We present in this review those that have been tested either in vitro or in vivo. 5. NPs and the anterior segment of the eye The bioavailability of an instilled conventional drug onto the ocular surface is usually low. Most of it is lost due to physiological mechanisms, such as tear drainage and blinking, a few seconds after instillation (Bayens and Gurny, 1997). Thus, there is a short pre-corneal residence time, usually less than 2 min, and a nonproductive absorption thorough the well vascularised conjunctiva and the nasolachrymal drainage system (Jarvinen et al., 1995). Therefore, the picture we face includes a very limited absorption of drug, a potential although limited access to intraocular tissues through the conjunctivalescleral pathway, and the risk of systemic side effects. For those reasons, intensive research in recent decades has focused on increasing the corneal penetration of topically applied drugs (Schoenwald, 1990; Sasaki et al., 1999; Edwards and Prausnitz, 2001; Mannermaa et al., 2006). NPs were considered to offer the possibility of a more facile delivery and transport across tissues, and consequently their potential started to be studied.

Table 3 Main pathways for nanoparticle (NP) internalization and preferential degradation route. Pathway Target cells Endocytic vesicles size 0.25e10 mm z120 nm z60 nm Degradation route Examples of NPs studied for ocular delivery e SLNs loaded with pCMS-EGFP (del Pozo-Rodrguez et al., 2008) HAS NPs loaded with SOD1 gene (Mo et al., 2007) HA-CSO NPs loaded with pSEAP (Contreras-Ruiz et al., Submitted for publication) PLGA NPs loaded with 6-cumarin (Qaddoumi et al., 2003) e

Phagocytosis Clathrin-mediated endocytosis Caveolin-mediated endocytosis

Professional phagocytes All mammalian cells All mammalian cells

Early endosomes and lysosomes Early endosomes and lysosomes Endosomal accumulation; nondegradative pathway

Clathrin- and caveolin-independent endocytosis Micropynocytosis

All mammalian cells All mammalian cells

z90 nm >1 mm

Early endosomes Macropinosomes and lysosomes

SLNs, solid lipid NPs; HSA NPs, human serum albumin NPs; SOD1, Cu, Zn superoxide dismutase gene; pCMS-EGFP, plasmid DNA encoding enhanced green uorescent protein; HA-CSO, hyaluronic acid-chitosan oligomers; pSEAP, plasmid DNA encoding secreted alkaline phosphatise; PLGA, Poly-D-lactic-co-glycolide.

Y. Diebold, M. Calonge / Progress in Retinal and Eye Research 29 (2010) 596e609

601

In 1986, an early attempt to use NPs was done with poly(butylcyanoacrylate) (PBCA) NPs loaded with progesterone, a highly lipophilic molecule (Li et al., 1986). The NPs were topically applied to rabbit eyes and concentrationetime proles in different ocular tissues were assessed. The authors found a decreased progesterone concentration in tissues when compared to control solutions, suggesting a greater afnity of progesterone for the NP polymer than expected. As a consequence, progesterone was less available for absorption during its residence time in the pre-corneal area. The lesson learned from these experiments was that it is important to optimize the physicochemical relations between the polymers and drug to obtain an efcient carrier. Later, Losa et al. (1993), used poly (isobutyl-cyanoacrylate) (PIBCA) and poly-e-caprolactone (PECL) nanocapsules for the ocular delivery of metipranolol, a beta-blocker used to treat certain types of glaucoma. In that work, the polymer coating was not responsible for the differences observed regarding the controlled release of the encapsulated drug. Instead, the drug release prole was mainly inuenced by the type of oil core in the nanocapsule. However, a certain contribution of the polymer coating on emulsion stability was acknowledged. Our group later studied PECL nanocapsules loaded with 1% cyclosporine (CsA) in a rat model of corneal transplantation rejection (Juberas et al., 1998). This formulation had been developed to improve the corneal penetration of CsA applied topically (Calvo et al., 1996). This well-known immunosuppressive drug, widely used in transplantation patients, has severe systemic side effects such as nephrotoxicity, hepatotoxicity, and hypertension. These side effects have limited its use in the local management of immune rejection of corneal grafts. By using PECL nanocapsules, toxic effects of systemic administration of CsA were avoided along with improved ocular penetration. While corneal graft rejection was not prevented by the CsA-loaded PECL NPs, the failure was not considered to be a consequence of negative interactions between the polymer and the drug. Improved formulations using polymers with known biocompatibility and biodegradability, such as poly-D-lactic acid (PLA) and its copolymer glycolic acid (PLGA), were developed later. Flurbiprofen-loaded PLGA NPs successfully reduced inammation in an in vivo rabbit model of ocular inammation (Vega et al., 2006). In comparison with commercial urbiprofen formulations, urbiprofen-loaded NPs were more effective in reducing inammation as evaluated by direct observation of clinical signs. More recently, we and others have developed applications using chitosan (CS)-based nanosystems. CS is a natural polysaccharide with interesting features, such as biocompatibility and biodegradability, mucoadhesiveness, and the ability to transiently enhance the permeability of mucosal barriers. These features have made it quite useful in the development of drug delivery systems for transmucosal administration (for review see Alonso and Snchez, 2003; Paolicelli et al., 2009). NPs made of CS and carbopol, a cross-linked polymer of acrylic acid (Kao et al., 2006), combined properties of both polymers, such as stability in aqueous solution, small size, improved biocompatibility, and positive charge to facilitate interaction with the negatively charged biological membranes. CS/carbopol NPs can be loaded with pilocarpine, a parasympathomimetic drug used as treatment for open-angle glaucoma. Patients with this form of glaucoma need to frequently instil pilocarpine eye-drops, which increases aqueous humour outow for only 4e8 h Kao et al. (2006) compared the efcacy of pilocarpine-loaded NPs against liposomes, gel, and eye-drops formulations of pilocarpine in normal rabbits. Pilocarpine induces a miotic response measured as a decrease in pupil diameter, and the effect of the CS/carbopol NPs lasted up to 24 h, much longer than the other formulations. The release prole, previously studied in vitro, showed an initial burst release followed by a sustained

release for at least 24 h. These results are very promising, and further evaluation in a glaucoma animal model is now warranted. Recently, another CS-based nanosystem, CS/sodium alginate NPs, has been reported as a potential reservoir for topical delivery of gatioxacin, a potent fourth-generation uoroquinolone with low intraestromal penetration that is mainly used for microbial keratitis (Motwani et al., 2008), but so far, the NP characterization has been only physicochemical, and no in vitro or in vivo experiments have been reported. The transport pathways by which NPs penetrate the ocular surface tissues are of great interest. Zimmer et al. (1991) studied the ocular transport pathway of uorescein-labeled PBCA NPs in rabbits. They found the uorescence signal localized inside conjunctival and corneal epithelial cells, and observed differences in depth of tissue penetration. They proposed a variety of pathways to explain their data, including NP endocytosis, lysis of cell membrane by NP metabolic degradation products, and a transcellular route. The transcellular route was also proposed for coated PECL nanocapsules (de Campos et al., 2003). A critical question with regard to NP delivery of drugs is the concentration and duration of the drug in the target and surrounding tissues. Losa et al. (1991) tested PBCA NPs with different stabilizer agents to improve the binding of the antibiotic amikacin sulphate to the NPs. One of the formulations, using Dextran 70000 as a stabilizer, resulted in a signicant increase of the amikacin concentration not only in the cornea but also in the aqueous humour. de Campos et al. (2001) studied the distribution of CsA-loaded CS NPs in different rabbit ocular tissues. In that study, therapeutic concentrations of this immunosuppressant drug adequate to modulate the local immune response were maintained in the cornea and conjunctiva for 48 h post-administration. However, those concentrations were not achieved using a formulation consisting on a CsA suspension in either a chitosan aqueous solution or a CsA suspension in water. The concentration of CsA in aqueous humour, iris, and ciliary body were extremely low. In addition, no detectable CsA levels were measured in plasma. Therefore, a prolonged local drug delivery was achieved using the CS NPs with no signicant accumulation in intraocular tissues. The surface characteristics of the nanocarriers also have an inuence on the interaction with the ocular surface structures. For instance, CS-coated PECL nanocapsules enter the corneal epithelium in vivo more efciently than uncoated PECL or polyethylene glycol-coated PECL nanocapsules (de Campos et al., 2003). We have already mentioned the exceptional biological features of CS, especially mucoadhesiveness and the ability to transiently enhance the permeability of mucosal barriers. Our group is most interested in this approach using mucoadhesive NPs able to reach the anterior structures of the eye after topical administration. We started testing CS NPs because of the great potential envisioned for this polymer in the ophthalmology eld (Alonso and Snchez, 2003). We reported that CS NPs did not cause toxicity-related alterations in several cell lines derived from human conjunctiva epithelium (de Campos et al., 2004; Enrquez-de-Salamanca et al., 2006). We were able to identify albumin-loaded CS NPs inside the cells by uorescence microscopy in a time-dependent manner. With in vivo experiments using albino rabbits, we conrmed that CS NPs were well-tolerated by ocular surface structures, causing no harm or inammation, as determined by histopathological analysis. The corneal and conjunctival tissues that took up the CS NPs showed interesting tissue-related distribution patterns (Fig. 3). Both cornea and conjunctiva incorporated the NPs; however, the conjunctiva was more permeable to the nanocarrier as shown by the deeper penetration into the epithelium and underlying stroma. Other authors have also explored similar strategies. For instance, Yenice et al. (2008) used hyaluronic acid (HA)-coated PECL nanospheres for

602

Y. Diebold, M. Calonge / Progress in Retinal and Eye Research 29 (2010) 596e609

Fig. 3. CS NP in vivo uptake. Fluorescence microscopy of ocular surface structures of sham-treated (A, D), CS NP-treated (B, E) and contralateral control (C, F) rabbit eyes. Representative corneal (A, B, C) and conjunctival (D, E, F) sections are shown. No uorescence was detected in sham control corneas (A) or conjunctivas (D). (B) Corneal epithelial cells of the CS NP-treated rabbits were uniformly uorescent. Inset in (B): zoom area showing details of the corneal epithelium uorescence pattern. (E) Fluorescence in conjunctival epithelial cells was intense in the apical cell membranes and positive along the basolateral cell membrane. Inset in (E): zoom area showing the basolateral membrane uorescence staining in goblet and non-goblet cells. (C, F) Some uorescence was detected in corneal and conjunctival epithelial cells from contralateral control eye (OS), although much less intense than in the treated (OD) eye. Scale bar (AeF) 50 mm; Inset scale bar 10 mm. (Taken from Enrquez de Salamanca et al., Invest. Ophthalmol. Vis. Sci., 2006, with permission of the Association for Research in Vision and Ophthalmology).

corneal CsA delivery. The concentrations in rabbit cornea were 10e15-fold higher than that achieved when CsA was administered as solution in castor oil. These results moved us to explore different biomaterial combinations intended specically for the ocular surface tissues. Consequently we joined liposomes and CS NPs to form a new nanosystem that we termed liposomeechitosan nanoparticles (LCS-NPs) (Diebold et al., 2007). The theoretical advantages of these complexes are the combination liposome biocompatibility with biological membranes and the demonstrated properties of CS NPs. We tested three different complex formulations, showing again their potential for ocular administration. Using mucus-producing primary cultures of conjunctival epithelium, we observed that the three nanosystems were rst retained in the in vitro mucus layer

and then entered the epithelial cells, depending on the particular NP composition (Fig. 4). We consider this feature a potential advantage that can be used to modulate the retention time of the nanocarrier in the tear lm. Different in vivo retention times would be required depending on the encapsulated drug. Use of different lipid carriers has had renewed interest. An example is the recent paper by Attama et al. (2009) that presents phospholipid nanoparticles made in theobroma oil. These were designed to incorporate timolol hydrogen maleate, a water-soluble drug used as a rst-choice treatment for glaucoma. Using a modied Franz diffusion cell apparatus, they did drug permeation experiments using a bio-engineered cornea construct, composed of human immortalized corneal endothelial cells, stromal broblasts, and epithelial cells. This in vitro study, although showing promising

Fig. 4. Confocal microscopy images of primary cultures of human conjunctival epithelium. Control cultures had normal morphology when viewed with transmitted light (TL). During 30 min of incubation, LCS-NP complexes (green) passed through the mucus layer and were present in deeper cell layers. They formed aggregates with different patterns for each type of LCS-NP tested. Z-series prole images (lowest panels) are projections of stacked image proles from optical sections captured along the Z-axis. These conrmed the intracellular presence of LCS-NPs (green), which were localized among the actin bres in the cytoskeleton (red) stained with phalloidin. Scale bar 25 mm. Images are representative of at least three independent experiments. (Taken from Diebold et al., 2007 (Biomaterials), with permission of Elsevier).

Y. Diebold, M. Calonge / Progress in Retinal and Eye Research 29 (2010) 596e609

603

results in terms of sustained release and timolol permeation, lacks a complementary toxicity study. We have recently started working with NPs composed of hyaluronic acid and chitosan (HA-CS NPs) (de la Fuente et al., 2008a). This new development is capable of encapsulating macromolecules of both hydrophilic nature, such as the protein bovine serum albumin, and of hydrophobic nature, such as the polypeptide CsA. Also, HA-CS NPs can carry plasmid DNA and may be suitable for gene delivery as shown with ocular surface-derived cell lines (de la Fuente et al., 2008b). The most important aspect is the good in vivo tolerance of these nanosystems (Contreras-Ruiz et al., 2009), which opens the possibility of testing them in actual disease models. By exploiting the known properties of HA-CS NPs, our group is currently working to design specic nanomedicines that utilize gene therapy to target certain ocular surface inammatory diseases. It is also worth mentioning that primary open-angle glaucoma is the most widespread neuropathy, affecting 60.5 million people in the world (Quigley and Broman, 2006). Aside from surgery, the only partially effective treatment for this disease is IOP-reducing drugs applied topically onto the cornea. The necessity of numerous daily applications of eye-drops for life is a serious issue for patients, besides the risk of side effects in the anterior structures of the eye. While there was an early interest in drug delivery systems, such as the Ocusert in 1974 for pilocarpine sustained release, there are surprisingly few reports testing NPs intended for glaucoma treatment. Zimmer et al. (1994) tested different pharmaceutical aspects of pilocarpine-loaded PBCA NPs versus a standard pilocarpine solution in an elevated IOP rabbit model. The NPs were better in reducing IOP and maintaining miosis, especially at lower drug concentrations, than the drug solution. The NPs induced maximum reduction of IOP at 2e3 h, whereas the drug solution maximum response was at 1e2 h. Epinephrine-loaded poly-N-isopropylacrylamide NPs were also tested in rabbit to evaluate IOP-lowering effect (Hsiue et al., 2002). The polymer in this nanosystem is thermosensitive and undergoes a phase transition when the temperature increases to about 32  C. This allows the progressive release of epinephrine after being topically administered. The NPs had six times more longlasting effect compared to conventional eye-drops. Finally, Wadhawa et al. (2009) reported a signicant IOP reduction in rabbit eyes exposed to CS-HA NPs loaded with timolol compared to standard timolol eye-drops or blank NPs. There were no irritant effects. Although not properly NPs but a nanodevice, it is worth mentioning here the development of a prototype for a nanodrainage system to be implanted in the sclera as a bypass route for humour aqueous outow (Pan et al., 2006). This new concept might revolutionize glaucoma treatment in the coming years. 6. NPs and the posterior segment of the eye Even though the cornea constitutes one of the most selective barriers to foreign molecules for the eye, transcorneal penetration of topically administered ophthalmic medicines intended for the posterior segment is persistently sought. The reason is that currently, the best way to treat intraocular inammation, either infectious or non-infectious, is by injecting drugs into the vitreous. The vitreous is a gelatinous, cell-free structure that is capable of retaining molecules and also delivering them to nearby structures, such as the ciliary body or the RPE, a vital component of the retina. Frequent intraocular injections are needed to treat retinal disorders. With these injections come potential undesired side effects, higher risk of infections, and poor acceptance by the patient. More frequent side effects associated with repeated intravitreal injections include increased risk of cataract development, vitreous hemorrhage, retinal detachment, and endophthalmitis.

The prospect of frequent intravitreal injections to treat serious intraocular disorders affecting the choroid and retina has moved researchers to look for better solutions derived from the use of NPs as drug carriers. However, the scenario is quite challenging for NPs. Typically, the cornea is penetrated by less than 5% of drugs applied as liquid eye-drops (Keister et al., 1991) because of the limitations mostly imposed by tear turnover associated with blinking and the nasolacrymal drainage system. In addition, the drugs must diffuse a great distance between the ocular surface and the intraocular targets. Therefore, usually eye-drops do not provide sufcient drug concentration in the posterior ocular tissues. On the other hand, systemic drug administration delivered through the blood vascular system is not very effective because of the uveal bloodeaqueous and blooderetina barriers. As a consequence, a poor doseeresponse prole for vitreoretinal diseases is generally achieved, and a large amount of the drug is needed to maintain therapeutic levels, usually for insufcient amounts of time (Geroski and Edelhauser, 2001). Additionally, the high concentration of drugs needed to penetrate the bloodeocular barriers is often associated with systemic side effects. Thus, much effort has been invested in the last decade to optimize drug delivery systems for intraocular treatment. Alternatives to intravitreal or periocular injections, including scleral implants and devices, transdermal patches, and different iontophoretic devices including hydrogel reservoirs, have been explored with variable results (for review, see del Amo and Urtti, 2008). However, the ability to achieve long-term drug delivery in the retina and nearby tissues while reducing the number of intraocular injections to just one seems feasible at this time. Several kinds of NPs carrying diverse active molecules, including genetic material, are currently in pre-clinical studies using the above mentioned approaches (Bourges et al., 2003; Bejjani et al., 2005; Normand et al., 2005; Irache et al., 2005; Farjo et al., 2006; Paasonen et al., 2007). The underlying idea is to take advantage of the vitreous capacity for retaining and delivering molecules to tissues with which it is in direct contact and to use it as a biological reservoir once the NPs are placed inside. If therapeutic levels of a drug can be maintained for months after a single intravitreal injection, that would be considered an enormous improvement for the quality of life of many patients. There are promising studies reported in the recent literature on the use of intravitreally injected NPs. Ganciclovir-loaded albumin NPs are an interesting example. Ganciclovir is one of the standard treatments for cytomegalovirus retinitis, a prevalent infectious retinal disease in immunosuppressed patients, such as those with AIDS. In vitro experiments demonstrated that albumin NPs released ganciclovir in a sustained way (Merodio et al., 2001), with a signicant improvement of drug uptake by cytomegalovirusinfected human cells (Merodio et al., 2002). For single intravitreal injections in rats, these NPs were safe, well-tolerated carriers not only for ganciclovir, but also for the anti-cytomegaloviral oligonucleotide analog formivirsen. They were present in the vitreous and ciliary body for at least two weeks (Irache et al., 2005). RPE cells have the capacity to take up different kinds of NPs (Bourges et al., 2003; Bejjani et al., 2005; Normand et al., 2005) opening the possibility of using them to treat retinal disorders associated with ageing or photoreceptor dystrophies. The purpose will be to target these cells with specic molecules or genetic material capable of reversing or stopping the processes leading to these diseases. Bourges et al. (2003) tested in rats intravitreal PLA NPs loaded with uorochromes and showed a preferential localization at the RPE cells after 24 h. The most interesting achievement was that RPE cells retained the NPs, which continuously delivered the uorochrome for months after the single injection. Fluorescence diffusing from the NPs was observed in distant parts of

604

Y. Diebold, M. Calonge / Progress in Retinal and Eye Research 29 (2010) 596e609

retinal tissue, ganglion cells, and rod outer segments for up to four months after the injection. In contrast, detection after injection of the free uorochrome lasted barely one week. Later, Bejjani et al. (2005) studied in vitro and in vivo PLA and PLGA NPs loaded not only with uorochromes but also with model plasmids. NPs encapsulating a plasmid encoding red nuclear uorescent protein were localized in the RPE cells 24 h after intravitreal injection in rats. Effective plasmid expression was achieved after four days of injection and expression-associated red uorescence remained detectable in RPE cells during the following three weeks, with no apparent tissue damage or toxicity. In all of these studies, the association of the delivered molecules with the NPs and kinetic release rate proles were determined prior to the in vivo experiments. From a therapeutic point of view, not only is the reduction in the number of intravitreal injections by the use of NPs a goal, but the improvement in intraocular availability of topically applied drugs is an important and remarkably challenging goal. An illustrative example is the case of steroidal and non-steroidal anti-inammatory drugs. There are many clinical situations in which these drugs are normally used. However, because they are almost completely insoluble in water and because they are effectively excluded from intraocular sites by the various bloodeocular barriers, the ability to reach intraocular targets is quite low. For instance, surgical traumas, such as cataract surgery, cause miosis that is treated with non-steroidal anti-inammatory drugs. Pignatello et al. (2002) developed ibuprofen-loaded NPs made of inert polymeric resins (Eudragit RS100) optimized as a pharmaceutical preparation. The NPs showed a controlled ibuprofen release prole in vitro, and they had a high efcacy in reducing miosis with an in vivo model of ocular trauma. The therapeutic effect was achieved with lower drug concentration than in an eye-drop formulation and without any toxic effect on ocular tissues. Recently, Kassem et al. (2007) studied different nanosuspensions prepared by high-pressure homogenization of three insoluble glucocorticoids: hydrocortisone, prednisolone, and dexamethasone. Using normotensive rabbits, they determined if the glucocorticoid-associated NPs instilled into the lower cul-de-sac demonstrated enhanced absorption and improved intensity of drug action. Based on the measured increase in intraocular pressure (IOP), they reported not only improvements in both, but also a signicant extension of the glucocorticoid action. Interestingly, intravenously injected PLA NPs encapsulating betamethasone phosphate effectively controlled inammation in a rat model of experimental autoimmune uveoretinitis (Sakai et al., 2006). The studies described above have opened a new perspective for the treatment of retinal and uveal disorders. Nevertheless, to implement this kind of delivery system in a clinical setting, more functional studies are needed to exclude any impairment of the retinal function and vision and the development of accompanying chronic inammatory processes. Surely, in coming years we shall see reports dealing with all of these topics. 7. NPs and gene delivery/therapy The Roadmap for Nanomedicine (http://nihroadmap.nih.gov/ nanomedicine/) released by the NIH presents NPs as a strategy to improve non-viral gene transfer. The eye is an excellent candidate for gene therapy for two main reasons: it has immune privilege, and it is affected by many well understood genetic-based diseases. Immune privilege means that the immune system is driven towards tolerance to foreign antigens rather than rejection and inammation, the normal responses. Immune tolerance serves to protect vision by avoiding the collateral inammation that is associated with the immune response against any antigen. Also, it is

a small and closed organ, with very limited diffusion of locally applied active molecules to the bloodstream because of the bloodetissue barriers. Hence, there is a growing interest in exploring the suitability of gene delivery strategies in ocular therapy since the 90s (Nussenblatt and Csaky, 1997; Tanelian et al., 1997). Genetic-based therapies can be developed using different nucleic acids such as DNA, antisense oligonucleotides (AS-ODNs), small interfering RNA (siRNA), and aptamers. AS-ODNs are synthetic molecules of short sequences, 13e25 nucleotides, that bind to specic mRNAs. By binding to the mRNA molecules, AS-ODNs are capable of stopping translation of the mRNA and, consequently, block the protein synthesis of the targeted gene (Loke et al., 1989). siRNAs share with AS-ODNs the capacity of blocking protein synthesis from a given mRNA. However, the gene silencing mechanism by which it is performed, called RNA interference (RNAi), is different (Leung and Whittaker, 2005; Bumcrot et al., 2006). RNAi is induced in mammalian cells by means of synthetic double-stranded siRNAs. These molecules have small sequences, 21e23 nucleotides, are highly selective and sequence-specic, and have better stability compared to that of AS-ODNs. Finally, aptamers are synthetic singlestranded DNA or RNA molecules with a unique 3-D structure. They are able to specically bind other molecules and are particularly prone to bind the functional domains. This feature makes them useful as modulators of the targeted molecule (Proske et al., 2005; Nimjee et al., 2005). Therapeutic applications of all of these molecules in the eye have been extensively reviewed (Borrs, 2003; Henry et al., 2004; Campochiaro, 2006; Fattal and Bochot, 2006, 2008; Levy-Nissenbaum et al., 2008). Delivery of genetic material is quite a challenge from a pharmaceutical point of view. It is unstable in biological uids and has poor cell penetration due to its size or charge. For instance, plasmid DNA is large; however, siRNA is quite small. These facts imply that suitable carriers to deliver it to ocular tissues are needed. A partially successful approach in recent years has been the use of viral vectors such as adenovirus, retrovirus, lentivirus, and mainly recombinant adeno-associated virus (rAAV), as gene carriers (Snyder, 1999). rAVV vectors carry single-stranded DNA which is inserted into the genome of the targeted cell. In general, gene delivery using rAAV shows a lack of pathogenicity, good long-term transgene expression, and no toxicity. However this technology has limitations such as lack of effective transduction in some cell types or presence of neutralizing antibodies for some rAVV serotypes (Rabinowitz and Samulski, 1998; Grimm and Kay, 2003). Also, progress has been made to deliver naked DNA to cells (Herweijer and Wolff, 2003). Different studies have established that some viral vectors can efciently deliver transgenes to ocular tissues while others cannot. For instance, several rAAV serotype vectors appear unsuitable for anterior segment delivery; however, rAAVs appear to have a selective tropism for retinal ganglion cells (Borrs et al., 2002). There are examples of effective rescue of genetic decits in the eye using viral and rAAV-mediated gene therapy in different in vitro and in vivo models (Campochiaro, 2002; Martin et al., 2002; Borrs, 2003; Ziche et al., 2004; Ralph et al., 2006; Roy et al., 2010). One remarkable use of this technology is the recovery of visual function in RPE65-decient dogs. Genetic deciency of RPE65, a protein involved in retinoid metabolism in RPE cells, results in blindness similar to human Lebers Congenital Amaurosis. Recovery of visual function in the dogs was achieved after subretinal injection of rAAV encoding RPE65 (Acland et al., 2001). These important results have led to a gene therapy clinical trial for the RPE65-decient form of the human disease (Bainbridge et al., 2008; Cideciyan et al., 2008). However, the use of viral vectors poses risks for the safety of patients. Additionally, the effectiveness in the eye may be limited due to several factors, including cell tropism, the size of the

Y. Diebold, M. Calonge / Progress in Retinal and Eye Research 29 (2010) 596e609

605

sequence to be carried and expressed, and most importantly, host immunity (Bennett, 2003). Despite the promising results of virallymediated ocular gene therapy, non-viral carriers would be the preferred choice. Still, the issue of an optimal delivery system remains to be solved. Features that gene delivery systems should provide for optimal activity of the delivered nucleic acid molecule include improved stability, increased half-life, specic tissue and cellular targeting, improved cellular penetration, and release of the molecule in the right intracellular compartment (i.e., DNA in the nucleus, however, siRNAs in the cytoplasm). Hence, gene delivery using non-viral carrier systems holds promise of being safer and more effective, as those limiting factors cited above do not inuence them so strongly. Several investigations have started using different approaches, including naked DNA delivery by means of electroporation and the formation of lipoplexes by DNA condensation with cationic lipids among others (Kachi et al., 2005; Bloquel et al., 2006; Johnson et al., 2008). Regarding genetic material incorporation to NPs, there are examples using compacted DNA NPs. Regarding the anterior segment of the eye, our group has obtained promising results using HA-chitosan oligomers (CSO) NPs loaded with a model plasmid encoding for alkaline phosphatase (manuscript submitted). We are currently studying the intracellular trafcking of those nanoparticles and contents in human epithelial cell lines derived from the ocular surface. Our preliminary results show an effective delivery of the plasmid to the cell interior and alkaline phosphatase expression 48 h after transfection using HACSO NPs. Another recent example is the work published by Klausner et al. (2010) in which ultrapure CSO (NOVAFECT) was used to complex with a model plasmid encoding for green uorescent protein (pGFP) to form NPs. Transgene expression of the pGFP was detected in cell cultures. In rat corneas pGFP was expressed 5.4 times that of control polyethylenimine-pGFP NPs. It is worth mentioning that in vivo transgene expression was identied in corneal stroma but not in the epithelium or the endothelium. Another approach is DNA condensed with polycationic polymers. Farjo et al. (2006) packaged compacted DNA in PEGsubstituted lysine peptides that formed the NPs. The compacted DNA was a model plasmid encoding for enhanced green uorescent protein (pEGFP), and its expression was studied in mice. Two days after an intravitreal injection, uorescence was present in the lens, cornea, trabecular meshwork, sclera, choroid, RPE, and other retinal cells. However, two days after a subretinal injection, uorescence was restricted to retina and RPE cells, choroid, and sclera, with a minimal presence in the lens. The authors reported no alteration in visual function, evaluated by electroretinography, and no evidence of inammation in the histological analysis of the ocular tissues. In this study, the duration of DNA expression was not reported. Ding et al. (2009) recently reported the preparation of single molecule of pEGFP compacted with PEG-substituted polysysine (CK30PEG). The formed NPs were subretinally injected in mice and resulted in efcient retinal cells transfection. This work was more focused on toxicity issues related with that technology. There were no signs of local inammatory response in terms of inltration of inammatory cells or chemokine marker expression. Even though this method of gene therapy is potentially applicable for multiple ocular diseases, a more directed targeting is desirable. A quite recent concept is the use of light-sensitive NPs made by combining a structural protein of the Herpes simplex virus, VP22, with AS-ODNs bound through the C-terminal end of the viral protein. This leads to the formation of spherical NPs of 0.3e1 mm in diameter named vectosomes. AS-ODNs selectively modulate the expression of a given gene by displaying a base sequence that is complementary to a specic mRNA (Helene and Toulme, 1990). The

precise local delivery of bound ODNs in target cells is controlled by illuminating them (Normand et al., 2001; Zavaglia et al., 2003). This technology, used for tumour cells, is particularly interesting in ophthalmology due to the frequent use of lasers for therapeutic purposes. Light-induced delivery of AS-ODNs from vectosomes has been studied in vitro, using human melanoma and RPE cell lines, and in vivo with rats (Normand et al., 2005). Interestingly, the vectosomes followed a rapid transretinal migration pattern after intravitreal injection in rats, and they were internalized by RPE cells and other cell types. Transscleral illumination of injected eyes induced disruption of the light-sensitive vectosomes and migration of released AS-ODNs to the cell nucleus, where they were localized in a light-dependent manner. Not only were the AS-ODNs localized in the RPE cells, they were also detected in ganglion cells, inner nuclear layer, and even in the choroid. Much work is needed to learn about the light wavelengths and energy produced as a consequence of vectosome rupture, which can potentially harm the retina. However, this approach offers great therapeutic potential. A nal example of potential gene therapy treatment of retinal diseases is the recent work of Park et al. (2009). They encapsulated in PLGA NPs an expression plasmid for a natural angiogenic inhibitor, plasminogen kringle 5 (K5) that inhibits ischemiainduced neovascularisation in a rat model of oxygen-induced retinopathy (OIR). After administering K5-loaded NPs intravitreally to animals with OIR, the authors reported high K5 expression in the inner retina for four weeks. Also, retinal vascular leakage and retina neovascularisation were reduced in those K5-NP injected eyes when compared to fellow control eyes. Even though these possibilities make those of us who work in the development of new therapies in ophthalmology dream of the end of many handicapping disorders, many challenges remain. For instance, we must consider the potential capacity of the vitreous to act as a barrier for gene delivery, mainly due to its composition and biological characteristics that affect diffusion of large molecules (Xu et al., 2000; Peeters et al., 2005). We need a better understanding of the biological processes affecting intraocular structures to help design more specic solutions. The identication of more target genes involved in the development of each pathological condition is necessary. We must learn about potential immune responses derived from the use of different non-viral vectors. For pre-clinical studies, it is important to have available and select carefully appropriate animal models of the target disease. Finally, we will need to identify those elements that control the long-term expression of the delivered transgenes or the permanent shutting off of genes with delivered AS-ODNs. 8. Nanoparticle safety: toxicity and interaction with the immune system Not just ocular, but any biomedical application of NPs as a therapeutic agent requires biocompatibility. This means that NPs, both the components and the assembled NP itself, need to be biologically compatible with living tissues by not producing toxic, injurious, or immunological responses in them. Key aspects inuencing the biocompatibility of NPs are the physicochemical characteristics, such as size, shape, charge, solubility, and chemical groups on the surface that provide particle charge and lipo- or hydro-phobic features (McNeil, 2005). However, the same properties that make NPs attractive for biomedical applications may make them reactive in biological systems and develop toxicity. For instance, smaller size NPs are preferred for better interactions at the cellular level. However, smaller NPs have larger surface area per unit mass, which may mean higher reactivity and consequently, cell or tissue toxicity (Kipen and Laskin, 2005).

606

Y. Diebold, M. Calonge / Progress in Retinal and Eye Research 29 (2010) 596e609

Risks posed by both organic and inorganic NPs include aggregation, tissue accumulation, and adsorption of plasma proteins onto the surface. This latter consideration is particularly important when thinking about an intravenous NP administration or a potential access of the blood system using other non-systemic administration routes, e.g., intraocular administration to treat posterior segment diseases that involve blooderetina barrier impairment. NP aggregation may block cell metabolism or even impair tissue function. For instance, aggregation of topically applied NPs onto the ocular surface may block the lachrymal drainage punctum and impair tear lm recycling. Additionally, indiscriminate NP accumulation in ocular tissues may distort tissue architecture and consequently, alter function. Finally, there may be toxic effects due to the presence of high levels of the loaded drug in a non-target tissue. Potential cytotoxic activity of NPs may include alterations of cell membranes such as membrane disruption, as has been described for carbon nanotubes (Panessa-Warren et al., 2009). However, sometimes this particular property may be sought, especially for gene delivery (Kiang et al., 2004; Akagi et al., 2010). For instance, Kiang et al. (2004) used poly (propyl acrylic acid) to formulate chitosan-DNA NPs with enhanced in vitro transfection efciency. That polymer was specically designed to disrupt the lipid bilayer in cell membranes. By changes in the pH, it triggered the release of DNA from the endosomal compartment. Very recently, Akagi et al. (2010) evaluated the relationship of different physicochemical characteristics of 200 nm-size NPs composed of poly(gamma-glutamic acid). The protein-loaded NPs had signicant haemolytic activity in erythrocytes, depending on NP hydrophobicity and pH, with the greatest activity present at pH 7 to 5.5 and absent at physiological pH. Noxious effects of different kinds of NPs have been reported in several organ systems (for review please see Medina et al., 2007). One of the main mechanisms described by which NPs may harm cells and tissues is oxidative stress generation, which in turn, may lead to the activation of different transcription factors (Medina et al., 2007). Generally, NPs can be taken up by lymphatic nodes and distributed through the lymphatic system in parallel with the blood vascular system. The ocular mucosa possesses lymphoid

tissue that drains to different face and neck lymphatic ganglia. TW Prow (2009) recently published a very comprehensive review about the toxicity of nanomaterials in the eye. He nicely summarized in vitro and in vivo relevant studies accomplished since 1996 for about thirty different kinds of nanoparticles and other nanocarriers tested for ocular applications. The types of toxicity reported included cell morphology and viability, clinical signs evaluation, gross tissue examination, irritation test, histology and functional analyses, and inammatory response (Table 4). That review highlights the importance of accomplishing toxicity testing of newly developed drug carrier candidates. Another consideration is the potential inammatory, immunostimulatory, and immunosuppressive properties described for different kinds of NPs (Dobrovolskaia and McNeil, 2007; Zolnik et al., 2010). There are limited data available about this topic (Zolnik et al., 2010), but it is known that some NPs are antigenic themselves. The antigenicity depends on particle size, especially those of ultra-small size (25 nm or smaller) that improves lymphatic uptake, and surface charge (Reddy et al., 2007; Manolova et al., 2008). Allergic or hypersensitivity reactions can be induced or aggravated in animal models and humans by dendrimers (Toyama et al., 2008), carbon nanotubes (Nygaard et al., 2009), lipid-based NPs (Szebeni et al., 2007), titanium dioxide NPs (Yanagisawa et al., 2009), and polystyrene NPs (Yanagisawa et al., 2010). However, it is important to bear in mind that sometimes NPs are specically designed to target the immune system, and interactions with immune cells are considered benecial. Such is the case for those NPs intended for vaccine development, in which the immunogenic properties are exploited. NPs can serve as adjuvants as they can be conjugated with antigens. Currently, a major issue related to the development of NP-based novel therapies is the rigorous evaluation of the potential immunotoxic effects. Researchers in the eld of nanomedicine agree that the potential environmental and health-related risks should be carefully analyzed. Importantly, we, in the eye community, test NPs for toxic effects less than we should. The reasons for that are many, but the most important one is the absence of a common regulatory framework. Technological advancements develop faster than

Table 4 Summary of key questions to be addressed during early phase pre-clinical evaluation of nanoparticles (NPs) intended for use in biomedical applications. Assay category In vitro Hemolysis Platelet aggregation Coagulation time Complement activation Colony-forming unit granulocyte macrophage Leukocyte proliferation Uptake by macrophages Cytokine induction Nitric oxide production Cytotoxicity of natural killer cells Endotoxin contamination Microbial/viral/mycoplasma contamination In vivo Single-dose toxicity study: Standard toxicity tests (blood chemistry, hematology, histopathology, gross pathology) T cell dependent antibody response (TDAR) Host resistance studies; evaluation of cell-mediated immunity Questions to address Do NPs change integrity of red blood cells? Do NPs interfere with cellular components of the blood coagulation cascade? Do NPs cause changes in the function of the coagulation factors? Do NPs activate the complement system? Do NPs cause myelosuppression (toxicity to bone marrow precursors)? Do NPs have adverse effects on leukocyte proliferative responses? Are NPs internalized by specialized phagocytes? Do NPs activate immune cells to elicit cytokine production or interfere with that caused by known immunogens? Do NPs induce oxidative stress? Indirect test for potential endotoxin contamination Do NPs interfere with the ability of natural killer cells to recognize and kill tumour target cells? Pyrogen contamination test Sterility test These tests aim at answering the following questions: Do NPs cause toxicity to immune cells and organs? Are there any indications for additional toxicity studies? Additional studies are conducted on a case-by-case basis using weight-of-evidence approach This test is recognized for its high predictability for human models These tests are recommended for 1) testing the potential effects that particles might have on host resistance towards pathogens and tumour cells, and 2) to check for contact sensitization and delayed type hypersensitivity reactions Do NPs elicit particle specic immune response?

Repeated dose toxicity study; immunogenicity According to Dobrovolskaia & McNeil (Nature Nanotech., 2007).

Y. Diebold, M. Calonge / Progress in Retinal and Eye Research 29 (2010) 596e609

607

regulations in the nanomedicine eld. Standardization of procedures to study immunological properties and toxicology-related issues would also help. There certainly are regulations in Europe, U.S., and Japan intended to assess the immunotoxic potential of newly developed pharmaceuticals (Putman et al., 2003; Snodin, 2004). However, there are no specic protocols for those nanotechnology-based tests because the properties of the NPs may interfere with the established tests (Stone et al., 2009). A rst step towards that purpose has been taken by the National Characterization Laboratory, U.S. National Cancer Institute (http:// ncl.cancer.gov/working_assay-cascade.asp), whose mission is to perform and standardize the pre-clinical characterization of nanomaterials intended for cancer therapeutics and diagnostics. As an example, in a recent review Dobrovolskaia and McNeil (2007) suggested the most important parameters that need to be addressed during an initial evaluation of new nanotechnologyderived pharmaceuticals (Table 4). 9. Future directions A novel direction for nanomedical applications involves nanoceria particles. This unique type of NP is made of nanocrystalline cerium oxide, CeO2, also known as ceria. It is a rare earth oxide from the lanthanide series of the periodic table. Properly speaking, nanoceria particles do not deliver any drug; rather they are the drugs. These NPs can scavenge free radicals and reactive oxygen species (Heckert et al., 2008). Interest in them for biomedical applications derives from several in vitro studies in which they increased the lifespan of cultured brain cells (Rzigalinski et al., 2003), protected non-tumoral cells from radiation therapy effects (Tarnuzzer et al., 2005), and protected in vitro rat spinal cord neurons from oxidative stress (Das et al., 2007). For ocular applications, there are only a few studies with promising results. One of them, by Chen et al. (2006), showed that nanoceria particles were effective in the inhibition of the reactive oxygen intermediate-induced photoreceptor cell death. As macular degeneration and retinitis pigmentosa, among other blinding diseases, are thought to generate reactive oxygen species, the use of nanoceria may be a useful therapeutic strategy. In another study, Pierscionek et al. (2010) showed the potential of antioxidant nanoceria particles in cataract treatments. It is certain that more studies will arise in the near future to explore the potential benet of nanoceria particles in different eye diseases. Other inorganic NPs made of noble metals, such as gold (Hayashi et al., 2009) or silver (Gurunathan et al., 2009), are generating much interest due to their small size, about 20 nm, and the great potential of traversing the blooderetina barrier (Kim et al., 2009). There are very few reports, and most of them show preliminary results. More and deeper studies using these kinds of small NPs are necessary to understand the pharmacokinetics and clearance mechanisms (Amrite et al., 2008) and the actual therapeutic potential. Also, the use of inorganic NPs as novel contrast agents for molecular imaging in other tissues suggests that this technology can be applied to ocular imaging. A few interesting examples of NPs used as imaging agents are gadolinium-loaded nanoemulsions for brain imaging, gold NPs for optical coherence tomography imaging, and superparamagnetic iron oxide NPs with gold shells for magnetic resonance imaging. Regarding ocular tissues, Yamamoto et al. (2007) have shown that quantum dots can be used for imaging the vitreous. No doubt, the coming years will bring exciting developments in this eld. 10. Summary and conclusions It is currently possible to design nanocarriers with specic delivery requirements for ocular administration. Those carriers are

able to safely deliver the loaded therapeutic molecule while preventing damage or deactivation of it. The loaded agents can act more efciently and with fewer side effects when compared to the same agents administered without the nanocarrier. However, many questions still remain. It is an urgent matter to resolve them so that new and more efcient drug formulations based on NP technology for ocular therapy can be made available for patient care. Author disclosure statement The authors report no nancial interest in any of the materials or nanosystems presented in this review. Acknowledgements This work was supported by a Spanish Ministry of Science and Technology Grant (MAT2007-64626-C02) and the Networking Research Centre on Bioengineering, Biomaterials and Nanomedicine (CIBER-BBN), Spain. CIBER-BBN is an initiative funded by the VI National R&D&I Plan 2008e2011, Iniciativa Ingenio 2010, Consolider Program, CIBER Actions and nanced by the Instituto de Salud Carlos III with assistance from the European Regional Development Fund. References
Acland, G.M., Aguirre, G.D., Ray, J., Zhang, Q., Aleman, T.S., Cideciyan, A.V., PearceKelling, S.E., Anand, V., Zeng, Y., Maguire, A.M., Jacobson, S.G., Hauswirth, W.W., Bennet, J., 2001. Gene therapy restores vision in a canine model of childhood blindness. Nat. Genet. 28, 92e95. Aggarwal, P., Hall, J.B., McLeland, C.B., Dobrovolskaia, M.A., McNeil, S.E., 2009. Nanoparticle interaction with plasma proteins as it relates to particle biodistribution, biocompatibility and therapeutic efcacy. Adv. Drug Deliv. Rev. 61, 428e437. Akagi, T., Kim, H., Akashi, M., 2010. pH-dependent disruption of erythrocyte membrane by amphiphilic poly(amino acid) nanoparticles. J. Biomater. Sci. Polym. Ed. 21, 315e328. Alonso, M.J., Snchez, A., 2003. The potential of chitosan in ocular drug delivery. J. Pharm. Pharmacol. 55, 1451e1463. Amrite, A.C., Edelhauser, H.F., Singh, S.R., Kompella, U.B., 2008. Effect of circulation on the disposition and ocular tissue distribution of 20 nm nanoparticles after periocular administration. Mol. Vis. 14, 150e160. Attama, A.A., Reichl, S., Mller-Goymann, C.C., 2009. Sustained release and permeation of timolol from surface-modied solid lipid nanoparticles through bioengineered human cornea. Curr. Eye Res. 34, 698e705. Bareford, L.M., Swaan, P.W., 2007. Endocytic mechanisms for targeted drug delivery. Adv. Drug Deliv. Rev. 59, 748e758. Bayens, R., Gurny, R., 1997. Chemical and physical parameters of tears relevant for the design of ocular drug delivery formulations. Pharm. Acta Helv. 72, 191e202. Bejjani, R.A., BenEzra, D., Cohen, H., Rieger, J., Andrieu, C., Jeaanny, J.-C., Golomb, G., Behar-Cohen, F.F., 2005. Nanoparticles for gene delivery to retinal pigment epithelial cells. Mol. Vis. 11, 124e132. Bennett, J., 2003. Immune following intraocular delivery of recombinant viral vectors. Gene Ther. 10, 977e982. Bloquel, C., Bourges, J.L., Touchard, E., Berdugo, M., BenEzra, D., Behar-Cohen, F.F., 2006. Non-viral ocular gene therapy: potential ocular therapeutic avenues. Adv. Drug Deliv. Revs. 58, 1224e1242. Borrs, T., 2003. Recent developments in ocular gene therapy. Exp. Eye Res. 76, 643e652. Borrs, T., Brandt, C.R., Nickells, R., Ritch, R., 2002. Gene therapy for glaucoma: treating a multifaceted, chronic disease. Invest. Ophthalmol. Vis. Sci. 43, 2513e2518. Bourges, J.L., Gautier, S.E., Delie, F., Bejjani, R.A., Jeanny, J.-C.-, Gurny, R., BenEzra, D., Behar-Cohen, F.F., 2003. Ocular drug delivery targeting the retina and retinal pigment epithelium using polylactide nanoparticles. Invest. Ophthalmol. Vis. Sci. 44, 3562e3569. Bainbridge, J.W., Smith, A.J., Barker, S.S., Robbie, S., Henderson, R., Balaggan, K., Viswanathan, A., Holder, G.E., Stockman, A., Tyler, N., Petersen-Jones, S., Bhattacharya, S.S., Thrasher, A.J., Fitzke, F.W., Carter, B.J., Rubin, G.S., Moore, A.T., Ali, R.R., 2008. Effect of gene therapy on visual function in Lebers congenital amaurosis. N. Engl. J. Med. 358, 2231e2239. Bumcrot, D., Manoharan, M., Koteliansky, V., Sah, D.W., 2006. RNAi therapeutics: a potential new class of pharmaceutical drugs. Nat. Chem. Biol. 2, 711e719. Calvo, P., Snchez, A., Martnez, J., Lpez, M.I., Calonge, M., Pastor, J.C., Alonso, M.J., 1996. Polyester nanocapsules as new topical ocular delivery systems for cyclosporin A. Pharm. Res. 13, 311e315.

608

Y. Diebold, M. Calonge / Progress in Retinal and Eye Research 29 (2010) 596e609 Gurunathan, S., Lee, K.J., Kalishwaralal, K., Sheikpranbabu, S., Vaidyanathan, R., Eom, S.H., 2009. Antiangiogenic properties of silver nanoparticles. Biomaterials 30, 6341e6350. Heckert, E.G., Karakoti, A.S., Seal, S., Self, W.T., 2008. The role of cerium redox state in the SOD mimetic activity of nanoceria. Biomaterials 29, 2705e2709. Hmlinen, K.M., Kananen, K., Auriola, S., Kontturi, K., Urtti, A., 1997. Characterization of paracellular and aqueous penetration routes in cornea, conjunctiva, and sclera. Invest. Ophthalmol. Vis. Sci. 38, 627e634. Harush-Frenkel, O., Altschuler, Y., Benita, S., 2008. Nanoparticle-cell interactions: drug delivery implications. Crit. Rev. Ther. Drug Carrier Syst. 25, 485e544. Hayashi, A., Naseri, A., Pennesi, M.E., de Juan Jr., E., 2009. Subretinal delivery of immunoglobulin G with gold nanoparticles in the rabbit eye. Jpn. J. Ophthalmol. 53, 249e256. Helene, C., Toulme, J.J., 1990. Specic regulation of gene expression of by antisense, sense and antigenic nucleic acids. Biochim. Biophys. Acta 1049, 99e125. Henry, S.P., Marcusson, E.G., Vincent, T.M., Dean, N.M., 2004. Setting sights on the treatment of ocular angiogenesis using antisense oligonucleotides. TRENDS Pharm. Sci. 25, 523e527. Herweijer, H., Wolff, J.A., 2003. Progress and prospects: naked DNA gene transfer and therapy. Gene Ther. 10, 453e458. Hillaireau, H., Couvreur, C., 2009. Nanocarriers entry into the cell: relevance to drug delivery. Cell. Mol. Life Sci. 66, 2873e2896. Hsiue, G.H., Hsu, S.H., Yang, C.C., Lee, S.H., Yang, I.K., 2002. Preparation of controlled release ophthalmic drops, for glaucoma therapy using thermosensitive polyN-isopropylacrylamide. Biomaterials 23, 457e462. Irache, J.M., Merodio, M., Arnedo, A., Campanero, M.A., Mirshahi, M., Espuelas, S., 2005. Albumin nanoparticles for the intravitreal delivery of anticytomegaloviral drugs. Mini Rev. Med. Chem. 5, 293e305. Jarvinen, K., Jarvinen, T., Urtti, A., 1995. Ocular absorption following topical delivery. Adv. Drug Deliv. Rev. 16, 3e19. Johnson, L.N., Cashman, S.M., Kumar-Singh, R., 2008. Cell-penetrating peptide for enhanced delivery of nucleic acids and drugs to ocular tissues including retina and cornea. Mol. Ther. 16, 107e114. Juberas, J.R., Calonge, M., Gmez, S., Lpez, M.I., Calvo, P., Herreras, J.M., Alonso, M.J., 1998. Efcacy of topical cyclosporine-loaded nanocapsules on keratoplasty rejection model in the rat. Curr. Eye Res. 17, 39e46. Kachi, S., Oshima, Y., Esumi, N., Kachi, M., Rogers, B., Zack, D.J., Campochiaro, P.A., 2005. Nonviral ocular gene transfer. Gene Ther. 12, 843e851. Kao, H.-J., Lin, H.-R., Lo, Y.-L., Yu, S.-P., 2006. Characterization of pilocarpine-loaded chitosan/carbopol nanoparticles. J. Pharmacokinet. Pharmacodyn. 58, 179e186. Kassem, M.A., AbdelRahman, A.A., Ghorab, M.M., Ahmed, M.B., Khalil, R.M., 2007. Nanosuspensions as an ophthalmic delivery system for certain glucocorticoid drugs. Int. J. Pharm. 340, 126e133. Kaur, I.P., Kanwar, M., 2002. Ocular preparations: the formulation approach. Drug Dev. Ind. Pharm. 28, 473e493. Keister, J.C., Cooper, E.R., Missel, P.J., Lang, J.C., Hager, D.F., 1991. Limits on optimizing ocular drug delivery. J. Pharm. Sci. 80, 50e53. Kiang, T., Brigth, C., Cheung, C.Y., Stayton, P.S., Hoffman, A.S., Leong, K.W., 2004. Formulation of chitosan-DNA nanoparticles with poly(propyl acrylic acid) enhances gene expression. J. Biomater. Sci. Polym. Ed. 15, 1405e1421. Kim, J.H., Kim, J.H., Kim, K.W., Kim, M.H., Yu, Y.S., 2009. Intravenously administered gold nanoparticles pass through the blooderetinal barrier depending on the particle size, and induce no retinal toxicity. Nanotechnology 20, 505101. Kipen, H.M., Laskin, D.L., 2005. Smaller is not always better: nanotechnology yields nanotoxicology. Am. J. Physiol. Lung Cell Mol. Physiol. 289, L696eL697. Klausner, E.A., Zhang, Z., Chapman, R.L., Multack, R.F., Volin, M.V., 2010. Ultrapure chitosan oligomers as carriers for corneal gene transfer. Biomaterials 31, 1814e1820. Leung, R.K.M., Whittaker, P.A., 2005. RNA interference: from gene silencing to genespecic therapeutics. Pharmacol. Ther. 107, 222e239. Levy-Nissenbaum, E., Radovic-Moreno, A.F., Wang, A.Z., Langer, R., Farokhzad, O.C., 2008. Nanotechnology and aptamers: applications in drug delivery. Trends Biotechnol. 26, 442e449. Li, V.H., Wood, R.W., Kreuter, J., Harmia, T., Robinson, J.R., 1986. Ocular drug delivery of progesterone using nanoparticles. J. Microencapsul. 3, 213e218. Loke, S.L., Stein, C.A., Zhang, X.H., Mori, K., Nakanishi, M., Subasinghe, C., Cohen, J.S., Neckers, L.M., 1989. Characterization of oligonucleotide transport into living cells. Proc. Natl. Acad. Sci. U.S.A. 86, 3474e3478. Losa, C., Marchal-Heussler, L., Orallo, F., Vila-Jato, J.L., Alonso, M.J., 1993. Design of new formulations for topical ocular administration: polymeric nanocapsules containing metipranolol. Pharm. Res. 10, 80e87. Losa, C., Calvo, P., Castro, E., Vila-Jato, J.L., Alonso, M.J., 1991. Improvement of ocular penetration of amikacin sulphate by association to poly(butylcyanoacrylate) nanoparticles. J. Pharm. Pharmacol. 43, 548e552. Manolova, V., Flace, A., Bauer, M., Schwarz, K., Saudan, P., Bachmann, M.F., 2008. Nanoparticles target distinct dendritic cell populations according to their size. Eur. J. Immunol. 38, 1404e1413. Mannermaa, E., Vellonen, K.S., Urtti, A., 2006. Drug transport in corneal epithelium and blooderetina barrier: emerging role of transporters in ocular pharmacokinetics. Adv. Drug Deliv. Rev. 58, 1136e1163. Martin, K.R., Klein, R.L., Quigley, H.A., 2002. Gene delivery to the eye using adenoassociated viral vectors. Methods 28, 267e275. McNeil, S.E., 2005. Nanotechnology for the biologist. J. Leukoc. Biol. 78, 585e594.

Campochiaro, P.A., 2002. Gene therapy for retinal and choroidal diseases. Expert Opin. Biol. Ther. 2, 537e544. Campochiaro, P.A., 2006. Potential applications for RNAi to probe pathogenesis and develop new treatments for ocular disorders. Gene Ther. 13, 559e562. Cideciyan, A.V., Aleman, T.S., Boye, S.L., Schwartz, S.B., Kaushal, S., Roman, A.J., Pang, J.J., Sumaroka, A., Windsor, E.A., Wilson, J.M., Flotte, T.R., Fishman, G.A., Heon, E., Stone, E.M., Byrne, B.J., Jacobson, S.G., Hauswirth, W.W., 2008. Human gene therapy for RPE65 isomerase deciency activates the retinoid cycle of vision but with slow rod kinetics. Proc. Natl. Acad. Sci. U.S.A. 105, 15112e15117. Chen, J., Patil, S., Seal, S., McGinnis, J.F., 2006. Rare earth nanoparticles prevent retinal degeneration induced by intracellular peroxides. Nat. Nanotechnol. 1, 142e150. Conti, S., Polonelli, L., Frazzi, R., Artusi, M., Bettini, R., Cocconi, D., Colombo, P., 2000. Controlled delivery of biotechnological products. Curr. Pharm. Biotechnol. 1, 313e323. Contreras-Ruiz, L., de la Fuente, M., Garca-Vzquez, C., Sez, V., Seijo, B., Alonso, M.J., Calonge, M., Diebold, Y., 2010. Ocular tolerance to a topical formulation of hyaluronic acid and chitosan-based nanoparticles. Cornea 29, 550e558. Couvreur, P., Tulkenst, P., Roland, M., Trouet, A., Speiser, P., 1977. Nanocapsules: a new type of lysomotropic carrier. FEBS Lett. 84, 323e326. Das, M., Patil, S., Bhargava, N., Kang, J.-F., Riedel, L.M., Seal, S., et al., 2007. Autocatalytic ceria nanoparticles offer neuroprotection to adult rat spinal cord neurons. Biomaterials 28, 1918e1925. de Campos, A.M., Diebold, Y., Carvalho, E.L.S., Snchez, A., Alonso, M.J., 2004. Chitosan nanoparticles as new ocular drug delivery systems: in vitro stability, in vivo fate, and cellular toxicity. Pharm. Res. 21, 803e810. de Campos, A.M., Snchez, A., Alonso, M.J., 2001. Chitosan nanoparticles: a new vehicle for the improvement of the delivery of drugs to the ocular surface. Application to cyclosporine A. Int. J. Pharm. 224, 159e168. de Campos, A.M., Snchez, A., Gref, R., Calvo, P., Alonso, M.J., 2003. The effect of a PEG versus a chitosan coating on the interaction of drug colloidal carriers with the ocular mucosa. Eur. J. Pharm. Sci. 20, 73e81. de la Fuente, M., Seijo, B., Alonso, M.J., 2008a. Novel hyaluronan-based nanocarriers for transmucosal delivery of macromolecules. Macromol. Biosci. 8, 441e450. de la Fuente, M., Seijo, B., Alonso, M.J., 2008b. Novel hyaluronic acid-chitosan nanoparticles for ocular gene delivery. Invest. Ophthalmol. Vis. Sci. 49, 2016e2024. del Amo, E.M., Urtti, A., 2008. Current and future ophthalmic drug delivery systems. A shift to the posterior segment. Drug Discov. Today 13, 135e143. del Pozo-Rodrguez, A., Delgado, D., Solins, M.A., Gascn, A.R., Pedraz, J.L., 2008. Solid lipid nanoparticles for retinal gene therapy: transfection and intracellular trafcking in RPE cells. Int. J. Pharm. 360, 177e183. Diebold, Y., Jarrn, M., Sez, V., Carvalho, E.L.S., Orea, M., Calonge, M., Seijo, B., Alonso, M.J., 2007. Ocular drug delivery by liposomeechitosan nanoparticle complexes (LCS-NP). Biomaterials 28, 1553e1564. Ding, X.Q., Quiambao, A.B., Fitzgerald, J.B., Cooper, M.J., Conley, S.M., Naash, M.I., 2009. Ocular delivery of compacted DNA-nanoparticles does not elicit toxicity in the mouse retina. PLoS ONE 4, 1e11. Dobrovolskaia, M.A., Aggarwal, P., Hall, J.B., McNeil, S.E., 2008. Preclinical studies to understand nanoparticles interaction with the immune system and its potential effects on nanoparticles biodistribution. Mol. Pharmacol. 5, 487e495. Dobrovolskaia, M.A., McNeil, S.E., 2007. Immunological properties of engineered nanomaterials. Nat. Nanotechnol. 2, 469e478. Degim, I.T., Celebi, N., 2007. Controlled delivery of peptides and proteins. Curr. Pharm. Des. 13, 99e117. Edwards, A., Prausnitz, M.R., 2001. Predicted permeability of the corneal to topical drugs. Pharm. Res. 18, 1497e1508. Enrquez-de-Salamanca, A., Diebold, Y., Calonge, M., Garca-Vzquez, C., Callejo, S., Vila, A., Alonso, M.J., 2006. Chitosan nanoparticles as a potential drug delivery system for the ocular surface: toxicity, uptake mechanism and in vivo tolerance. Invest. Ophthalmol. Vis. Sci. 47, 1416e1425. European Science Foundation Forward Look on Nanomedicine. http://www.esf.org/ publications/forward-looks.html. Farjo, R., Skaggs, J., Quiambao, A.B., Cooper, M.J., Naash, M.I., 2006. Efcient nonviral ocular gene transfer with compacted DNA nanoparticles. PLoS ONE 1 e38. Fattal, E., Bochot, A., 2006. Ocular delivery of nucleic acids: antisense oligonucleotides, aptamers and siRNA. Adv. Drug Dev. Rev. 58, 1203e1223. Fattal, E., Bochot, A., 2008. State of the art and perspectives for the delivery of antisense oligonucleotides and siRNA by polymeric nanocarriers. Intl. J. Pharm. 364, 237e248. Frenkel, V., 2008. Ultrasound mediated delivery of drugs and genes to solid tumors. Adv. Drug Deliv. Rev. 30, 1193e1208. Gaudana, R., Jwala, J., Boddu, S.H.S., Mitra, A.K., 2009. Recent perspectives in ocular drug delivery. Pharm. Res. 26, 1197e1216. Geroski, D.H., Edelhauser, H.F., 2001. Transscleral drug delivery for posterior segment diseases. Adv. Drug Deliv. Revs. 52, 37e48. Gref, R., Minamitake, Y., Peracchia, M.T., Trubetskoy, V., Torchilin, V., Langer, R., 1994. Biodegradable long-circulating polymeric nanospheres. Science 263, 1600e1603. Grimm, D., Kay, M.A., 2003. From virus evolution to vector revolution: use of naturally occurring serotypes of adeno-associated virus (AAV) as novel vectors for human gene therapy. Curr. Gene Ther. 3, 281e304.

Y. Diebold, M. Calonge / Progress in Retinal and Eye Research 29 (2010) 596e609 Medina, C., Santos-Martnez, M.J., Radomski, A., Corrigan, O.I., Radomski, M.W., 2007. Nanoparticles: pharmacological and toxicological signicance. Br. J. Pharmacol. 150, 552e558. Merodio, M., Arnedo, A., Renedo, M.J., Irache, J.M., 2001. Ganciclovir-loaded nanoparticles: characterization and in vitro release properties. Eur. J. Pharm. Sci. 12, 251e259. Merodio, M., Espuelas, M.S., Mirshahi, M., Arnedo, A., Irache, J.M., 2002. Efcacy of ganciclovir-loaded nanoparticles in human cytomegalovirus (HCMV)-infected cells. J. Drug Target. 10, 231e238. Mo, Y., Barnett, M.E., Takemoto, D., Davidson, H., Kompella, U.B., 2007. Human serum albumin nanoparticles for efcient delivery of Cu, Zn superoxide dismutase gene. Mol. Vis. 23, 746e757. Motwani, S.K., Chopra, S., Talegaonkar, S., Kohli, K., Ahmad, F.J., Khar, R.K., 2008. Chitosan-sodium alginate nanoparticles as submicroscopic reservoirs for ocular delivery: formulation, optimisation and in vitro characterisation. Eur. J. Pharm. Biopharm. 68, 513e525. National Characterization Laboratory, U.S. National Cancer Institute. http://ncl. cancer.gov/working_assay-cascade.asp, (accessed June 2010). National Institutes of Health of the U.S.A., Nanomedicine Roadmap Initiative. http://nihroadmap.nih.gov/nanomedicine/. Nimjee, S.M., Rusconi, C.P., Sullenger, B.A., 2005. Aptamers: an emerging class of therapeutics. Annu. Rev. Med. 56, 555e583. Normand, N., Valamanesh, F., Savoldelli, M., Mascarelli, F., BenEzra, D., Courtois, Y., Behar-Cohen, F.F., 2005. VP22 light controlled delivery of oligonucleotides to ocular cells in vitro and in vivo. Mol. Vis. 11, 184e191. Normand, N., van Leeuwen, H., OHare, P., 2001. Particle formation by a conserved domain of the Herpes simplex virus protein VP22 facilitating protein and nucleic acid delivery. J. Biol. Chem. 276, 15042e15050. Nussenblatt, R.B., Csaky, K., 1997. Perspectives on gene therapy in the treatment of ocular inammation. Eye (Lond) 11, 217e221. Nygaard, U.C., Hansen, J.S., Samuelsen, M., Alberg, T., Marioara, C.D., Lovik, M., 2009. Single-walled and multi-walled carbon nanotubes promote allergic immune responses in mice. Toxicol. Sci. 109, 113e123. Paasonen, L., Laaksonen, T., Johans, C., Yliperttula, M., Kontturi, K., Urtti, A., 2007. Gold nanoparticles enable selective light-induced contents release from liposomes. J. Control Release 11, 86e93. Pan, T., Brown, D.J., Ziaie, B., 2006. An articial nano-drainage implant (ANDI) for glaucoma treatment. In: Conf. Proc. IEEE Eng. Med. Biol. Soc., vol. 1, pp. 3174e3177. Panessa-Warren, B.J., Maye, M.M., Warren, J.B., Crosson, K.M., 2009. Single walled carbon nanotube reactivity and cytotoxicity following extended aqueous exposure. Environ. Pollut. 157, 1140e1151. Paolicelli, P., de la Fuente, M., Snchez, A., Seijo, B., Alonso, M.J., 2009. Chitosan nanoparticles for drug delivery to the eye. Expert Opin. Drug Deliv. 6, 239e253. Park, K., Chen, Y., Hu, Y., Mayo, A.S., Kompella, U.B., Longeras, R., Ma, J.-X., 2009. Nanoparticle-mediated expression of an angiogenic inhibitor ameliorates ischemia-induced retinal neovascularisation and diabetes-induced retinal vascular leakage. Diabetes 58, 1902e1913. Peeters, L., Sanders, N.N., Braeckmans, K., Boussery, K., van de Voorde, J., de Smedt, S.C., Demeester, J., 2005. Vitreous: a barriers to nonviral ocular gene therapy. Invest. Ophthalmol. Vis. Sci. 46, 3553e3561. Pierscionek, B.K., Li, Y., Yasseen, A.A., Colhoun, L.M., Schachar, R.A., Chen, W., 2010. Nanoceria have no genotoxic effect on human lens epithelial cells. Nanotechnology 21, 035102. Pignatello, R., Bucolo, C., Ferrara, P., Maltese, A., Puleo, A., Puglisi, G., 2002. Eudragit RS100 nanosuspensions for the ophthalmic controlled delivery of ibuprofen. Eur. J. Pharm. Sci. 16, 53e61. Pinto-Reis, C., Neufeld, R.J., Ribeiro, A.J., Veiga, F., 2006. Nanoencapsulation I. Methods for preparation of drug-loaded polymeric nanoparticles. Nanomedicine 2, 8e21. Proske, D., Blank, M., Buhmann, R., Resch, A., 2005. Aptamers-basic research, drug development, and clinical applications. Appl. Microbiol. Biotechnol. 69, 367e374. Prow, T.W., 2009. Toxicity of nanomaterials to the eye. WIREs Nanomed. Nanobiotechnol. Dec 11. Putman, E., van der Laan, J.W., van Loveren, H., 2003. Assessing immunotoxicity: guidelines. Fundam. Clin. Pharmacol. 17, 615e626. Qaddoumi, M.G., Gukasyan, H.J., Davda, J., Labhasetwar, V., Kim, K.J., Lee, V.H.L., 2003. Clathrin and caveolin-1 expression in primary pigmented rabbit conjunctival epithelial cells: Role in PLGA nanoparticle Endocytosis. Mol. Vis. 9, 559e568. Quigley, H.A., Broman, A.T., 2006. The number of people with glaucoma worldwide in 2010 and 2020. Br. J. Ophthalmol. 90, 262e267. Rabinowitz, J.E., Samulski, J., 1998. Adeno-associated virus expression systems for gene transfer. Curr. Opin. Biotechnol. 9, 470e475. Ralph, G.S., Binley, K., Wong, L.F., Azzouz, M., Mazarakis, N.D., 2006. Gene therapy for neurodegenerative and ocular diseases using lentiviral vectors. Clin. Sci. (Lond) 110, 37e46. Reddy, S.T., van der Vlies, A.J., Simeoni, E., Angeli, V., Randolph, G.J., ONeil, C.P., Lee, L.K., Swartz, M.A., Hubbell, J.A., 2007. Exploiting lymphatic transport and complement activation in nanoparticle vaccines. Nat. Biotechnol. 25, 1159e1164. Roy, K., Stein, L., Kaushal, S., 2010. Ocular gene therapy: an evaluation of recombinant adeno-associated virus-mediated gene therapy interventions for the treatment of ocular disease. Hum. Gene Ther. 21, 915e927.

609

Rzigalinski, B.A., Bailey, D., Chow, L., Kuiry, S.C., Patil, S., Merchant, S., et al., 2003. Cerium oxide nanoparticles increase the lifespan of cultured brain cells and protect against free radical and mechanical trauma. FASEB J. 17, A606. Sahoo, S.K., Dilnawaz, F., Krishnakumar, S., 2008. Nanotechnology in ocular drug delivery. Drug Discov. Today (3/4), 144e151. Sahoo, S.K., Parveen, S., Panda, J.J., 2007. The present and future of nanotechnology in human health care. Nanomedicine 3, 20e31. Sakai, T., Kohno, H., Ishihara, T., Higaki, M., Saito, S., Matsushima, M., Mizushima, Y., Kitahara, K., 2006. Treatment of experimental autoimmune uveoretinitis with poly(lactic acid) nanoparticles encapsulating betamethasone phosphate. Exp. Eye Res. 82, 657e663. Snchez, A., Alonso, M.J., 2006. Nanoparticular carriers for ocular drug delivery. In: Torchilin, Vladimir (Ed.), Nanoparticulates as Drug Carriers. World ScienticImperial College Press, ISBN 1-86094-630-5, pp. 1e30 (Chapter 27). Sasaki, H., Yamamura, K., Mukai, T., Nishida, K., Nakamura, J., Nakashima, M., Ichikawam, M., 1999. Enhancement of ocular drug penetration. Crit. Rev. Ther. Drug Carrier Syst. 16, 85e146. Schoenwald, R.D., 1990. Ocular drug delivery: pharmacokinetics considerations. Clin. Pharmacokinet. 18, 255e269. Snodin, D.J., 2004. Regulatory immunotoxicology: does the published evidence support mandatory nonclinical immune function screening in drug development? Regul. Toxicol. Pharmacol. 40, 336e355. Snyder, R.O., 1999. Adeno-associated virus-mediated gene delivery. J. Gene Med. 1, 166e175. Soppimath, K.S., Aminabhavi, T.M., Kulkarni, A.R., Rudzinski, W.E., 2001. Biodegradable polymeric nanoparticles as drug delivery devices. J. Control Release 70, 1e20. Stone, V., Johnston, H., Schins, R.P., 2009. Development of in vitro systems for nanotoxicology: methodological considerations. Crit. Rev. Toxicol. 39, 613e626. Szebeni, J., Alving, C.R., Rosivall, L., Bnger, R., Barany, L., Bedcs, P., Tth, M., Barenholz, Y., 2007. Animal models of complement-mediated hypersensitivity reactions to liposomes and other liposome-based nanoparticles. J. Liposome Res. 17, 107e117. Tarnuzzer, R.W., Colon, J., Patil, S., Seal, S., 2005. Vacancy engineered ceria nanostructures for protection from radiation-induced cellular damage. Nano Lett. 12, 2573e2577. Tanelian, D.L., Barry, M.A., Johnston, S.A., Le, T., Smith, G., 1997. Controlled gene gun delivery and expression of DNA within the cornea. Biotechniques 23, 484e488. Toyama, T., Matsuda, H., Ishida, I., Tani, M., Kitaba, S., Sano, S., Katayama, I., 2008. A case of toxic epidermal necrolysis-like dermatitis evolving from contact dermatitis of the hands associated with exposure to dendrimers. Contact Derm. 59, 122e123. Urtti, A., 2006. Challenges and obstacles of ocular pharmacokinetics and drug delivery. Adv. Drug Deliv. Revs. 58, 1131e1135. Vanderwoot, J., Ludwig, A., 2007. Ocular drug delivery: nanomedicines application. Nanomedicine 2, 11e21. Vega, E., Egea, M.A., Valls, O., Espina, M., Garca, M.L., 2006. Flurbiprofen loaded biodegradable nanoparticles for ophthalmic administration. J. Pharm. Sci. 95, 2393e2405. Wadhawa, S., Paliwal, R., Paliwal, S.R., Vyas, S.P., 2009. Hyaluronic acid modied chitosan nanoparticles for effective management of glaucoma: development, characterization, and evaluation. J. Drug Target.. doi:10.3109/10611860903450023 Nov 27 (Epub ahead of print). Wood, R.W., Lee, H.K., Kreuter, J., Robinson, J.R., 1985. Ocular disposition of polyhexyl-2-cyano(3-14C)acrylate nanoparticles in the albino rabbit. Int. J. Pharm. 23, 175e183. Xu, J., Heyes, J.J., Barocas, V.H., Randolph, T.W., 2000. Permeability and diffusion in vitreous humour: implications for drug delivery. Pharm. Res. 17, 664e669. Yamamoto, S., Manabe, N., Fujioka, K., Hoshino, A., Yamamoto, K., 2007. Visualizing vitreous using quantum dots as imaging agents. IEEE Trans. Nanobioscience 6, 94e98. Yanagisawa, R., Takano, H., Inoue, K.I., Koike, E., Sadakane, K., Ichinose, T., 2009. Titanium dioxide nanoparticles aggravate atopic dermatitis-like skin lesions in NC/Nga mice. Exp. Biol. Med. 234, 314e322. Yanagisawa, R., Takano, H., Inoue, K.I., Koike, E., Sadakane, K., Ichinose, T., 2010. Size effects of polystyrene nanoparticles on atopic dermatitis-like skin lesions in NC/NGA mice. Int. J. Immunopathol. Pharmacol. 23, 131e141. Yenice, I., Mocan, M.C., Palaska, E., Bochot, A., Bilensoy, E., Vural, I., Irke, M., Hincal, A.A., 2008. Hyaluronic acid coated poly-epsilon-caprolactone nanospheres deliver high concentrations of cyclosporine A into the cornea. Exp. Eye Res. 87, 162e167. Zavaglia, D., Normand, N., Brewis, N., OHare, P., Favrot, M.C., Coll, J.L., 2003. VP22mediated and light-activated delivery of anti-c-raf1 antisense oligonucleotide improves its activity after intratumoral injection in nude mice. Mol. Ther. 8, 840e845. Ziche, M., Donnini, S., Morbidelli, L., 2004. Development of new drugs in angiogenesis. Curr. Drug Targets 5, 485e4493. Zimmer, A., Kreuter, J., Robinson, J.R., 1991. Studies on the transport pathway of PBCA nanoparticles in ocular tissues. J. Microencapsul. 8, 497e504. Zimmer, A., Mutschler, E., Lambrecht, G., Mayer, D., Kreuter, J., 1994. Pharmacokinetic and pharmacodynamic aspects of an ophthalmic pilocarpine nanoparticle-delivery-system. Pharm. Res. 11, 1435e1442. Zolnik, B.S., Gonzlez-Fernndez, A., Sadrieh, N., Dobrovolskaia, M.A., 2010. Minireview: nanoparticles and the immune system. Endocrinology 151, 458e465.

You might also like