221
PENTABLOCK COPOLYMER BASED CONTROLLED RELEASE FORMULATIONS OF SMALL AND MACROMOLECULES FOR OPHTHALMIC APPLICATIONS A DISSERTATION IN Pharmaceutical Sciences and Chemistry Presented to the Faculty of the University of Missouri - Kansas City in partial fulfillment of the requirements for the degree DOCTOR OF PHILOSOPHY by VIRAL M. TAMBOLI B. Pharm., Hemchandracharya North Gujarat University, India, 2006 Kansas City, Missouri 2012

PENTABLOCK COPOLYMER BASED CONTROLLED RELEASE …

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

PENTABLOCK COPOLYMER BASED CONTROLLED RELEASE FORMULATIONS

OF SMALL AND MACROMOLECULES FOR OPHTHALMIC APPLICATIONS

A DISSERTATION IN

Pharmaceutical Sciences

and

Chemistry

Presented to the Faculty of the University

of Missouri - Kansas City in partial fulfillment of

the requirements for the degree

DOCTOR OF PHILOSOPHY

by

VIRAL M. TAMBOLI

B. Pharm., Hemchandracharya North Gujarat University, India, 2006

Kansas City, Missouri

2012

iii

PENTABLOCK COPOLYMER BASED CONTROLLED RELEASE FORMULATIONS

OF SMALL AND MACROMOLECULES FOR OPHTHALMIC APPLICATIONS

Viral M. Tamboli, Candidate for the Doctor of Philosophy degree

University of Missouri Kansas City, 2012

ABSTRACT

Pentablock copolymers comprised of multi-polymer blocks such as polyethylene

glycol (PEG), polycaprolactone (PCL) and poly-lactide (PLA) were developed for

fabrication of nanoparticles and thermosensitive hydrogel formulations for long term delivery

of small and macromolecules. Different pentablock copolymer compositions were evaluated

to optimize drug release profile from nanoparticles and thermosensitive gel formulations. Our

composite approach i.e. pentablock copolymers based nanoparticles suspended in

thermosensitive gel provided sustained zero-order delivery of encapsulated therapeutic

agents without producing any significant burst effect.

Different compositions of pentablock copolymers (polylatide- polycaprolactone-

polyethylene glycol- polycaprolactone- polylatide) (PLA-PCL-PEG-PCL-PLA) and (PEG-

PCL-PLA-PCL-PEG) were synthesized and characterized to prepare nanoparticle and

thermosensitive hydrogel formulations, respectively. The effect of poly (L-lactide) (PLLA) or

poly (D, L-lactide) (PDLLA) incorporation on crystallinity of pentablock copolymers and in

vitro release profile of triamcinolone acetonide (selected as model drug) from nanoparticles was

also evaluated. Pentablock polymer with proper ratio PDLLA/PCL was amorphous in nature

whereas PLLA containing polymer has semicrystalline nature. Release of triamcinolone

acetonide from nanoparticles was significantly affected by crystallinity of the copolymers.

Burst release of triamcinolone acetonide from nanoparticles was significantly minimized

with incorporation of proper ratio of PDLLA in the existing triblock (PCL-PEG-PCL)

iv

copolymer. Moreover, pentablock copolymer based nanoparticles exhibited continuous

release of triamcinolone acetonide for longer duration. The release profile of various steroids

commonly utilized for chronic ocular diseases was also evaluated with optimized pentablock

polymer. We found that steroids with different log P values did not exhibited significant

difference in release profile. These results could be attributed with the fact that drug release

from the nanoparticles is mainly diffusion mediated process and small molecules can easily

diffuse out from the pore the polymer matrix. However, we found that pentablock copolymer

based nanoparticles can be utilized to achieve continuous near zero-order delivery of small

molecules from nanoparticles without any burst effect. Further, release profile of timolol

from composite formulations was evaluated for glaucoma therapy. We observed that

composite approach could provide sustain release of timolol for longer duration. Successful

accomplishment of this project may lead to application of this strategy for the treatment of

other chronic ocular diseases such as age related macular degeneration and diabetic macular

edema. Treatment of these diseases requires frequent intravitreal injections to maintain

therapeutic levels of antibodies at retina/choroid. Frequent administrations can cause

potential complications like endophthalmitis, retinal detachment and retinal hemorrhage.

Sustained intraocular therapeutic drug concentration can be achieved by suspending the

therapeutic macromolecules in thermosensitive hydrogel. Considering this we have

characterized the release kinetics of various macromolecules from pentablock copolymer

based thermosensitive hydrogel. We observed that release kinetics of macromolecules from

thermosensitive hydrogel was depended on the size of a molecule. Our studies indicate that

pentablock copolymer based delivery systems can provide sustained drug release profile for

longer duration, and thereby eliminate the need for repeated intravitreal injections.

v

APPROVAL PAGE

The faculty listed below, appointed by the Dean of the School of Graduate Studies have

examined a dissertation titled “Pentablock Copolymers Based Controlled Release

Formulations of Small and Macromolecules for Ophthalmic Applications” presented by Viral

M. Tamboli, candidate for the Doctor of Philosophy degree, and certify that in their opinion

it is worthy of acceptance.

Supervisory Committee

Ashim K. Mitra, Ph.D., Committee Chair

Department of Pharmaceutical Sciences

Chi Lee, Ph.D.

Department of Pharmaceutical Sciences

Kun Cheng, Ph.D.

Department of Pharmaceutical Sciences

Kenneth S. Schmitz, Ph.D.

Department of Chemistry

Andrew Holder, Ph.D.

Department of Chemistry

vi

CONTENTS

ABSTRACT ............................................................................................................................. iii

LIST OF ILLUSTRATIONS .................................................................................................. vii

LIST OF TABLES ................................................................................................................. xii

ACKNOWLEDGEMENTS ................................................................................................... xiii

Chapters

1. LITERATURE REVIEW ......................................................................................................1

Mechanism and barriers of ocular drug absorption ...................................................................1

Biodegradable polymers for ocular drug delivery .....................................................................8

Controlled release carriers for ocular drug delivery ................................................................31

2. INTRODUCTION ...............................................................................................................50

Statement of the problem .........................................................................................................50

Objectives ................................................................................................................................53

3. NOVEL PENTABLOCK COPOLYMER (PLA-PCL-PEG-PCL-PLA)

BASED NANOPARTICLES FOR CONTROLLED DRUG DELIVERY: EFFECT OF

COPOLYMER COMPOSITIONS ON POLYMER CRYSTALLINITY AND DRUG

RELEASE PROFILE FROM NANOPARTICLES.................................................................55

Rationale ..................................................................................................................................55

Materials and methods .............................................................................................................58

Results and discussion .............................................................................................................64

Conclusions ..............................................................................................................................96

4. A NOVEL PENTABLOCK COPOLYMER BASED COMPOSITE DRUG DELIVERY

SYSTEM FOR TIMOLOL MALEATE DELIVERY .............................................................97

Rationale ..................................................................................................................................97

Materials and methods ...........................................................................................................100

vii

Results and discussion ...........................................................................................................126

Conclusions ............................................................................................................................131

5. HYDROLYTIC AND ENZYMATIC DEGRADATION OF PENTABLOCK

POLYMERS ..........................................................................................................................132

Rationale ................................................................................................................................132

Materials and methods ...........................................................................................................132

Result and discussion .............................................................................................................134

Conclusions ............................................................................................................................144

6. PENTABLOCK COPOLYMER BASED THERMOSENSITIVE HYDROGEL FOR

MACRO MOLECULE DELIVERY .....................................................................................146

Rationale ................................................................................................................................146

Materials and methods ...........................................................................................................149

Results and discussion ...........................................................................................................154

Conclusions ............................................................................................................................171

7. SUMMARY AND RECOMMENDATIONS....................................................................172

Summary ................................................................................................................................172

Recommendations ..................................................................................................................176

APPENDIX ............................................................................................................................177

REFERENCES ......................................................................................................................181

VITA ......................................................................................................................................207

viii

LIST OF ILLUSTRATIONS

Figure Page

1.1 Anatomical Structure of the eye ..........................................................................................2

1.2 Corneal barriers ....................................................................................................................4

1.3 The blood retinal barrier ......................................................................................................5

1.4 Schematic representations of polymer degradation mechanisms ......................................10

1.5 Structures of different biodegradable polymers .................................................................13

1.6 Chemical structures of different polymers of polyester class ............................................18

1.7 Chemical structures of different Poly (ortho esters) ..........................................................27

1.8 Schematic representation of basic structures and different types of Liposomes ...............39

3.1 Synthetic scheme of pentablock copolymer (PLA-PCL-PEG-PCL-PLA) ........................66

3.2 1H-NMR spectra of PCL-PEG-PCL copolymer in CDCl3 ................................................70

3.3 1H-NMR spectra of PLA-PCL-PEG-PCL-PLA copolymer in CDCl3 ...............................71

3.4 FTIR spectra of copolymers (A) P-2 (B) P-3 (C) P-5 ........................................................73

3.5A X-ray diffraction diagrams of polymers P-1, P-4, P-6 ....................................................76

3.5B X-ray diffraction diagrams of polymers P-2, P-3, P-5 ....................................................77

3.6 DSC thermograms of polymers (A) First heating (B) Second heating ..............................78

3.7 Particle size distributions ...................................................................................................81

3.8 SEM image of nanoparticles prepared from pentablock polymers ....................................85

3.9 Figure 3.9 Release of triamcinolone acetonide from A , B , and C

triblock copolymers nanoparticles in PBS buffer (pH 7.4) at 37 °C. The values are

represented as mean ± standard deviation of n=3 ....................................................................89

a

b c

ix

3.10 Release of triamcinolone acetonide from P-2 , P-3 and P-5

copolymers nanoparticles in PBS buffer (pH 7.4) at 37 °C. The values are represented as

mean ± standard deviation of n=3 ............................................................................................90

3.11 Release of triamcinolone acetonide from P-2 , P-4 and P-6

copolymers nanoparticles in PBS buffer (pH 7.4) at 37 °C. The values are represented as

mean ± standard deviation of n=3 ............................................................................................91

Figure 3.12 Release of triamcinolone acetonide from P-6 nanoparticles alone and P-6

nanoparticles suspended in gel copolymers nanoparticles in PBS buffer (pH 7.4) at 37

°C. The values are represented as mean ± standard deviation of n=3 .....................................92

4.1 Synthetic scheme of PLA-PCL- PEG -PCL-PLA ...........................................................102

4.2 Synthetic scheme of PEG-PCL-PLA-PCL-PEG .............................................................103

4.3 1H-NMR spectra of PLA-PCL-PEG-PCL-PLA copolymer in CDCl3 .............................106

4.4 1H-NMR spectra of PEG-PCL-PLA-PCL-PEG copolymer in CDCl3 .............................107

4.5 FT-IR spectra (A) PLA-PCL-PEG-PCL-PLA (B) PEG-PCL-PLA-PCL-PEG ...............108

4.6 Sol-gel transition study of P-3 ........................................................................................110

4.7 Particle size distribution for P-1 ......................................................................................112

4.8 XRD analysis of nanoparticles loaded with timolol ........................................................115

4.9 Release of timolol from P-1 and P-2 copolymers nanoparticles in PBS buffer (pH

7.4) at 37 °C. The values are represented as mean ± standard deviation of n=3 ...................118

4.10 Release of timolol from P-1 Nanoparticles alone and P-1 nanoparticles in gel

copolymers nanoparticles in PBS buffer (pH 7.4) at 37 °C. The values are represented as

mean ± standard deviation of n=3 ..........................................................................................119

x

4.11 Cell viability studies (MTS assay) on P-1 and P-2 nanoparticles. The values are

represented as mean ± standard deviation of n=6. .................................................................123

4.12 Cell cytotoxicity studies (LDH assay) on P-1 and P-2 nanoparticles. The values are

represented as mean ± standard deviation of n=6. .................................................................124

4.13 Cell viability studies (MTS assay) on P-3 hydrogel. The values are represented as mean

± standard deviation of n=6. ..................................................................................................125

5.1 XRD analysis of PCL2500-PEG2000-PCL2500 after hydrolytic degradation .......................135

5.2 XRD analysis of PLLA2500-PCL2500-PEG2000-PCL2500-PLLA2500 after hydrolytic

degradation .............................................................................................................................136

5.3 XRD analysis of PDLLA2500-PCL2500-PEG2000-PCL2500-PDLLA2500 after hydrolytic

degradation .............................................................................................................................137

5.4 XRD analysis of PLLA2500-PCL7500-PEG1000-PCL7500-PLLA2500 after enzymatic

degradation .............................................................................................................................138

5.5 XRD analysis of PLLA2500-PCL2500-PEG2000-PCL2500-PLLA2500 after enzymatic

degradation .............................................................................................................................139

5.6 XRD analysis of PDLLA2500-PCL2500-PEG2000-PCL2500-PDLLA2500 after enzymatic

degradation .............................................................................................................................140

5.7 XRD analysis of PEG550-PCL825-PLA550-PCL825-PLA550 after enzymatic degradation ..141

6.1 Synthetic scheme of PEG-PCL-PLA-PCL-PEG .............................................................156

6.2 1HNMR spectra of P-2 .....................................................................................................157

6.3 FTIR spectrum of P-1 ......................................................................................................158

6.4 Phase diagram of thermosensitive hydrogel formulations of P-1 and P-2 .......................162

xi

6.5 Lysozyme release from two different thermosensitive gels PB-1 and PB-2 (20 wt %) at

37 °C. The values are represented as mean ± standard deviation of n=3 (p< 0.05) ..............164

6.6 Lysozyme release from PB-2 thermosensitive gel at two different concentrations at 37

°C. The values are represented as mean ± standard deviation of n=3 (p< 0.05) ...................165

6.7 Release profile of FITC-Dextran of different Mw was suspended in thermosensitive gel

PB-2 (20 wt %) at 37 °C. The values are represented as mean ± standard deviation of n=3 (p<

0.05) .......................................................................................................................................166

6.8 CD spectra of released lysozyme sample from P-1 formulation .....................................170

xii

LIST OF TABLES

Table Page

1.1 Polymeric drug delivery systems used in ophthalmic research .........................................15

1.2 Physico-chemical properties of polymers ..........................................................................20

1.3 Size of different types of liposomes...................................................................................40

1.4 Application of liposomes for the delivery of various drug molecules ...............................46

3.1 Characterization of triblock polymers ...............................................................................68

3.2 Characterization of pentablock copolymers .......................................................................69

3.3 Characterization of triblock nanoparticles .........................................................................83

3.4 Characterization of pentablock nanoparticles ....................................................................84

3.5 Kinetic parameters for drug release from triblock nanoparticles .......................................94

3.6 Kinetic parameters for drug release from pentablock nanoparticles .................................95

4.1 Characterization of copolymers .......................................................................................104

4.2 Characterization of nanoparticles ....................................................................................114

4.3 Kinetic parameters for drug release .................................................................................120

5.1 Degradation of PLA2500-PCL7500-PEG1000-PCL7500-PLA2500 ...........................................142

5.2 Degradation of PLLA2500-PCL2500-PEG2000-PCL2500-PLLA2500 ......................................142

5.3 Degradation of PDLLA2500-PCL2500-PEG2000-PCL2500-PDLLA2500 ................................142

5.4 Degradation of PEG550-PCL825-PLA550-PCL825-PLA550 ..................................................142

6.1 Molecular weight characterization of polymers ..............................................................159

6.2 Rheological measurement ................................................................................................160

6.3 Biological activity of lysozyme in the released samples .................................................169

xiii

ACKNOWLEDGEMENTS

I would like to express my sincere regards to everyone contributed in the successful

completion of my research work. I am extremely thankful to my advisor Dr. Ashim K. Mitra

for his constant support, motivation and exceptional guidance throughout my graduate

studies. He is a good mentor, whose constructive feedback always inspired me to navigate

my project in correct direction. I am thankful to Drs. Chi Lee, Kun Cheng, Kenneth S.

Schmitz and Andrew Holder for their time and contribution as dissertation committee

members. I am also thankful to Dr. Dhananjay Pal for teaching me cell culture experiments

and also for his constant moral support. I would also love to express my special regards to

Mrs. Ranjana Mitra for constant support throughout my stay at UMKC. I am also thankful to

Gyan Mishra for constant discussions and his help me in many experiments performed

during my graduate study. I am also thankful to Sulabh Patel for helping in all the cell culture

work.

I am also thankful to other investigators from UMKC for their time and support in my

dissertation project. In this regard, I am grateful to Dr. James Murowchick (Department of

Geosciences, University of Missouri-Kansas City) for conducting XRD experiments, Dr.

Gabriel L. Converse (Cardiac Surgical Research Laboratories, The Children's Mercy

Hospital) for conducting DSC experiments, Dr. Elisabet Nalvarte (Division of

Pharmacology, University of Missouri-Kansas City) for DLS experiments and Dr. Vladimir

M. Dusevich (School of Dentistry, University of Missouri-Kansas City) for helping me in

SEM studies. I am thankful to Dr. Sarah Hook, (School of Pharmacy, University of Otago)

for conducting rheological experiments. I am especially thankful to Joyce Johnson and

Sharon Self for their constant support and assistance in administrative work. I appreciate

xiv

support from Nancy Hoover and Connie Mahone from School of Graduate Studies. I am

thankful to National Institute of Health and School of Graduate Studies for constant funding.

Last, but the most important, I thank my parents and my brother for invaluable support and

immense faith in me.

1

CHAPTER 1

LITERATURE REVIEW

Mechanism and barriers of ocular drug absorption

Numerous efforts have been made to improve the bioavailability of ocular

therapeutics. Development of effective ocular drug delivery systems is an interesting and

complicated task for scientists in the area of ophthalmology. The eye is characterized by its

complex structure, both the anterior and posterior segments of the eye are in juxtaposition to

each other (Fig. 1.1) but different in anatomical and physiological aspects, which possess

unique challenges in delivering therapeutic agents (Ghate and Edelhauser, 2006; Mishra et

al., 2010). Both compartments are protected by anatomical and physiological barriers, which

prevent the entry of pathogens and foreign substances into the inner structure of the eye. The

major challenge for drug delivery scientists is to circumvent ocular barriers without causing

permanent tissue damage while maintaining the therapeutic concentration at the site of

action. Topical drug delivery is the most preferred and effective method to treat anterior

segment diseases. This route avoids the first-pass metabolism in the intestine and liver and

selectively targets the drug to the anterior segment tissues. Topical eye-drops currently

represent 90% of the marketed formulations. However, most of the applied dose is easily

drained away from the ocular surface due to defensive mechanism, resulting in low ocular

bioavailability (only 1-7 %) and sub therapeutic concentration in the posterior segment

tissues (Ghate and Edelhauser, 2006). Drug delivery to the posterior segments of the eye is of

a greater challenge than to the anterior segments due to various biological barriers i.e static

and dynamic in nature. There are two potential pathways for molecules to reach posterior

segment eye tissues following topical administration:

2

Figure 1.1 Anatomical Structure of the eye [adapted with permission from ref (Tamboli et al.,

2011)]

3

Role of corneal route in ocular drug absorption

Upon topical administration, therapeutic molecules are mainly subjected to the

corneal route of absorption. Most of the medication is taken away by the lacrimal drainage

system and some is removed by tear turnover. Drugs are also eliminated by systemic

absorption through conjunctival sac or nasolacrimal duct. A small fraction of drug is

available to enter the cornea and the inner region of the eye that has to face several

membrous barriers located in the cornea (Fig. 1.2), conjunctiva, iris-ciliary body and retina

(Fig. 1.3). Blood-Aqueous barrier (BAB), also known as anterior chamber barrier, primarily

prevents the entry of exogenous compound into the aqueous humor. It is formed from the

endothelial cells of uvea and restricts the movement of hydrophilic drug from plasma into the

anterior chamber. The drug molecules can cross the cornea either by transcellular

(intracellular) or paracellular (intercellular) pathway. The intracellular route dominates for

the entry hydrophilic compound or ions of small molecules, while the paracellular route

dominates for lipophilic molecules. The corneal epithelium is a lipophilic tissue that provides

resistance to hydrophilic compounds, whereas the stroma having an aqueous environment

controls the entry of hydrophobic compounds. The Bowman’s and descemet’s membrane do

not act as a barrier to drug absorption. The corneal endothelium is a monolayer of polygonal

cells with large intercellular junction that form cellular barrier between the stroma and

aqueous humor. The leaky hydrophilic gates of the endothelium allow the entry of

macromolecules from the stroma into the aqueous humor. It does not act as a rate limiting for

hydrophilic molecules but controls the entry of lipophilic molecules to some extent.

4

Figure 1.2 Corneal barriers [adapted with permission from ref (Tamboli et al., 2011)]

5

Figure 1.3 The blood retinal barrier [adapted with permission from ref (Tamboli et al., 2011)]

6

Role of non-corneal route in ocular drug absorption

Although the corneal route is considered to be a major route for penetration of topically

applied drugs, the conjunctival/sclera pathway is also a competitive and parallel route of

absorption. Conjunctival epithelium is leakier and has approximately 20 times more area than

cornea. It allows easy permeation of hydrophilic macromolecules. This non-corneal route has

a significant contribution for the absorption of large hydrophilic compounds such as insulin,

peptides and proteins. Moreover, small molecules such as timolol maleate, gentamycin and

prostraglandin PGF2 diffuse through conjunctiva and sclera because of their poor corneal

permeability (Koevary, 2002; Lee, 1990; Lehr et al., 1994). In general the non-corneal route

is comparatively less productive because the limbal area is full of blood vessels. The large

amount of the administered drug dissipates in the systemic circulation while crossing the

conjunctiva. The remaining drug diffuses through sclera and enters the posterior segment

while some penetrates across the conjunctiva, sclera and ciliary body and enter into the

anterior chamber. Blood-Retinal Barrier (BRB) primarily prevents the entry of molecules in

the posterior segment. It is formed by RPE and retinal endothelium. Due to limited blood

flow through choroid and presence of this barrier only a limited percentage of orally

administered drugs reach to the retina.

There are many influx and efflux transporters present in various tissues of the eye. Those

transporters also play an important role in absorption of drugs. Efflux pump are the

transporters which are responsible for extrusion of drug by transporting them out of the cell.

These pumps can act as barrier for both segments of eye. These pumps are the transport

protein, mainly responsible for the efflux of drugs from the cornea. The efflux pumps such as

P-glycoprotein (P-gp), Multiple drug resistance protein (MRP) and Breast cancer resistance

7

protein (BCRP) are found on the rabbit pigmented corneal epithelial cells (Mitra, 2009): a

brief overview of these transporters is given below:

1. P-glycoprotein (P-gp)

It is an ABC type transporter which can interact with various drug molecules. These drug

substrates are HIV protease, anti cancer, antibiotic, antihypertensive, cytotoxic agents. P-

gp is mainly localized on the apical side of the cell membrane although its mechanism of

extrusion is still not clear.

2. Multiple drug resistance protein family (MRP)

It is an ATP dependent efflux pump. Previously it is commonly known as canalicular

multispecific organic anion transporter as it mainly efflux lipophilic anions into the bile.

Its expression has been reported in various cancers such as colorectal, breast and ovarian.

Its substrate specificity is different from P-gp. Studies by Karla et al, has shown the

expression of MRP-2 on the rabbit pigmented corneal epithelial cells (Karla et al.,

2007a).

3. Breast Cancer Resistant Protein (BCRP)

It is a half transporter present in homodimer form in plasma membrane and extrudes

various structurally diverse compounds. BCRP is commonly expressed on the apical

membrane of small intestine and colon. Recent studies by Karla et al, showed the

presence of this transporter on cornea (Karla et al., 2009).

To overcome these barriers and to increase contact time of the drug on the eye

surface, absorption enhancers and/or viscosity enhancers are generally used in the ocular

formulations. So far, these approaches have limited success to address the problem of poor

bioavailability from topical route to treat anterior segment diseases. On the other hand, for

8

the treatment of posterior segment diseases either systemic or local route is preferred because

of the poor corneal drug permeation. Systemic administration requires higher dosage and

frequent administration that results in severe adverse effects. Local injections, particularly

intravitreal and subconjunctival injections are alternate strategies to achieve therapeutic

concentration in the vitreo-retinal disorders. However, to maintain the effective concentration

repeated injections are required, which causes clinical complications or patient discomfort

(Kimura and Ogura, 2001).

Biodegradable polymers for ocular drug delivery (Tamboli et al., 2012)

Many approaches have been evaluated to improve ophthalmic drug delivery.

Application of controlled drug delivery systems was anticipated as an effective approach to

circumvent all these limitations. Controlled drug delivery systems release the drug in a

sustained and controlled manner by which the therapeutic concentration is maintained for the

prolonged period of time. These systems provide many practical advantages: they avoid

frequent administration, which is a major non-compliance with many chronic eye disorders.

The delivery of emerging therapeutic macromolecules having very short biological half-lives

could be possible as these systems protect the protein drugs in situ and have an ability to

deliver them at desire rate by overcoming anatomical and biochemical barriers of drug

transport (Daugherty and Mrsny, 2003; Shell, 1984). These systems can be based on either

erodible or non-erodible matrices. In the early 1960s, first polymeric device was developed

for controlled drug delivery. Synthetic biodegradable polymers such as poly (glycolic acid)

(PGA) and poly (lactic acid) (PLA) had gained attention for biomedical applications. After

five-years, poly (lactide-co-glycolide) (PLGA) sutures emerged on the market. Since then, a

wide variety of biodegradable polymers were explored for the drug delivery (U. Adlund,

9

2002). In the past two decades the development and application of synthetic biodegradable

polymers for ocular drug delivery have gained significant momentum. Polymeric devices

such as micro and nanoparticles, microspheres, liposomes, hydrogels and ocular implants

have been designed to deliver the therapeutic agents in the controlled manner. The release

rate of the drug molecules from these polymeric devices depends on many factors such as,

molecular weight and degradation mechanisms of the polymer, physicochemical properties

of the drug, thermodynamic compatibility between the drug and polymer and the shape and

size of the devices (Park et al., 2005).

Biodegradation is an enzymatic or non-enzymatic hydrolysis of the polymeric

backbone into water soluble or insoluble products. Biodegradation involves two

complementary processes, degradation and erosion. In the degradation process cleavage of

the polymeric backbone into low molecular weight fractions takes place, whereas the erosion

mechanism refers to the physical phenomena such as dissolution and diffusion of low

molecular weight fractions from the polymer matrix. The degradation products are eventually

eliminated from the body via normal metabolic pathway (Katti et al., 2002).

Types of biodegradation

Heller has described three basic mechanisms of polymer degradation and classified the

polymers based on the degradation mechanisms (Heller, 1984). Schematic representation of

polymer degradation mechanisms is shown in Figure. 1.4.

10

Figure 1.4 Schematic representations of polymer degradation mechanisms [adapted with

permission from ref (Tamboli et al., 2012)]

11

Type I biodegradation

Cross linked water soluble polymers generally follow type-I erosion. Polymers such

as gelatin, collagen, polyacrylamides, poly (vinyl alcohol) (PVA) and poly (N-vinyl

pyrrolidone) (PVP) upon crosslinking form hydrogel, which is a water insoluble three

dimensional structure that undergoes type I hydrolysis. On the basis of hydrolysis product

generated, type I erosion mechanism can be further subdivided into type IA and IB. Type IA

erosion mechanism produces high molecular weight water soluble polymers, whereas type IB

generates low molecular weight polymers. Polymers having type IA erosion kinetics are best

suited for topical applications because of faster elimination of high molecular weight water

soluble polymers from the ocular surface. Polymers following type IB degradation kinetics

are generally utilized for designing implants. Polymeric systems that undergo type I erosion

are highly water permeable therefore; they are not suitable for the delivery of low molecular

weight compounds with appreciable water solubility. However, the crosslinked polymeric

matrix physically entangles the macromolecules and restricts them to diffuse out of the

matrix. Therefore, these polymers are well suited for the delivery of macromolecules such as

enzymes and antigens or sparingly water soluble molecules, which are released from

hydrogels initially via diffusion followed by degradation of the polymer (Heller, 1984).

Type II biodegradation

Conversion of water insoluble linear polymers into water soluble moiety through

hydrolysis, ionization, or protonation of pendant groups is defined as type II

erosion. However, since no backbone degradation is involved during erosion process overall

molecular weight of polymers does not change significantly. These polymers are generally

employed for the topical applications. Copolymers of alkyl vinyl ether and maleic anhydride

12

follow type II erosion mechanism where the degradation rate of the copolymers is affected by

the size of the alkyl substitute, pH of the degradation medium and pKa of the carboxylic

group (Woodruff et al., 1972).

Type III biodegradation

Type III erosion produces low molecular weight water soluble molecules by the

hydrolytic cleavage of water insoluble high molecular weight polymers. The polymers

demonstrating type III erosion kinetics produce non-toxic degradation products and thus

advantageous for topical and systemic administrations. These polymers are employed for

wound healing after surgery and also for chronic ocular diseases. These polymers are

available in a wide range of molecular weights having different physico-chemical properties,

which can be modulated to formulate drug delivery systems. PLA, PGA and their

copolymers PLGA, polyanhydrides, polyurathanes, polycaprolactone (PCL) and its

copolymers, poly (ortho esters) and poly (alkyl cynoacrylates) (PACA) exhibit type III

erosion mechanisms because of the characteristic hydrolytic instability in the polymer

backbone (Kimura and Ogura, 2001). The representative structures of these polymers are

shown in Figure. 1.5.

13

C

CN

COOR

* CH2 *

n

Poly(cyanoacrylates)

C O

O

C

O

*R*n

Poly(anhydrides)

C

OR'

R''

O O R* *

n

Poly(ortho esters)

* O

*

O

n

Polyester

Figure 1.5 Structures of different biodegradable polymers

14

Advantages of biodegradable polymers

Biodegradable polymers offer several advantages over non-biodegradable polymers

for controlled drug delivery. They do not require surgical removal after application, being the

most important advantage in ophthalmic drug delivery as it can circumvent surgical

complications associated with non-biodegradable implanted devices. The natural and

synthetic biodegradable polymers have many favorable properties such as biocompatibility

with ocular tissues, biodegradability and mechanical strength. They provide negligible

toxicity and also their degradation products are non-toxic in terms of both local and systemic

response. Due to the adequate mechanical properties, they can be tailored to wide range of

properties. Natural biodegradable polymers such as gelatin, albumin, chitosan, hyaluronic

acid and synthetic biodegradable polymers such as PVP, PACA, PCL, PEO, polyanhydrides

and thermoplastic aliphatic polyesters like PLA, PGA and PLGA have been thoroughly

explored for ocular delivery systems as summarized in Table 1.1. These polymers are

approved by FDA for human applications.

15

Table 1.1 Polymeric drug delivery systems used in ophthalmic research [adapted with

permission from ref (Tamboli et al., 2012)]

Polymer Example of

bioactivates Dosage form Model Ref.

Gelatin Timolol maleate Microsphere Rabbits (Bonferoni et al.,

2004)

Collagen Mitomycin c Implant In-vitro (Zimmerman, 2004)

Chitosan Indomethacin Nanoemulsions Rabbits (Badawi et al.,

2008)

PLA Ganciclovir Implant Rabbits (Kunou et al., 2000)

PLGA Vancomycine Microsphere Rabbits (Gavini et al., 2004)

PCL Dexamethasone Implants Rabbits (Silva-Cunha et al.,

2009)

PiBCA Pilocarpine Nanocapsules Rabbit (Desai and

Blanchard, 2000)

PECA Ganciclovir Nanoparticles Rabbit (EL- Samaligy,

1996)

Polyanhydride 5-fluorouridine Disks Monkeys (Jampel et al., 1990)

POE 5-chlorouracil and

fluorouracil

Injectable

solution Rabbits (Polak et al., 2008)

16

Synthetic biodegradable polymers used for ocular drug delivery

Poly N-vinylpyrrolidone (PVP)

PVP is a synthetic and biocompatible polymer widely utilized for vitreo-retinal drug

delivery. It is mainly employed for the preparation of hydrogels that exhibit viscoelastic

properties (Bruining et al., 1999). The decomposition products of PVP-based hydrogels are

easily eliminated from the vitreous through phagocytosis (Hong et al., 1998; Vijayasekaran et

al., 1996). Hydrogel prepared from cross linked PVP was used as a vitreous substitute

(Colthurst et al., 2000). Hong et al. evaluated biodegradation of poly (1-vinyl-2-

pyrrolidinone) cross-linked with 1% 14

C-methyl methacrylate. They observed that cross-

linked PVP hydrogel did not degrade in vitro in presence of proteolytic enzymes such as

trypsin or collagenase. However, in vivo half of the hydrogel disappeared from the rabbit

vitreous cavity within 4 weeks by phagocytosis (Hong et al., 1998). PVP based hydrogels

were transparent materials and remain at the site of injection for several weeks. However,

fragmentation of the hydrogels triggers an inflammatory response resulting in the vacuole

formation in the retinal pigment epithelium (Vijayasekaran et al., 1996). In addition, clinical

studies have shown that PVP-based hydrogels cause intravitreal opacity, hazy corneas and

inflammation and might not be suitable as vitreous substitutes (Colthurst et al., 2000).

Degradation kinetics of this biomaterial could be easily modulated by varying crosslinking

density. Niu et al. investigated injectable hydrogel of acrylamide/N-vinylpyrrolidone

copolymer crosslinked with reversible disulfide bond for ophthalmic applications. This

hydrogel showed characteristic in-situ sol-gel transition that facilitated the designing of

complex shapes, which was advantageous as artificial vitreous substance and scaffold for

lens regeneration (Niu et al., 2009). Hacker et al. explored the matrices composed of

17

photocrosslinked poly (propylenefumarate) (PPF)/ (PVP) for a long term delivery of

antiglaucoma drugs, such as acetazolamide (AZ), dichlorphenamide (DP) and timolol

maleate (TM). Authors suggested that the use of PVP based implants could be a valuable

strategy for controlled release of drugs over a period of 300 days in glaucoma therapy

(Hacker et al., 2009).

Poly (lacticide) (PLA), poly (glycolide), and their copolymers polylactide-co- glycolide

(PLGA)

PLA and PLGA are the most promising biodegradable polymers (Hyon, 2000). PGA

alone is highly prone to hydrolysis and remains insoluble in common organic solvents

therefore it is not widely acceptable for the fabrication of controlled drug delivery systems.

PLA alone and in combination with PGA with different ratios are mostly utilized in the

formulations. These polymers are synthesized by two methods. First involves direct

condensation reaction of monomers, which results in low molecular weight polymers and the

other method is based on ring opening polymerization of cyclic dimmers, which yields high

molecular weight polymers. These polymers upon non-enzymatic or enzymatic hydrolysis

produce water soluble metabolic products, which are not harmful to living tissues (Cam et

al., 1995; Hyon et al., 1998). These polymers belong to polyester class and degrade mainly

through bulk erosion. In vitro degradation of polyesters primarily occurs through hydrolytic

cleavage. However, in vivo, enzymes play an important role to initiate the degradation

process. The degradation products lactic acid and glycolic acid are nontoxic and eliminate in

the form of CO2 and water via Krebs cycle (Yasukawa et al., 2005). Chemical structures of

different polymers of polyester class are shown in Figure. 1.6.

18

Figure 1.6 Chemical structures of different polymers of polyester class

HO

O

O

O

H

O

y

x

PLGA

o

CH3

OH

O

H

n

PLA

oOH

O

H

n

PGA

C

O

H2C OHOH5 n

Polycaprolactone

19

Polymer degradation rate can be easily modulated by changing the molecular weight,

composition, conformation and crystallinity of the polymers (Thassu D., 2007). For example,

by varying the ratio of lactide and glycolide a wide range of diffusion and degradation

profiles can be obtained in PLGAs. PLGA with 50:50 ratio of lactic and glycolic acid

degrades faster than either PLA or PGA alone (Ogawa et al., 1988; Yasukawa et al., 2006).

The presence of methyl group provides more hydrophobicty to PLA and it degrades slowly in

comparison to PGA. These polymers have glass transition temperature ranging from 45 to 65

°C (Park P., 2006). Physico-chemical properties of different grades of PLGAs and other

polymers are summarized in Table 1.2.

20

Table 1.2 Physico-chemical properties of polymers [adapted with permission from ref

(Tamboli et al., 2012)]

Polymers Glass transition

temp, Tg (˚C)

Melting

temp,Tm(˚C)

Biodegradation

time(months)

Approximate

strength

( Modulus )

Poly (glycolic acid ) 35-40 225-230 6-12 7.0 Gpa

Poly ( l- lactic acid ) 60-65 173-178 >24 2.7 Gpa

Poly ( d, l- lactic

acid) 55-60 None 12-16 1.9 Gpa

Poly(ε-caprolactone) -65 to -60 58-63 >24 0.4 Gpa

Poly(d,l-lactic-co-

glycolic acid) [85/15] 50-55 None 5-6 2.0 Gpa

Poly (d,l-lactic-co-

glycolic acid) [50/50] 45-50 None 1-2 2.0 Gpa

Poly1,6[-bis

(carboxyphenoxy)

hexane]

-- -- 12 (in-vitro) 1.3 Mpa

21

Drug release from the PLGA system depends on the proportion of two monomer used,

porosity, and surface area of the carrier and physico-chemical properties of the incorporated

drug (Deshpande et al., 1998; Vega et al., 2008). These polymeric materials have been used

as surgical sutures due to their good biocompatibility and rapid clearance (Visscher et al.,

1985). PLA and PLGA are widely utilized in ocular drug delivery systems such as implants,

injectable microspheres and nanoparticles. PLA and PLGA microspheres have been

evaluated to reduce the intravitreal administration frequency for various chronic eye diseases

such as cytomegalovirus retinitis and endophthalmitis (Duvvuri et al., 2007; Moritera et al.,

1991). Dillen et al. developed cationic Eudragit®

coated PLGA nanoparticles loaded with

ciprofloxacin, a most commonly used fluoroquinolone for ocular infections. These authors

found that positively charged drug loaded nanoparticles can adhere to the negatively charged

bacterial surface. In addition, particulate systems enhanced the therapeutic drug

concentration at the target site by providing prolonged diffusion controlled release (Dillen et

al., 2006). Kunou et al. achieved pseudo zero-order release kinetics of GCV over a period of

one year by employing two monomers of PLA with different molecular weights and ratios

for the preparation of biodegradable scleral implant (Kunou et al., 2000). The PEG- coated

PLA nanospheres were more efficient for sustaining the drug release and improving the

ocular bioavailability of ACV in the treatment of viral infections (Giannavola et al., 2003).

PLGA was also utilized to encapsulate anti-VEGF RNA aptamer (EYE001) in microspheres.

In contrast to a characteristic triphasic release pattern of microsphere, the release of EYE001

from PLGA microspheres was a diffusion-controlled process that exhibited drug release in

continuous manner over a period of 20 days. Furthermore, the bioactivity of the aptamer was

retained in the formulation during the entire release period (Carrasquillo et al., 2003).

22

Duvvuri et al. discussed the conventional triphasic release pattern from PLGA microspheres

and optimized the drug release kinetics by employing various PLGA polymer blends

(Duvvuri et al., 2006). However, particulate systems such as microspheres and nanospheres

may cause vision obstruction or irritation to the retinal tissues after intravitreal injections. In

addition, most of the PLGA based drug delivery systems have initial burst release phase.

Authors investigated the composite approach to minimize the particulate system related

drawbacks. This dual approach involved the use of PLGA-PEG-PLGA triblock

thermogelling polymer to suspend the particulate system (Duvvuri et al., 2005; Zentner et al.,

2001). The thermosensitive polymer exists in the liquid state at room temperature and forms

gel upon contact with eye tissue i.e. 34 0C. Thermosensitive hydrogel holds particles at the

site of administration and avoids vision interference. In addition, gel matrix protects the

microspheres from enzymatic and cellular degradation. This dual system showed release of

drug in a more controlled manner for prolonged period of time (Duvvuri et al., 2007).

Poly- ε-caprolactone (PCL)

PCL is an aliphatic polyester synthesized form monomer ε- caprolactone through ring

opening polymerization catalyzed by stannous octoate at 140 0C (Sinha et al., 2004). It is a

tough semi crystalline polymer having the melting point in the range of 59 and 64 °C and a

glass transition temperature of -60 °C (Murthy R., 1997). Permeability and crystallinity of

the PCL can be modified by co-polymerization with PLA or PGA (Sinha et al., 2004).

Degradation of PCL occurs in two phases. First phase involves molecular weight (Mn) loss

up to 5000 due to cleavage of ester linkage in the polymer backbone (chain scission), that

produces ε- hydroxyl caproic acid and decreases the intrinsic viscosity of polymer. In the

second phase (commonly observed in vivo), chain scission of low molecular weight polymer

23

produces small fragments, which diffuse out of the polymer bulk and break the polymer in

small particles that undergo phagocytosis (Deshpande et al., 1998). PCL is utilized for

sustained drug delivery due to its higher permeability to various drug molecules and slower

degradation in comparison to other polymers (Murthy R., 1997). Degradation rate of PCL

can be improved by co-polymerizing with other fast degradating polymers. PCL implant

loaded with dexamethasone had released the drug within the therapeutic range over the

period of more than one year and was well tolerated in the rabbit eye (Fialho et al., 2008).

Rod shaped PCL implant loaded with triamcinolone was also well tolerated in sub-retinal

space of rabbit eye and had released the drug over the period of 4 weeks without any clinical

complications (Beeley et al., 2005). Yenice et al. evaluated hyaluronic acid coated PCL

nanospheres loaded with cyclosporine. Investigators found that bioavailability of

cyclosporine nanospheres was 10-15 fold higher than the drug solution in castor oil. PCL can

be utilized to prepare in situ gel-forming sustained drug delivery system. PCL based triblock

polymer was recently characterized for ophthalmic applications. Gong et al. evaluated the

toxicity of PEG-PCL-PEG triblock copolymer hydrogel after intracameral injections. This

hydrogel was biocompatible with ocular tissues and appeared to be a promising controlled

release systems for chronic ocular diseases (Yin et al., 2010).

Poly (alkyl cynoacrylates) (PACA)

PACA is synthesized from monomer alkyl cynoacrylate, which exhibits bioadhesive

properties. It can form a strong bond with polar surfaces including skin and living tissues.

Polymethylacrylate composed of smaller alkyl chain is not applicable to drug delivery due to

tissue toxicity and inflammation. Therefore, larger alkyl chains such as n-butyl, octyl

cyanoacrylates are used for clinical applications. In practice anionic or zwitterionic

24

polymerization are commonly used for synthesis of PACA due to rapid initiation at ambient

temperature. Polymer degradation occurs by enzymatic hydrolysis of alkyl side chain

producing an alkyl alcohol and poly (cyanoacrylic acid). The degradation products are

soluble in water and eliminate via kidney filtration (Vauthier et al., 2003). This polymer was

mostly explored for the preparation of biodegradable nanoparticles. Layre et al. reported the

encapsulation of highly crystalline alkylating drug, busulfan, utilizing five different PACA

polymers. The highest encapsulation of busulfan was found in poly (isobutyl cyanoacrylate)

(PIBCA) and poly (ethyl cyanoacrylate) due to the specific interaction between the drug and

the polymers. They suggested that this nanoparticulate formulation when given intravenously

have nigligible toxicity and also minimize the variability in bioavailability (Layre et al.,

2006). Peracchia et al. suggested the potential use of PEG-coated PIBCA nanoparticles as

drug delivery carriers, which are rapidly biodegradable. The covalently bound PEG avoids

interaction with blood components and prevents recognition by macrophages of the

mononuclear phagocyte system after intravenous injection (Peracchia et al., 1997). PEG-

coated polyethyl-2-cyanoacrylate nanospheres had increased the ocular bioavailability of

ACV by 25 fold when instilled in the conjunctival sac of rabbit eyes. The improved drug

bioavailability was attributed to the colloidal nature of nanospheres that can facilitate the

transport of drug paracellularly. In addition, presence of PEG provided better mucoadhesion

on the corneal surface and improved dug permeation (Fresta et al., 2001).

Polyanhydrides

In 1930s, Hill and Carothers proposed polyanhydrides as substitutes of polyesters for

textile applications. However, they were not useful for textile industry due to faster

hydrolytic cleavage of anhydride linkage in the polymer backbone. Instead, due to this

25

intrinsic property polyanhydrides are considered as an ideal candidate for formulation of

controlled drug delivery systems. Hydrolytically labile linkages of polyanhydrides provide

bio-degradability and regulate degradation rate. For example, poly [bis (p- carboxyphenoxy)

alkane anhydrides] degradation rate can be adjusted from 10-1

to 10-4

mg/hr/cm2 upon

changing the methyl group to the hexyl group. Polyanhydrides can be synthesized by three

methods: melt condensation, dehydrochlorination and dehydrative coupling. Melt

condensation method can produce high molecular weight polymer (up to 50,000) while other

two methods are useful for the synthesis of low molecular weight polymers (Leong, 1987).

Polyanhydrides show pH dependent degradation, which can be modulated by additives. Basic

additives primarily promote bulk erosion, whereas acidic additives favor surface erosion and

produces acetic acid upon degradation (Leong et al., 1985). Further, upon changing the

polymeric backbone drug release rate can be modulated over a thousand fold (Jain et al.,

2005). Mostly copolymer of bis(p-carboxyphenoxy propane) and sebacic acid is utilized for

drug delivery applications. Release of drug from this polymeric delivery system occurs

mainly by surface erosion rather than drug diffusion. It has a potential to provide almost

zero-order drug release rate and also undergoes relatively faster in-vivo biodegradation (Jain

et al., 2005). Rosen et al. demonstrated near zero order degradation and drug release kinetics

from poly [bis (p-carboxyphenoxy methane andride] for several months at two different

temperatures i.e. 37 and 60˚ C (Rosen et al., 1983). Polyanhydride microspheres have been

employed to avoid repeated intravitreal injections for the treatment of vitreoretinal diseases

(Jain et al., 2005). Microspheres prepared from poly (adipic anhydride) (PAA) exhibited

surface degradation. Release of timolol maleate from these microspheres was sustained for 7

hrs and mainly controlled by polymer degradation. Further, to improve ocular bioavailability

26

of timolol maleate PAA-microspheres were suspended in the Gelrite@ (an in situ

polysaccharide gel) (Albertsson, 1996). In another study by Lee et al., 5-fluorouracil (5-FU)

was incorporated in 3mm bio-erodible disc of bis (p-carboxyphenoxy) propane and sebacic

acid. They reported that 5-FU was delivered in a sustained manner and maintained intra-

ocular pressure for 3 weeks (Lee et al., 1988).

Poly (orthoester) (PEOs)

Since early 1970s, four families of poly (ortho esters) have been synthesized. POEs

are hydrophobic polymers having hydrolytically labile ortho ester bonds. The amount of

water available to react with these bonds is very less under physiological conditions that

make the polymer extremely stable. POEs undergo surface erosion and provide zero order

release rate for a longer period of time (Einmahl et al., 2003). POEs based formulations have

been proven promising in the treatment of ocular diseases such as glaucoma filtration surgery

and proliferative vitreoretinopathy (PVR) (Bernatchez et al., 1993). They can be injected

directly into the eye with a needle of appropriate size. Chemical structures of four types of

POEs are shown in Figure 1.7.

27

Figure 1.7 Chemical structures of different Poly (ortho esters) [adapted with permission from

ref (Tamboli et al., 2012)]

28

POE I is synthesized by transesterification reaction between a diol and diethoxy

tetrahydrofuran. It is a hydrophobic solid polymer having acid sensitive nature and easily

hydrolyzed in an aqueous environment. Basic ingredients such as sodium carbonate are

generally utilized to prevent autocatalytic hydrolysis. This polymer has been widely explored

for orthopedic applications, treatment of burns and for the delivery of narcotic antagonist and

contraceptive steroids. It is not explored much for ocular drug delivery (Bernatchez et al.,

1993; Heller, 2005; Heller et al., 2002).

POE II is synthesized by simple addition reaction between diol and di keteneacetal

3,9-di(ethylidene 2,4,8,10-tetraoxaspiro[5.5]undecane) (Heller et al., 2002). Monomers are

required to dissolve in tetrahydrofuran and trace of acidic catalyst is used to initiate polymer

synthesis instantaneously. Polymer hydrolysis occurs in two steps, unlike POE I there is an

absence of autocatalytic hydrolysis. This polymer is also synthesized by crosslinking a triol,

either alone or as a mixture with diols. It forms a dense polymer upon crosslinking, which

biodegrades to small water soluble fragments. The cross linked density can be adjusted by

varying the ratio of diol to triol. POE II can be fabricated as a hard glassy material to semi

solid material; mechanical and thermal properties can be controlled by using diols having

different degrees of chain flexibility (Heller, 2005; Heller et al., 2002). POE II have been

extensively explored for the release of 5-FU, which is mainly utilized as an adjunct to

glaucoma filtration surgery. The erosion rate of POE II can be controlled by incorporating

the acidic excipients such as suberic, adipic and itaconic acids in the polymer matrix. Nearly

zero–order release of 5-FU was obtained by incorporating different amount of suberic acid in

POE polymeric matrix (Einmahl et al., 2003). According to the United States of

29

Pharmacopoeia, this generation of polymer is nontoxic for cellular, subcutaneous,

intramuscular and systemic implant applications.

POE III is semi-solid at room temperature and synthesized via transesterification of 1,

2, 6- and trimethyl orthoacetate (Heller, 2005). This polymer has a very flexible backbone

and allows incorporation of therapeutic agents at room temperature without using organic

solvent. Therefore, it can be used for thermo labile and solvent sensitive drugs. The release

rate of incorporated drug can be controlled by modulating the molecular weight of the

polymer. Merkli et al. observed that the release of 5-FU occurred within 1 day from 3500 Da

and was sustained for 1 week from 33,300 Da POE III. (Merkli, 1994). Drug delivery

systems fabricated from this polymeric material do not show any burst release and the drug

release rate was governed by the polymer degradation rate (Einmahl et al., 2001). Sintzel et

al. reported that the drug release rate from POE III can be controlled by modulating the

hydrophobicity of the polymer by substituting triol from 1,2,6 hexanetriol to 1,2,10-

decanetriol (Sintzel, 1997). According to Einmahl et al., this new generation of POE has a

potential for application in glaucoma filtering surgery for the patients with higher risk of

surgery failure. This injectable polymer can provide sustained release of 5-FU for 2 weeks

after subconjunctival injection that can avoid frequent administrations and minimize the

adverse effects (Einmahl et al., 2001). These authors have also described that after

subconjunctival administration, polymer degradation products follow several pathways. One

major pathway involves direct entry into the anterior chamber through the fistula, to the

ciliary body, into the vitreous body, and then into the retina (Einmahl et al., 2003). Therefore

they evaluated the biocompatibility of this polymer in different parts of the eye including

anterior chamber and suprachoroidal space. They found that the anterior chamber of the

30

rabbit eye can tolerate up to 50 µl of polymer solution, which degrades within 1 week

(Einmahl et al., 2000). In addition, after suprachoroidal injection the retinal pigmented

epithelial (RPE) cells, retinal and choroidal vasculatures were not affected by the polymeric

formulation (Einmahl et al., 2000; Einmahl et al., 2002). POE III demonstrated excellent

biocompatibility to the different parts of the rabbit eye. Difficulties in synthesis and lack of

reproducibility have limited the use of POE III in biomedical applications (Heller, 2005).

POE IV is synthesized by reacting diols with the diketeneacetal 3, 9-diethylidene-2,

4, 8, 10-tetraoxaspiro [5.5]undecane. It is a modified form of POE II, which contains latent

acid in the polymer backbone that regulates the erosion rate. The latent acid is generally

composed of glycolic acid or lactic acid. POE IV does not require external acidic excipients

to control the erosion rate, unlike POE II. When the polymer is exposed to an aqueous

solution, the latent acid will hydrolyze to give lactic acid or glycolic acid that will further

assist in the hydrolysis of polymer. POE IV can be fabricated as solid or gel-like material by

changing the nature of diols. POE IV-based devices generally undergo surface erosion and

produce acidic degradation products which readily diffuse out from the device. Lactic acid

based fourth generation POEs are biocompatible and have long residence times following

intracameral, subconjunctival, intravitreal and suprachoroidal injections in the rabbit eyes

(Einmahl et al., 2003). Polak et al. evaluated the efficacy of 5-chlorouracil (5-CU) loaded

POE IV formulation in the glaucoma filtration surgery. They found that 5-CU suspended in

POE IV has maintained low IOP in the rabbit eye for 5 months (Polak et al., 2008).

Polymeric materials contribute a significant role in the controlled drug delivery. In particular,

biodegradable polymers have been extensively explored for ocular therapeutics in the recent

years. In this chapter we have summarized mainly the properties and applications of

31

biodegradable polymers having natural and synthetic origins. We have exemplified the

applications of biodegradable polymers for the delivery small molecules to the different parts

of the eye. Two major advantages of polymeric drug delivery devices, enhancing drug

bioavailability and minimizing side effects, are significant in ocular drug delivery. The

development of new biodegradable block polymers has gained significant momentum in the

recent years. These polymers would be advantageous in the delivery of newer therapeutic

agents including genes, therapeutic antibodies and bioactive proteins. It is challenging to

deliver these macromolecules to targeted tissues of the eye. Therefore, design of novel

biodegradable polymeric devices is currently under investigations for targeted delivery of

macromolecules.

Controlled release carriers for Ocular drug delivery

Various sight-threatening and chronic ocular diseases such as age-related macular

degeneration, proliferative vitreoretinopathy, chronic cytomegalovirus retinitis (CMV)

diabetic macular edema and other ocular inflammatory conditions require sustained levels of

therapeutic agents for longer duration (Mishra et al., 2011b). Treatment of these diseases

requires frequent drug administrations. Conventional routes fail to achieve required

therapeutic levels in the eye due to the presence of ocular barriers. In recent years,

nanotechnology has attained wide acceptance in ocular drug delivery. Novel nanocarriers

such as nanomicelles, microsphere, nanoparticles, liposomes and surface modified nano

formulations are very efficient in circumventing various ocular barriers (Sahoo et al., 2008;

Vandervoort and Ludwig, 2007). These systems can avoid frequent administrations and

release therapeutic molecules at the targeted site in a controlled manner for prolonged

periods. Nanotechnology based drug delivery systems may provide many advantages such as

32

enhanced cellular uptake, stimuli sensitive release and targeted delivery to specific ocular

tissues. Nanocarriers are particularly beneficial in many angiogenic ocular diseases such as

diabetic retinopathy, choroidal neovascularisation (CNV), central retinal vein occlusion and

intraocular solid tumors, where enhance permeation retention (EPR) effect can be achieved

by targeting drugs with nanocarriers (Mishra et al., 2010; Yasukawa et al., 2004). In addition,

polymeric nanocarriers are effective in gene delivery to specific ocular which can overcome

issues regarding short intravitreal half-life and transient gene expression. An ideal

nanocarrier for ophthalmic applications should possess appropriate size and narrow

distribution that ensure low irritation to ocular tissues and provide adequate ocular

bioavailability (Sahoo et al., 2008). Our research group has developed and evaluated

different nanocarriers for ophthalmic applications.

Microparticles and microspheres

PLGA polymer based controlled drug delivery systems such as microspheres and

microparticles have gained attention in ophthalmic applications. Availability of different

grades of PLGA polymers provides an excellent platform for formulation scientists to

develop a tailor made sustained release formulations according to drug of choice (Duvvuri et

al., 2006). An ideal sustained release microsphere formulation should provide continuous

release of entrapped drug for a desire period. Drug release from PLGA microsphere usually

follows three stages (Duvvuri et al., 2005). Initial burst release phase (Phase I) is attributed

to diffusion of surface absorbed and poorly encapsulated drug. Slow or no release phase

(Phase– II) can be attributed to possible drug-polymer interactions, which results in low drug

levels at the target site. Third phase of rapid release (Phase-III) is mainly attributed to faster

diffusion of drug molecules from polymer matrix due to degradation of polymer (Duvvuri et

33

al., 2006). PLGA polymers are available in wide range of molecular weights with different

lactide/glycolide ratios, the amount of drug release during each phase could vary

considerably depending upon type of PLGA utilized in microsphere formulation.

Particularly, phase-II release duration is dependent on the hydrophilicity of the matrix. In

addition, all these release phases are most likely evident for hydrophilic molecules compared

to lipophilic molecules (Duvvuri et al., 2006). The extent of drug release during different

release phases can be modulated by adding polymeric additives (Mishra et al., 2010), release

modifying agents such as tweens and polyethylene glycols or by utilizing polymeric blends

(Duvvuri et al., 2005). In an attempt to modify GCV release from PLGA microsphere,

different PLGA blends were used in the formulation. A small molecular weight hydrophilic

PLGA polymer, Resomer RG 502H (d,l-lactide : glycolide::50: 50, Mw=8000 Da) was

blended with PLGA 65/35 (d,l-lactide : glycolide::65: 35, Mw=45,000–75, 000 Da) to

modulate drug release kinetics from microspheres (Duvvuri et al., 2005). Drug entrapment

efficiency was also increased from 47.13% with Resomer RG 502H microspheres to 72.67 %

for polymer blend microspheres. GCV release from both Resomer RG 502H microspheres

and blend microsphere followed triphasic pattern. However, drug release rate constant of

each release phase was significantly lowered in the case of blend microsphere compared to

Resomer RG 502H microspheres. Although microsphere can provide controlled delivery of

entrapped molecules for longer time periods major obstacles anticipated with its application

are retinal irritation and vision obstruction due to movement of particulate system (floaters)

in the vitreous following intravitreal injection. Moreover, due to hydrophobic nature PLGA

microsphere can agglomerate in aqueous buffer and poses a challenge for formulation

development. One promising strategy is the development of composite formulation

34

comprising of nanoparticles in suspended thermosensitive hydrogel. Thermogelling polymer

solution remains in sol state at room temperature and forms gel at physiological temperature.

We observed that thermogelling polymer can form a depot upon administration in to vitreous

cavity and minimize the movement of GCV microspheres in the vitreous. PLGA-PEG-PLGA

copolymer was utilized as thermosensitive hydrogel (Duvvuri et al., 2005). Our research

group evaluated the intravitreal pharmacokinetics of GCV following single intravitreal

injection of GCV solution and equivalent GCV microsphere formulation (Resomer RG 502H

and a blend of Resomer RG 502H: PLGA 65/35::1:3, were physically mixed in a 1:1 ratio

and suspended in a 23% w/w aqueous solution of PLGA-PEG-PLGA) in male New Zealand

white rabbits (Duvvuri et al., 2007). Following administration of GCV solution, the vitreous

level of GCV was maintained above its minimum inhibitory concentration (MIC) for only 54

hrs with vitreous elimination half-life of 6.45 hrs. In contrast, with gel formulation the

vitreous level of GCV was maintained above its MIC for 14 days. These results suggest that

composite formulations can be used to deliver GCV to the vitreous and retina/choroid

following intravitreal administration continuously over 4–5 weeks. In addition, these

formulations is biodegradable and biocompatible in nature and do not require surgical

removal following completion of drug release. These studies also suggest that sustained

delivery of GCV can be tailor made by polymer blending strategy.

Nanoparticles

Nanoparticles are defined as sub-micron sized carriers (10 to 1000 nm) in which drug

molecules are dissolved, entrapped or adsorbed. Nanoparticles can be further divided in to

nanospheres and nanocapsules. Nanospheres have a solid matrix of polymers or lipids on

which drug molecules are simply adsorbed on the surface. Nanocapsules have polymeric

35

shell and aqueous core like structure in which molecules are mostly entrapped and dissolved

in the aqueous core. Nanoparticle surfaces can be functionalized with specific receptor

targeted moiety to achieve site specific delivery and to enhance transport of therapeutic

molecules across the various physiological ocular barriers such as corneal and blood retinal

barrier. Various natural and synthetic biodegradable polymers used in the development of

nanoparticles have demonstrated promising results in ophthalmic drug delivery. The most

common ones are polycaprolactone (PCL), polylactide (PLA) and poly(lactide-co-glycolide)

(PLGA). These are USFDA-approved polymers that degrade in vivo via the Kreb’s cycle,

resulting in biocompatible by-products (lactic and glycolic acids). These polymers can be

fabricated or formulated into devices such that they provide controlled drug release from a

few days to years (Janoria et al., 2007). PLGA nanoparticles loaded with different steroids

such as dexamethasone, hydrocortisone acetate, and prednisolone acetate were developed as

a vehicle to provide sustained levels of corticosteroids in the posterior ocular segment

(Janoria et al., 2007). PLGA nanoparticles prepared from single emulsion solvent

evaporation method resulted in uniform size particles with significant increase in entrapment

efficiency and drug loading relative to nanoparticles prepared by dialysis method. Different

grades of PLGA i.e. PLGA 50/50 and PLGA 65/35 were evaluated for nanoparticle

preparation. Nanoparticles prepared from PLGA 65/35 provided more prolonged release of

steroids in comparison to PLGA 50/50. This effect may be explained by the fact that PLGA

65/35 is comparatively more hydrophobic in nature that in turn retards the rate of water

penetration into polymer matrices. Initial burst release of steroids was also found to be less in

nanoparticles prepared from PLGA 65:35. Based on the in vitro release data, the researchers

concluded that PLGA 65/35 provides promising results in comparison to PLGA 50/50 for

36

sustained delivery of steroids from nanoparticles (Janoria et al., 2007). However,

nanoparticles prepared from PLGA polymer produce high amounts of lactic and glycolic

acids after degradation, which may cause local tissue irritation and degradation of therapeutic

molecules (Kang and Schwendeman, 2002; Meyer et al., 2012). In addition, PLGA

nanoparticles produce high initial burst release followed by no release phase (Boddu et al.,

2010b; Budhian et al., 2005). Considering this, we developed novel polymeric materials for

preparation of controlled release formulations that can address these limitations.

Micelles

Drug delivery to the posterior segments of the eye presents a greater challenge

because of the selective functionality of the biological barriers. Particularly, delivery of

hydrophobic molecules to the posterior segments via topical delivery is ineffective.

Moreover, the low aqueous solubility of many steroids limits the feasibility of formulating a

highly concentrated aqueous eye drop formulation, making the topical route inefficient in

delivering adequate drug levels to the retina (Forrest et al., 2006; Loftsson and Hreinsdottir,

2006). However, many investigators are now exploring the potential of topical delivery for

the back-of-the-eye diseases. Polymeric micelles are exploited as pharmaceutical

nanocarriers for the delivery of poorly water-soluble drugs, which can be solubilized in the

hydrophobic inner core of a nanomicelle. In this regard, LX214, a nanomicellar formulation,

10 to 15 nm in size containing up to 0.2% voclosporin, was developed by our research group

in collaboration with Lux Biosciences Inc (Mitra et al., 2010; Mitra et al., 2009, 2011).

Nanomicelles are colloidal particles in nanometer size ranges (10-100 nm), forming spherical

structures of amphiphilic molecules in water (Aliabadi and Lavasanifar, 2006; Torchilin,

2007). The nanomicellar corona is comprised of hydrophilic chains extended outwards (see

37

Figure 1). Micelles can therefore serve to improve solubility and bioavailability of various

hydrophobic (water-insoluble) drugs. LX214, mixed nanomicelles are composed of two non-

ionic surfactants, D-alphatocopheryl polyethylene glycol 1000 succinate (Vitamin E TPGS)

stabilized with octyl phenol ethoxylate (octoxynol-40) in a defined ratio. LX214, packaged in

single-use sterile low-density polyethylene, blow-fill -sealed vials, has demonstrated stability

for at least 1 year under refrigeration and for 2 months at room temperature. Hydrophobic

molecules, such as voclosporin, encapsulated in 15 nm nanomicelles, form spherical

structures in water. Due to their hydrophilic corona, these micellar nanocarriers can pass

through the aqueous channels/pores of the sclera, which range from about 30 nm to about

300 nm in size. Nanomicelles may then be absorbed onto the basolateral side of the retinal

pigment epithelium (RPE) through endocytosis. The contents of the micellar nanocarriers are

discharged inside the cell after fusion with the cell membrane. It is also believed that during

the transit, the hydrophilic nanomicellar corona helps drug molecules to evade wash-out into

the systemic circulation by the conjunctival/choroidal blood vessels and lymphatics. In recent

years, polymeric micelles composed of amphiphillic block copolymers such as PLGA-PEG,

PCL-PEG, polyethylene oxide (PEO)-b-poly (ester) and Pluronics have gained attention in

drug delivery. Chemical structure of block copolymers can be easily modified to achieve

high drug payload, improve stability, modulate drug release kinetics and achieve target

specific delivery (Croy and Kwon, 2006). In an attempt to evaluate the role of folate

receptor-targeted drug delivery system for the treatment of retinoblastoma doxorubicin

(DOX) loaded PLGA-PEG-folate micelles were developed by our research group (Boddu et

al., 2010a). The uptake of DOX was found to be 4 times higher in the presence of targeted

DOX micelles compared to DOX alone. Development of novel micellar formulation targeted

38

to specific receptor/transporter opens doors for non invasive delivery of hydrophobic drugs

for treatment of diseases affecting the anterior and/or posterior ocular segments.

Liposomes

In 1965 liposomes were first introduced as drug delivery carriers (Bangham et al.,

1965). Liposomes are usually within the size range of 10 nm to 1 µm or greater. These

vesicular systems are composed of an aqueous core enclosed by phospholipid bilayers of

natural or synthetic origin. Liposomes are structurally classified on the basis of lipid bilayers

such as small unilamellar vesicles (SUVs) or multilamellar vesicles (MLVs). Furthermore,

on the basis of size liposomes are classified into small unilamellar vesicles (SUVs), giant

unilamellar vesicles (GUVs) and large unilamellar vesicles (LUVs) (Fig.1.8). Unilamellar

vesicles are composed of single layer of lipid such as lecithin or phopshotidylglycerol

encapsulating aqueous interior core. Multilamellar vesicle is composed of various layers of

lipid bilayers (Elbayoumi and Torchilin, 2010; Kaur et al., 2004; Mainardes et al., 2005).

MLVs are metastable energy configuration having different facets depending upon the

polydispersity of the liposomal formulation. Various types of liposomes with size are

summarized in table 1.3.

39

Figure 1.8 Schematic representation of basic structures and different types of Liposomes

[adapted with permission from ref (Mishra et al., 2011a)]

40

Table 1.3 Size of different types of liposomes (Jesorka and Orwar, 2008)

Vesicle type Size

SUVs ~ 20nm to ~ 200 nm

LUVs ~ 200 nm to ~ 1 μm

MLVs > 0.5 μm

GUVs > 1 μm

41

Drug loading capacity of liposomes depends on many factors such as size of

liposomes, types of lipid utilized for preparation and physicochemical properties of

therapeutic agent itself. For example, entrapping efficiency for SUVs are poor in comparison

to MLVs. However, LUVs provide a balance between size and drug loading capacity.

Liposomes are advantageous in encapsulating both lipophilic and hydrophilic molecules.

Hydrophilic drugs are entrapped in the aqueous layer while hydrophobic drugs are held in the

lipid bilayers. Loading capacity of ionic molecules can be further improved by using cationic

or anionic lipids for the preparation of liposomes (Ding X et al.).

Majority of liposomal formulations utilize phosphotidylcholine (PC) and other

constituents such as cholesterol and lipid conjugated hydrophilic polymers as the main

ingredients. Incorporation of cholesterol enhances the stability by improving the rigidity of

the membrane. Stability of liposomes depends upon the various properties such as surface

charge, size, surface hydration and fluidity of lipid bilayers. Surface charge determines

interaction of liposomes with ocular membrane. Positively charged liposomes display better

corneal permeation than the neutral and negatively charged liposomes. Neutral liposomes

upon systemic administration evade the elimination by reticuloendothelial system (RES).

However, these vesicles possess higher self aggregation tendency. In contrast, negative and

positively charged liposomes exhibit lower aggregation tendency but undergo rapid clearance

by RES cells due to higher interaction with serum proteins. In addition, size of the liposomes

can also regulate the clearance by RES. Liposomes of size less than 100 nm generally exhibit

significantly higher circulation time due to decrease in opsonization of liposomes with serum

protein (Lian and Ho, 2001).

42

Amphiphilic nature of phospholipids allows these molecules to form lipid bilayers.

This unique feature is utilized for the preparation of liposomes. In general, hydration of

phospholipids results into formation of MLVs, which can be processed into SUVs with

proper sonication. However, addition of aqueous solution of surfactant above the critical

micelle concentration results in the formation of phospholipids micelles. After the dialysis of

surfactant aggregation of micelles form LUVs Critical micelle concentrations of amphiphiles

which can form micelles are four to five orders of magnitude higher than the phospholipids

which form liposomes (Jesorka and Orwar, 2008). Numerous methods have been reported to

prepare liposomes. Most commonly solvent evaporation method, reverse phase evaporation

method and detergent dialysis method are employed. (Niesman, 1992). The encapsulated

drug from liposome can be released either through passive diffusion, vesicle erosion or

vesicle retention. In passive diffusion, drug molecules tend to penetrate through the lipid

layers of liposome to reach extra vesicular layer either by diffusion or convection

mechanism. The rate of diffusion depends on the size, lipid composition, and the properties

of the drug itself (Guy et al., 1983; Papahadjopoulos et al., 1973; Tsukada K et al., 1984).

Unilamellar liposomes exhibit faster release rate than multilamellar ones because in multi

layered liposome, drug diffusion occurs through a series of barriers hence the drug release is

delayed. Phospholipase and high density lipoprotein present in blood plasma can damage

phospholipid layers of liposome and thus results in vesicle erosion and releases the

encapsulated drug into the cell. The drug release rate depends on the extent of liposomal

membrane damage (Zeng L and X, 2010). Liposome-cell interactions depend on several

factors like size, surface charge, composition of liposomes, targeting ligand on the surface of

liposome and biological environment. Liposomes can interact with cells by four different

43

mechanisms: adsorption, fusion, lipid exchange and endocytosis (receptor mediated).

Liposomes can be specifically or nonspecifically adsorbed onto the cell surface or can be

fused with cell membranes, and release encapsulated drug inside the cell. During adsorption,

liposomes can release encapsulated drug in front of cell membrane and released drug can

enter cell via micropinocytosis. They can also be engulfed inside the cell by specific or

nonspecific endocytosis process. Negatively charged liposomes have been found to be more

efficient than neutral liposomes for internalization into the cells by endocytosis process.

Liposomes bind to the receptor present in the invaginations of cellular membrane and are

internalized into the cell by endocytotic pathway. After endocytosis, they can fuse with the

endosomal membrane to form endosome which can be delivered to lysosomes. In lysosomes,

the presence of peptidase and hydrolase degrade the liposomes and their content. To avoid

this degradation and thus to increase cytoplasmic bioavailability stimuli responsive

liposomes (such as pH or temperature) have been developed. pH sensitive liposomes can

undergo fusion with endosomal membrane and release their content directly into cytosol. In

some cases liposomes become destabilized inside the endosome and release their content or

they destabilize endosomal membrane resulting in leakage of encapsulated content into

cytosol (Bonacucina et al., 2009; Torchilin, 2005).

Application of liposomes in ophthalmic drug delivery

Liposomes have been investigated for ophthalmic drug delivery since it offers

advantages as a carrier system. It is a biodegradable and biocompatible nanocarrier. It can

enhance the permeation of poorly absorbed drug molecules by binding to the corneal surface

and improving residence time. It can encapsulate both hydrophilic and hydrophobic drug

molecules. In addition, liposomes can improve pharmacokinetic profile, enhance therapeutic

44

effect and reduce toxicity associated with higher dose. Owing to their versatile nature,

liposomes have been widely investigated for the treatment of both anterior and posterior

segment eye disorders. Current approaches for the anterior segment drug delivery are focused

on improving corneal adhesion and permeation by incorporating various bioadhesive and

penetration enhancing polymers. However, in the case of posterior segment disorders

improvement of intravitreal half life and targeted drug delivery to the retina are necessary.

Currently verteporfin is being used clinically in photodynamic therapy for the treatment of

subfoveal choroidal neovascularization (CNV), ocular histoplasmosis, or pathological

myopia effectively. Verteporfin is a light-activated drug which is administered by

intravenous infusion. In photodynamic therapy, after the drug is injected, a low energy laser

is applied to the retina through the contact lens in order to activate verteporfin that results in

closure of the abnormal blood vessels. Unfortunately, photodynamic therapy usually does not

permanently close the abnormal vessels and choroidal neovessels reappear after several

months. Another liposomal photosensitizing agent, rostaporfin, was evaluated for the

treatment of age related macular degeneration. It is now under Phase 3 clinical trial.

Rostaporfin requires less frequent administration compared to verteporfin. Liposome

technology has been explored for ophthalmic drug delivery. However, there are some issues

to be addressed such as formulation and storage of liposomes is very difficult and they are

known to cause long term side effects. Intravitreal administration of liposomes has resulted in

vitreal condensation, vitreal bodies in the lower part of eye, and retinal abnormalities.

Therefore all these factors should be taken into account while developing liposomal

formulation for ophthalmic application (2004; Chakravarthy et al., 2006; Lazzeri et al., 2011;

45

Ruiz-Moreno et al., 2006; Sachdeva et al., 2010).Recent applications of liposomal

formulations encapsulating various therapeutic molecules are summarized in table 1.4

46

Table 1.4 Application of liposomes for the delivery of various drug molecules [adapted with

permission from ref (Mishra et al., 2011a)]

Drug Formulation Result Ref

GCV Liposomes

In vitro transcorneal permeation and in vivo ocular

pharmacokinetics was improved

(Shen and

Tu, 2007)

Ciprofloxacin Liposomal

hydrogel

Fivefold higher transcorneal permeation than the

liposomes alone

(Hosny,

2010)

Levofloxacin

Liposomes

attached to

the contact

lense

Drug was released following first order kinetics for

more than 6 days and formulation had showed

activity against S. aureus

(Danion et

al., 2007)

Herpes simplex

virus antigens

Periocular

vaccine

Treated rabbits showed anti-gB immune response

and protected against reactivation of HSV infection

(Cortesi et

al., 2006)

Acetazolamide

Neutral and

surface

charged

liposomes

Positively charged liposomes reduced IOP and

ehhibited prolonged effect than negatively charged

liposomes

(Hathout et

al., 2007)

Tacrolimus Liposomes

more than 50 ng/ ml vitreous concentration was

maintained for 2 weeks and reduced drug related

toxicity

(Zhang et

al., 2010)

Vasoactive

intestinal

peptide

Rhodamine

conjugated

liposomes

Liposomes were internalized by retinal Müller glial

cells, resident macrophages

Majority of the liposomes reached the cervical

lymph nodes and resulted in slower release and

long term expression inside the eye

(Camelo et

al., 2007)

Clodronate Liposomes Effectively inhibit infiltration of ED2-positive

macrophages

(Fukushima

et al., 2005)

Plasmid DNA Cationic

liposomes

Significantly increased transfection efficiency of

pDNA

(Kawakami

et al., 2004)

Therapeutic

DNA

Cationic

lipoplexes

Achieved good vitreous mobility with moderately

pegylated cationic lipoplexes with size less than

500 nm

(Peeters et

al., 2005)

47

Niosomes

Niosomes are non-ionic surfactant vesicles physically similar to liposomes. This

vesicular system is formed by hydration of cholesterol and a single-alkyl chain, non-ionic

surfactant. Niosomes are preferred in topical ocular drug delivery because they are

chemically stable, biodegradable, biocompatible, non-immunogenic and have low toxicity

because of non-ionic nature. They can entrap both hydrophobic and hydrophilic drugs.

Niosomes do not require special handling techniques and show flexibility in structural

characterization such as size, composition and fluidity. Discomes are a modified form of

niosomes with higher amount of non-ionic surfactant. These carrier because of their shape

and larger size (12-60 µm) do not wash out quickly form the cul-de-sac. It has the capacity of

encapsulating a higher amount of hydrophilic drug than niosomes (Mainardes et al., 2005).

Many researchers reviewed significance of niosomes for glaucoma therapy. Niosomal

formulation of timolol meleate showed improved bioavailability and better efficacy for intra

ocular pressure lowering effect as a result of controlled release of drug from nisomes

(Aggarwal and Kaur, 2005; Kaur et al., 2010).

Hydrogel

Hydrogel is a hydrophilc homopolymer or copolymer network capable of imbibing

large amounts of water or biological fluids (Klouda and Mikos, 2008). It forms a three-

dimensional cross linked network by chemical (covalent bonds) and physical (secondary

forces, crystalline formation or chain entanglements) forces (Qiu and Park, 2001). Hydrogels

are mainly divided in to two groups, preformed gels and stimuli responsive hydrogels.

Preformed hydrogels are highly viscous in nature, which do not undergo any change after

administration. A stimuli responsive hydrogel which exists as an aqueous solution or

48

suspension before administration and undergo gelation in response to the physiological

environment is known as in-situ forming hydrogel (He et al., 2008). Stimuli-sensitive

hydrogels are called smart hydrogels. They undergo a reversible phase-transition or a sol-gel

transition in response to physical stimuli such as temperature, pressure, light, electromagnetic

radiation, sound or chemical stimuli such as pH and ionic strength (Nanjawade et al., 2007;

Qiu and Park, 2001). Solution nature allows them to incorporate various therapeutic agents

by simple physical mixing (Ruel-Gariepy and Leroux, 2004). These in-situ forming

hydrogels offer several advantages for application to ophthalmic drug delivery. For example

they can apply as eye drops, which undergo phase-transition in the ocular cul-de-sac to form

a viscoelastic gel and provide prolonged contact time of drugs on the eye surface. In addition,

they can be injected into the eye in a less invasive manner before they solidify (Nanjawade et

al., 2007). Hydrogel based formulations that do not create any problem with vision and

reduce the dosing frequency. Hydrogel can be useful for the delivery of different therapeutic

agents ranging from small hydrophilic and hydrophobic drugs to the gene and protein

molecules. It can be tailor-made for specific use (Klouda and Mikos, 2008). Many

investigators have evaluated the significance of hydrogel based ocular drug delivery systems.

For example, PEG hydrogel was evaluated for controlled delivery of pilocarpine. The in situ

gel forming solution was applied topically, which was able to resist the shear forces in cul-

de-sac. The hydrogel provided sustained drug release and shown prolonged pharmacological

response in comparison to eye drops (Anumolu et al., 2009). Rapid sol-gel transition,

transparent nature, biocompatibility and low toxicity have attracted pharmaceutical scientists

towards the utilization of biodegradable block polymers in ocular drug delivery. Recently,

PLGA-PEG-PLGA block polymer was utilized as thermosensitive in situ gel forming

49

solution for dexamethasone acetate delivery. Significantly enhanced therapeutic

concentration of DXA was achieved in the anterior chamber of rabbit cornea in comparison

the eye drops (Gao et al., 2010).

50

CHAPTER 2

INTRODUCTION

Statement of problem

Topical administration is the preferred and most convenient route when it comes to

treating eye disorders. However, the ocular bioavailability of most of the topically applied

drugs is less than 1% (Macha S et al., 2003). Poor bioavailability mainly results from

precorneal factors such as blinking, transient residence time in cul-de-sac, and nasolacrimal

drainage. In addition, the lipoidal nature of the corneal epithelium restricts the entry of

hydrophilic drug molecules and the water- laden stroma acts as a rate limiting membrane for

lipophilic molecules. Moreover, physicochemical properties of a drug entity itself determine

the diffusion resistance and relative impermeability offered by various ocular tissues (Ghate

and Edelhauser, 2008; Mitra, 2009; Watsky et al., 1988). Also, Multidrug efflux pumps such

as the P-glycoprotein (P-gp) and multi-drug resistance protein (MRP) have been reported to

limit the in vivo ocular bioavailability of topically applied drugs (Dey et al., 2003; Karla et

al., 2007a; Karla et al., 2007b).

Glaucoma is a progressive optic nerve disease often associated with elevated

intraocular pressure and characterized by optic disc cupping and visual field loss. Glaucoma

is the second most common cause of blindness in the United States affecting approximately

2.5 million people (Quigley and Broman, 2006). Almost 50% of the patients affected by this

disease are unaware because they do not experience symptoms and therefore do not seek

treatment (Hoyng and Van beek, 2000). Early diagnosis and treatment play a vital role in

halting the progression of this vision threatening condition. A major cause for glaucoma is

the relative obstruction to the outflow of aqueous humor from the anterior segment, which

51

leads to increased intraocular pressure. The goal of glaucoma therapy is to preserve vision by

reducing intraocular pressure to a level thought to be safe for the optic nerve, while

preserving the patient’s quality of life. This can involve any or all of the following: (i) topical

ocular therapy (ii) laser therapy (iii) surgical therapy. Currently, the only non invasive mode

of treating glaucoma is with eye drops taken on a daily basis. Amongst the various β- blocker

drugs used for glaucoma, direct clinical comparisons have demonstrated that timolol is the

most efficacious. Timolol acts by decreasing aqueous humor production from the cilliary

body. Timolol is available as marketed eye-drops formulations (example- TIMOPTIC).

However, the effectiveness of eye-drops is less than 1% and frequent administration is

required to maintain the therapeutic levels of drug. Glaucoma generally affects older people

and taking eye drops twice a day is difficult for these patients. Administering frequent eye

drops plays a major role in non-compliance with patients (Guttman, 2005). Poor compliance

with medications is a important reason for vision loss in glaucoma patients. There is a need

for a better delivery strategy which can provide constant levels of drug over a long period of

time and thus helps in reducing the frequent application of eye drops.

Our hypothesis is to develop novel pentablock polymers, based on optimization of

both diffusion and degradation mediated drug release from polymer matrix. These polymers

will be comprised of multi-polymer blocks such as polyethylene glycol (PEG),

polycaprolactone (PCL) and polylactide (PLA) which are FDA approved for human use. In

comparison to PLGA, these novel polymers will release lesser amount of lactic acid upon in

vivo degradation due to low molecular weight of PLA block. Existing PCL or PLA based

block polymers primarily cause diffusion mediated drug release due to extremely slow

degradation (Gou et al., 2009a). So our approach may lead to development of novel polymers

52

with optimum degradation rate to achieve constant (zero order) drug release via both

diffusion and degradation pathway. These polymers will be utilized for the development of

drug delivery systems such as of nanoparticles embedded in thermosensitive gel in order to

attain long term constant release over a period of one week. Pentablock polymers with

different block ratios of PEG and PLA will be prepared and utilized for the nanoparticle

preparation. These polymers have similar composite molecular structures with different

molecular weights and block ratios. Pentablock polymers based thermosensitive gel

formulation will remain in liquid aqueous state at room temperature (25 °C) and will form a

thin transparent film on the surface of cornea upon contact with eye temperature i.e. 34 °C.

The rationale behind developing novel biodegradable and biocompatible pentablock based

biomaterial is to achieve controlled drug delivery over a period of several days from a single

drop. Our dual approach of nanoparticles suspended in a thermogelling system could

minimize burst release due to longer diffusion pathway of entrapped molecules from the

system. Such a novel technology upon instillation as an eye drop will result in a prolonged

duration of action of the drug and thereby eliminates the need for repeated administrations.

Objectives

Therefore objective of this research are as follows:

1. To prepare nanoparticle formulation of model hydrophobic drug using PEG-PCL-

PEG based triblock copolymers with different molecular weight of PEG

This objective will be studied by synthesizing triblock copolymers with PEG of molecular

weight 2000-8000. In the next step we will characterize the copolymer compositions for

molecular weight, and crystallinity using 1HNMR, GPC and XRD. We will evaluate the

effect of copolymer compositions on release profile of triamcinolone acetonide from

53

nanoparticles. Nanoparticles will be characterized for size, surface morphology, drug loading

and in vitro release kinetics. Nanoparticles size, surface charge and morphology will be

characterized by DLS, Zetasizer and SEM. In-vitro release studies will be performed by

dialysis method and amount of drug release at different time intervals will be measured by

HPLC analysis.

2. To design and evaluate Pentablock polymers for controlled drug delivery and

evaluate the nanoparticle formulation of various steroids.

This objective will be studied by synthesizing novel PLA based pentablock polymers of

different molecular weight and block ratios. The polymers will be optimized by changing the

block ratio of PEG, PCL and PLA for the preparation of nanoparticles. The objective behind

changing the block length is to achieve near zero-order drug release kinetics from the

nanoparticles. Pentablock copolymers will be characterized for molecular weight, molecular

weight distribution, crystallinity and biocompatibility of the polymers. Triamcinolone

acetonide loaded nanoparticles utilizing different pentablck polymers will be prepared and

characterized for degradation, drug release profile, size, surface morphology and charge.

Nanoparticles encapsulating different steroids will be prepared from optimized pentablock

polymer. Drug release from nanoparticles alone and nanoparticles suspended in

thermosensitive gel are performed by standard dialysis method in phosphate buffer saline

(pH 7.4) at 37 °C.

3. To develop novel composite strategy for the treatment of glaucoma

Timolol loaded nanoparticles will be prepared using different polymer compositions and

characterized for surface morphology, charge, particle size and drug entrapment efficiency.

Different pentablock copolymers synthesized in objective 2 will be utilized for preparation of

54

timolol loaded nanoparticles. Timolol nanoparticles will be characterized for drug release

profile, size and drug entrapment efficiency. In vitro release kinetics from nanoparticle alone

and suspended in thermosensitive gel will be evaluated. Two different compositions of

thermosensitive gel will be utilized to optimize drug release profile from composite

formulations. In vitro cell cytotoxicity and cell viability will be evaluated in various ocular

cell-lines.

4. To develop controlled release formulations of macromolecules using thermosensitive

hydrogel.

Drug release kinetics of various large molecules of different molecular weight from

thermosensitive hydrogel will be evaluated. FITC-Dextran of different molecular weight will

be screened to evaluate effect of drug molecular size on release profile from thermosensitive

hydrogel. Release profile of lysozyme (model protein molecule) will be evaluated from

different compositions of thermosensitive hydrogels. Physical and chemical stability of

lysozyme in released samples will be determined by CD. Biological activity of the protein

molecule will be determined by enzyme activity assay.

55

CHAPTER 3

NOVEL PENTABLOCK COPOLYMER (PLA-PCL-PEG-PCL-PLA)

BASED NANOPARTICLES FOR CONTROLLED DRUG DELIVERY: EFFECT OF

COPOLYMER COMPOSITIONS ON POLYMER CRYSTALLINITY AND DRUG

RELEASE PROFILE FROM NANOPARTICLES

Rationale

Chronic posterior segment diseases such as diabetic retinopathy, age-related macular

degeneration (AMD), macular edema and retinovascular diseases require sustained levels of

corticosteroids (Jonas et al., 2002). Direct intravitreal injections provide therapeutic levels in

ocular tissues and avoid systemic toxicity (Tamura et al., 2005; Young et al., 2001).

However, high bolus dose and repeated intravitreal injections are required to maintain

therapeutic concentrations. The common side effects associated with frequent high

dose intravitreal injections are development of endophthalmitis, blurred vision, increase in

intraocular pressure, cataract formation and increased risk of retinal detachment (van Kooij et

al., 2006). Moreover, higher doses of steroids may cause retinal toxicity (Kwak and

D'Amico, 1992). To overcome these problems and to avoid direct tissue exposure of high

concentration of steroids, encapsulation of the drug molecules in polymeric nanoparticles

may be an ideal strategy. In this regard, biodegradable polymeric nanoparticles hold

significant promise in the treatment of chronic ocular diseases. Unlike implants

biodegradable polymeric nanoparticles do not require surgical removal and thus may avoid

autoimmune responses and disorganization of ocular tissues. In the last decade, numerous

investigators evaluated the significance of nanoparticles for ocular drug delivery (Araujo et

al., 2009; Cai et al., 2008; Diebold and Calonge, 2010). Biodegradable nanoparticles with

56

appropriate size and narrow size distribution can provide adequate ocular bioavailability

(Sahoo et al., 2008). Biodegradable polymers such as poly (DL-glycolide-co-lactide)

(PLGA) (Gupta et al., 2010), poly (lactide) (PLA) (Liang et al., 2005) and poly

(caprolactone) (PCL) (Marchal-Heussler et al., 1993) are widely studied for the preparation

of nanoparticles. Particularly, micelle- like nanoparticles having amphiphillic block with poly

ethylene glycol (PEG) such as PLA-PEG (Venkatraman et al., 2005), PCL-PEG (Li et al.,

2009), PCL-PEG-PLA (Ahmed and Discher, 2004), PLGA-PEG (Dalwadi and Sunderland,

2009), and PCL-PEG-PCL have gained attention in ocular drug delivery. In these block

copolymers PEG block forms the outer shell of nanoparticles whereas PCL or PLA due their

hydrophobic nature forms nanoparticle core. PEG is well known for its non-antigenicity and

non immunogenicity (Hu et al., 2003; Hu et al., 2007). Because of its hydrophilic nature it

facilitates the diffusion of water into nanoparticles matrix and provides diffusion mediated

drug release from nanoparticles. PCL is hydrophobic biodegradable polyester which enables

high permeability for small molecules (Ghoroghchian et al., 2006). Being hydrophobic in

nature it provides good encapsulation efficiency to lipophilic drugs via hydrophobic

interactions. However, it exhibits very slow degradation because of its hydrophobic and

crystalline nature. Existing PCL and PEG based di or triblock copolymers such as PEG-PCL

or PCL-PEG-PCL have limitations of burst release from nanoparticles due to crystallinity of

PCL block. Nanoparticles prepared from PCL based block polymers primarily exhibit

diffusion mediated drug release due to extremely slow degradation of PCL (Frank et al.,

2005; Mishra et al., 2010). In addition, hydrophobic drugs often are trapped in the

hydrophobic core of the nanoparticles and achieve no or limited release from nanoparticles.

Previous reports suggest that high molecular weight PCL needs more than 1 year for

57

complete degradation in vitro (Lam et al., 2008). Therefore, there is a need of controlled

delivery system that releases drugs via both diffusion and degradation mechanism.

Crystallinity and slow degradation of PCL can be improved by conjugating with relatively

faster degrading PLA. Block ratio of PCL to PLA can be adjusted to optimize release of

hydrophobic drugs via both diffusion and degradation mediated pathway. In vitro drug

release may be further optimized by adjusting the molecular weight of each polymeric block.

Considering these facts we have developed novel pentablock copolymers (PLA-PCL-PEG-

PCL-PLA) by conjugating faster degrading hydrophobic segment (PLA) to the triblock

copolymer (PCL-PLA-PCL) to achieve near zero order Triamcinolone acetonide release from

nanoparticles.

Triamcinolone acetonide was selected as a model hydrophobic drug to evaluate the

potential of PLA-PCL-PEG-PCL-PLA based nanoparticles as carrier of hydrophobic

molecules. Triamcinolone acetonide is a water insoluble synthetic corticosteroid with

longer intravitreal half-life compared to other steroids (Chin et al., 2005). It is very effective

and being used off-label for the treatment of many sight threatening ocular disorders such as

AMD and vitreoratinopathy (Jermak et al., 2007; Jonas et al., 2002). Marketed formulations

of triamcinolone acetonide are available in the form of injectable suspension and require

repeated administrations to maintain therapeutic levels at the target sites. (Jermak et al.,

2007; Jonas et al., 2002). High dose intravitreal injections of triamcinolone acetonide have

many injection-related complications as described above. In addition, drug release rate

cannot be controlled adequately from the suspension. Therefore, there is a need for sustained-

release drug delivery system for lipophilic drugs such as triamcinolone acetonide, which can

58

provide therapeutic levels in ocular tissues over longer periods and avoid repeated

administrations (Okabe et al., 2003).

Hence, the aim of this study was to develop triamcinolone acetonide loaded

nanoparticles utilizing novel pentablock copolymers (PLA-PEG-PCL-PCL-PLA) and to

investigate the effect of polymeric compositions on drug release profile from nanoparticles.

Pentablock copolymers were characterized by 1HMNR, GPC, FT-IR, XRD and DSC.

Nanoparticles were characterized for size, surface morphology and surface charge. Change in

crystallinity of copolymers with change in ratio of PCL/PLA was discussed. Moreover, effect

of copolymer compositions on triamcinolone release profile was discussed.

Materials and Methods

Material

Triamcinolone acetonide, poly (ethylene glycol) (PEG, Mw: 2 kDa), ε- caprolactone,

stannous octoate, poly (vinyl alcohol) (PVA) were purchased from sigma aldarich company

(St. Louis, MO; USA). D, L lactide and L-lactide were purchased from Acros organics

(Morris Plains, NJ; USA). All other chemicals purchased were of analytical grade and used

as received.

Methods

Synthesis of Triblock copolymers

Triblock copolymers (PCL-PEG-PCL) were synthesized by ring opening polymerization

of ε-caprolactone as reported in the literature. PEG of Mw 2000-8000 was selected as

initiator and stannous octoate was added as catalyst (Gou et al., 2009a). After purification

predetermined amount of PCL-PEG-PCL was used as initiator for the synthesis of D, L-

lactide or L-lactide containing pentablock copolymers.

59

Synthesis of pentablock copolymers

Briefly, for synthesis of triblock copolymer (P-1) PEG (Mw 2000) was dried under

vacuum for 3 h before copolymerization. Calculated amount of PEG (0.001 mol), ε-

Caprolactone (0.001 mol) and stannous octoate (0.5 wt %) were added in the round bottom

flask and degassed for 30 minutes. Then the flask was purged with nitrogen and the reaction

was performed for 24 h at 130 °C. The resulting crude product was dissolved in methylene

chloride and precipitated with cold petroleum ether to remove un-reacted monomers. The

precipitated polymer was filtered and vacuum dried for 24 h.

For the synthesis of pentablock copolymer (P-4), purified PCL-PEG-PCL (P-1) was

utilized as initiator for copolymerization with L-lactide. Triblock copolymer and L-lactide

(0.001 mol) monomer were added in round bottom flask and stannous octoate (0.5 wt %) was

added as a catalyst. Then the flask was purged with nitrogen and reaction was performed for

24 h at 130 °C. The final product was again purified by dissolving in methylene chloride

followed by precipitation with cold petroleum ether and the precipitate was vacuum dried for

24 h. Same procedure was followed for the synthesis of different compositions of

pentablockcopolymer.

Characterization of copolymers

NMR

1H NMR spectroscopy was performed to characterize copolymer compositions. Spectra

were recorded by dissolving polymeric material in deuterated chloroform (CDCl3) and then

analyzed the proton NMR spectra recorded with a Varian-400 NMR instrument.

60

Gel Permeation Chromatography (GPC) analysis

GPC analysis was performed with the Shimadzu refractive index detector to

determine molecular weight and its distribution. 1.5 mg polymeric material was dissolved in

1.5 ml Tetrahydrofuran (THF). Polystyrenes standards of different molecular weights were

utilized as reference. THF was used as eluting solvent at a flow rate of 1 ml/min and Styragel

HR-3 column maintained at 35 °C was utilized for separation.

Fourier transform infrared spectroscopy (FTIR)

Fourier transform infrared spectroscopy (FT-IR) spectra were recorded with a

Nicolet-100 infrared spectrophotometer at a resolution of 4 sec-1

. Polymer was dissolved in

methylene chloride and casted on KBr plates.

X-ray diffraction analysis of copolymers

X-ray diffraction (XRD) analysis was performed for tri block and pentablock

copolymers, by MiniFlex automated X-ray diffractometer (Rigaku, The Woodlands, Texas)

with Ni-filtered Cu-kα radiation (30 kV and 15 mA) at room temperature.

DSC analysis of copolymers

Differential scanning calorimetry (DSC) analysis was performed for tri and

pentablock copolymers by Perkin Elmer, Diamond DSC thermal analyzer. Approximately 5

mg sample mass of each copolymer was taken. DSC heating and cooling run were performed

at the rate of 5 °C/min. Samples were heated till 100 °C and kept for 2 min at that

temperature. Then samples were cooled down to 10 °C and followed by a second heating till

100 °C.

61

Preparation of blank and drug loaded nanoparticles

Triamcinolone acetonide loaded as well as blank nanoparticles were prepared as per a

previous reported method with minor modifications (Li et al., 2009). Concisely, single o/w

emulsion solvent evaporation method was used to prepare triblock and pentablock

nanoparticles. Copolymer and drug in the ratio of 1:10 were dissolved in 1 ml methylene

chloride at room temperature. The organic phase containing drug and copolymer was

emulsified in an aqueous solution of 2 wt. % PVA by sonication for 5 minutes at 65W to

obtain o/w emulsion. The organic solvent was evaporated by continuous stirring for 2 h to

generate nanoparticles. The resulting nanoparticle suspension was then centrifuged at 21,000

rpm, 4 °C for 1h. The prepared nanoparticles were washed three times with distilled water to

remove un-entrapped drug and excess PVA. Blank nanoparticles were prepared by the same

method without drug incorporation in organic phase. Prepared nanoparticle suspensions were

lyophilized with 5 wt % mannitol (act as lyoprotectant) for 24h and stored at 4 °C for further

studies.

Characterization of nanoparticles

Size and zeta potential determination of nanoparticles

The mean diameter and size distribution of nanoparticles were determined by

dynamic light scattering (DLS) with a 90 Plus Particle Size Analyzer (Brookhaven

Instruments Corporation). Nanoparticle suspension was diluted with distilled water and

analysis was performed over 3 minutes at 25 °C and 90° scattering angle. All measurements

were made in triplicate and average particle size and size distribution data were reported ±

SD.

62

Zeta potential of nanoparticles was determined by Malvern Zeta sizer (Nano-ZS, Malvern

Instrument, UK). All analyses were performed at 25 °C. Values were reported as mean value

± SD of 3 test runs.

Drug loading and entrapment efficiency

Drug loaded freeze dried nanoparticles were employed for the determination of drug

loading and entrapment efficiency. Briefly, 5 mg of nanoparticles were dissolved in 0.5 ml of

dichloromethane and diluted with deionized water. Amount of drug extracted in water was

analyzed by high performance liquid chromatography (HPLC) instrument (Shimadzu, Japan)

at a wavelength of 240 nm. The mobile phase containing acetonitrile / water (40/60) mixture

was used as eluent solvent and C18 reverse phase column was utilized for separation

(Havlikova et al., 2008). The total amount of drug in nanoparticles was calculated and drug

loading and entrapment efficiency were determined by the following equations 1 and 2:

Drug loading =

Entrapment Efficiency =

Surface morphology study

The surface morphology of lyophilized nanoparticles was observed by scanning

electron microscopy (SEM). Lyophilized nanoparticles were coated with gold/palladium at

0.6 kV. Samples were examined by FEG ESEM XL 30 electron microscope.

In vitro release studies of TA-loaded nanoparticles

Drug loaded nanoparticles were suspended in 500 µl 10mM phosphate buffer saline

(PBS) of pH 7.4 and this nanosuspension was placed in dialysis bag with molecular weight

cutoff of 6500 Da. (Sigma Aldrich, MO; USA). The dialysis bag was immersed in tube

containing 10 ml of 10 mM PBS (pH 7.4) at 37 °C with continuous shaking (60 rpm). At pre-

63

determined time intervals samples were withdrawn and the entire release medium was

replaced with fresh buffer to maintain sink conditions. The amount of triamcinolone

acetonide released at each time point was determined by HPLC analysis. The experiments

were repeated 3 times and mean value ± SD was expressed as cumulative % drug released

with time.

Drug release kinetics

Drug release parameters were computed by two different methods utilizing Higuchi and

Korsmeyer equations. Drug release mechanism was evaluated by model equations as

described below (Mishra et al., 2010).

Higuchi equation:

Mt = KHt1/2

KH indicates the Higuchi release rate constant obtained by plotting cumulative percent drug

released against the square root of time.

Korsmeyer-peppas equation:

Mt/ M∞ = ktn

Mt and M∞ denote the cumulative amount of drug released at time t and at the equilibrium,

respectively. The constant k represents the kinetic constant and n is the release exponent that

indicates the drug release mechanism. Values of n< 0.5 indicates Fickian (ideal) diffusion

mechanism and values of 0.5<n<1.0 suggest non-Fickian diffusion. When value of n is

greater than 1.0, it represents case II transport or zero order release kinetics. Drug release

data is employed to obtain the release exponent for Mt/ M∞ ≤ 0.6 (Mishra et al., 2010).

64

Statistical analysis

All release studies were performed in triplicate. The results were reported as mean ±

standard deviation. Statistical analysis of the effect of copolymer compositions on

triamciolone acetonide release profile from nanoparticles were compared by one-way

ANOVA. Statistical package for social science (SPSS) version 11 was applied to compare

mean of each group. A level of P < 0.05 was considered statistically significant in all cases.

Results and discussion

Polymer synthesis and characterization

Pentablock copolymers were synthesized by sequential ring opening polymerization of ε-

caprolactone and L-lactide or D, L-lactide. In the first step triblock copolymers with different

ratios of PEG/PCL were synthesized by ring opening polymerization of ε-caprolactone in the

presence of PEG and a small amount of stannous octoate. In the second step these tribolck

copolymers served as initiators for the synthesis of pentablock (PLA-PCL-PEG-PCL-PLA)

copolymers by ring opening polymerization of D, L-lactide or L-lactide (Fig. 3.1). Different

compositions of pentablock copolymers were synthesized by changing the molar ratios of

PEG/PCL/PLA, as listed in Table 3.1. We attempted to modify the properties of copolymers

such as molecular weight and crystallinity by changing the hydrophobic segment of

amphiphillic block copolymers. In this study, we want to evaluate the effect of hydrophobic

segment of copolymer on the release profile of a hydrophobic drug. Pentablock copolymers

are composed of relatively low molecular weight PLA in comparison to PLGA polymers.

PLGA polymers are conventionally utilized for the preparation of nanoparticles and known

for producing high concentration of lactic and glycolic acid after degradation that may cause

irritation to the local tissue or degradation of entrapped molecules (Meyer et al., 2012). In

65

contrast, due to low molecular weight of PLA, pentablock copolymers will produce low

amounts of lactic or glycolic acids upon degradation and may not cause tissue irritation at the

target site and are advantageous in the preparation of controlled release formulations.

Moreover, tailor-made drug release profile can be obtained by changing the compositions of

pentablock copolymers.

66

Sn(Oct)2

130 oC

O

O

O

O

C OH2C C

O

O CH2CH2O C

O

CH2 O C5 5Y XX

HC

O

O HZ

HCOH

O

Z

Lactide

CH3 CH3

PLA-PCL-PEG-PCL-PLA

Y

O

O

Sn(Oct)2

130 oC

H OH2C C

O

O CH2CH2O C

O

CH2 O H5

5Y XX

PEG Caprolactone PCL-PEG-PCL

H OCH2CH2 OH

H OH2C C

O

O CH2CH2O C

O

CH2 O H5

5Y XX

PCL-PEG-PCL

Figure 3.1 Synthetic scheme of pentablock copolymer (PLA-PCL-PEG-PCL-PLA)

67

PEG of molecular weight 2000 with two hydroxyl terminals was utilized for the

synthesis of tri and pentablock copolymers. Compositions of copolymers, their molecular

weight and hydrophobicity index are summarized in Table 3.1 and 3.2. 1HNMR spectrum of

PCL-PEG-PCL (Fig. 3.1) and PLLA-PCL-PEG-PCL-PLLA (Fig. 3.2) in CDCl3 are

described in respective figures. Typical signals of PEG, PCL and PLA components was

utilized to calculate the molecular weight of copolymers. Signal at 3.65 ppm (–CH2–) was

assigned to PEG block. Signals at 1.28, 1.6, 2.3 and 4.09 ppm were assigned to different

methylene protons (–CH2–) of PCL blocks and 1.4 (–CH3) and 5.19 ppm (–CH) were

assigned to PLA block. The molar ratio of PEG/PCL/PLA was determined by integrating

peak intensities of methylene protons from PEG block at 3.65 ppm, PCL block at 4.09 ppm,

and to PLA block at 5.10 ppm. The number average molecular weight (Mn) of copolymers

were calculated as per the reported equations (Huang et al., 2003). The Mn and Mw (weight

average molecular weight) values obtained by GPC analysis are summarized in Table 3.1 and

3.2. The Mn values of copolymers determined by GPC analysis were lower than the Mn

values calculated from 1HNMR analysis. This result was attributed to the change in

hydrodynamic volume of block copolymers as compared with parent homopolymers (Huang

et al., 2003).

68

Table 3.1 Characterization of triblock polymers

Code Copolymers Mna Mn

b Mw

b PDI

b

A PCL5000-PEG2000-

PCL5000 12450 8975 11630 1.50

B PCL5000-PEG4000-

PCL5000 14099 9795 13115 1.33

C PCL5000-PEG8000-

PCL5000 18027 14652 18068 1.23

a. Values calculated from 1HNMR spectra

b. Values obtained by GPC analysis

69

Table 3.2 Characterization of pentablock copolymers

Code Copolymers PEG/PCL/PLA

ratio Mn

a Mn

b Mw

b PDI

b

P-1 PCL-PEG-PCL 1/2.5 7105 4083 6080 1.63

P-2 PCL-PEG-PCL 1/5 12450 8975 11630 1.50

P-3 PLLA-PCL-PEG-

PCL-PLLA 1/5/2.5 19072 10964 16260 1.48

P-4 PLLA-PCL-PEG-

PCL-PLLA 1/2.5/2.5 11062 6186 9526 1.54

P-5 PDLLA-PCL-

PEG-PCL-PDLLA 1/5/2.5 18803 11739 16246 1.38

P-6 PDLLA-PCL-

PEG-PCL-PDLLA 1/2.5/2.5 11665 6165 10160 1.64

a. Values calculated from 1HNMR spectra

b. Values obtained by GPC analysis

70

Figure 3.2 1H-NMR spectra of PCL-PEG-PCL copolymer in CDCl3

71

Figure 3.3 1H-NMR spectra of PLA-PCL-PEG-PCL-PLA copolymer in CDCl3

72

Figure 3.4 depicts FT-IR spectra of P-2, P-3 and P-5 block copolymers. On the spectrum of

PCL-PEG-PCL (P-2) (Fig. 3.4A), band for C=O stretching appeared at 1732 cm-1

and

bands for C-H stretching appeared at 2941 cm-1

and 2860 cm-1

for PCL block. Absorption

band at 1140 cm-1

appeared because of C-O-C stretching vibrations of the repeated

OCH2CH2 units of PEG and band at 1279 cm-1

was attributed to the -COO- stretching

vibrations (Li et al., 2009). On the P-3 (Fig. 3.4B) and P-5 (Fig. 3.4C) spectra another C=O

stretching band was observed at 1757 cm-1

, which can be attributed to PLA block.

73

Figure 3.4 FTIR spectra of copolymers (A) P-2 (B) P-3 (C) P-5

C

B

A

74

To improve the crystallinity and degradation of PCL based triblock copolymers, PLA block

was introduced in the copolymers. Earlier reports suggest that by changing the

hydrophobic/hydrophilic constituents, crystallinity and degradation of the copolymers can be

modulated (Frank et al., 2005; Gou et al., 2010; Huang et al., 2004; Li et al., 2005; Li et al.,

2002). We synthesized four compositions of pentablock copolymers with different ratios of

PCL/PLA in the hydrophobic segments. We also evaluated the effect of different isomeric

forms of PLA block on the crystallinity of copolymers. L-lactide is semicrystalline in nature

where as D, L-lactide has amorphous structure. Both polymers may act differently in

reducing the crystallinity of PCL block, which eventually may affect drug release profile.

The X-ray diffractograms obtained for different triblock and pentablock copolymer

compositions are shown in figure 4. For triblock copolymer (P-1) (Fig. 3.5A), diffraction

pattern exhibited two characteristic crystalline peaks of PCL block at 2θ = 21.5° and 23.8°.

L-lactide containing pentablock copolymer (P-4) Fig. 4a, showed less intense peaks for PCL

and two another crystalline peaks at 2θ = 16.5° and 19.05° for PLA block. These peaks

suggest that incorporation of L-lactide in the copolymers has reduced the crystallinity of PCL

and because of its semicrystalline nature PLA exhibited its own crystalline peaks. However,

P-6 represented a wide amorphous peak suggesting that D, L-lactide containing pentablock

copolymer may exist in amorphous state. This result suggested that D, L-lactide exhibited

significant effect in reducing crystallinity of PCL in comparison to L-lactide. However, there

is no significant effect on crystalline peak of PCL in P-3 and P-5 (Fig. 3.5B). The PCL/PLA

block ratio in the case of P-3 and P-5 was higher as compared to P-4 and P-6, respectively.

Therefore, effect of PLA in reducing the PCL crystallinity was not as intense as observed in

75

the case of P-4 and P-6. From the XRD results it can be concluded that a proper ratio of

PCL/PLA in the copolymer compositions can alter the crystallinity of PCL.

76

Figure 3.5A X-ray diffraction diagrams of polymers P-1, P-4, P-6

77

Figure 3.5B X-ray diffraction diagrams of polymers P-2, P-3, P-5

78

B

A

Figure 3.6 DSC thermograms of polymers (A) First heating (B) Second heating

79

Figure 3.6A shows DSC results of first heating scan and 4.5B shows the DSC scan of second

heating cycle of all prepared copolymers. The endotherms located in between 30 °C to 60 °C

correspond to the melting of PCL component in the copolymers. The melting points of PCL

in copolymers were clearly decreased with incorporation of PLLA or PDLLA in the polymer.

This finding suggests that PLA block has reduced the crystallization ability of PCL block.

Further, in first heating scan melting point of PDLLA containing copolymer is slightly lower

than PLLA containing copolymer. These were expected results since it is widely reported

that PDLLA contains random mixture of L-lactide and D-lactide and interrupts the

crystallization of PCL more significantly than PLLA. Also, second heating scan shows

similar decrease in melting points of pentablock copolymer in comparison to triblock

copolymers. It is interesting to observe that second heating cycle shows two melting peaks

for PCL. The second melting peaks were attributed to the imperfect crystallization of PCL

chains during cooling and second heating process. Our DSC results further confirmed that

crystallization ability of PCL in copolymer compositions was altered by incorporation of

PLLA or PDLLA. Hydrolytic degradation of polymers depends on the crystallinity.

Amorphous polymers degrade at a faster rate than the crystalline polymers. Earlier reports

suggest that higher the PLA content in PCL-PLA block copolymers or PCL/PLA blends,

lower the crystallinity of PCL (Newman et al., 2009). Also, degradation rate of block

copolymer was faster in comparison to PCL and PLA homopolymers (Haung et al., 2006).

However, no literature available that suggests the effect of block copolymer crystallinity on

drug release profile from nanoparticles. In this study, two feed ratios of PCL/PLA i.e. 1: 0.5

and 1:1 were selected for the synthesis of D, L-lactide and L-lactide containing pentablock

copolymers in which PLA content was more than 50 % in each prepared copolymer. We

80

attempted to improve the crystallinity of triblock copolymers by incorporating PLA segment.

In addition, we observed that effect of D, L-lactide on reducing the crystallinity of PCL was

more pronounced in comparison the L-lactide. A change in crystallinity of PCL has also

affected release rate of triamcinolone acetonide from nanoparticles which was discussed in

another section. The hydrophobicity index of copolymers also changed with hydrophobic

segment of copolymer which was determined by the ratio of molecular weight of

hydrophobic segment of copolymer over the total molecular weight of copolymer.

Preparation and characterization of nanoparticles

Particle size and zeta potential

Triblock and pentablock copolymers based nanoparticles were prepared by single o/w

emulsion solvent evaporation method. Table 2 summarizes the properties of each

nanoparticle sample. There was not much difference on size of nanoparticles prepared from

different copolymer compositions. Nanoparticles were in the range of 160- 280 nm and were

monodispersed. Figure 3.7 shows the particle size distribution. It appears entrapped drug may

not significantly alter the size of nanoparticles. Zeta potential of each prepared nanoparticles

sample was also listed in Table 2. It appears that nanoparticles have very high negative zeta

potential, indicating stable nature of the particles in water due to repulsion. However,

hydrophobic nanoparticles tend to aggregate upon storage due to hydrophobic attraction

(Soppimath et al., 2001).

81

Figure 3.7 Particle size distributions

82

Drug loading content and entrapment efficiency

Nanoparticles drug loading and entrapment efficiency depend mainly on copolymer

composition. It can be modulated by many factors, such as molecular weight, ratio of

hydrophobic to hydrophilic segment of copolymers, crystallinity, drug solubility and drug-

polymer interactions. Since triamcinolone acetonide has very low aqueous solubility, it can

precipitate in water during fabrication of nanoparticles that resulted in higher drug loading in

all nanoparticle batches. As shown in Table 3.4, molecular weight or hydrophobic segment of

the copolymer affected drug loading and percentage entrapment efficiency of nanoparticles.

Higher drug loading in P-3 and P-5 in comparison to P-4 and P-6, respectively could be

attributed to longer chain length of PCL segment that provides more hydrophobicity to the

polymers.

Surface morphology of nanoparticles

To investigate the morphology of prepared nanoparticles SEM analysis was

conducted. Fig. 6B shows SEM graph of drug loaded nanoparticles. It can be seen that

nanoparticles are spherical in shape and possess smooth surface morphology.

83

Table 3.3 Characterization of triblock nanoparticles

Polymer Size

(nm) PDI % EE Drug loading

Zeta

Potential

A 229 ± 2 0.294 74.4 ± 2.4 % 6.7 ± 1.2 % -26.4 ± 4.5

B 219.4 ± 8 0.297 66.9 ± 3.9 % 6.1 ± 0.8 % -24.5 ± 4.8

C 199.4 ± 5 0.284 60.6 ± 1.2 % 5.5 ± 0.3 % -22.9 ± 4.9

84

Table 3.4 Characterization of pentablock nanoparticles

Polymer Size (nm) PDI EE Drug loading

Zeta potential

P-2 229 ± 2 0.294 74.4 ± 2.4 % 6.7 ± 1.2 % -26.4 ± 4.5

P-3 279 ± 2 0.257 76.8 ± 3.5 % 6.5 ± 0.6 % -34.2 ± 6.0

P-4 272 ± 12 0.237 63.1 ± 2.7 % 5.7 ± 0.3 % -29.2 ± 3.9

P-5 310 ± 14 0.286 80.6 ± 2.9 % 7.7 ± 0.2 % -29.9 ± 4.8

P-6 281 ± 8 0.322 68.3 ± 2.5 % 6.2 ± 0.3 % -30.3 ± 4.8

85

Figure 3.8 SEM image of nanoparticles prepared from pentablock polymers

86

In vitro drug release study

Figures 3.10 and 3.11 show the comparison of triamcinolone acetonide release profile from

triblock (P-2) and pentablock based nanoparticles. Different pentablock copolymer

compositions with various block ratios of PCL/PLA were used to prepare nanoparticles. In

vitro studies revealed that drug release from nanoparticles depended on length of

hydrophobic segment, molecular weight and crystallinity of the copolymers. Also,

pentablock copolymer based nanoparticles exhibited very low initial burst release phase in

comparison to triblock nanoparticles. These results were in agreement from other reported

results and explained by the fact that pentablock copolymer based nanoparticles have very

low surface absorbed drug in comparison to triblock nanoparticles (Hu et al., 2003). Figure

3.10 suggests that in comparison to triblock copolymer (P-2), drug release rate from both

pentablock copolymers (P-3 and P-5) was slower. This observation may be explained by the

fact that pentablock copolymers (P-3 and P-5) were of higher molecular weight. Although all

three copolymers have same PCL block length P-3 and P-5 have additional PLA segment in

their composition, which made their structure more hydrophobic in comparison to triblock

copolymers. However, there is no significant difference in triamcinolone acetonide release

profile from nanoparticles prepared from P-3 and P-5 copolymers. P-3 contain L-lactide in its

structure whereas, P-5 contain D, L- lactide. In both P-3 and P-5 copolymer block length of

PCL polymer is higher in comparison to PLA block. Also, in copolymers P-3 and P-5

crystallinity of PCL block was not much affected with incorporation of PLA (as per our XRD

results). Moreover, triamcinolone acetonide is hydrophobic in nature which has high

tendency to remain with hydrophobic segment of polymer. Therefore, we can conclude that

PCL segment has predominant effect on drug release profile of nanoparticles prepared from

87

P-3 and P-4 copolymers. In addition, because of high hydrophobicity of PCL block release of

lipophilic drug such as triamcinolone acetonide was much slower in later phase. Lipophilic

drugs get trapped in the hydrophobic core of the nanoparticles and achieve no or limited

release from nanoparticles as time passes. We also observed no release phase from P-3 and

P-5 nanoparticles after 9 days. Moreover, due to crystalline nature of polymers nanoparticles

exhibited initial burst release phase with almost 37 cumulative % of drug released in 2 days.

However, our aim was to modulate the drug release kinetics from nanoparticles to achieve

near zero-order release and reduce the initial burst release phase. Drug release rate from

crystalline copolymers is higher than the amorphous matrix of copolymers (Miyajima et al.,

1997). Therefore, we tried to optimize the drug release rate by changing the molecular state

of the copolymers. In copolymers P-4 and P-6, we decreased the block length of PCL in

comparison to P-3 and P-5. Pentablock copolymers P-4, P-6 and triblock copolymer P-2 have

same hydrophobic/hydrophilic block ratio. However, P-4 and P-6 contains both PLA and

PCL segment in their structure in comparison to P-2 that contains only PCL. The overall

block ratio of PCL/PLA was changed in P-4 and P-6 as compared to P-3 and P-5,

respectively to reduce the crystallinity of copolymers. Our XRD results also suggested that

incorporation of L-lactide has reduced the crystallinity of PCL block in P-4, whereas in the

case of P-6, PCL was present in amorphous state due to conjugation of D, L-lactide.

Diffusion of drug from the pores of crystalline copolymers is much faster than the amorphous

structure of the polymers. Therefore, we observed slow release of triamcinolone acetonide

from P-4 and P-6 (fig. 3.11) in comparison to other copolymers. P-4 and P-6 copolymer

based nanoparticles exhibited continuous release of triamcinolone acetonide for 14 days

without any slow release phase because of low crystallinity and hydrophobicity of

88

copolymers. Also, drug release from P-6 was much slower in comparison to P-4 because of

amorphous nature of P-6. Initial burst release profile of triamcinolone acetonide was also

decreased in P-6 nanoparticles with only 27 cumulative % drug release observed in 2 days.

Therefore, we were successful in modulating the release profile of hydrophobic drug from

pentablock copolymer based nanoparticles by changing the copolymer compositions. Also,

optimized pentablock copolymer based nanoparticles exhibited continuous zero-order

delivery of triamcinolone acetonide without producing any significant burst effect. Therefore,

pentablock copolymers are very advantageous to prepare nanoparticle formulations in

comparison to PLGA polymers which show very high burst release (Gomez-Gaete et al.,

2007).

89

Figure 3.9 Release of triamcinolone acetonide from A , B , and C

triblock copolymers nanoparticles in PBS buffer (pH 7.4) at 37 °C. The values are

represented as mean ± standard deviation of n=3

90

Figure 3.10 Release of triamcinolone acetonide from P-2 , P-3 and P-5

copolymers nanoparticles in PBS buffer (pH 7.4) at 37 °C. The values are represented as

mean ± standard deviation of n=3

91

Figure 3.11 Release of triamcinolone acetonide from P-2 , P-4 and P-6

copolymers nanoparticles in PBS buffer (pH 7.4) at 37 °C. The values are represented as

mean ± standard deviation of n=3

92

Figure 3.12 Release of triamcinolone acetonide from P-6 nanoparticles alone and P-6

nanoparticles suspended in gel copolymers nanoparticles in PBS buffer (pH 7.4) at 37

°C. The values are represented as mean ± standard deviation of n=3

93

Drug release kinetics

Drug release from nanoparticles generally follows diffusion/degradation or a combination of

diffusion and degradation mediated release phenomena (Wu and Chu, 2008). Copolymer

compositions have significant effect on the drug release profile from nanoparticles. Analysis

of drug release kinetics correlated well with Korsmeyer model. This observation suggested

that drug release depended primarily on diffusion from the nanoparticles matrix rather than

erosion process of the copolymer. In addition, analysis of first 60% release data according to

Korsmeyer model suggested that release rate of triamcinolone acetonide was slowest and

followed near zero-order kinetics from the nanoparticles prepared from P-6 pentablock

copolymer. A diffusion exponent 0.685<n<0.869 suggest anomalous diffusion mechanism

from all nanoparticles matrixes.

94

Table 3.5 Kinetic parameters for drug release from triblock nanoparticles

Polymer

Higuchi Korsmeyer-Peppas

r2 kH (day

-1/2)

r2 KKP(day

-n) n

A 0.711 18.78 0.998 40.70 0.744

B 0.848 18.85 0.980 30.23 0.558

C 0.882 19.76 0.986 24.88 0.698

95

Table 3.6 Kinetic parameters for drug release from pentablock nanoparticles

Polymer

Higuchi Korsmeyer-Peppas

r2 kH (day

-1/2)

r2 KKP(day

-n) n

P-2 0.711 18.79 0.998 40.70 0.744

P-3 0.891 17.43 0.983 23.57 0.583

P-4 0.974 15.71 0.920 15.97 0.706

P-5 0.913 17.05 0.989 19.27 0.634

P-6 0.996 14.86 0.952 8.22 0.847

96

Conclusions

PLA-PCL-PEG-PCL-PLA pentablock copolymer with different block ratios of

PEG/PCL/PLA was synthesized. Pentablock copolymer crystallinity can be modulated by

changing the PCL/PLA ratio. Triamcinolone acetonide nanoparticles were successfully

prepared from pentablock copolymers for ocular delivery. The copolymer compositions,

molecular weight and crystallinity can influence the drug release kinetics from nanoparticles.

Drug release rate from nanoparticles prepared from amorphous pentablock copolymers is

slower than crystalline copolymers. Novel pentablock copolymers are excellent biomaterials

that could serve as a vehicle for ophthalmic drug delivery as well as for other disorders where

sustained levels of corticosteroids are required.

97

CHAPTER 4

A NOVEL PENTABLOCK COPOLYMER BASED COMPOSITE DRUG DELIVERY

SYSTEM FOR TIMOLOL MALEATE DELIVERY

Rationale

Glaucoma is the second leading cause of blindness in the world, affecting

approximately 70 million patients aged 60 and over. It is an ocular disorder that

progressively causes degeneration of retinal ganglion cells causing damage to the optic nerve

head. Although there are many risk factors associated with the development of glaucoma

(age, race, myopia, family history, and injury), development of the disease is mainly

associated with elevated intra ocular pressure (IOP) caused by blockage of ocular fluid

outflow via trabecular meshwork pathway and/or uveoscleral pathway (Ghate and

Edelhauser, 2008). Progression of the disease leads to partial vision loss ultimately

progressing to complete blindness if remained untreated. Early diagnosis and effective

treatment play a vital role in halting the progression of this vision threatening condition.

Timolol maleate, a non-selective β-adrenergic blocker, is considered a gold standard

drug for the treatment of glaucoma based on its excellent pressure lowering efficacy (Kaur et

al.). It reduces the IOP by blocking the production of aqueous humor from ciliary bodies.

Timolol maleate eye drop is the primary mode of drug administration. Topical administration

is the most preferred and convenient route to treat anterior segment diseases. However,

ocular bioavailability of eye-drops is less than 1% and frequent administration is required to

maintain therapeutic levels. Poor bioavailability mainly results from precorneal factors such

as blinking, transient residence time in cul-de-sac and nasolacrimal drainage. Corneal

epithelium is composed of 6-7 layers of epithelial cells with tight junctions, which is highly

98

lipophilic in nature. It restricts the entry of hydrophilic molecules such as timolol maleate

(Ghate and Edelhauser, 2008). Frequent eye drop administrations play a major role in patient

non-compliance. Poor compliance with medications is an important reason for vision loss in

glaucoma patients. In addition, a majority of the topically applied timolol maleate eye drops

exhibit systemic absorption that leads to severe side effects such as cardiac arrhythmia,

congestive heart failure, bronchospasm and status asthmaticus (Kaur et al.). Therefore, there

is a need for a better delivery system which can avoid the frequent administration of eye

drops and reduce the systemic side effects of timolol maleate while maintaining sustained

drug levels over a longer time periods.

In recent years, nanoparticles have gained significant attention in ocular drug delivery

(Araujo et al., 2009; Cai et al., 2008; Diebold and Calonge). Biodegradable nanoparticles

with appropriate size and distribution that can provide adequate bioavailability are always

desirable for ocular drug delivery. However, previous reports suggest that subconjunctival

administration of nanoparticles do not maintain prolonged therapeutic concentration in ocular

tissues (Kompella et al., 2003). The major concern is that nanoparticles observe high burst

release in initial time points followed by slow release phase. Because of the smaller diameter

and large surface area, these exhibit rapid release of surface adsorbed drug followed by slow

release from the nanoparticle core (Kompella et al., 2003). Moreover, blood and lymphatic

circulations in the periocular region can washout a large fraction of drug loaded nanoparticles

from the subconjunctival space (Amrite et al., 2008). Drug loaded nanoparticles can be

suspended in themosensitive hydrogel to modify overall drug release rate. Thermosensitive

hydrogel can be utilized in the development of in situ depot forming injectable controlled

release formulations (Lee et al.; Packhaeuser and Kissel, 2007). Thermosensitive hydrogel

99

exhibits sol-gel transition at physiological temperature. Drug loaded nanoparticles can be

incorporated in aqueous polymer solution at room temperature, that gels to form depot

system in subconjunctival space. Thermosensitive gel can hold the drug loaded nanoparticles

at injection site that minimizes nanoparticle loss due to periocular circulation. Composite

drug delivery system (biodegradable nanoparticles suspended in injectable thermosensitive

hydrogel) will provide more sustained release over an extended period of time compared to

nanoparticles without thermosensitive gel. Drug release from the polymeric depot system

takes place in sustained manner by diffusion and degradation mechanism. Nanoparticles in

gel will also significantly reduce the initial burst release by proving additional layer for drug

diffusion. Therefore, it is very attractive to evaluate the novel pentablock polymer based

composite polymeric system for sustained delivery of timolol maleate in glaucoma therapy.

In the present work, we evaluated release of timolol from nanoparticles alone and

nanoparticles suspended in gel, which were prepared from novel biodegradable pentablock

copolymers. PLA-PCL-PEG-PCL-PLA (polylatide- polycaprolactone- polyethylene glycol-

polycaprolactone- polylatide) was utilized for preparation of nanoparticles and polyethylene

glycol- polycaprolactone- polylatide- polycaprolactone- polyethylene glycol (PEG-PCL-

PLA-PCL-PEG) was utilized for the preparation of thermosensitive hydrogel. It is interesting

to observe that both types of copolymers are composed of PEG, PCL and PLA; however they

are different in molecular weight and block arrangements. PLA block in the outer segment

makes the polymer more hydrophobic that can be utilized as nanoparticle biomaterial. Whereas

PEG block in the outer segment makes the polymer hydrophilic that can be solubilized in water

and can be utilized to prepare thermosensitive hydrogel. Drug release from nanoparticles

depended on the molecular weight of the copolymers. Burst release of timolol was significantly

100

reduced by suspending nanoparticles in thermosensitive gel. Polymeric blocks utilized for the

synthesis of pentablock copolymers are FDA approved. However, earlier investigations

suggest that biocompatibility also depends on the processing method and properties of final

polymeric biomaterial. Therefore, cytotoxicity of nanoparticles and thermosensitive hydrogel

was investigated utilizing rabbit primary corneal epithelial culture cells (rPCEC) and human

retinal pigmented epithelial cells (ARPE-19) to evaluate the suitability of these polymeric

systems for ocular drug delivery. These delivery systems were biocompatible in nature. The

composite formulation i.e. timolol loaded nanoparticles suspended in thermosensitive gel can

be injected in to the subconjunctival space to achieve long term delivery of timolol for more

than 1 month.

Method and materials

Materials

Poly (ethylene glycol) (PEG, Mw = 2000), methoxy poly(ethylene glycol) (mPEG,

Mw = 550), ε- caprolactone, stannous octoate, poly (vinyl alcohol) (PVA), were obtained

from sigma aldrich company (St. Louis, MO; USA). Hexamethylene diisocyanate and L-

lactide were purchased from Acros organics (Morris Plains, NJ; USA). Analytical grade

petroleum ether, dichloromethane (DCM) and Acetonitrile (ACN) were procured from Fisher

Scientific (Morris Plains, NJ; USA). Timolol maleate was purchased from sigma aldrich

company (St. Louis, MO; USA).

Methods

Synthesis of pentablock copolymers

PLA-PCL-PEG-PCL-PLA copolymer was synthesized for the preparation of

nanoparticles. Briefly, in the first step, PCL-PEG-PCL triblock copolymer was synthesized

101

by ring opening polymerization of ε-caprolactone. In round bottom flask first calculated

amount of PEG, (Mw=2000; 0.001 mol) was vacuum dried for 3 h then ε-caprolactone

(0.001 mol) and stannous octoate (0.5 wt. %) were added. The reaction mixture was degassed

for 30 min and purged with nitrogen gas. The reaction was performed at 130 °C for 24 h. The

resulting triblock copolymer PCL-PEG-PCL was purified by precipitation in cold petroleum

ether after dissolving in DCM. The purified copolymer was vacuum dried to remove residual

solvent. In the second step, calculated amount of purified triblock copolymer (0.001 mol)

was copolymerized with L-lactide (0.001 mol) with stannous octoate (0.5 wt. %) as catalyst.

The flask was purged with nitrogen gas and the reaction was performed at 130 °C for 24 h.

The final product PLA-PCL-PEG-PCL-PLA pentablock copolymer was purified by solvent

precipitation method as described previously.

PEG-PCL-PLA-PCL-PEG copolymer was synthesized for thermosensitive gel

formulation. In the first step, mPEG-PCL diblock copolymer was synthesized by ring

opening polymerization of ε-caprolactone (0.005 mol) with mPEG (Mw = 550; 0.005 mol) as

a initiator and stannous octoate (0.5 wt.%) as a catalyst at 130 °C for 24 h . In the second

step, calculated amount of L-lactide (0.005 mol) and stannous octoate (0.5 wt. %) were added

into the reaction mixture to prepare mPEG-PCL-PLA triblock copolymer. In the third step,

mPEG-PCL-PLA copolymer coupled with hexamethylene diisocyanate (HMDI; 0.01 mol) at

80 °C for 8 h to prepare PEG-PCL-PLA-PCL-PEG pentablock copolymer. The final product

was purified by solvent precipitation method as described above and vacuum dried to remove

residual solvent.

102

Sn(Oct)2

130 oC

O

O

O

O

C OH2C C

O

O CH2CH2O C

O

CH2 O C5 5Y XX

HC

O

O HZ

HCOH

O

Z

Lactide

CH3 CH3

PLA-PCL-PEG-PCL-PLA

Y

O

O

Sn(Oct)2

130 oC

H OH2C C

O

O CH2CH2O C

O

CH2 O H5

5Y XX

PEG Caprolactone PCL-PEG-PCL

H OCH2CH2 OH

H OH2C C

O

O CH2CH2O C

O

CH2 O H5

5Y XX

PCL-PEG-PCL

Figure 4.1 Synthetic scheme of PLA-PCL- PEG -PCL-PLA

103

Figure 4.2 Synthetic scheme of PEG-PCL-PLA-PCL-PEG

104

Table 4.1 Characterization of copolymers

Code Copolymers Mna Mn

b Mw

b PDI

b

P-1

PLA2500-PCL5000-PEG2000-PCL5000-

PLA2500

19072 10964 16260 1.48

P-2

PLA2500-PCL2500-PEG2000-PCL2500-

PLA2500

11062 6186 9526 1.54

P-3

PEG550-PCL825-PLA550-PCL825-

PEG550

3300 3872 5260 1.35

a. Values calculated from 1HNMR spectra

b. Values obtained by GPC analysis

105

Characterization of pentablock copolymers

NMR

Both pentablock copolymers were characterized by 1H NMR spectroscopy. Each

copolymer (10 mg/mL) was dissolved in deuterated chloroform (CDCl3) and proton NMR

spectra were recorded with a Varian-400 NMR spectrophotometer.

Gel permeation chromatography (GPC)

GPC analysis was performed with the Shimadzu refractive index detector to

determine molecular weight and its distribution. Polymeric material (1.5 g) was dissolved in

1.5 ml tetrahydrofuran (THF). Polystyrenes standards of different molecular weight were

utilized for GPC analysis. THF was used as eluting solvent at a flow rate of 1 ml/min and

Styragel HR-3 column maintained at 35 °C was utilized for separation.

Fourier transforms infrared spectroscopy (FTIR)

Fourier transform infrared spectroscopy (FT-IR) spectra were recorded with a

Nicolet-100 infrared spectrophotometer at a resolution of 8 sec-1

. Each copolymer was

dissolved in DCM and polymer films were casted on KBr plates.

106

Figure 4.3 1H-NMR spectra of PLA-PCL-PEG-PCL-PLA copolymer in CDCl3

107

Figure 4.4 1H-NMR spectra of PEG-PCL-PLA-PCL-PEG copolymer in CDCl3

108

Figure 4.5 FT-IR spectra (A) PLA-PCL-PEG-PCL-PLA (B) PEG-PCL-PLA-PCL-PEG

A

B

109

Phase transition study of thermosensitive gel

The sol-gel transition behavior of aqueous co-polymeric solution was investigated by

test tube inverting method. The pentablock copolymer was dissolved in 10 mM PBS (pH 7.4)

at different concentrations (15- 40 wt. %) and then incubated for 12 h at 4 °C. Then 1 ml of

polymer solution was taken in 4 ml glass vial with an inner diameter of 12 mm and placed in

water bath. The water bath temperature was raised slowly from 20 to 60 °C at the rate of 5

°C increment per 5 min. The gel formation was characterized visually by inverting the tube.

A physical state of flow was characterized as sol phase, whereas a state of no flow was

characterized as gel phase (Mishra et al.).

Preparation of timolol loaded PLA-PCL-PEG-PCL-PLA nanoparticles

Timolol maleate is a highly water soluble salt form of the drug and incorporation of

salt in the oil phase is generally difficult (Li et al., 2007), which results in less entrapment

efficiency. Since salt form is not suitable for nanoparticle preparation, we converted timolol

maleate salt into the base form of timolol as per the earlier reported method (Li et al., 2007).

Timolol base is comparatively lipophilic in nature and mostly preferred for the preparation of

controlled release formulations (Bertram et al., 2009). Briefly, 83 mg timolol maleate was

dissolved in 6 ml of 0.77 M NaOH solution. A majority of salt was converted to the base

form when pH of the solution was adjusted above the pKa of timolol i.e. (pKa ~ 9.2).

Timolol base was an oily liquid phase which was allowed to be separated from aqueous

phase at room temperature. After phase separation, 5 ml of aqueous phase was removed and

500 μl of ethyl butyrate was added into the mixture. It was again allowed to phase separate

into timolol base rich ethyl butyrate phase and aqueous phase. The timolol base containing

ethyl butyrate phase was pipette out and

110

Figure 4.6 Sol-gel transition study of P-3

111

vacuum dried for 30 min. The timolol base was an oily liquid at room temperature, which

was utilized for the nanoparticle preparation (Li et al., 2007).

Timolol nanoparticles were prepared by o/w single emulsion solvent evaporation

method. One hundred mg polymer and 10 mg drug were dissolved in 2 ml DCM. The organic

phase was emulsified in aqueous phase containing 1 % PVA by tip sonication for 5 minutes

at 65W to obtain o/w emulsion. The organic phase was evaporated by continuous stirring at

room temperature to form nanoparticles. The resulting nanoparticle suspension was then

centrifuged at 21,000 rpm, 4 °C for 1 h. The prepared nanoparticles were washed three times

with distilled water to remove un-entrapped drug and excess PVA. Nanoparticle were

lyophilized with 5 wt.% mannitol (act as lyoprotectant) for 24 h and stored at 4 °C for further

studies.

Characterization of nanoparticles

Size and zeta potential determination of nanoparticles

Nanoparticles were analyzed for mean size and distribution by dynamic light

scattering (DLS) with a 90 Plus Particle Size Analyzer (Brookhaven Instruments

Corporation). The lyophilized nanoparticles were suspended in distilled water and analysis

was performed for 3 minutes at 25 °C and 90° scattering angle. All measurements were

made in triplicate and average particle size and size distribution data were reported as mean ±

SD.

Zeta potential of nanoparticles was determined by Nano-ZS Malvern Zeta sizer. All

analyses were performed at 25 °C. Values were reported as mean value ± SD of 3 test runs.

112

Figure 4.7 Particle size distribution for P-1

113

Drug loading and entrapment efficiency

Five mg of timolol loaded freeze dried nanoparticles were dissolved in 0.5 ml of

dimethyl sulfoxide (DMSO) and diluted approximately with deionized water. Total amount

of drug in solution was analyzed by high performance liquid chromatography (HPLC) at a

wavelength of 295 nm. The mobile phase containing acetonitrile / water/ triethylamine

(18:81:1, v/v/v) mixture was used as eluent solvent at the rate of 1 mL/min and C18 reverse

phase column was utilized for separation (El-Laithy, 2009). Drug loading and entrapment

efficiency of nanoparticles were determined by the equations:

Drug loading =

Entrapment Efficiency =

114

Table 4.2 Characterization of nanoparticles

Polymer Size PDI EE Drug loading Zeta potential

P-1 250.2 ± 2.6 0.211 48.62 ± 1.6 % 5.3 ± 1.6 % -32.4 ± 5.2

P-2 220.8 ± 3.7 0.218 37.77 ± 1.8 % 2.4 ± 0.8 % -30.5 ± 4.6

115

Figure 4.8 XRD analysis of nanoparticles loaded with timolol

116

In vitro release studies of timolol from nanoparticles alone, gel alone and nanoparticles

suspended in gel

Drug loaded nanoparticles (corresponding to 0.5 mg timolol) were suspended in 500

µl 10mM phosphate buffer saline (PBS) of pH 7.4 and this nanosuspension was placed in

dialysis bag with molecular weight cutoff of 6500 Da. (Sigma Aldrich, MO; USA). The

dialysis bag was immersed in a tube containing 10 ml of 10 mM PBS (pH 7.4) at 37 °C with

continuous shaking (60 rpm). At pre-determined time intervals samples were withdrawn and

the entire release medium was replaced with fresh buffer to maintain sink conditions. To

perform the release study of nanoparticles suspended in thermosensitive gel, 0.5 mg drug

equivalent nanoparticles were first suspended in 0.5 ml hydrogel solution by simple physical

mixing. Nanoparticle suspended gel was then placed in dialysis bag with mol. wt cutoff of

6500 Da. The rest of the procedure was followed as described above. The amount of timolol

released at each time point was determined by HPLC analysis as mentioned above. The

experiments were repeated three times and mean value ± SD was expressed as a cumulative

% drug released with time.

Drug release kinetics

Drug release parameters were computed by two different methods utilizing Higuchi

and Korsmeyer equations. Drug release mechanism was evaluated by model equations as

described below.

Higuchi equation:

Mt = KHt1/2

KH indicates the Higuchi release rate constant obtained by plotting cumulative percent drug

released against the square root of time.

117

Korsmeyer-peppas equation:

Mt/ M∞ = ktn

Mt and M∞ denote the cumulative amount of drug released at time t and at the equilibrium,

respectively. The constant k represents the kinetic constant and n is the release exponent that

indicates the drug release mechanism. Values of n< 0.5 indicates Fickian (ideal) diffusion

mechanism and values of 0.5<n<1.0 suggest non-Fickian diffusion. When value of n is

greater than 1.0, it represents case II transport or zero order release kinetics. Drug release

data is employed to obtain the release exponent for Mt/ M∞ ≤ 0.6 (Mishra et al.,

2011c)(Mishra et al., 2011c)(Mishra et al., 2011c).

118

Figure 4.9 Release of timolol from P-1 and P-2 copolymers nanoparticles in

PBS buffer (pH 7.4) at 37 °C. The values are represented as mean ± standard deviation of

n=3

119

Figure 4.10 Release of timolol from P-1 Nanoparticles alone and P-1 nanoparticles in

gel copolymers nanoparticles in PBS buffer (pH 7.4) at 37 °C. The values are

represented as mean ± standard deviation of n=3

120

Table 4.3 Kinetic parameters for drug release

Polymer

Higuchi Korsmeyer-Peppas

r2 kH (day

-1/2)

r2 KKP(day

-n) n

P-1 0.964 20.18 0.973 31.54 0.263

P-2 0.964 23.60 0.983 39.03 0.294

P-3 0.961 14.17 0.983 1.18 1.3061

121

Cell line experiments

Rabbit primary corneal epithelial culture cells (rPCEC) and human retinal pigment

epithelial cells (ARPE-19) were cultured in Dulbecco's modified essential medium (DMEM,

Sigma Aldarich, USA) containing 10% fetal bovine serum, HEPES (4-(2-hydroxyethyl)-1-

piperazineethanesulfonic acid), 100 U/ml streptomycin-penicillin. The cells were cultivated

in humidified environment of 5 % CO2 at 37 °C and sub cultured after reaching 80 %

confluences.

Cytotoxicity experiments

MTS Assay

Cells were seeded in 96-well culture plates at a density of 10,000 cells per well. At 70

% confluency, 200 µl fresh medium containing nanoparticles or fresh medium containing

hydrogel of different concentrations ranging from 1 mg/mL to 20 mg/mL were added. Cells

were incubated for 48 h at the humidified atmosphere at 37° C and 5% CO2. Following

incubation, nanoparticles suspension was removed and wells were washed three times with

PBS then 20 μL of MTS (3-(4, 5-dimethylthiazol-2-yl)-5-(3-carboxymethoxyphenyl)-2-(4-

sulfophenyl)-2H-tetrazolium) solution and 80 μL of serum free medium were added into each

well. After 4 h incubation, absorbance of each well was measured at 450nm by ELISA plate

reader. Absorbance is directly proportional to viability of cells. In MTS assay, Trioton X

treated cells were served as negative control and untreated cells served as positive control.

Cell viability (%) was calculated by following equation.

Cell viability % =

Lactate Dehydrogenase (LDH) assay

122

To estimate the cytotoxicity of polymeric nanoparticles to the exposed cells lactate

Dehydrogenase (LDH) assay was performed. Cell growth conditions were kept same as the

MTS assay. Cells were seeded in 96-well culture plates at a density of 10,000 cells per well.

At 70 % confluency, 200 µl fresh medium containing nanoparticles or fresh medium

containing hydrogel of different concentrations ranging from 1 mg/mL to 20 mg/mL were

added. Cells were incubated for 48 h at the humidified atmosphere at 37° C and 5% CO2.

Damage of the cell membrane or cell death releases the lactate dehydrogenase (LDH) in the

supernatant. Concentration of LDH released in the supernatant of the rPCEC and ARPE-19

cells was measured by Cytotoxicity Detection Kit (Takara Bio Inc., Japan). Amount of the

LDH release is directly proportional to the number of damaged cells. Less than 10% LDH

release was considered as non toxic in our experiment and spectrophotometric determination

was carried out at 450 nm. LDH (%) release was calculated by following equation.

Cell viability % =

123

Figure 4.11 Cell viability studies (MTS assay) on P-1 and P-2 nanoparticles. The values are

represented as mean ± standard deviation of n=6.

124

Figure 4.12 Cell cytotoxicity studies (LDH assay) on P-1 and P-2 nanoparticles. The values

are represented as mean ± standard deviation of n=6.

125

Figure 4.13 Cell viability studies (MTS assay) on P-3 hydrogel. The values are represented as

mean ± standard deviation of n=6.

126

Results and discussion

Currently, eye drops are the primary method for reducing intraocular pressure in

glaucoma therapy. However, 80% of the topically applied drug passes through the

nasolacrimal canal and absorbs systemically. Only less than 1% of administered drug reaches

the aqueous humor. To control the intraocular pressure frequent high dose eye drops are

required which is the major reason for non-compliance in the glaucoma patients. It is

reported that to 80% of glaucoma patients do not take their medication as prescribed which

leads to the treatment failure (Olthoff et al., 2005). Poor compliance with medications is an

important reason for vision loss in glaucoma patients. Moreover, topical administration of

timolol maleate in the form of eye drops is reported to be extensively absorbed into the

systemic circulation, which can result in severe cardiopulmonary side effects (Schuman,

2000). Thus, a formulation that provides sustain levels of timolol maleate while reducing the

systemic side effects could increase the patient compliance in glaucoma therapy. Controlled

release formulation(s) of timolol maleate are reported to be a better alternative. In the current

study, we studied novel pentablock copolymers based controlled release formulations of

timolol maleate. PLA-PCL-PEG-PCL-PLA copolymer was utilized for nanoparticle

preparation and PEG-PCL-PLA-PCL-PEG copolymer was utilized for thermosensitive

hydrogel formulation. PCL and PEG based triblock copolymers (i.e. PCL-PEG-PCL or PEG-

PCL-PEG) are widely explored for drug delivery. In attempt to sustain the drug release

profile investigators generally increase the molecular weight of PCL block. However, high

molecular weight PCL block increases the overall crystallinity and hydrophobicity of the

polymer. Nanoparticles prepared from these copolymers generally exhibit high burst release,

exhibiting almost 60 % drug release during initial time points (Gou et al., 2009b; Jia et al.,

127

2008; Li et al., 2009). Moreover, drugs often get trapped in the hydrophobic core of the

nanoparticles and achieve no or limited release in the later time intervals (Gou et al., 2009a;

Jia et al., 2008). PCL exhibits very slow degradation because of its hydrophobic and

crystalline nature. Crystallinity and degradation profile of the PCL can be modulated by

conjugating with PLA (Chen et al., 2003). To modulate the crystallinity of PCL based

triblock copolymers PLA block was introduced in the pentablock copolymers.

Pentablock copolymers were synthesized by sequential ring opening polymerization

of ε-caprolactone and DL-lactide. In the first step triblock copolymers with different ratios of

PEG/PCL were synthesized by ring opening polymerization of ε-caprolactone in the presence

of PEG (Mw=2000) and a small amount of stannous octoate (Gou et al., 2009a). In the

second step these tribolck copolymers served as initiators for the synthesis of pentablock

(PLA-PCL-PEG-PCL-PLA) copolymers by ring opening polymerization of D, L-lactide (Fig.

1). For the synthesis of thermosensitive polymer mPEG of Mw= 550 was used as initiator.

Figure 4.1 and 4.2 represents the synthesis schemes for PLA-PCL-PEG-PCL-PLA (P-1) and

PEG-PCL-PLA-PCL-PEG (P-3) copolymers. Figure 4.3 and 4.4 shows the 1H NMR spectra

of both polymers. Typical signals of PEG, PCL and PLA components were utilized to

calculate the molecular weight of copolymers. Signal at 3.65 ppm (–CH2–) was assigned to

PEG block. Signals at 1.28, 1.6, 2.3 and 4.09 ppm were assigned to different methylene

protons (–CH2–) of PCL blocks and 1.4 (–CH3) and 5.19 ppm (–CH) were assigned to PLA

block. The molar ratio of PEG/PCL/PLA was determined by integrating peak intensities of

methylene protons from PEG block at 3.65 ppm, PCL block at 4.09 ppm, and to PLA block

at 5.10 ppm. The number average molecular weight (Mn) of copolymers were calculated as

per the previously reported equations (Huang et al., 2003). The Mn and Mw (weight average

128

molecular weight) values obtained by GPC analysis are summarized in Table 1. The Mn

values of copolymers determined by GPC analysis were lower than the Mn values calculated

from 1H NMR analysis. This result was attributed to the change in hydrodynamic volume of

block copolymers as compared with parent homopolymers (Huang et al., 2003).

Figure 4.5 depicts FT-IR spectra of P-1 and P-3 block copolymers. On the spectrum

of P-1 (Fig. 3A), band for C=O stretching appeared at 1732 cm-1

and bands for C-H

stretching appeared at 2941 cm-1

and 2860 cm-1

for PCL block. Absorption band at 1140 cm-1

appeared because of C-O-C stretching vibrations of the repeated OCH2CH2 units of PEG

and band at 1279 cm-1

was attributed to the -COO- stretching vibrations (Li et al., 2009). For

copolymer P-3 (Fig. 3B) another stretching band at 1526 cm-1

confirms the formation of

urethane group in pentablock copolymer.

Phase Transition Studies

Aqueous solution of pentablock copolymer P-3 exhibited sol-gel transition response

upon increasing the temperature in a concentration range of 10-35 wt%. The phase diagram

(Fig. 4) revealed critical gel concentration (CGC) for solution to gel conversion at 37 °C. The

phase transition behavior of the pentablock polymer aqueous solution was similar to other

triblock polymers. A clear aqueous solution of polymer is observed due to self assembly of

polymeric chains into micellar structure which exhibit aggregation with rise in temperature

resulting in gel formation. However, upon further heating gel-sol conversion takes place due

to increased molecular motion of hydrophobic chain of PCL and PLA. Further, elevation in

temperature results in dehydration of mPEG chains that leads to syneresis. Earlier

compositions of PEG-PCL-PEG exhibited lower CGC than the polymer synthesized in this

129

study due to the fact that incorporation of PLA lowers hydrophobicity of central block

(Ghoroghchian et al., 2006; Hariharan et al., 2009).

Preparation and characterization of timolol loaded PLA-PCL-PEG-PCL-PLA copolymer

based nanoparticles

Table 2 summarizes the properties of nanoparticles prepared by utilizing two different

polymer compositions. Our study suggests that nanoparticles prepared from P-1 have larger

particle size in comparison to P-2. Nanoparticles size was increased with increase in the PCL

chain length in the polymer. All prepared nanoparticles have polydispersity in the range of

0.218-0.231. Figure 4.7 shows the particle size distribution. Entrapment efficiency of timolol

was also depended on the pentablock copolymer composition. In the case of P-1 entrapment

efficiency was around 48.62 ± 1.6 % whereas copolymer P-2 based nanoparticles have

entrapment efficiency around 37.77 ± 1.8 %. This difference in entrapment efficiency can be

attributed to the difference in molecular weight polymers. Therefore, polymer molecular

weight is a dominant factor that regulates the size and encapsulation efficiency of

nanoparticles. Zeta potential of prepared nanoparticles sample was also listed in Table 2. It

appears that nanoparticles have very high negative zeta potential, indicating stable nature of

the particles in water due to repulsion (Soppimath et al., 2001).

In vitro drug release study

Figure 4.9 represents the drug release data of timolol from two different

nanoparticles. It is interesting to notice that molecular weight of polymer does not have

significant effect on drug release rate from nanoparticles. Timolol is a small hydrophilic

molecule, which can easily diffuse out from the nanoparticles pores upon diffusion of water

molecules inside the nanoparticles matrix. However, suspension of nanoparticles in to the

130

thermosensitive hydrogel significantly restricts the diffusion of water molecules in to

polymeric matrix. In the case of P-1 nanoparticles alone, 88.0 ± 9.4% drug release was

observed in 12 days whereas in composite formulation 90.9 ± 1.9 % of drug release was

observed in 25 days (Fig. 4.10). Moreover, nanoparticles alone exhibited very high burst

release from nanoparticles due to surface adsorbed drug, whereas in the case of nanoparticles

suspended in thermosensitive gel burst effect was significantly minimized due the additional

diffusion layer of thermosensitive polymer. These results suggest that our strategy has

minimized the burst release phase and significantly prolonged the drug release duration.

Drug Release Kinetics

We have evaluated the drug release profile of composite formulations through

Higuchi, and Korsmeyer model. In the case of nanoparticles alone release was primarily

depended (Table 3) on diffusion mechanisms whereas in the case of composite formulations

diffusion exponent value is greater than 0.5 which suggest that release process was mediated

by diffusion and degradation mechanisms. We also observed that release rate constant was

significantly decreased in the case of composite formulations in comparison to nanoparticles

alone.

Cytotoxicity studies

Cell viability responses suggested that nanoparticles prepared from both P-1 and P-2

copolymers as well as thermogelling polymer were not cytotoxic to ARPE-19 and rPCEC

cells. Figs. 8 and 9 suggest that of even at higher polymer concentration (i.e., 10 mg/ml) cell

viability was maintained almost 95 % for both ocular cell-lines However, other investigators

reported that PCL-PEG-PCL triblock copolymers are cytotoxic to cancer cell-line at

concentration of 5 mg/ml. Also, as shown in Figure 4.11 and 4.13, MTS assay results

131

suggested that nanoparticles and thermosensitive gel were not cytotoxic or negligible

cytotoxic to both cell-lines at all Therefore, MTS and LDH assay suggested that PLA

containing copolymers based nanoparticles might be a safe vector for ocular drug delivery.

Conclusions

We have successfully prepared composite formulation of nanoparticles suspended in

the hydrogel matrix for timolol delivery. This formulation had successfully minimized the

burst release phase and provided sustained zero-order release kinetics of timolol. Such novel

technology will result in a prolonged duration of action of the drug and thereby eliminates

the need for repeated administrations. Composite formulation has potential for clinical

applications based on their excellent biocompatible nature.

132

CHAPTER 5

HYDROLYTIC AND ENZYMATIC DEGRADATION OF PENTABLOCK POLYMERS

Rationale

Biodegradable polyesters such as PGA, PLA and PCL and their copolymers have been

widely explored for the various therapeutic applications such as tissue engineering scaffolds,

sutures, osteosythetic devices and controlled release drug delivery systems. These polymers

are FDA approved for human use and are biocompatible and biodegradable in nature. The

degradation products are pharmacologically inactive and can be excreted by human body.

PLA upon degradation produces lactic acid and PCL produces 6-hydroxycaproic acid, which

are re-absorbable by the human body or removable by metabolism (Zhao et al., 2004).

Hydrolytic degradation of these homopolymers and copolymers is widely explored by

various researchers. However, most of the studies dealt with either bulk materials or polymer

films. We synthesized pentablock polymers for the preparation of nanoparticles and

thermosensitive hydrogel formulations. In-vitro degradation data generally do not match with

in-vivo degradation process. We selected pseudomonas lipase enzyme to evaluate the

enzymatic degradation study. Pseudomonas lipase is an esterase capable of cleaving ester

bonds on hydrophobic substrates.

Materials and methods

Materials

Poly (ethylene glycol) (PEG, Mw = 2000), methoxy poly (ethylene glycol) (mPEG, Mw =

550), ε- caprolactone, stannous octoate, poly (vinyl alcohol) (PVA), Pseudomonas lipase (40

IU/mg) were purchased from sigma aldrich company (St. Louis, MO; USA). Hexamethylene

diisocyanate and L-lactide were purchased from Acros organics (Morris Plains, NJ; USA).

133

Analytical grade petroleum ether and dichloromethane (DCM) were purchased from Fisher

Scientific (Morris Plains, NJ; USA).

Methods

Synthesis of pentablock copolymers

Synthesis methods for PLA-PCL-PEG-PCL-PLA and PEG-PCL-PLA-PCL-PEG were

described in the previous chapter. PLA-PCL-PEG-PCL-PLA of different molecular weight

was synthesized for nanoparticle preparation and PEG-PCL-PLA-PCL-PEG was synthesized

for thermosensitive gel preparation.

Preparation of nanoparticles and thermosensitive gel

We followed the similar methods for the preparation of nanoparticles and hydrogel as

described in the chapter 4.

Hydrolytic and enzymatic degradation studies of nanoparticles

Degradation studies were performed at 37 °C in 0.05 M phosphate buffer pH 7.4 with and

without 0.2 mg/mL pseudomonas lipase enzyme. Sodium azide (0.02 wt. %) was added in

the solution to prevent the growth of microorganisms. Nanoparticles prepared from different

polymers were suspended in 5 mL of PBS. The buffer solution was replaced at 24 h to

maintain the enzyme activity. At definite time intervals, the nanoparticles were collected and

washed thoroughly with distilled deionied water by ultracentrifugation. The washed

nanoparticles were freeze dried over night for further analysis (Kulkarni et al., 2008; Zeng et

al., 2004).

134

Characterization of nanoparticles

NMR

1H NMR spectroscopy was performed to characterize copolymer compositions.

Spectra were recorded with a Varian-400 NMR instrument by dissolving polymeric material

in deuterated chloroform (CDCl3).

Gel Permeation Chromatography (GPC) analysis

GPC analysis was performed with the Shimadzu refractive index detector to

determine molecular weight and its distribution. 1.5 mg polymeric material was dissolved in

1.5 ml Tetrahydrofuran (THF). Polystyrenes standards of different molecular weights were

utilized as reference. THF was used as eluting solvent at a flow rate of 1 ml/min and Styragel

HR-3 column maintained at 35 °C was utilized for separation.

XRD analysis of nanoparticles

X-ray diffraction (XRD) analysis was performed for tri block and pentablock

copolymers, by MiniFlex automated X-ray diffractometer (Rigaku, The Woodlands, Texas)

with Ni-filtered Cu-kα radiation (30 kV and 15 mA) at room temperature.

Results and discussion

Degradation of copolymers depends on many factors such as molecular weight,

composition and hydrophobicity. We found that both hydrolytic and enzymatic degradation

of pentablock copolymer depended on the composition of polymers rather than molecular

weight or hydrophobicity of the polymers. Reports suggest that surface area of the polymer is

also a main factor responsible for degradation of copolymers. For example, rate of

degradation for nanoparticles is much higher compare to polymeric films or intact polymers

(Singh et al., 2010). Therefore, we prepared nanoparticles of different pentablock polymers

135

Figure 5.1 XRD analysis of PCL2500-PEG2000-PCL2500 after hydrolytic degradation

136

Figure 5.2 XRD analysis of PLLA2500-PCL2500-PEG2000-PCL2500-PLLA2500 after hydrolytic

degradation

137

Figure 5.3 XRD analysis of PDLLA2500-PCL2500-PEG2000-PCL2500-PDLLA2500 after

hydrolytic degradation

138

Figure 5.4 XRD analysis of PLLA2500-PCL7500-PEG1000-PCL7500-PLLA2500 after enzymatic

degradation

139

Figure 5.5 XRD analysis of PLLA2500-PCL2500-PEG2000-PCL2500-PLLA2500 after enzymatic

degradation

140

Figure 5.6 XRD analysis of PDLLA2500-PCL2500-PEG2000-PCL2500-PDLLA2500 after

enzymatic degradation

141

Figure 5.7 XRD analysis of PEG550-PCL825-PLA550-PCL825-PLA550 after enzymatic

degradation

142

Table 5.1 Degradation of PLA2500-PCL7500-PEG1000-PCL7500-PLA2500

Table 5.2 Degradation of PLLA2500-PCL2500-PEG2000-PCL2500-PLLA2500

Table 5.3 Degradation of PDLLA2500-PCL2500-PEG2000-PCL2500-PDLLA2500

Table 5.4 Degradation of PEG550-PCL825-PLA550-PCL825-PLA550

Time (h) Mnb Mw

b PDI

b

0 12932 19195 1.48

96 5385 11685 2.17

Time

(h) Mn

b Mw

b PDI

b

0 6186 9526 1.54

96 4112 5289 1.26

Time (h) Mnb Mw

b PDI

b

0 6165 10160 1.64

96 6177 9348 1.51

Time (h) Mnb Mw

b PDI

b

0 3872 5260 1.35

96 2763 3932 1.42

143

that are subjected to hydrolytic and enzymatic degradation. Moreover, we evaluated the

degradation of thermosensitive gel also. Our XRD results suggested that degradation of

pentablock copolymers first occurs in the amorphous region of the polymer followed by

crystalline region. Therefore, we observed the increase in polymer crystallinity with

degradation. Earlier reports by other investigators also suggest the similar pattern of

degradation (Zhao et al., 2007). Moreover, we observed that hydrolytic degradation of

polymers is much slower in comparison to enzymatic degradation. Due to high

hydrophobicity and crystallinity hydrolytic degradation of PCL is very slow (Liu et al.,

2000). That has decrease the overall degradation rate of copolymer. Change in crystallinity

was very prominent in the case of enzymatic degradation in comparison to hydrolytic

degradation. We evaluated the hydrolytic degradation of triblock and pentablock copolymers

having same molecular weight of PCL. We found that degradation of pentablock copolymer

was faster compared to triblock copolymers. These results support our hypothesis that in

pentablock copolymer incorporation of PLA block reduces the crystallinity of PCL block

thereby increase the degradation rate. Our previous XRD results suggest that the crystallinity

of pentablock copolymers is lower than the triblock copolymers. Enzyme is a macromolecule

and cannot enter the polymer matrix very easily. Therefore, enzymatic degradation of

nanoparticles or thermosensitive gel depends on how fast enzyme can enter the polymer

matrix (Zhao et al., 2007). Water uptake by the nanoparticles and hydrogel structure is

considered a major factor for enzymatic degradation process since water uptake could cause

swelling of the polymer and facilitate the enzymatic attack. As per release studies results

water uptake in PDLLA containing pentablock polymer is much slower in compared to

PLLA containing pentablock copolymers. Therefore, we observed the negligible change in

144

molecular weight and crystallinity of copolymers containing PDLLA. Moreover, we

observed more decrease in the molecular weight for our most hydrophobic pentablock

copolymer with PCL mw= 15000 compared to hydrophilic pentablock polymer with PCL

mw=5000. These results also can be explained by the fact that high molecular weight of PCL

provides higher crystallinity to copolymer which can lead to faster enzyme uptake through

the pore of the nanoparticles compared to less crystalline polymer. Therefore, we conclude

that crystallinity of polymer could affect the degradation of polymers. However, degradation

of copolymers is very complex process and many factors can affect the rate of degradation of

polymers. Reports suggest that although lipase is the PCL specific enzyme it can also

degrade the amorphous PLLA block. Therefore we observed the higher change in the

molecular weight of hydrophobic pentablock copolymer (Liu et al., 2000; Zhao et al., 2007).

In this scenario degradation of pentablock polymers with similar PCL content should be

similar. However, as described in table 5.2 and 5.3 there is a lot of difference in the

molecular weight change of two pentablock with similar molecular weight of PCL. These

results further support our hypothesis that the water uptake in the polymer uptake is the

possible factors that affect the degradation of copolymers. Pentablock copolymer having

PDLLA in the structure is amorphous in nature while PLLA containing copolymer has

crystalline structure. Diffusion of water across the crystalline matrix is much faster the

amorphous structure. This could be the possible reason for less change in the molecular

weight of polymer containing PDLLA.

Conclusions

Our results suggest that incorporation of PLA block not only improves the drug release rate

from nanoparticles but also increased the degradation rate of polymers. Hydrolytic

145

degradation of copolymer occurs at slower rate compared to enzymatic degradation.

Copolymer degradation depends on the composition of polymer. However, degradation of

copolymers depends on the many factors and advance polymer characterization techniques

are required to delineate the effect of each factor responsible for degradation of copolymers.

146

CHAPTER 6

PENTABLOCK COPOLYMER BASED THERMOSENSITIVE HYDROGEL FOR

MACRO MOLECULE DELIVERY

Rationale

Age-related macular degeneration (AMD) is a vision threatening ocular disease. The

most common cause of central vision loss in AMD is development of choroidal

neovascularization (CNV) characterized by fluid and blood leakage into the subretinal space

and fibrovascular scar formation (wet AMD) (Kulkarni and Kuppermann, 2005). The disease

affects the macular region of the retina, RPE, Bruch’s membrane and choriocapillaris. It may

cause the irreversible blindness in older age patients (Munoz et al., 2000). Vascular

endothelial growth factor (VEGF), a naturally occurring large lipoprotein molecule is

responsible for the growth of blood vessels (Leung et al., 1989). Elevated levels of VEGF

have been reported in many sight threatening eye diseases such as AMD and diabetic

retinopathy. It is a potent mitogen for endothelial cells. This growth factor increases vascular

permeability by leukocyte-mediated endothelial cell injury, formation of fenestration, and

dissolution of tight junctions. These manifestations cause intra-retinal fluid accumulation

which has a negative effect on visual acuity. Moreover, VEGF can also cause release of

inflammatory cytokines which can further reinforce the cycle of inflammation and

angiogenesis (Daughaday, 1992). Thus, anti-VEGF agents seem to be attractive therapeutic

molecules for the treatment of wet AMD. Anti-VEGF molecules of different sizes such as

Avastin (a 149 kDa antibody), Lucentis (a 48 kDa antibody, fab fragment) and Macugen (a

28-base ribonucleic acid aptamer covalently linked to two branched 50 kDa PEG moieties)

are being used for the treatment of different ocular conditions. Direct intra-vitreal injection of

147

antibodies appears found to be a more effective treatment of wet-AMD (McVey et al., 2008).

Monthly injection of Lucentis has become the standard care for wet AMD (El Sanharawi et

al., 2010). Since wet AMD is the chronic ocular disease therapeutic concentration of anti

VEGF proteins must be maintained for longer durations, which requires repeated intravitreal

injections. This thereby can be costly and inconvenient to patients. Moreover, various side

effects are associated with repeated intravitreal injections as summarized earlier. Delivery of

large molecules to the posterior segment of the eye has always been a challenge. Various

reports suggest that trasscleral route may be a viable option for delivery of large molecules to

the posterior segment (Ambati et al., 2000; Pescina et al., 2011). Sclera has large accessible

surface area and relatively high permeability that does not decrease with age. Different ex-

vivo studies reported the permeability of various macromolecules across the sclera (Ambati

et al., 2000). Human sclera is permeable to 70 kDa dextran (Olsen et al., 1995) whereas

rabbit sclera is permeable up to 150 kDa dextran and IgG (Ambati and Adamis, 2002;

Ambati et al., 2000). Literature suggests that molecular radius of the molecule is more

important factor than the molecular weight because reports mention that the globular proteins

has higher permeability through sclera than the liner protein (Ambati et al., 2000; Boubriak et

al., 2000; Olsen et al., 1995). Recently, in-situ depot forming drug delivery systems have

gained significant attention for sustained delivery of various therapeutic macromolecules.

This novel delivery system provides the ease of administration, delivery of accurate dose and

enhanced contact time with the target tissue. We have synthesized novel block copolymers

for the preparation of macromolecules loaded thermosensitive hydrogel formulation.

Thermosensitive hydrogels are environment-sensitive polymeric systems, in which sol-gel

transition is triggered by change in temperature. The ideal critical temperature for this

148

delivery system is in the range of physiological temperature where the source of heat to form

a gel is supplied by the body, which makes this system more flexible (Rathore, 2010). The

thermosensitive gels are liquid polymeric clear aqueous solution at room temperature. Large

molecules can be easily dispersed in liquid polymeric solution. Drug loaded polymeric

solution can be injected at subconjunctival space to form a depot system. A subconjunctival

injection can deliver up to 500μl of drug volume underneath the conjunctiva. Sustained

delivery via the subconjunctival route offers easier access and a safer administration route.

Though there are minor risks associated with subconjunctival injections, proper selection and

utilization of advanced techniques can minimize such risk (Ghate et al., 2007; Raghava et al.,

2004).

Pentablock copolymers (PEG-PCL-PLA-PCL-PEG) were synthesized as ingredients

of thermosensitive gel formulation. Each block of these pentablock copolymers is FDA

approved for human use. PLGA polymer is most commonly incorporated in the preparation

of controlled release formulation. Drug delivery systems prepared from this polymer

following degradation produce high molar mass of lactic and glycolic acids, which may

cause tissue irritation and toxicity by lowering pH (Kang and Schwendeman, 2002). PLGA

particles can cause rapid release (40-50% of payload) of the cargo within one to two days

(Budhian et al., 2005; Jwala et al., 2011). Pentablock copolymers are composed of relatively

low molecular weight of PLA block. Upon degradation of copolymer, due to low molecular

weight of PLA, the polymer will produce negligible amount of lactic acid. Therefore,

pentablock polymers are more suitable than the existing PLGA based gel for delivery of

macromolecules, particularly proteins and antibodies. We have synthesized two different

compositions of copolymers with different ratios of PCL/PLA blocks. The copolymers were

149

characterized for molecular weight and distribution with GPC and 1HNMR. The functional

group modification was confirmed by FT-IR. Crystallinity was characterized by XRD

analysis. Effect of polymer composition and concentration on drug release profile of

Lysozyme were also evaluated. Effect of size of the large molecule on release profile from

hydrogel also evaluated utilizing FITC-dextran of different molecular weight.

Materials and methods

Methoxy poly(ethylene glycol) (mPEG, Mw = 550), ε- caprolactone, stannous octoate,

Lysozyme from chicken egg white and Micrococcus luteus were purchased from Sigma

Aldrich (St. Louis, MO; USA). Hexamethylene diisocyanate and L-lactide were obtained

from Acros organics (Morris Plains, NJ; USA). Analytical grade petroleum ether,

dichloromethane (DCM) and Micro BCA protein assay kit were procured from Fisher

Scientific (Morris Plains, NJ; USA).

Synthesis of pentablock copolymers

Two compositions of PEG-PCL-PLA-PCL-PEG copolymer were synthesized for

thermosensitive gel formulation. In the first step, mPEG-PCL diblock copolymer was

synthesized by ring opening polymerization of ε-caprolactone (0.01 mol) using mPEG (Mw

= 550; 0.01mol) as a initiator and stannous octoate (0.5 wt.%) as a catalyst at 130 °C for 24 h

. In the second step, calculated amount of L-lactide (0.01 mol) and stannous octoate (0.5

wt.%) was added into the reaction mixture to prepare mPEG-PCL-PLA triblock copolymer.

In the third step, mPEG-PCL-PLA copolymer coupled with hexamethylene diisocyanate

(HMDI; 0.01 mol) at 80 °C for 8 h to prepare PEG-PCL-PLA-PCL-PEG pentablock

copolymer. The final product was purified by precipitation in cold petroleum ether after

dissolving in DCM. The purified copolymer was vacuum dried to remove residual solvent

150

Characterization of pentablock copolymers

NMR

Pentablock copolymers were characterized by 1H NMR spectroscopy. Each

copolymer (10 mg/mL) was dissolved in deuterated chloroform (CDCl3) and proton NMR

spectra were recorded with a Varian-400 NMR spectrophotometer.

Gel permeation chromatography (GPC)

GPC analysis was performed with the Shimadzu refractive index detector to

determine molecular weight and its distribution. The polymeric material (1.5 mg) was

dissolved in 1.5 ml tetrahydrofuran (THF). Polystyrenes standards of different molecular

weight were employed for GPC analysis. THF as eluting solvent was pumped at a flow rate

of 1 ml/min through Styragel HR-3 column maintained at 35 °C for separation.

Fourier transforms infrared spectroscopy (FTIR)

Fourier transform infrared spectroscopy (FT-IR) spectra were recorded with a

Nicolet-100 infrared spectrophotometer at a resolution of 8 sec-1

. Each copolymer was

approximately dissolved in DCM and polymer films were casted on KBr plates.

Phase transition study of thermosensitive gel

The sol-gel transition behavior of aqueous co-polymeric solution was investigated by

test tube inversion method. The pentablock copolymer was dissolved in 10 mM PBS (pH 7.4)

at different concentrations (15- 40 wt. %) and then incubated for 12 h at 4 °C. One ml of

polymer solution was taken in 4 ml glass vial with an inner diameter of 12 mm and placed in

water bath. The water bath temperature was raised from 20 to 60 °C at the rate of 5 °C

increment over 5 min. The gel formation was observed visually by inverting the tube. A

151

physical state of flow was characterized as sol phase, whereas a state of no flow was

considered as gel phase (Mishra et al., 2010).

Rheological Measurement

Rheology of polymer solution was conducted in Oscillatory HR-3 rheometer, TA

instruments with cone and plate geometry. The angle of the cone was 2˚ with a diameter of

60mm. Two mL sample was placed on the temperature controlled peltier plate of rheometer

at 5˚C and further analysis was conducted at a angular frequency of 1 rad. All the rheological

studies were conducted within the linear viscoelastic transition range of the polymer. Sol- gel

phase transition was performed at temperatures ranging from 10-42 °C with 1 °C intervals.

The gelation time was studied by time sweep measurements at 37 °C and the angular

frequency was kept at 1 rad/s.

Preparation of in situ gel forming formulations for lysozyme

Pentablock copolymers were dissolved in PBS solution at a concentration of 20 wt%.

Calculated amount of lysozyme was suspended into the polymeric solution and the vial was

vortexed until a homogeneous solution was formed. A 500 µL volume of lysozyme loaded

polymeric solution was injected into the glass tube maintained at 37 °C water bath. After 10

min the polymeric solution turned into a gel.

In vitro release of lysozyme

Five mL PBS solution (0.01 M, pH 7.4, 0.02% Sodium azide) was added onto the top

of the gel inside a tube. The tube was incubated in the shaker bath at 30 rpm and 37 °C. To

maintain the sink condition samples were withdrawn at predetermined time interval and

replaced with fresh PBS solution. The amount of lysozyme released at each time points was

calculated by Micro-BCA assay. Briefly, 150 µL of diluted sample was mixed with 150 µL

152

of Micro-BCA working reagent (Micro-BCA reagent A, B, C at the ratio of 50:48:2) and

poured into the 96-well microplate reader. The solution mixture was incubated at 37 °C for 2

h and allowed to cooled down to 25 °C. The micro plate reader analysis was performed at

595 nm. To prepare the standard curve freshly diluted lysozyme solution was prepared at

different concentration range of 25-200 µg/mL. The effect of polymer composition on the

release profile of lysozyme was evaluated by plotting the cumulative % lysozyme released

with time. To determine the effect of polymer concentration on the release profile of

lysozyme, two different concentrations of polymer (i.e. 20 wt. % and 25 wt. %) were

selected. The rest of the conditions were kept similar to previous condition.

In vitro release of dextran

FITC-dextran of different molecular weights (i.e. 9200, 38000 and 72000) were selected to

determine the effect of molecule size on the release profile from thermosensitive hydrogel.

PB-2 polymer was selected for this study. Pentablock copolymer was dissolved in PBS

solution at the concentration of 20 wt%. Calculated amount of FITC-dextran of different size

were suspended into the polymeric solutions and vial were vertexed until the homogeneous

solution was formed. A 500 µL of polymeric solution was injected into the glass tube. The

tube was put into the 37 °C water bath. After 10 min the polymeric solution became the gel.

A 5 mL of PBS solution (0.01 M, pH 7.4, 0.02% Sodium azide) was added on the top of the

gel in the tube. The tube was incubated in the shaker bath at 30 rpm and 37 °C. To maintain

the sink condition the samples were withdrawn at predetermined time interval and replaced

with fresh PBS solution. The amount of FITC-dextran released at each time points was

calculated using plate reader with excitation wavelength of 485 nm and emission wavelength

of 535 nm.

153

Biological activity of lysozyme by enzyme activity assay

To determine enzymatic activity stock solution of Micrococcus luteus (0.01%, w/v)

was prepared in phosphate buffer (66 mM, pH 6.15) and diluted to obtain absorbance in

between 0.2 to 0.6 at 450 nm. A 2.5 ml of M. luteus solution was added to a 0.1 ml of

approximately diluted sample of lysozyme. The reaction mixture was then immediately

transferred to a spectrophotometer cell having path length of 1 cm and absorbance was

determined at 450 nm with UV spectrophotometer. A rate of decrease of absorbance at 450

nm was monitored over a period of 5 min at 25 °C. The slope of the linear portion of the plot

(obtained by plotting absorbance vs time) determined the amount of lysozyme in one enzyme

unit (EU). The units of active lysozyme were calculated as per the previously reported

formula (Tang and Singh, 2009).

Where, df is the dilution factor, 0.001 was obtained from the definition of lysozyme unit, and

0.1 is the volume (in ml) of sample or standard. The biological activity of lysozyme from

released sample was compared with the control lysozyme sample in phosphate buffer to

determine the effect of pentablock polymer based hydrogel formulation on lysozyme activity.

Secondary structure stability study

Structural stability of lysozyme was determined by Circular dichroism (CD)

measurement. CD analysis was performed using Jasco 720 spectropolarimeter at 25 °C. CD

spectra were recorded over a range of 260 to 200 nm using 1 cm cell, band width of 1 nm and

a scanning speed of 100 nm/min. A spectrum for the blank PBS solution was obtained at the

154

same wavelength range as the blank subtraction. CD measurements were reported as molar

ellipticity [θ]

Statistical analysis

All release studies were performed in triplicate. The results were reported as mean ±

standard deviation. Effects of copolymer compositions on lysozyme release profile from

thermosensitive hydrogel were compared by student’s t-test.

Results and discussion

Pentablock copolymers were synthesized by sequential ring opening polymerization

of ε-caprolactone and L-lactide. In the first step mPEG of Mw= 550 was copolymerized with

caprolactone. Figure 6.1 represents the synthetic scheme for PEG-PCL-PLA-PCL-PEG

copolymer. Figure 6.2 illustrates the 1H NMR spectra of both polymers. Typical signals of

PEG, PCL and PLA components were utilized to calculate the molecular weight of

copolymers. Signal at 3.65 ppm (–CH2–) was assigned to PEG block. Signals at 1.28, 1.6,

2.3 and 4.09 ppm were assigned to different methylene protons (–CH2–) of PCL blocks and

1.4 (–CH3) and 5.19 ppm (–CH) were assigned to PLA block. The molar ratio of

PEG/PCL/PLA was determined by integrating peak intensities of methylene protons from

PEG block at 3.65 ppm, PCL block at 4.09 ppm, and to PLA block at 5.10 ppm. The number

average molecular weight (Mn) of copolymers were calculated according to a previous report

(Huang et al., 2003). The Mn and Mw (weight average molecular weight) values obtained by

GPC analysis are summarized in Table 6.1. Mn values of copolymers determined by GPC

analysis are lower than Mn values calculated from 1H NMR analysis. This result was

attributed to the change in hydrodynamic volume of block copolymers relative to parent

homopolymers (Huang et al., 2003).

155

Figure 6.3 depicts FT-IR spectra of pentablock copolymer. On the spectrum of P-1,

band for C=O stretching at 1732 cm-1

and bands for C-H stretching at 2941 cm-1

and 2860

cm-1

for PCL block appeared. Absorption band at 1140 cm-1

signifies C-O-C stretching

vibrations of OCH2CH2 repeat units of PEG and band at 1279 cm-1

can be attributed to -

COO- stretching vibrations (Li et al., 2009). Another band for N-H stretching at 1526 cm-1

confirms the formation of urethane group in pentablock copolymer.

P-1 is more viscous than the other polymer P-2. The storage modulus (G’) values for

P-1 and P-2 are higher than the loss modulus (G’’) values, indicating the formation of more

elastic (solid) gels.

156

Figure 6.1 Synthetic scheme of PEG-PCL-PLA-PCL-PEG

157

Figure 6.2 1HNMR spectra of P-2

158

Figure 6.3 FTIR spectrum of P-1

159

Table 6.1 Molecular weight characterization of polymers

Code Copolymers Mna Mn

b Mw

b PDI

b

P-1 PEG550-PCL825-PLA550-PCL825-PEG550 2893 2810 4211 1.49

P-2 PEG550-PCL550-PLA1100-PCL550-PEG550 2705 2625 3763 1.43

160

Table 6.2 Rheological measurement

Code tan delta G’(Pa) G’’(Pa) Gelation

time (s)

P-1 0.143 2754.67 334.95 25

P-2 0.194 1526.13 296.71 38

161

Phase Transition Studies

Aqueous solution of pentablock copolymer P-1 and P-2 exhibited sol-gel transition

with rising temperature within a concentration range of 10-35 wt %. The phase diagram (Fig.

6.4) revealed that critical gel concentration (CGC) for P-1 is 18 wt %, whereas for P-2 it is 20

wt %. This change in the CGC may be attributed to the higher hydrophobicity of the

polymer. Below the gelling temperature a clear aqueous solution of polymer is observed due

to self assembly of polymeric chains into micellar structure. Polymeric micelles exhibit

aggregation with rise in temperature resulting in gel formation. However, upon further

heating at elevated temperature causes dehydration of mPEG chains leading to syneresis.

Results of sol-gel transition process is consistent with our previously reported data for

triblock polymers (Mishra et al., 2011c).

162

Figure 6.4 Phase diagram of thermosensitive hydrogel formulations of P-1 and P-2

163

In vitro release of lysozyme from thermosensitive hydrogel

Release of lysozyme from the hydrogel formulation appears to depend on the

copolymer composition as shown in Figure 6.5. Rate of lysozyme release is slower from P-1

copolymer formulation than P-2. In the first 24 h cumulative percentage release from P-1 was

20.5 ± 2.6 % whereas in P-2 it was 25.5 ± 1.5 %. This result can be attributed to the fact that

P-1 is more hydrophobic in nature compared to P-2 since it possess relatively longer PCL

chains that restrict entry of water molecule in the hydrogel matrix. Therefore, diffusion of

lysozyme molecules across the hydrogel matrix is slower in P-1 than P-2. These results

suggest that release rate of lysozyme from the hydrogel formulation may be affected by

hydrophilic/hydrophobic balance of the copolymer. Figure 6.6 suggests that copolymer

concentration has significant effect on the release profile of lysozyme from P-2. Release of

lysozyme from 20 wt. % is more rapid than 25 wt. % formulation due to volume effect of the

gel matrix. Since, 25 wt. % gel forms dense matrix compared to 20 wt. % it restricts the

diffusion of lysozyme molecules.

As shown in figure 6.7 release of FITC-dextran from thermosensitive gel P-2 was depended

on the size of the molecule. Molecular chains of large molecules get trapped in the hydrogel

matrix that provide hinder for the diffusion of molecules from the hydrogel and result delay

in release. Therefore, we observed the release of FITC-dextran the following order

(Mw=70000, size=58 °A) < (Mw=40000, size=44.5 °A) < (Mw=10000, size=23.6 °A). From

our studies we can conclude that release of large molecules from the thermosensitive gel is

mainly depended on the size of the molecules. These data are in consistent with literature and

with our previous study with the small molecules, where we did not found any significant

164

Figure 6.5 Lysozyme release from two different thermosensitive gels PB-1 and PB-2 (20 wt

%) at 37 °C. The values are represented as mean ± standard deviation of n=3 (p< 0.05)

165

Figure 6.6 Lysozyme release from PB-2 thermosensitive gel at two different concentrations

at 37 °C. The values are represented as mean ± standard deviation of n=3 (p< 0.05)

166

Figure 6.7 Release profile of FITC-Dextran of different Mw was suspended in

thermosensitive gel PB-2 (20 wt %) at 37 °C. The values are represented as mean ± standard

deviation of n=3 (p< 0.05)

167

difference in the release profile with small molecules having same size but different

lipophilicity (Lu and Anseth, 2000).

Biological Activity

Table 6.3 displays biological activity of lysozyme in released samples. Protein

activity of freshly prepared lysozyme sample is 22.2 ± 2.0 (EU/mg) ×103. Enzymatic activity

in released samples taken at different time points is higher than the respective controls.

Thermosensitive hydrogel polymers are water soluble and hydrogel formulations are

prepared by vertexing lysozyme solution with the polymer in a phosphate buffer solution.

There is no harsh treatment or organic solvent required to prepare the thermosensitive gel

formulations. Therefore, we conclude that protein activity of the entrapped lysozyme

remained unaltered for a long duration in comparison to control samples which are in

phosphate buffer for the entire duration. However, we observed that enzymatic activity of

lysozyme at each time points is higher in P-2 formulation than P-1. This result can be

explained by the fact that P-1 is more hydrophobic than the P-2. Hydrophobic residues can

cause protein adsorption on the polymer surface that will result in loss of biological activity

of the entrapped protein. However, pentablock polymers are amphiphilic in nature which are

more suitable for protein delivery. As shown in table 3 protein activity in the released sample

diminished with time. It may be explained by the fact that protein molecules remained in the

release media for a specific time interval. Control samples have been collected at respective

time intervals by suspending lysozyme in phosphate buffer at the same concentration as in

formulations. A decrease in enzymatic activity can be attributed to storage conditions which

will not be the case under in vivo conditions. Under in vivo conditions proteins would be

absorbed immediately after release from the gel. The thermosensitive gel slows diffusion of

168

water molecules to the core of the hydrogel and protects the protein (Crotts and Park, 1997;

Singh et al., 2007).

Secondary structure stability by circular dichroism (CD)

Figure 6.8 represents the CD spectra of released sample from P-1 formulation and

native lysozyme. CD spectra from the sample at day 20 and day 50 displayed a strong

negative band at 208 nm and 222 nm, which were similar to the native sample. CD is a very

sensitive method and can be utilized to investigate the secondary structure of proteins.

Previous work suggests that copolymer degradation products such as lactic acid can degrade

the protein (Determan et al., 2006). Following degradation pentablock copolymer will release

lactic acid but as per the CD spectra the released sample did not exhibit any significant

difference from native lysozyme indicating stable nature of the protein inside the hydrogel.

169

Table 6.3 Biological activity of lysozyme in the released samples

Days Specific enzyme activity (EU/mg)×103

Control PB-1 PB-2

1 19.0 ± 2.0 19.3 ± 0.5 21.8 ± 1.1

20 12.5 ± 1.2 16.8 ± 1.2 18.8 ± 1.1

30 9.2 ± 0.6 13.6 ± 1.4 15.3 ± 1.8

50 5.5 ± 0.3 11.1 ± 2.4 12.1 ± 1.2

170

Figure 6.8 CD spectra of released lysozyme sample from P-1 formulation

171

Conclusions

Novel pentablock polymers for the preparation of hydrogel formulations have been

synthesized. Release kinetics of lysozyme from the hydrogel depends on the pentablock

copolymer compositions and concentration. In addition, release profile of FITC-dextran

depends on the size of the molecules. Results from our laboratory also suggest that

pentablock copolymers are biocompatible with the different ocular cell-lines (results are not

shown). Therefore, pentablock copolymers have the potential for providing sustained release

profile for protein molecules without affecting its structure and activity. Pentablock

copolymers appear to be excellent biomaterials for ocular protein delivery.

172

CHAPTER 7

SUMMARY AND RECOMMENDATIONS

Summary

Different compositions of pentablock copolymers were synthesized and characterized to

modulate the drug release profile from nanoparticles. We evaluated the effect of PEG on the

release profile of triamcinolone acetonide by synthesizing triblock copolymers PCL-PEG-

PCL with different molecular weights of PEG in the range from 2000- 8000 and found that

the hydrophilic segment does not have significant effect on the release profile of a

hydrophobic drug. However, a relatively slower release rate was observed from PEG of

mw=2000 in comparison to PEG of mw=8000. Therefore, we seleacted PEG=2000

molecular weight for the synthesis of pentablock polymers.

We synthesized pentablock copolymers containing PEG, PCL and PLA blocks. Each

block has its unique properties. Literature suggests that the block copolymers can retain the

properties of individual blocks. PEG is well known for its non-antigenic and non

immunogenic nature. In addition, because of its hydrophilic nature it facilitates the diffusion

of water into nanoparticles matrix and provides diffusion mediated drug release from

nanoparticles. PCL is hydrophobic biodegradable polyester which enables high permeability

for small molecules. Being hydrophobic in nature it also provides good encapsulation

efficiency to lipophilic drugs via hydrophobic interactions. However, it exhibits very high

crystallinity that is responsible for high burst release. PLA can influence the crystallinity of

PCL. We wanted to have all these properties in a single polymer to prepare an ideal drug

delivery system for delivery of steroids. Therefore, we optimize block ratio of PCL/PLA by

synthesizing a series of pentablock polymers. We attempted to improve the limitations of

triblock polymers i.e. PCL-PEG-PCL by incorporating the additional block PLA at the end

173

by covalent conjugation, i.e. PLA-PCL-PEG-PCL-PLA. Triblock copolymers have limitation

of burst release of drug from the nanoparticles. Burst release mainly responsible for the high

crystallinity of PCL and surface adsorbed drug. For the first time we described a strategy,

using covalently conjugated PLA blocks to reduce the crystallinity of PCL-PEG-PCL and

affect a drug release profile from nanoparticles. For reference, we synthesized the triblock

copolymers as well with same molecular weight of PCL, and characterized the triblock

copolymers for crystallinity and drug release profile. We evaluated and compared the effect

on crystallinity and drug release profile of pentablock polymers having the same molecular

weight of PCL. We were successful in changing the crystalline properties of triblock

polymers by synthesizing pentablock polymers, which has altered the release profile of

triamcinolone acetonide from nanoparticles. Moreover, pentablock copolymer based

nanoparticles exhibited minimal burst release profile. Therefore, we concluded that different

compositions of pentablock polymer (PLA-PCL-PEG-PCL-PLA) have better properties

compared to respective triblock polymers and can; provide sustained release profile for

longer duration without producing significant burst release.

Further we evaluated the pentablock copolymers for long term delivery of timolol.

Our earlier attempt of modulating the sustained release profile of timolol from PEG-PCL-

PEG hydrogel by adding the polymeric additives did not show significant improvement

(Mishra et al., 2011c). Therefore, we synthesized a pentablock copolymer based

thermosensitive hydrogel i.e. PEG-PCL-PLA-PCL-PLA. We found that pentablock

copolymer based hydrogel had better reproducibility in terms of sol-gel transitive reversion

compared to triblock polymer based hydrogels. However, to further improve the long term

release profile we evaluated the composite approach. We suspended the nanoparticles in to

174

thermosensitive hydrogel solution. We found that release profile of timolol from both

nanoparticles and thermosensitive gel alone was faster compared to the combination of both

formulations. Since, timolol is a hydrophilic molecule diffusion of molecules from the gel

was faster. In addition, nanoparticle performance suffers from the surface adsorbed drug

causing initial burst release. Our combination approach has significantly reduced the drug

release rate because of the fact that drug has to travel longer through the polymeric matrix in

the case of composite formulation.

Further, we evaluated the degradation kinetics of different pentablock copolymers.

We performed the hydrolytic and enzymatic degradation of pentablock copolymer based

nanoparticles and thermosensitive hydrogel. We found that hydrolytic degradation of these

formulations was much slower compared to enzymatic degradation. These results could be

explained by the fact that due to high hydrophobicity, hydrolytic degradation of PCL is very

slow. However, lipase enzyme can specifically degrade the PCL chains and low molecular

weight PLLA chains, but not the PDLLA. Therefore, we found faster degradation of PLLLA

containing pentablock polymers compared to PDLLA containing structure.

Finally, we evaluated the potential of thermosensitive hydrogel copolymers for long

term delivery of sensitive protein molecules. Thermosensitive hydrogels and nanoparticles

are porous structures and size of the molecules has highest impact on the drug release rate

compared to other parameters. We evaluated the effect of size of the macromolecules on the

release profile by utilizing FITC-dextran of molecular weights. We found that size of the

molecules have significant effect on the release profile, with smaller sized molecules

diffusing more rapidly than larger sized. These results were in agreement with our previous

results with small molecules. We found that various small molecules having different

175

hydrophobicity and log P values did not show significant difference in the release pattern

because of the similar smaller size. In contrast, large molecules get trapped in the polymer

matrix that provides hindrance to the diffusion of molecules. Moreover, we evaluated the

biological activity and structural stability of lysozyme (selected as model protein molecule)

released from the thermosensitive hydrogel formulation. Since, hydrogel preparation does not

involve any harsh treatment such as use of organic solvent or sonication, protein molecules

can retain structural integrity and activity during preparation of the formulation. In addition,

hydrogel can minimize the direct exposure of protein molecules with water molecules,

further retarded the degradation of protein molecules inside the formulation. We also found

that structural integrity and biological activity of lysozyme was maintained in the release

medium during the entire release period. Therefore, we conclude that our novel pentablock

copolymers based thermosensitive hydrogel and nanoparticles have potential for long term

delivery of small as well as macromolecules in the treatment of chronic ocular diseases

where sustained levels of therapeutic molecules are required.

176

Recommendations

Pentablock copolymers can be further evaluated for the delivery of sensitive molecules such

as proteins and antibodies. Preliminary results from our laboratory suggest that PEG is more

compatible to hydrophilic molecules such as antibodies and proteins. Determination of the

effect of PEG on the release profile of large hydrophilic molecules would be helpful to

optimize the controlled release formulations for proteins. Further, our composite approach of

nanoparticles suspended in a thermosensitive gel can be applicable to large molecules which

will further prolong the release of therapeutic molecules. Since, our results suggest size of the

large molecules has impact on the release profile from gel, different compositions of

copolymers should tried. Pentablock copolymer based-drug delivery systems can be

introduced in to subconjunctival space or in the intravitreal cavity. Such a delivery system

may result in prolonged duration of action, thereby completely eliminating the need for

repeated intravitreal injections. Moreover, target specific drug delivery can be obtained by

using the targeted polymers along with pentablock copolymers in the formulation of

controlled release delivery systems. For example, peptide transporter present on the

basolateral side of the RPE and Bruch’s membrane. These novel polymers could be utilize

for delivery of siRNA, protein, antibody and fragments. Moreover, determination of

intravitreal pharmacokinetics of various small and macromolecules after administration in to

rabbit eye could provide new mechanistic approaches to biomaterial design.

177

Confirmation Number: 110:57:59:5 Order ~te: 01/ 02/2013

Customer Jnfonnation

Customer. wI! tfmboli "<COIInl Number. 3000578120 Or'9aniZitlon: viral tamboli Email: vtamboliOma~.um~c.edu I'tlOIll:: +1 (551)5805580

~ Pnnlthos~ Pnnl terms &. concICJons Pnnl atabon Iflfonnatioll (Wbtt's thos?)

Search Ofde r dehils by: I C/Ioo$e One v II

Order Details

Th.~ptutiC delivery

Order ddaillO: 63300113

ISSN: 2()4I·sm Publication Typt! : )oIIrnal Volume: ISlIIt:: Start INlge: Publil her:

~rmission Stalul: ti Grlnled

Permillion type: Republish or disolay content

Bill irlg Status: .,. Typt! of ust: Republish in . thestsldissertltion

Order lic~nst Id: 3061091178202

Nole: This ~em wts wwoictd $tPlfatr/y through our Rightslink 'leNice. MoI'e Info So.OO

178

I Copy order > I

Confi rm~ lion Number: 11052144 Order O~te: 12/07/20 12

Cuslome r Informlltion

( u~tomer: virlltlmboli Account Number: 3000576120 01'1l"nIZlll lon: viral tlmboli Em"II: vtlmbcH iOm.il ,um kc.edu Phorn:: +1 (551)5805580

Search ord" d.tl il. by : I Choose One

Order Deill ii s

Colloid & polymer $eitnee

Order detai l 10:

Art icle Tille:

AulhlH"('):

001: Dale: ISS N:

63219057

I-Iovel pentltH<xk wPQiym er (M- PCl -PEG-PCL-PL.A.)-bl..,d nl nopartide. for controlled drug delivery: effed of wpoiymer oompo<ilion. on the (t)':St/lllinity 01 (oPQiymers and in vitro drug reluse pro~le from nanopartid"" Tlmboli, Virll: Hishrl, GYln p, M ~ ... Ad, im K. 10. 1007 f S00396-0 12 -2854-0 I-Iov 21, 2012 1435-153£

Publication Type: e-)ournal Vo lume : In ue: Sla rt 1'"11": Publioher: SPRINGER-VERlAG

BERLIWHEIDELBERG

ta Print tIli. paQ" Print terms'" conditi""s Print c~~on inform.tion (What'. tIli'?)

Perminlon Slalu., ~ Grl nted

Pe rminion type: ReputHish or display oontent Type of u..,: u.., in. the~diner1.tion

Order Liu n ... Id: 3043770565542

I%l View detlils

Nn ' ,,· ThO. ".m .... ino"" ..... "" •••• .." . h ..... 'nh .... ., ....... . 'n • ....... r . Mn ... in'n

179

Re: Permission for reuse the figure and text of my article in my dissertation

Noran EI-Zoheary [noran.el [email protected][

To: 1II 1~ m ~ol\ Vir~1 M, {UMKC->ludenll

• F~ lorlollow up. StaJt by Monday. ~mberOS. 2012. Due by Monday. ~mberOS. 2012.

Dear Dr . Tamboli ,

Thank you for your email and for your i nterest i n Journal of Drug Delivery, which is an open access journal . Open Access authors retain the copyrights of their papers , and all open access articles are dist ribut ed under t he terms of the creat ive Commons Attribution license, which permit s unrest rict ed use, dist ribut ion and reproduct ion i n any medi um, provided that the original work is properly cited . Accordingly, any reader/visitor is allowed to reuse any of our open access publications without getting a permission, provided that the original work is properly cited .

Please fee l free to contact me i f I can be of any further help .

Best regards,

Noran

******************************* Noran El-Zoheary Editorial Office Hi ndawi Publishi ng Corporation http : //www.hindawi.com

180

Re: Policy on reuse the material

Shankar Pandalai [email protected]

To: T~l!Ibcl\ Vir.1 M. (IIMlC.Studml)

October 6, 2012

il<!ar Dr. Viral Tamboli,

RESEARCH SIGNPOST TRANSWORLD RESEARCH NETWORK

Thank you for you e-mail dated October 4, 2012. I have talked to our management. They have given permission to use your own chapter for free. Kindly quote source,

Thanking you

Sincerely yours,

A. Gayathri Publication Manager.

181

References

Aggarwal, D., Kaur, I.P., 2005. Improved pharmacodynamics of timolol maleate from a

mucoadhesive niosomal ophthalmic drug delivery system. Int J Pharm 290, 155-159.

Ahmed, F., Discher, D.E., 2004. Self-porating polymersomes of PEG-PLA and PEG-PCL:

hydrolysis-triggered controlled release vesicles. J Control Release 96, 37-53.

Albertsson, A., C., Carlfors, J., and sturesson, C., 1996. J Applied polymer sci. 62, 695.

Aliabadi, H.M., Lavasanifar, A., 2006. Polymeric micelles for drug delivery. Expert Opin

Drug Deliv 3, 139-162.

Ambati, J., Adamis, A.P., 2002. Transscleral drug delivery to the retina and choroid. Progress

in retinal and eye research 21, 145-151.

Ambati, J., Canakis, C.S., Miller, J.W., Gragoudas, E.S., Edwards, A., Weissgold, D.J., Kim,

I., Delori, F.C., Adamis, A.P., 2000. Diffusion of high molecular weight compounds through

sclera. Investigative ophthalmology & visual science 41, 1181-1185.

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, 150-160.

Anumolu, S.S., Singh, Y., Gao, D., Stein, S., Sinko, P.J., 2009. Design and evaluation of

novel fast forming pilocarpine-loaded ocular hydrogels for sustained pharmacological

response. J Control Release 137, 152-159.

Araujo, J., Gonzalez, E., Egea, M.A., Garcia, M.L., Souto, E.B., 2009. Nanomedicines for

ocular NSAIDs: safety on drug delivery. Nanomedicine 5, 394-401.

Badawi, A.A., El-Laithy, H.M., El Qidra, R.K., El Mofty, H., El dally, M., 2008. Chitosan

based nanocarriers for indomethacin ocular delivery. Arch Pharm Res 31, 1040-1049.

182

Bangham, A.D., Standish, M.M., Watkins, J.C., 1965. Diffusion of univalent ions across the

lamellae of swollen phospholipids. J Mol Biol 13, 238-252.

Beeley, N.R., Rossi, J.V., Mello-Filho, P.A., Mahmoud, M.I., Fujii, G.Y., de Juan, E., Jr.,

Varner, S.E., 2005. Fabrication, implantation, elution, and retrieval of a steroid-loaded

polycaprolactone subretinal implant. J Biomed Mater Res A 73, 437-444.

Bernatchez, S.F., Merkli, A., Tabatabay, C., Gurny, R., Zhao, Q.H., Anderson, J.M., Heller,

J., 1993. Biotolerance of a semisolid hydrophobic biodegradable poly(ortho ester) for

controlled drug delivery. J Biomed Mater Res 27, 677-681.

Bertram, J.P., Saluja, S.S., McKain, J., Lavik, E.B., 2009. Sustained delivery of timolol

maleate from poly(lactic-co-glycolic acid)/poly(lactic acid) microspheres for over 3 months.

J Microencapsul 26, 18-26.

Boddu, S.H., Jwala, J., Chowdhury, M.R., Mitra, A.K., 2010a. In vitro evaluation of a

targeted and sustained release system for retinoblastoma cells using Doxorubicin as a model

drug. J Ocul Pharmacol Ther 26, 459-468.

Boddu, S.H., Jwala, J., Vaishya, R., Earla, R., Karla, P.K., Pal, D., Mitra, A.K., 2010b. Novel

nanoparticulate gel formulations of steroids for the treatment of macular edema. J Ocul

Pharmacol Ther 26, 37-48.

Bonacucina, G., Cespi, M., Misici-Falzi, M., Palmieri, G.F., 2009. Colloidal soft matter as

drug delivery system. J Pharm Sci 98, 1-42.

Bonferoni, M.C., Chetoni, P., Giunchedi, P., Rossi, S., Ferrari, F., Burgalassi, S., Caramella,

C., 2004. Carrageenan-gelatin mucoadhesive systems for ion-exchange based ophthalmic

delivery: in vitro and preliminary in vivo studies. Eur J Pharm Biopharm 57, 465-472.

183

Boubriak, O.A., Urban, J.P., Akhtar, S., Meek, K.M., Bron, A.J., 2000. The effect of

hydration and matrix composition on solute diffusion in rabbit sclera. Experimental eye

research 71, 503-514.

Bruining, M.J., Edelbroek-Hoogendoorn, P.S., Blaauwgeers, H.G., Mooy, C.M., Hendrikse,

F.H., Koole, L.H., 1999. New biodegradable networks of poly(N-vinylpyrrolidinone)

designed for controlled nonburst degradation in the vitreous body. J Biomed Mater Res 47,

189-197.

Budhian, A., Siegel, S.J., Winey, K.I., 2005. Production of haloperidol-loaded PLGA

nanoparticles for extended controlled drug release of haloperidol. J Microencapsul 22, 773-

785.

Cai, X., Conley, S., Naash, M., 2008. Nanoparticle applications in ocular gene therapy.

Vision Res 48, 319-324.

Cam, D., Hyon, S.H., Ikada, Y., 1995. Degradation of high molecular weight poly(L-lactide)

in alkaline medium. Biomaterials 16, 833-843.

Camelo, S., Lajavardi, L., Bochot, A., Goldenberg, B., Naud, M.C., Fattal, E., Behar-Cohen,

F., de Kozak, Y., 2007. Ocular and systemic bio-distribution of rhodamine-conjugated

liposomes loaded with VIP injected into the vitreous of Lewis rats. Mol Vis 13, 2263-2274.

Carrasquillo, K.G., Ricker, J.A., Rigas, I.K., Miller, J.W., Gragoudas, E.S., Adamis, A.P.,

2003. Controlled delivery of the anti-VEGF aptamer EYE001 with poly(lactic-co-

glycolic)acid microspheres. Investigative ophthalmology & visual science 44, 290-299.

Chakravarthy, U., Soubrane, G., Bandello, F., Chong, V., Creuzot-Garcher, C., Dimitrakos,

S.A., 2nd, Korobelnik, J.F., Larsen, M., Mones, J., Pauleikhoff, D., Pournaras, C.J.,

Staurenghi, G., Virgili, G., Wolf, S., 2006. Evolving European guidance on the medical

184

management of neovascular age related macular degeneration. Br J Ophthalmol 90, 1188-

1196.

Chen, C.C., Chueh, J.Y., Tseng, H., Huang, H.M., Lee, S.Y., 2003. Preparation and

characterization of biodegradable PLA polymeric blends. Biomaterials 24, 1167-1173.

Chin, H.S., Park, T.S., Moon, Y.S., Oh, J.H., 2005. Difference in clearance of intravitreal

triamcinolone acetonide between vitrectomized and nonvitrectomized eyes. Retina 25, 556-

560.

Colthurst, M.J., Williams, R.L., Hiscott, P.S., Grierson, I., 2000. Biomaterials used in the

posterior segment of the eye. Biomaterials 21, 649-665.

Cortesi, R., Argnani, R., Esposito, E., Dalpiaz, A., Scatturin, A., Bortolotti, F., Lufino, M.,

Guerrini, R., Cavicchioni, G., Incorvaia, C., Menegatti, E., Manservigi, R., 2006. Cationic

liposomes as potential carriers for ocular administration of peptides with anti-herpetic

activity. Int J Pharm 317, 90-100.

Crotts, G., Park, T.G., 1997. Stability and release of bovine serum albumin encapsulated

within poly(lactide-co-glycolide) microparticles. Journal of controlled release 44, 123-134.

Croy, S.R., Kwon, G.S., 2006. Polymeric micelles for drug delivery. Curr Pharm Des 12,

4669-4684.

Dalwadi, G., Sunderland, B., 2009. An ion pairing approach to increase the loading of

hydrophilic and lipophilic drugs into PEGylated PLGA nanoparticles. Eur J Pharm Biopharm

71, 231-242.

Danion, A., Arsenault, I., Vermette, P., 2007. Antibacterial activity of contact lenses bearing

surface-immobilized layers of intact liposomes loaded with levofloxacin. J Pharm Sci 96,

2350-2363.

185

Daughaday, W.H., 1992. Pituitary gigantism. Endocrinology and metabolism clinics of North

America 21, 633-647.

Daugherty, A.L., Mrsny, R.J., 2003. Emerging technologies that overcome biological barriers

for therapeutic protein delivery. Expert Opin Biol Ther 3, 1071-1081.

Desai, S.D., Blanchard, J., 2000. Pluronic F127-based ocular delivery system containing

biodegradable polyisobutylcyanoacrylate nanocapsules of pilocarpine. Drug Deliv 7, 201-

207.

Deshpande, A.A., Heller, J., Gurny, R., 1998. Bioerodible polymers for ocular drug delivery.

Crit Rev Ther Drug Carrier Syst 15, 381-420.

Determan, A.S., Wilson, J.H., Kipper, M.J., Wannemuehler, M.J., Narasimhan, B., 2006.

Protein stability in the presence of polymer degradation products: consequences for

controlled release formulations. Biomaterials 27, 3312-3320.

Dey, S., Patel, J., Anand, B.S., Jain-Vakkalagadda, B., Kaliki, P., Pal, D., Ganapathy, V.,

Mitra, A.K., 2003. Molecular evidence and functional expression of P-glycoprotein (MDR1)

in human and rabbit cornea and corneal epithelial cell lines. Investigative ophthalmology &

visual science 44, 2909-2918.

Diebold, Y., Calonge, M., Applications of nanoparticles in ophthalmology. Prog Retin Eye

Res 29, 596-609.

Diebold, Y., Calonge, M., 2010. Applications of nanoparticles in ophthalmology. Prog Retin

Eye Res 29, 596-609.

Dillen, K., Vandervoort, J., Van den Mooter, G., Ludwig, A., 2006. Evaluation of

ciprofloxacin-loaded Eudragit RS100 or RL100/PLGA nanoparticles. Int J Pharm 314, 72-

82.

186

Ding X, Alani WG, JR, R., Extended-release and targeted drug delivery systems, in: Troy,

D.B. (Ed.), Remington: the science and practice of pharmacy. Lippincott Williams and

wilkins.

Duvvuri, S., Gaurav Janoria, K., Mitra, A.K., 2006. Effect of polymer blending on the release

of ganciclovir from PLGA microspheres. Pharm Res 23, 215-223.

Duvvuri, S., Janoria, K.G., Mitra, A.K., 2005. Development of a novel formulation

containing poly(d,l-lactide-co-glycolide) microspheres dispersed in PLGA-PEG-PLGA gel

for sustained delivery of ganciclovir. J Control Release 108, 282-293.

Duvvuri, S., Janoria, K.G., Pal, D., Mitra, A.K., 2007. Controlled delivery of ganciclovir to

the retina with drug-loaded Poly(d,L-lactide-co-glycolide) (PLGA) microspheres dispersed in

PLGA-PEG-PLGA Gel: a novel intravitreal delivery system for the treatment of

cytomegalovirus retinitis. J Ocul Pharmacol Ther 23, 264-274.

Einmahl, S., Behar-Cohen, F., D'Hermies, F., Rudaz, S., Tabatabay, C., Renard, G., Gurny,

R., 2001. A new poly(ortho ester)-based drug delivery system as an adjunct treatment in

filtering surgery. Investigative ophthalmology & visual science 42, 695-700.

Einmahl, S., Behar-Cohen, F., Tabatabay, C., Savoldelli, M., D'Hermies, F., Chauvaud, D.,

Heller, J., Gurny, R., 2000. A viscous bioerodible poly(ortho ester) as a new biomaterial for

intraocular application. J Biomed Mater Res 50, 566-573.

Einmahl, S., Ponsart, S., Bejjani, R.A., D'Hermies, F., Savoldelli, M., Heller, J., Tabatabay,

C., Gurny, R., Behar-Cohen, F., 2003. Ocular biocompatibility of a poly(ortho ester)

characterized by autocatalyzed degradation. J Biomed Mater Res A 67, 44-53.

187

Einmahl, S., Savoldelli, M., D'Hermies, F., Tabatabay, C., Gurny, R., Behar-Cohen, F., 2002.

Evaluation of a novel biomaterial in the suprachoroidal space of the rabbit eye. Investigative

ophthalmology & visual science 43, 1533-1539.

EL- Samaligy, M., S., Rojanasakul, Y., Charlton, J., F., Weinstein, G. W., and Lim J., K.,,

1996. Ocular disposition of nanoencapsulated acyclovir and ganciclovir via intravitreal

injection in rabbit’s eye. Drug Delivery 3, 93-97.

El-Laithy, H.M., 2009. Novel transdermal delivery of Timolol maleate using sugar esters:

preclinical and clinical studies. Eur J Pharm Biopharm 72, 239-245.

El Sanharawi, M., Kowalczuk, L., Touchard, E., Omri, S., de Kozak, Y., Behar-Cohen, F.,

2010. Protein delivery for retinal diseases: from basic considerations to clinical applications.

Progress in retinal and eye research 29, 443-465.

Elbayoumi, T.A., Torchilin, V.P., 2010. Current trends in liposome research. Methods Mol

Biol 605, 1-27.

Fialho, S.L., Behar-Cohen, F., Silva-Cunha, A., 2008. Dexamethasone-loaded poly(epsilon-

caprolactone) intravitreal implants: a pilot study. Eur J Pharm Biopharm 68, 637-646.

Forrest, M.L., Won, C.Y., Malick, A.W., Kwon, G.S., 2006. In vitro release of the mTOR

inhibitor rapamycin from poly(ethylene glycol)-b-poly(epsilon-caprolactone) micelles. J

Control Release 110, 370-377.

Frank, A., Rath, S.K., Venkatraman, S.S., 2005. Controlled release from bioerodible

polymers: effect of drug type and polymer composition. J Control Release 102, 333-344.

Fresta, M., Fontana, G., Bucolo, C., Cavallaro, G., Giammona, G., Puglisi, G., 2001. Ocular

tolerability and in vivo bioavailability of poly(ethylene glycol) (PEG)-coated polyethyl-2-

cyanoacrylate nanosphere-encapsulated acyclovir. J Pharm Sci 90, 288-297.

188

Fukushima, A., Ozaki, A., Ishida, W., van Rooijen, N., Fukata, K., Ueno, H., 2005.

Suppression of macrophage infiltration into the conjunctiva by clodronate liposomes in

experimental immune-mediated blepharoconjunctivitis. Cell Biol Int 29, 277-286.

Gao, Y., Sun, Y., Ren, F., Gao, S., 2010. PLGA-PEG-PLGA hydrogel for ocular drug

delivery of dexamethasone acetate. Drug development and industrial pharmacy 36, 1131-

1138.

Gavini, E., Chetoni, P., Cossu, M., Alvarez, M.G., Saettone, M.F., Giunchedi, P., 2004.

PLGA microspheres for the ocular delivery of a peptide drug, vancomycin using

emulsification/spray-drying as the preparation method: in vitro/in vivo studies. Eur J Pharm

Biopharm 57, 207-212.

Ghate, D., Brooks, W., McCarey, B.E., Edelhauser, H.F., 2007. Pharmacokinetics of

intraocular drug delivery by periocular injections using ocular fluorophotometry.

Investigative ophthalmology & visual science 48, 2230-2237.

Ghate, D., Edelhauser, H.F., 2006. Ocular drug delivery. Expert Opin Drug Deliv 3, 275-287.

Ghate, D., Edelhauser, H.F., 2008. Barriers to glaucoma drug delivery. J Glaucoma 17, 147-

156.

Ghoroghchian, P.P., Li, G., Levine, D.H., Davis, K.P., Bates, F.S., Hammer, D.A., Therien,

M.J., 2006. Bioresorbable Vesicles Formed through Spontaneous Self-Assembly of

Amphiphilic Poly(ethylene oxide)-block-polycaprolactone. Macromolecules 39, 1673-1675.

Giannavola, C., Bucolo, C., Maltese, A., Paolino, D., Vandelli, M.A., Puglisi, G., Lee, V.H.,

Fresta, M., 2003. Influence of preparation conditions on acyclovir-loaded poly-d,l-lactic acid

nanospheres and effect of PEG coating on ocular drug bioavailability. Pharm Res 20, 584-

590.

189

Gomez-Gaete, C., Tsapis, N., Besnard, M., Bochot, A., Fattal, E., 2007. Encapsulation of

dexamethasone into biodegradable polymeric nanoparticles. Int J Pharm 331, 153-159.

Gou, M., Gong, C., Zhang, J., Wang, X., Gu, Y., Guo, G., Chen, L., Luo, F., Zhao, X., Wei,

Y., Qian, Z., 2010. Polymeric matrix for drug delivery: honokiol-loaded PCL-PEG-PCL

nanoparticles in PEG-PCL-PEG thermosensitive hydrogel. J Biomed Mater Res A 93, 219-

226.

Gou, M., Zheng, L., Peng, X., Men, K., Zheng, X., Zeng, S., Guo, G., Luo, F., Zhao, X.,

Chen, L., Wei, Y., Qian, Z., 2009a. Poly(epsilon-caprolactone)-poly(ethylene glycol)-

poly(epsilon-caprolactone) (PCL-PEG-PCL) nanoparticles for honokiol delivery in vitro. Int

J Pharm 375, 170-176.

Gou, M., Zheng, X., Men, K., Zhang, J., Zheng, L., Wang, X., Luo, F., Zhao, Y., Zhao, X.,

Wei, Y., Qian, Z., 2009b. Poly(epsilon-caprolactone)/poly(ethylene glycol)/poly(epsilon-

caprolactone) nanoparticles: preparation, characterization, and application in doxorubicin

delivery. J Phys Chem B 113, 12928-12933.

Gupta, H., Aqil, M., Khar, R.K., Ali, A., Bhatnagar, A., Mittal, G., 2010. Sparfloxacin-

loaded PLGA nanoparticles for sustained ocular drug delivery. Nanomedicine 6, 324-333.

Guttman, C., 2005. Multiple factors underlie non-compliance with treatment.,

Ophthalmology Times.

Guy, R.H., Hadgraft, J., Taylor, M.J., Kellaway, I.W., 1983. Release of non-electrolytes from

liposomes. J Pharm Pharmacol 35, 12-14.

Hacker, M.C., Haesslein, A., Ueda, H., Foster, W.J., Garcia, C.A., Ammon, D.M., Borazjani,

R.N., Kunzler, J.F., Salamone, J.C., Mikos, A.G., 2009. Biodegradable fumarate-based drug-

delivery systems for ophthalmic applications. J Biomed Mater Res A 88, 976-989.

190

Hariharan, S., Minocha, M., Mishra, G.P., Pal, D., Krishna, R., Mitra, A.K., 2009. Interaction

of ocular hypotensive agents (PGF2 alpha analogs-bimatoprost, latanoprost, and travoprost)

with MDR efflux pumps on the rabbit cornea. J Ocul Pharmacol Ther 25, 487-498.

Hathout, R.M., Mansour, S., Mortada, N.D., Guinedi, A.S., 2007. Liposomes as an ocular

delivery system for acetazolamide: in vitro and in vivo studies. AAPS PharmSciTech 8, 1.

Haung, M., Li, S., Hutmacher, D.W., Coudane, J., Vert, M., 2006. Degradation

Characteristics of Poly(ε-caprolactone)-Based Copolymers and Blends. J Applied Poly Sci

102, 1681-1687.

Havlikova, L., Matysova, L., Hajkova, R., Satinsky, D., Solich, P., 2008. Advantages of

pentafluorophenylpropyl stationary phase over conventional C18 stationary phase--

application to analysis of triamcinolone acetonide. Talanta 76, 597-601.

He, C., Kim, S.W., Lee, D.S., 2008. In situ gelling stimuli-sensitive block copolymer

hydrogels for drug delivery. J Control Release 127, 189-207.

Heller, J., 1984. Biodegradable polymers in controlled drug delivery. Crit Rev Ther Drug

Carrier Syst 1, 39-90.

Heller, J., 2005. Ocular delivery using poly(ortho esters). Adv Drug Deliv Rev 57, 2053-

2062.

Heller, J., Barr, J., Ng, S.Y., Abdellauoi, K.S., Gurny, R., 2002. Poly(ortho esters): synthesis,

characterization, properties and uses. Adv Drug Deliv Rev 54, 1015-1039.

Hong, Y., Chirila, T.V., Vijayasekaran, S., Shen, W., Lou, X., Dalton, P.D., 1998.

Biodegradation in vitro and retention in the rabbit eye of crosslinked poly(1-vinyl-2-

pyrrolidinone) hydrogel as a vitreous substitute. J Biomed Mater Res 39, 650-659.

191

Hosny, K.M., 2010. Ciprofloxacin as ocular liposomal hydrogel. AAPS PharmSciTech 11,

241-246.

Hoyng, P.J., Van beek, L.M., 2000. Pharmacological therapy for glaucoma: a review. Drugs

59, 411-434.

Hu, Y., Jiang, X., Ding, Y., Zhang, L., Yang, C., Zhang, J., Chen, J., Yang, Y., 2003.

Preparation and drug release behaviors of nimodipine-loaded poly(caprolactone)-

poly(ethylene oxide)-polylactide amphiphilic copolymer nanoparticles. Biomaterials 24,

2395-2404.

Hu, Y., Xie, J., Tong, Y.W., Wang, C.H., 2007. Effect of PEG conformation and particle size

on the cellular uptake efficiency of nanoparticles with the HepG2 cells. J Control Release

118, 7-17.

Huang, M.H., Li, S., Coudane, J., Vert, M., 2003. Synthesis and Characterization of Block

Copolymers of e-Caprolactone and DL-Lactide Initiated by Ethylene Glycol or Poly(ethylene

glycol). Macromol Chem and Phys 204, 1994-2001.

Huang, M.H., Li, S., Hutmacher, D.W., Schantz, J.T., Vacanti, C.A., Braud, C., Vert, M.,

2004. Degradation and cell culture studies on block copolymers prepared by ring opening

polymerization of epsilon-caprolactone in the presence of poly(ethylene glycol). J Biomed

Mater Res A 69, 417-427.

Hyon, S.H., 2000. Biodegradable poly (lactic acid) microspheres for drug delivery systems.

Yonsei Med J 41, 720-734.

Hyon, S.H., K., J., Ikada, Y., 1998. Polymer Int 46, 196.

Jain, J.P., Modi, S., Domb, A.J., Kumar, N., 2005. Role of polyanhydrides as localized drug

carriers. J Control Release 103, 541-563.

192

Jampel, H.D., Leong, K.W., Dunkelburger, G.R., Quigley, H.A., 1990. Glaucoma filtration

surgery in monkeys using 5-fluorouridine in polyanhydride disks. Arch Ophthalmol 108,

430-435.

Janoria, K.G., Gunda, S., Boddu, S.H., Mitra, A.K., 2007. Novel approaches to retinal drug

delivery. Expert Opin Drug Deliv 4, 371-388.

Jermak, C.M., Dellacroce, J.T., Heffez, J., Peyman, G.A., 2007. Triamcinolone acetonide in

ocular therapeutics. Surv Ophthalmol 52, 503-522.

Jesorka, A., Orwar, O., 2008. Liposomes: technologies and analytical applications. Annu Rev

Anal Chem (Palo Alto Calif) 1, 801-832.

Jia, W., Gu, Y., Gou, M., Dai, M., Li, X., Kan, B., Yang, J., Song, Q., Wei, Y., Qian, Z.,

2008. Preparation of biodegradable polycaprolactone/poly (ethylene

glycol)/polycaprolactone (PCEC) nanoparticles. Drug Deliv 15, 409-416.

Jonas, J.B., Kreissig, I., Degenring, R., 2002. Repeated intravitreal injections of

triamcinolone acetonide as treatment of progressive exudative age-related macular

degeneration. Graefe's archive for clinical and experimental ophthalmology = Albrecht von

Graefes Archiv fur klinische und experimentelle Ophthalmologie 240, 873-874.

Jwala, J., Boddu, S.H., Shah, S., Sirimulla, S., Pal, D., Mitra, A.K., 2011. Ocular sustained

release nanoparticles containing stereoisomeric dipeptide prodrugs of acyclovir. Journal of

ocular pharmacology and therapeutics : the official journal of the Association for Ocular

Pharmacology and Therapeutics 27, 163-172.

Kang, J., Schwendeman, S.P., 2002. Comparison of the effects of Mg(OH)2 and sucrose on

the stability of bovine serum albumin encapsulated in injectable poly(D,L-lactide-co-

glycolide) implants. Biomaterials 23, 239-245.

193

Karla, P.K., Earla, R., Boddu, S.H., Johnston, T.P., Pal, D., Mitra, A., 2009. Molecular

expression and functional evidence of a drug efflux pump (BCRP) in human corneal

epithelial cells. Current eye research 34, 1-9.

Karla, P.K., Pal, D., Mitra, A.K., 2007a. Molecular evidence and functional expression of

multidrug resistance associated protein (MRP) in rabbit corneal epithelial cells. Experimental

eye research 84, 53-60.

Karla, P.K., Pal, D., Quinn, T., Mitra, A.K., 2007b. Molecular evidence and functional

expression of a novel drug efflux pump (ABCC2) in human corneal epithelium and rabbit

cornea and its role in ocular drug efflux. Int J Pharm 336, 12-21.

Katti, D.S., Lakshmi, S., Langer, R., Laurencin, C.T., 2002. Toxicity, biodegradation and

elimination of polyanhydrides. Adv Drug Deliv Rev 54, 933-961.

Kaur, I.P., Aggarwal, D., Singh, H., Kakkar, S., Improved ocular absorption kinetics of

timolol maleate loaded into a bioadhesive niosomal delivery system. Graefe's archive for

clinical and experimental ophthalmology = Albrecht von Graefes Archiv fur klinische und

experimentelle Ophthalmologie 248, 1467-1472.

Kaur, I.P., Aggarwal, D., Singh, H., Kakkar, S., 2010. Improved ocular absorption kinetics of

timolol maleate loaded into a bioadhesive niosomal delivery system. Graefe's archive for

clinical and experimental ophthalmology = Albrecht von Graefes Archiv fur klinische und

experimentelle Ophthalmologie 248, 1467-1472.

Kaur, I.P., Garg, A., Singla, A.K., Aggarwal, D., 2004. Vesicular systems in ocular drug

delivery: an overview. Int J Pharm 269, 1-14.

194

Kawakami, S., Harada, A., Sakanaka, K., Nishida, K., Nakamura, J., Sakaeda, T., Ichikawa,

N., Nakashima, M., Sasaki, H., 2004. In vivo gene transfection via intravitreal injection of

cationic liposome/plasmid DNA complexes in rabbits. Int J Pharm 278, 255-262.

Kimura, H., Ogura, Y., 2001. Biodegradable polymers for ocular drug delivery.

Ophthalmologica 215, 143-155.

Klouda, L., Mikos, A.G., 2008. Thermoresponsive hydrogels in biomedical applications. Eur

J Pharm Biopharm 68, 34-45.

Koevary, S.B., 2002. Trends in the noncorneal delivery of drugs into the eye. Archivos de la

Sociedad Espanola de Oftalmologia 77, 347-349.

Kompella, U.B., Bandi, N., Ayalasomayajula, S.P., 2003. Subconjunctival nano- and

microparticles sustain retinal delivery of budesonide, a corticosteroid capable of inhibiting

VEGF expression. Investigative ophthalmology & visual science 44, 1192-1201.

Kulkarni, A., Reiche, J., Hartmann, J., Kratz, K., Lendlein, A., 2008. Selective enzymatic

degradation of poly(epsilon-caprolactone) containing multiblock copolymers. Eur J Pharm

Biopharm 68, 46-56.

Kulkarni, A.D., Kuppermann, B.D., 2005. Wet age-related macular degeneration. Advanced

drug delivery reviews 57, 1994-2009.

Kunou, N., Ogura, Y., Yasukawa, T., Kimura, H., Miyamoto, H., Honda, Y., Ikada, Y., 2000.

Long-term sustained release of ganciclovir from biodegradable scleral implant for the

treatment of cytomegalovirus retinitis. J Control Release 68, 263-271.

Kwak, H.W., D'Amico, D.J., 1992. Evaluation of the retinal toxicity and pharmacokinetics of

dexamethasone after intravitreal injection. Arch Ophthalmol 110, 259-266.

195

Lam, C.X., Savalani, M.M., Teoh, S.H., Hutmacher, D.W., 2008. Dynamics of in vitro

polymer degradation of polycaprolactone-based scaffolds: accelerated versus simulated

physiological conditions. Biomed Mater 3, 034108.

Layre, A.M., Couvreur, P., Chacun, H., Aymes-Chodur, C., Ghermani, N.E., Poupaert, J.,

Richard, J., Requier, D., Gref, R., 2006. Busulfan loading into poly(alkyl cyanoacrylate)

nanoparticles: physico-chemistry and molecular modeling. J Biomed Mater Res B Appl

Biomater 79, 254-262.

Lazzeri, S., Figus, M., Di Bartolo, E., Rizzo, S., Nardi, M., 2011. Verteporfin photodynamic

therapy for retinal hemangioblastoma associated with Von Hippel-Lindau Disease in a 9-

year-old child. Clin Experiment Ophthalmol.

Lee, D.A., Leong, K.W., Panek, W.C., Eng, C.T., Glasgow, B.J., 1988. The use of

bioerodible polymers and 5-fluorouracil in glaucoma filtration surgery. Investigative

ophthalmology & visual science 29, 1692-1697.

Lee, J.Y., Kim, K.S., Kang, Y.M., Kim, E.S., Hwang, S.J., Lee, H.B., Min, B.H., Kim, J.H.,

Kim, M.S., In vivo efficacy of paclitaxel-loaded injectable in situ-forming gel against

subcutaneous tumor growth. Int J Pharm 392, 51-56.

Lee, V.H., 1990. New directions in the optimization of ocular drug delivery. Journal of

ocular pharmacology 6, 157-164.

Lehr, C.M., Lee, Y.H., Lee, V.H., 1994. Improved ocular penetration of gentamicin by

mucoadhesive polymer polycarbophil in the pigmented rabbit. Investigative ophthalmology

& visual science 35, 2809-2814.

196

Leong, K.W., Brott, B.C., Langer, R., 1985. Bioerodible polyanhydrides as drug-carrier

matrices. I: Characterization, degradation, and release characteristics. J Biomed Mater Res

19, 941-955.

Leong, K.W., Simonte, V., and langer R., 1987. Macromolecules 20, 705.

Leung, D.W., Cachianes, G., Kuang, W.J., Goeddel, D.V., Ferrara, N., 1989. Vascular

endothelial growth factor is a secreted angiogenic mitogen. Science 246, 1306-1309.

Li, C.C., Abrahamson, M., Kapoor, Y., Chauhan, A., 2007. Timolol transport from

microemulsions trapped in HEMA gels. J Colloid Interface Sci 315, 297-306.

Li, R., Li, X., Xie, L., Ding, D., Hu, Y., Qian, X., Yu, L., Ding, Y., Jiang, X., Liu, B., 2009.

Preparation and evaluation of PEG-PCL nanoparticles for local tetradrine delivery.

International journal of pharmaceutics 379, 158-166.

Li, S., Dobrzynski, P., Kasperczyk, J., Bero, M., Braud, C., Vert, M., 2005. Structure-

property relationships of copolymers obtained by ring-opening polymerization of glycolide

and epsilon-caprolactone. Part 2. Influence of composition and chain microstructure on the

hydrolytic degradation. Biomacromolecules 6, 489-497.

Li, S., Molina, I., Martinez, M.B., Vert, M., 2002. Hydrolytic and enzymatic degradations of

physically crosslinked hydrogels prepared from PLA/PEO/PLA triblock copolymers. J Mater

Sci Mater Med 13, 81-86.

Lian, T., Ho, R.J., 2001. Trends and developments in liposome drug delivery systems. J

Pharm Sci 90, 667-680.

Liang, H.F., Yang, T.F., Huang, C.T., Chen, M.C., Sung, H.W., 2005. Preparation of

nanoparticles composed of poly(gamma-glutamic acid)-poly(lactide) block copolymers and

evaluation of their uptake by HepG2 cells. J Control Release 105, 213-225.

197

Liu, L., Li, S., Garreau, H., Vert, M., 2000. Selective enzymatic degradations of poly(L-

lactide) and poly(epsilon-caprolactone) blend films. Biomacromolecules 1, 350-359.

Loftsson, T., Hreinsdottir, D., 2006. Determination of aqueous solubility by heating and

equilibration: a technical note. AAPS PharmSciTech 7, E4.

Lu, S., Anseth, K., 2000. Release Behavior of High Molecular Weight Solutes from

Poly(ethylene glycol)-Based Degradable Networks. Macromolecules 33, 2509-2515.

Macha S, Hughes PM, Mitra AK, 2003. Overview of Ocular Drug Delivery, in: Mitra, A.K.

(Ed.), Ophthalmic Drug delivery systems. Marcell Dekker, New york, USA.

Mainardes, R.M., Urban, M.C., Cinto, P.O., Khalil, N.M., Chaud, M.V., Evangelista, R.C.,

Gremiao, M.P., 2005. Colloidal carriers for ophthalmic drug delivery. Curr Drug Targets 6,

363-371.

Marchal-Heussler, L., Sirbat, D., Hoffman, M., Maincent, P., 1993. Poly(epsilon-

caprolactone) nanocapsules in carteolol ophthalmic delivery. Pharm Res 10, 386-390.

McVey, D., Hamilton, M.M., Hsu, C., King, C.R., Brough, D.E., Wei, L.L., 2008. Repeat

administration of proteins to the eye with a single intraocular injection of an adenovirus

vector. Molecular therapy : the journal of the American Society of Gene Therapy 16, 1444-

1449.

Merkli, A., Heller, J., Tabatabay, C., and Gurny, R., 1994. Semi-solid hydrophobic

bioerodible poly(ortho ester) for potential application in glaucoma filtration surgery. J

Control Release 29, 105-112.

Meyer, F., Wardale, J., Best, S., Cameron, R., Rushton, N., Brooks, R., 2012. Effects of

lactic acid and glycolic acid on human osteoblasts: A way to understand PLGA involvement

in PLGA/calcium phosphate composite failure. J Orthop Res.

198

Mishra, G., Gaudana, R., Tamboli, V., Mitra, A., 2010. Recent Advances in Ocular Drug

Delivery: Role of Transporters, Receptors, and Nanocarriers. , in: RI, M., AS, N. (Eds.),

Targeted Delivery of Small and Macromolecular Drugs. Taylor and Francis, New york, pp.

421-456.

Mishra, G.P., Bagui, M., Tamboli, V., Mitra, A.K., 2011a. Recent applications of liposomes

in ophthalmic drug delivery. Journal of drug delivery 2011, 863734.

Mishra, G.P., Tamboli, V., Jwala, J., Mitra, A.K., 2011b. Recent patents and emerging

therapeutics in the treatment of allergic conjunctivitis. Recent Pat Inflamm Allergy Drug

Discov 5, 26-36.

Mishra, G.P., Tamboli, V., Mitra, A.K., Effect of hydrophobic and hydrophilic additives on

sol-gel transition and release behavior of timolol maleate from polycaprolactone-based

hydrogel. Colloid Polym Sci 289, 1553-1562.

Mishra, G.P., Tamboli, V., Mitra, A.K., 2011c. Effect of hydrophobic and hydrophilic

additives on sol-gel transition and release behavior of timolol maleate from

polycaprolactone-based hydrogel. Colloid Polym Sci 289, 1553-1562.

Mitra, A.K., 2009. Role of transporters in ocular drug delivery system. Pharm Res 26, 1192-

1196.

Mitra, A.K., Velagaleti, P.R., Grau, U.M., 2010. Topical drug delivery systems for

ophthalmic use. Lux Biosciences, Inc.

Mitra, A.K., Velagaleti, P.R., Natesan, S., 2009. Ophthalmic compostions comprising

calcineurin inhibitors or mTOR inhibitors. Lux Biosciences, Inc.

Mitra, A.K., velagaleti, P.R., Natesan, S., 2011. Ophthalmic compositions comprising

calcineurin inhibitors or mTOR inhibitors. Lux Biosciences, Inc.

199

Miyajima, M., Koshika, A., Okada, J., Ikeda, M., Nishimura, K., 1997. Effect of polymer

crystallinity on papaverine release from poly(L-lactic acid) matrix. J Control Release 49,

207-215.

Moritera, T., Ogura, Y., Honda, Y., Wada, R., Hyon, S.H., Ikada, Y., 1991. Microspheres of

biodegradable polymers as a drug-delivery system in the vitreous. Investigative

ophthalmology & visual science 32, 1785-1790.

Munoz, B., West, S.K., Rubin, G.S., Schein, O.D., Quigley, H.A., Bressler, S.B., Bandeen-

Roche, K., 2000. Causes of blindness and visual impairment in a population of older

Americans: The Salisbury Eye Evaluation Study. Archives of ophthalmology 118, 819-825.

Murthy R., S.R., 1997. Controlled and Novel Drug Delivery. CBS Publisher, New Delhi.

Nanjawade, B.K., Manvi, F.V., Manjappa, A.S., 2007. In situ-forming hydrogels for

sustained ophthalmic drug delivery. J Control Release 122, 119-134.

Newman, D., Loredo, E., Bello, A., Grillo, A., eijoo, A., M ller, A., 2009. Molecular

Mobilities in Biodegradable Poly(dl-lactide)/Poly(ε-caprolactone) Blends. Macromol 42,

5219–5225.

Niesman, M.R., 1992. The use of liposomes as drug carriers in ophthalmology. Crit Rev Ther

Drug Carrier Syst 9, 1-38.

Niu, G., Yang, Y., Zhang, H., Yang, J., Song, L., Kashima, M., Yang, Z., Cao, H., Zheng, Y.,

Zhu, S., Yang, H., 2009. Synthesis and characterization of acrylamide/N-vinylpyrrolidone

copolymer with pendent thiol groups for ophthalmic applications. Acta Biomater 5, 1056-

1063.

Ogawa, Y., Okada, H., Yamamoto, M., Shimamoto, T., 1988. In vivo release profiles of

leuprolide acetate from microcapsules prepared with polylactic acids or

200

copoly(lactic/glycolic) acids and in vivo degradation of these polymers. Chem Pharm Bull

(Tokyo) 36, 2576-2581.

Okabe, K., Kimura, H., Okabe, J., Kato, A., Kunou, N., Ogura, Y., 2003. Intraocular tissue

distribution of betamethasone after intrascleral administration using a non-biodegradable

sustained drug delivery device. Investigative ophthalmology & visual science 44, 2702-2707.

Olsen, T.W., Edelhauser, H.F., Lim, J.I., Geroski, D.H., 1995. Human scleral permeability.

Effects of age, cryotherapy, transscleral diode laser, and surgical thinning. Investigative

ophthalmology & visual science 36, 1893-1903.

Olthoff, C.M., Schouten, J.S., van de Borne, B.W., Webers, C.A., 2005. Noncompliance with

ocular hypotensive treatment in patients with glaucoma or ocular hypertension an evidence-

based review. Ophthalmology 112, 953-961.

Packhaeuser, C.B., Kissel, T., 2007. On the design of in situ forming biodegradable

parenteral depot systems based on insulin loaded dialkylaminoalkyl-amine-poly(vinyl

alcohol)-g-poly(lactide-co-glycolide) nanoparticles. J Control Release 123, 131-140.

Papahadjopoulos, D., Jacobson, K., Nir, S., Isac, T., 1973. Phase transitions in phospholipid

vesicles. Fluorescence polarization and permeability measurements concerning the effect of

temperature and cholesterol. Biochim Biophys Acta 311, 330-348.

Park, J.H., Ye, M., Park, K., 2005. Biodegradable polymers for microencapsulation of drugs.

Molecules 10, 146-161.

Park P., J.S., 2006. J. Applied Poly. Sci. 100, 1983.

Peeters, L., Sanders, N.N., Braeckmans, K., Boussery, K., Van de Voorde, J., De Smedt,

S.C., Demeester, J., 2005. Vitreous: a barrier to nonviral ocular gene therapy. Investigative

ophthalmology & visual science 46, 3553-3561.

201

Peracchia, M.T., Vauthier, C., Puisieux, F., Couvreur, P., 1997. Development of sterically

stabilized poly(isobutyl 2-cyanoacrylate) nanoparticles by chemical coupling of

poly(ethylene glycol). J Biomed Mater Res 34, 317-326.

Pescina, S., Santi, P., Ferrari, G., Nicoli, S., 2011. Trans-scleral delivery of macromolecules.

Therapeutic delivery 2, 1331-1349.

Polak, M.B., Valamanesh, F., Felt, O., Torriglia, A., Jeanny, J.C., Bourges, J.L., Rat, P.,

Thomas-Doyle, A., BenEzra, D., Gurny, R., Behar-Cohen, F., 2008. Controlled delivery of 5-

chlorouracil using poly(ortho esters) in filtering surgery for glaucoma. Investigative

ophthalmology & visual science 49, 2993-3003.

Qiu, Y., Park, K., 2001. Environment-sensitive hydrogels for drug delivery. Adv Drug Deliv

Rev 53, 321-339.

Quigley, H.A., Broman, A.T., 2006. The number of people with glaucoma worldwide in

2010 and 2020. Br J Ophthalmol 90, 262-267.

Raghava, S., Hammond, M., Kompella, U.B., 2004. Periocular routes for retinal drug

delivery. Expert opinion on drug delivery 1, 99-114.

Rathore, K.S., 2010. Insitu gelling ophthalmic drug delivery system: An overview.

International Journal of Pharmacy and Pharmaceutical Sciences 2, 30-34.

Rosen, H.B., Chang, J., Wnek, G.E., Linhardt, R.J., Langer, R., 1983. Bioerodible

polyanhydrides for controlled drug delivery. Biomaterials 4, 131-133.

Ruel-Gariepy, E., Leroux, J.C., 2004. In situ-forming hydrogels--review of temperature-

sensitive systems. Eur J Pharm Biopharm 58, 409-426.

Ruiz-Moreno, J.M., Montero, J.A., Arias, L., Sanabria, M.R., Coco, R., Silva, R., Araiz, J.,

Gomez-Ulla, F., Garcia-Layana, A., 2006. Photodynamic therapy in subfoveal and

202

juxtafoveal idiopathic and postinflammatory choroidal neovascularization. Acta Ophthalmol

Scand 84, 743-748.

Sachdeva, R., Dadgostar, H., Kaiser, P.K., Sears, J.E., Singh, A.D., 2010. Verteporfin

photodynamic therapy of six eyes with retinal capillary haemangioma. Acta Ophthalmol.

Sahoo, S.K., Dilnawaz, F., Krishnakumar, S., 2008. Nanotechnology in ocular drug delivery.

Drug Discov Today 13, 144-151.

Schuman, J.S., 2000. Antiglaucoma medications: a review of safety and tolerability issues

related to their use. Clin Ther 22, 167-208.

Shell, J.W., 1984. Ophthalmic drug delivery systems. Surv Ophthalmol 29, 117-128.

Shen, Y., Tu, J., 2007. Preparation and ocular pharmacokinetics of ganciclovir liposomes.

AAPS J 9, E371-377.

Silva-Cunha, A., Fialho, S.L., Naud, M.C., Behar-Cohen, F., 2009. Poly-epsilon-

caprolactone intravitreous devices: an in vivo study. Investigative ophthalmology & visual

science 50, 2312-2318.

Singh, N.K., Das Purkayastha, B., Roy, J.K., Banik, R.M., Yashpal, M., Singh, G., Malik, S.,

Maiti, P., 2010. Nanoparticle-induced controlled biodegradation and its mechanism in

poly(epsilon-caprolactone). ACS applied materials & interfaces 2, 69-81.

Singh, S., Webster, D.C., Singh, J., 2007. Thermosensitive polymers: synthesis,

characterization, and delivery of proteins. International journal of pharmaceutics 341, 68-77.

Sinha, V.R., Bansal, K., Kaushik, R., Kumria, R., Trehan, A., 2004. Poly-epsilon-

caprolactone microspheres and nanospheres: an overview. Int J Pharm 278, 1-23.

Sintzel, M., B., Merkli, A., Heller, J., Tabatabay, C., and Gurny, R., 1997. Int J Pharm 155,

263.

203

Soppimath, K.S., Aminabhavi, T.M., Kulkarni, A.R., Rudzinski, W.E., 2001. Biodegradable

polymeric nanoparticles as drug delivery devices. J Control Release 70, 1-20.

Tamboli, V., Mishra, G.P., Mitra, A.K., 2011. Polymeric vectors for ocular gene delivery.

Ther Deliv 2, 523-536.

Tamboli, V., Mishra, G.P., Mitra, A.K., 2012. Biodegradable polymers for ocular drug

delivery, in: Mitra, A.K. (Ed.), Advances in Ocular Drug Delivery. Research Signpost, India.

Tamura, H., Miyamoto, K., Kiryu, J., Miyahara, S., Katsuta, H., Hirose, F., Musashi, K.,

Yoshimura, N., 2005. Intravitreal injection of corticosteroid attenuates leukostasis and

vascular leakage in experimental diabetic retina. Investigative ophthalmology & visual

science 46, 1440-1444.

Tang, Y., Singh, J., 2009. Biodegradable and biocompatible thermosensitive polymer based

injectable implant for controlled release of protein. International journal of pharmaceutics

365, 34-43.

Thassu D., D.M., and Pathak Y., 2007. Nanoparticulate Drug Delivery Systems

Informa Healthcare USA, Inc, New York.

Torchilin, V.P., 2005. Recent advances with liposomes as pharmaceutical carriers. Nat Rev

Drug Discov 4, 145-160.

Torchilin, V.P., 2007. Micellar nanocarriers: pharmaceutical perspectives. Pharm Res 24, 1-

16.

Tsukada K, Ueda S, R, O., 1984. Preparation of liposome-encapsulated antitumordrugs;

relation-ship between lipophilicity of drugs and in vitro drug release. Chem Pharm Bull 32,

1929-1935.

204

U. Adlund, A.C.A., 2002. Degradable polymer microspheres for controlled drug delivery.

Spinger-verlag berlin Heidelberg New york Germany.

van Kooij, B., Rothova, A., de Vries, P., 2006. The pros and cons of intravitreal

triamcinolone injections for uveitis and inflammatory cystoid macular edema. Ocul Immunol

Inflamm 14, 73-85.

Vandervoort, J., Ludwig, A., 2007. Ocular drug delivery: nanomedicine applications.

Nanomedicine (Lond) 2, 11-21.

Vauthier, C., Dubernet, C., Fattal, E., Pinto-Alphandary, H., Couvreur, P., 2003.

Poly(alkylcyanoacrylates) as biodegradable materials for biomedical applications. Adv Drug

Deliv Rev 55, 519-548.

Vega, E., Gamisans, F., Garcia, M.L., Chauvet, A., Lacoulonche, F., Egea, M.A., 2008.

PLGA nanospheres for the ocular delivery of flurbiprofen: drug release and interactions. J

Pharm Sci 97, 5306-5317.

Venkatraman, S.S., Jie, P., Min, F., Freddy, B.Y., Leong-Huat, G., 2005. Micelle-like

nanoparticles of PLA-PEG-PLA triblock copolymer as chemotherapeutic carrier. Int J Pharm

298, 219-232.

Vijayasekaran, S., Chirila, T.V., Hong, Y., Tahija, S.G., Dalton, P.D., Constable, I.J.,

McAllister, I.L., 1996. Poly(1-vinyl-2-pyrrolidinone) hydrogels as vitreous substitutes:

histopathological evaluation in the animal eye. J Biomater Sci Polym Ed 7, 685-696.

Visscher, G.E., Robison, R.L., Maulding, H.V., Fong, J.W., Pearson, J.E., Argentieri, G.J.,

1985. Biodegradation of and tissue reaction to 50:50 poly(DL-lactide-co-glycolide)

microcapsules. J Biomed Mater Res 19, 349-365.

205

Watsky, M.A., Jablonski, M.M., Edelhauser, H.F., 1988. Comparison of conjunctival and

corneal surface areas in rabbit and human. Current eye research 7, 483-486.

Woodruff, C.W., Peck, G.E., Banker, G.S., 1972. Dissolution of alkyl vinyl ether-maleic

anhydride copolymers and ester derivatives. J Pharm Sci 61, 1912-1916.

Wu, D.Q., Chu, C.C., 2008. Biodegradable hydrophobic-hydrophilic hybrid hydrogels:

swelling behavior and controlled drug release. J Biomater Sci Polym Ed 19, 411-429.

Yasukawa, T., Ogura, Y., Kimura, H., Sakurai, E., Tabata, Y., 2006. Drug delivery from

ocular implants. Expert Opin Drug Deliv 3, 261-273.

Yasukawa, T., Ogura, Y., Sakurai, E., Tabata, Y., Kimura, H., 2005. Intraocular sustained

drug delivery using implantable polymeric devices. Adv Drug Deliv Rev 57, 2033-2046.

Yasukawa, T., Ogura, Y., Tabata, Y., Kimura, H., Wiedemann, P., Honda, Y., 2004. Drug

delivery systems for vitreoretinal diseases. Prog Retin Eye Res 23, 253-281.

Yin, H., Gong, C., Shi, S., Liu, X., Wei, Y., Qian, Z., 2010. Toxicity evaluation of

biodegradable and thermosensitive PEG-PCL-PEG hydrogel as a potential in situ sustained

ophthalmic drug delivery system. J Biomed Mater Res B Appl Biomater 92, 129-137.

Young, S., Larkin, G., Branley, M., Lightman, S., 2001. Safety and efficacy of intravitreal

triamcinolone for cystoid macular oedema in uveitis. Clin Experiment Ophthalmol 29, 2-6.

Zeng, J., Chen, X., Liang, Q., Xu, X., Jing, X., 2004. Enzymatic degradation of poly(L-

lactide) and poly(epsilon-caprolactone) electrospun fibers. Macromolecular bioscience 4,

1118-1125.

Zeng L, X, W., 2010. Modeling the sustained release of lipophilic drugs from liposomes. .

Applied Physics Letters 97, 073701-073703.

206

Zentner, G.M., Rathi, R., Shih, C., McRea, J.C., Seo, M.H., Oh, H., Rhee, B.G., Mestecky,

J., Moldoveanu, Z., Morgan, M., Weitman, S., 2001. Biodegradable block copolymers for

delivery of proteins and water-insoluble drugs. J Control Release 72, 203-215.

Zhang, R., He, R., Qian, J., Guo, J., Xue, K., Yuan, Y.F., 2010. Treatment of experimental

autoimmune uveoretinitis with intravitreal injection of tacrolimus (FK506) encapsulated in

liposomes. Investigative ophthalmology & visual science 51, 3575-3582.

Zhao, Y., Fu, J., Ng, D.K., Wu, C., 2004. Formation and degradation of poly(D,L-lactide)

nanoparticles and their potential application as controllable releasing devices.

Macromolecular bioscience 4, 901-906.

Zhao, Z., Yang, L., Hu, Y., He, Y., Wei, J., Li, S., 2007. Enzymatic degradation of block

copolymers obtained by sequential ring opening polymerization of L-lactide and e-

caprolactone. Polymer Degradation and Stability 92, 1769-1777.

Zimmerman, C., Drewe, J., Flammer, J., and Shaarawy, T., 2004. Current eye research 21, 1.

207

VITA

Viral M. Tamboli was born on April 24, 1985, in Ahmedabad, Gujarat, India. She

completed her Bachelor of Pharmacy degree from Hemchandracharya North Gujarat

University in 2006. Following completion of her degree she joined Zydus Cadila as research

associate. She also worked at Sterling hospital in Ahmedabad as hospital pharmacist.

Viral Tamboli initiated her Doctoral studies at University of Missouri Kansas City in

the January, 2008. She is the active member of American Association of Pharmaceutical

Scientists (AAPS), Control Release Society (CRS), Association of Research in Vision and

Ophthalmology (ARVO) and Pharmaceutical Sciences Graduate Student Association

(PSGSA). She was awarded Chancellor’s doctoral fellowship from school of graduate studies

in the year 2011. She received travelship award from Astrazeneca to present her work at

AAPS 2010 annual meeting. She received travelship award from AAPS biotechnology

conference to present her work at AAPS 2011 annual meeting. She also received Amgen

travelship award for AAPS national biotechnology conference in 2012. She was awarded

Women’s council graduate assistance fund fellowship for her dissertation project. Viral

Tamboli has completed her doctoral studies in 2012 under the guidance of Dr. Ashim K.

Mitra and authored/co-authored several peer reviewed publications published in prestigious

books and journals