58
Centre for Drug Research Division of Biopharmaceutics and Pharmacokinetics Faculty of Pharmacy University of Helsinki Finland In vitro, in vivo, and in silico investigations of polymer and lipid based nanocarriers for drug and gene delivery Julia Lehtinen ACADEMIC DISSERTATION To be presented, with the permission of the Faculty of Pharmacy of the University of Helsinki, for public examination in Auditorium 2 at Viikki Korona Information Centre, Viikinkaari 11, on 7 th September 2013, at 12 noon. Helsinki 2013

In vitro, in vivo, and in silico investigations of polymer

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: In vitro, in vivo, and in silico investigations of polymer

Centre for Drug Research

Division of Biopharmaceutics and Pharmacokinetics

Faculty of Pharmacy

University of Helsinki

Finland

In vitro, in vivo, and in silico investigations of

polymer and lipid based nanocarriers

for drug and gene delivery

Julia Lehtinen

ACADEMIC DISSERTATION

To be presented, with the permission of the Faculty of Pharmacy of the University of

Helsinki, for public examination in Auditorium 2 at Viikki Korona Information Centre,

Viikinkaari 11, on 7th

September 2013, at 12 noon.

Helsinki 2013

Page 2: In vitro, in vivo, and in silico investigations of polymer

Supervisors: Professor Arto Urtti, Ph.D.

Centre for Drug Research

Faculty of Pharmacy

University of Helsinki

Finland

Professor Heike Bunjes, Ph.D.

Institute of Pharmaceutical Technology

Technische Universität Braunschweig

Germany

Mathias Bergman, Ph.D.

Karyon Ltd.

Helsinki

Finland

Reviewers: Docent Juha Holopainen, M.D., Ph.D.

Institute of Clinical Medicine

Faculty of Medicine

University of Helsinki

Finland

Professor Stefaan de Smedt, Ph.D.

Laboratory of General Biochemistry & Physical Pharmacy

Faculty of Pharmaceutical Sciences

University of Ghent

Belgium

Opponent: Academic Rector, Professor Jukka Mönkkönen, Ph.D.

University of Eastern Finland

Finland

© Julia Lehtinen 2013

ISBN 978-952-10-9039-4 (pbk.)

ISBN 978-952-10-9040-0 (PDF, http://ethesis.helsinki.fi)

ISSN 1779-7372

Helsinki University Print

Helsinki, Finland 2013

Page 3: In vitro, in vivo, and in silico investigations of polymer

3

Abstract

Nanomedicine research has expanded rapidly in the last decades. Several nanoparticle

formulations are accepted in clinical use, e.g. for the treatment of cancer, infections and

eye diseases, and also for diagnostics. Nanoparticle mediated drug delivery has many

potential advantages over the free drug, such as better pharmacokinetic profile, lowered

toxicity, and its possible use for cell-specific targeting and intracellular drug release.

Therapeutic genes can also be packed into nanocarriers to protect them from enzymatic

degradation and to mediate their cellular entry. The transfection efficacy of these synthetic

vectors is modest when compared to viral vectors, but they are considered to be safer.

Nonetheless, even though nanoparticles have so many advantages, there are many

extracellular and intracellular barriers to overcome before achieving successful drug or

gene delivery.

The focus of this research work was the formation and physico-chemical features of

lipid and polymer based nanoparticles for drug and gene delivery. In addition, two classes

of cancer cell targeting approaches were evaluated in biological and physical studies. First,

the effect of the polymeric gene carrier composition and structure on DNA condensation

efficacy, transgene expression, and cellular toxicity was examined. The linear architecture

and flexibility of poly(2-(dimethylamino)ethyl methacrylate) (PDMAEMA)-based block

co-polymers clearly enhanced DNA condensation and transfection efficiency. In addition,

by conjugating a membrane active protein, hydrophobin (HFBI) to DNA-binding cationic

dendrons, the transfection efficacy was increased compared to plain dendron. However,

cationic polymer-DNA complexes are prone to disruption by polyanions such as

glycosaminoglycans (GAG) in the extracellular space. We coated poly(ethyleneimine)

PEI/DNA complexes with anionic lipid mixture. The coating was able to protect the

contents against GAGs, and it could respond to the change of endosomal pH and release

the cargo inside the cells.

The next studies aimed to evaluate targeted liposomal cancer drug carriers in

physicochemical studies and in cancer cell models in vitro and in mice. A promising

activated endothelium targeting peptide (AETP) failed to target the liposomes to the cells.

Molecular modeling revealed that hydrophobic AETP was hidden in the PEG shield of the

liposomal surface thus it was not accessible for the target receptors. The last study

describes applicability of pre-targeting and local intraperitoneal administration of

liposomes for drug targeting to tumors located in peritoneal cavity. Epithelial growth

factor receptor (EGFR)-targeted liposomes bound specifically to ovarian cancer cells in

vitro. In the animal study, increased accumulation of liposomes in the xenograft tumors of

the mice was seen after intraperitoneal administration compared to intravenous

administration.

In conclusion, the composition and architecture of nanocarriers have a crucial impact

on DNA condensation, stability of the complexes and transfection efficacy. In liposomal

cancer drug targeting, polyethylene glycol (PEG) shield may hinder the targeting

efficiency of small molecular peptides. Intraperitoneal administration of liposomal drugs

seems to be promising route for targeting to tumors located in the peritoneal cavity.

Page 4: In vitro, in vivo, and in silico investigations of polymer

4

Acknowledgements

This study was carried out at the University of Kuopio, in the Department of

Pharmaceutics, (currently University of Eastern Finland, School of Pharmacy) during the

years 2003-2005 and it was continued at the University of Helsinki, Division of

Biopharmaceutics and Pharmacokinetics and the Centre for Drug Research (CDR) during

the years 2006-2013. This work has been financially supported by the Academy of

Finland, the Association of Finnish Pharmacies, the Finnish Cultural Foundation, the

Finnish Pharmaceutical Society, the National Agency of Technology (TEKES Finland),

the Science Foundation of Orion-Farmos, and the University of Helsinki. All financial

support is greatly acknowledged.

I want to express my deepest gratitude to my principal supervisor, Professor Arto Urtti

for his continuous support and optimistic attitude during these years. I am also very

grateful to my other supervisors, Professor Heike Bunjes for introducing me to the world

of nanoparticles and Mathias Bergman, Ph.D., for his skilful advice and guidance in

peptide targeting.

Professor Stefaan de Smedt and Docent Juha Holopainen are greatly acknowledged

for critical reading of this dissertation and for their valuable comments. I am honored that

Professor Jukka Mönkkönen has accepted the invitation to be my opponent in the public

defense of this thesis.

I would like to thank the current and former Deans of Faculty of Pharmacy in Kuopio

and in Helsinki, and Heads of Department of Pharmaceutics and Heads of Division of

Biopharmaceutics and Pharmacokinetics for providing excellent working facilities.

I wish to warmly thank my co-authors: Anu Alhoranta, M.Sc., Kim Bergström, Ph.D.,

Alex Bunker, Ph.D., Annukka Hiltunen, M.Sc., Zanna Hyvönen, Ph.D., Professor Olli

Ikkala, Raimo Ketola, Ph.D., Mauri Kostiainen, Ph.D., Katariina Lehtinen, M.Sc., Huamin

Liang, Ph.D., Aniket Magarkar, M.Sc., Ann-Marie Määttä, Ph.D., Jere Pikkarainen, Ph.D.,

Sari Pitkänen, M.Sc., Mari Raki, Ph.D., Tomasz Róg, Ph.D., Michał Stepniewski, M.Sc.,

Astrid Subrizi, M.Sc., Professor Heikki Tenhu, Päivi Uutela, Ph.D., Thomas Wirth, Ph.D.

and Professor Marjo Yliperttula for their valuable contribution to this work. It has been a

pleasure to collaborate with you all. I am also very grateful to Lea Pirskanen in Kuopio

and Leena Pietilä in Helsinki for their skilful and friendly assistance in laboratory. I also

wish to thank the personnel of Karyon ltd for welcoming me to do part of my Ph.D. work

in their laboratory.

My sincere thanks go to my friends and colleagues in the Faculty of Pharmacy in

Kuopio, and in the CDR and Drug Delivery and Nanotechnology (DDN) group in

Helsinki. Special thanks to the girls of the girls’ room: Astrid, Heidi, Johanna, Jonna,

Kati-Sisko, Mari, Marika, Martina, Melina and Polina for their friendship, joyful

company, and refreshing conversions.

Finally, I want to warmly thank my friends and relatives for their support and presence

Page 5: In vitro, in vivo, and in silico investigations of polymer

5

during these years. I owe my dearest gratitude to my husband and colleague Mika for his

love and support but also for scientific advice, and to our wonderful children Viljam and

Hilda for brightening up our everyday life.

Helsinki, July 2013

Julia Lehtinen

Page 6: In vitro, in vivo, and in silico investigations of polymer

6

Contents

Abstract 3

Acknowledgements 4

List of original publications 8

Abbreviations 9

1 Introduction 11

2 Review of the literature 13

2.1 Nanoparticles as drug and gene carriers 13

2.1.1 Liposomes 13

2.1.2 Polymeric carriers 15

2.1.3 Hybrid particles 17

2.2 Challenges in efficient drug and gene delivery 18

2.2.1 Stability of the vector 18

2.2.2 Tissue distribution and elimination 19

2.2.3 Cellular uptake 20

2.2.4 Intracellular distribution and cargo release 22

2.2.5 Diffusion in cytoplasm and nuclear import 23

2.3 Targeted cancer therapy 25

2.3.1 Passive targeting 25

2.3.2 Active targeting 26

2.3.2.1 Cancer cell targeting in solid tumors 27

2.3.2.2 Targeting to the tumor vasculature 27

3 Aims of the study 30

4 Overview of the methods 31

10 Summary of the main results 34

11 General discussion 36

Page 7: In vitro, in vivo, and in silico investigations of polymer

7

11.1 Structure-activity relationship of polymeric DNA carriers on DNA-complex

formation, transfection efficacy, and toxicity 36

11.2 Lipid-coated DNA-complexes as stable gene delivery vectors 37

11.3 Hindering effect of liposomal PEG on the targeting efficiency of a small

hydrophobic peptide, AETP 39

11.4 Pre-targeting and local administration of liposomes as potential approaches in

tumor targeting 40

12 Conclusions 43

13 Future prospects 44

Nanoparticles – drugs of the future? 44

References 45

Page 8: In vitro, in vivo, and in silico investigations of polymer

8

List of original publications

This thesis is based on the following publications:

I Anu M. Alhoranta, Julia K. Lehtinen, Arto O. Urtti, Sarah J. Butcher,

Vladimir O. Aseyev and Heikki J. Tenhu. Cationic amphiphilic star and

linear block copolymers: synthesis, self-assembly, and in vitro gene

transfection. Biomacromolecules 2011, 12, 3213-3222

II Mauri A. Kostiainen, Géza R. Szilvay, Julia Lehtinen, David K. Smith,

Markus B. Linder, Arto Urtti and Olli Ikkala. Precisely defined protein-

polymer conjugates: construction of synthetic DNA binding domains of

proteins by using multivalent dendrons. ACS NANO 2007, 1, 103-113

III Julia Lehtinen, Zanna Hyvönen, Astrid Subrizi, Heike Bunjes and Arto Urtti.

Glycosaminoglycan-resistant and pH-sensitive lipid-coated DNA complexes

produced by detergent removal method. Journal of Controlled Release 2008,

131, 145-149

IV Julia Lehtinen, Aniket Magarkar, Michał Stepniewski, Satu Hakola, Mathias

Bergman, Tomasz Róg, Marjo Yliperttula, Arto Urtti and Alex Bunker.

Analysis of course of failure of new targeting peptide in PEGylated

liposome: molecular modeling as a rationale design tool for nanomedicine.

European Journal of Pharmaceutical Sciences 2012, 46, 121-130

V Julia Lehtinen, Mari Raki, Kim A. Bergström, Päivi Uutela, Katariina

Lehtinen, Annukka Hiltunen, Jere Pikkarainen, Huamin Liang, Sari

Pitkänen, Ann-Marie Määttä, Raimo A. Ketola, Marjo Yliperttula, Thomas

Wirth and Arto Urtti. Pre-targeting and direct immunotargeting of liposomal

drug carriers to ovarian carcinoma. PLoS ONE 2012, 7(7):e41410

The publications are referred to in the text by their roman numerals

Page 9: In vitro, in vivo, and in silico investigations of polymer

9

Abbreviations

Ab antibody

AETP activated endothelium targeting peptide

αvβ3, αvβ5 integrins upregulated in proliferating endothelial cells

bp base pair

BSA bovine serum albumin

CHEMS cholesteryl hemisuccinate

CMC critical micelle concentration

CPP cell penetrating peptide

CMV cytomegalovirus

DMPG 1,2-dimyristoyl-sn-glycero-3-phospho-(1'-rac-glycerol)

DNA deoxyribonucleic acid

DOPE 1,2-dioleyl-sn-glycerol-3-phosphoethanolamine

DOTAP 1,2-dioleyl-3-trimethylammonium-propane

DPPC 1,2-dipalmitoyl-sn-glycero-3-phosphocholine

DSPE-PEG 1,2-distearoyl-sn-glycero-3-phosphoethanolamine-N-amino(polyethylene

glycol)

DSPG 1,2-distearoyl-sn-glycero-3-phospho-(1'-rac-glycerol)

EGFR epidermal growth factor receptor

Egg PC egg phosphatidyl choline

Egg SM egg sphingomyelin

EMA-DNA ethidiummonoazide

EPR enhanced permeability and retention

Fab´ antigen-binding fragment of antibody

FACS fluorescence-activated cell sorting

FITC fluorescein isothiocyanate

Fmoc 9-fluorenylmethyloxycarbonyl

GAG glycosaminoglycan

HII inverted hexagonal structure of lipid membrane

HDL high density lipoprotein

HER-2 human epidermal growth factor receptor 2

HFBI hydrophobin

HIV-1 human immunodeficiency virus 1

HPMA N-(2-hydroxypropyl)-methacrylamide copolymer

HSA human serum albumin

HSPC fully hydrogenated phosphatidyl choline

HUVEC human umbilical vein endothelial cell

IgG immunoglobulin G

IgM immunoglobulin M

Kd dissociation constant

Lα multilamellar structure or liquid crystalline phase of lipid membranes

Lβ gel phase of lipid membranes

LCDC lipid-coated DNA complexes

Page 10: In vitro, in vivo, and in silico investigations of polymer

10

LC-MS liquid chromatography - mass spectrometry

LDL low density lipoprotein

LUV large unilamellar vesicle

mAb monoclonal antibody

miRNA microRNA

MLV multilamellar vesicles

MPS mononuclear phagocyte system

mRNA messenger RNA

MTT (3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide

NCE new chemical entity

n/p nitrogen/phosphate ratio

ONPG ortho-nitrophenyl-β-D-galactopyranoside

PAMAM poly(amidoamine)

PBuA poly(n-butyl acrylate)

PC phosphatidylcholine

PDMAEMA poly(2-(dimethylamino)ethyl methacrylate)

PDP N-(3´-(pyridyldithio)propionoylamino

PE phosphatidylethanolamine

PEG polyethylene glycol

PEI poly(ethylene imine)

PG phosphatidylglycerol

PGA poly-L-glutamic acid

PL phospholipid

PLL poly-L-lysine

PS polystyrene

RGD arginine, lysine and aspartic acid containing peptide

Rho rhodamine

RNA ribonucleic acid

scFv single chain variable fragment of antibody

siRNA small interfering RNA

SPECT-CT single-photon emission computed tomography - computed tomography

TAT trans-acting activator of transcription

TRF time-resolved fluorescence

VEGFR vascular endothelial growth factor receptor

Page 11: In vitro, in vivo, and in silico investigations of polymer

11

1 Introduction

Discovery of new potent drug molecules has significantly improved the treatment of

serious illnesses, such as cancer and cardiovascular diseases. However, the development

of new compounds to serve as clinical drugs has become more and more difficult. This is

evident by the decreasing numbers of new chemical entities (NCE) that are introduced

annually for clinical use. Many new compounds face significant development challenges,

such as poor water-solubility or short half-life in the blood circulation. In many cases,

adverse effects do hamper treatment, particularly in the case of anti-cancer drugs. Drug

delivery systems can be used to modify the drug properties, for example by increasing

solubility, modulation of pharmacokinetics, and improving the safety of drug treatment.

Gene therapy was launched in the 1990s as an alternative to traditional drug therapy. It

presents a promising alternative for the correction of genetic deficiencies, e.g. haemophilia

or cystic fibrosis, but also for the treatment of acquired diseases, such as cancer and

cardiovascular conditions. Gene medicines can either induce protein translation in the

target cells (gene therapy) or silence the expression of the target protein (oligonucleotide

drugs, e.g. siRNA). These compounds (plasmid DNA, siRNA) cannot be delivered as

such, because they undergo rapid enzymatic degradation and do not reach target tissues.

The negative charges of DNA and RNA, and a large size of plasmid DNA prevent their

passage through cell membranes.

Nanocarriers are one possible solution to overcome the pharmacokinetic challenges in

drug and gene delivery. Nanocarriers, often generally referred to as nanoparticles, are

typically below one micrometer in diameter, usually consisting of lipids (e.g. liposomes,

lipid-DNA complexes), polymers (e.g. polymeric nanoparticles and micelles, polymer-

DNA complexes), peptides, proteins and/or metallic nanoparticles. Liposomes were

already developed in the 1960s by Alec Bangham and polymer nanoparticles in the 1970s

by Peter Speiser. Thereafter, the field of nanomedicines has expanded tremendously:

currently almost 20 000 publications are found in PubMed database with the search terms

‘nanoparticle and drug’.

By formulating a drug in a nanocarrier, the solubility and the pharmacokinetic profile

can be dramatically improved (Shi et al. 2010). Nanoparticles may also decrease the toxic

effects of the drug from off-target sites. For example, liposomal encapsulation of

doxorubicin lowered the risk of cardiotoxicity seven times compared to the free drug

(O'Brien et al. 2004). Nanocarriers can be tailored to release drugs in a controlled manner

or triggered by a change in environmental conditions and they can be targeted to desired

cell type expressing the target protein. In addition, they can be used to deliver several

drugs simultaneously as a combination therapy (Zhang et al. 2008). Complexation of DNA

and RNA into small nanoparticles masks their negative charges and protects them against

enzymatic degradation. When a massive molecule of plasmid DNA (mw of millions) is

condensed to a nanosized particle it is more suitable for systemic delivery. In this case,

nanoparticle has dual function of DNA protection from enzymatic catalysis and

augmenting cellular entry.

Around 40 nanoparticle formulations are now accepted in clinical use, mainly for the

treatment of cancer, but also for the treatment of infections, anemia, hypercholesterolemia,

Page 12: In vitro, in vivo, and in silico investigations of polymer

12

hepatitis, age-related macular degeneration, and in diagnostics (Duncan, Gaspar 2011).

The marketed products are so-called first generation nanomedicines that are not targeted.

The second generation targeted nanomedicines can bind to specific cellular antigens, but

they have not reached the market. Although targeted therapeutics have a lot of potential,

there are plenty of challenges and risks when more complicated formulations are designed

(Cheng et al. 2012).

Efficient but safe gene delivery vectors are still under development. For 500 million

years, viruses have developed a very efficient way to carry genetic material into cells.

Viral vectors are effective in DNA delivery, but their safety has not been totally verified.

In clinical studies, viral vectors have shown severe immunological reactions and even

caused patient death (Marshall 1999, Giacca, Zacchigna 2012). To date, three virus-based

gene therapy products have received market authorization from regulatory agencies in

China (Gendicine® and Oncorine

® for the treatment of cancer) and Europe (Glybera

®, for

the treatment lipoprotein lipase deficiency). Although non-viral polymer- and lipid-based

gene carriers lack the efficiency of the viral vectors, they are considered to be safer. In

addition, they are easier to synthesize and produce in large scale, and their DNA loading

capacity is higher than in the viral vectors (Kreiss et al. 1999). By learning from viruses,

more efficient synthetic carriers might also be developed.

In this study, the effect of the composition and architecture of polymer- and lipid-

based gene carriers on DNA condensation efficacy, transgene expression and cellular

toxicity was investigated. Furthermore, two types of liposomal cancer cell targeting

approaches were evaluated in physical and biological studies.

Page 13: In vitro, in vivo, and in silico investigations of polymer

13

2 Review of the literature

2.1 Nanoparticles as drug and gene carriers

The family of nanocarriers includes lipid-based carriers, such as liposomes and micelles,

polymer-based carriers, such as polymer conjugates, polymeric nanoparticles and

dendrimers, gold nanoparticles and carbon nanotubes. The size of the nanocarriers usually

varies between a few nanometers (polymer-drug conjugates, micelles and dendrimers) to

some hundreds of nanometers (liposomes and polymeric nanoparticles). Examples of

nanocarriers used for drug and gene delivery are presented in Figure 1. In the following

review, liposomes and polymeric nanocarriers for drug and gene delivery are discussed in

more detail.

Figure 1 Schematic illustration of different kinds of nanoparticles used for drug and gene

delivery.

2.1.1 Liposomes

Liposomes are spherical, self-assembling vesicles formed by one or several lipid bilayers

leaving an aqueous core inside. The lipid bilayer is composed of amphiphilic lipids,

derived from or based on the structure of biological membrane lipids. The hydrophobic

part of the lipid is usually formed of two hydrocarbon chains, which typically vary from 8

to 18 carbons in length and can be either saturated or non-saturated. Long and saturated

acyl chains form a membrane in gel phase (Lβ), resulting in increased stability and rigidity

of the liposomes. On the contrary, the use of short and/or unsaturated acyl chains results in

more fluid, liquid crystalline (Lα) bilayers. Incorporation of cholesterol into the lipid

bilayer minimizes the membrane permeability and improves the mechanical strength of

the liposomes. Surface charge of the liposome can be affected by varying the hydrophilic

head group of the lipid: being either zwitterionic (e.g. phosphatidylcholine (PC) and

Page 14: In vitro, in vivo, and in silico investigations of polymer

14

phosphatidylethanolamine (PE)), negatively charged (e.g. phosphatidylglycerol (PG)), or

positively charged (e.g. 3-trimethylammonium-propane (TAP)) (Figure 2) (Ulrich 2002).

Figure 2 Chemical structures of some phospholipids (fully hydrogenated soy

phosphatidylcholine (HSPC), 1,2-dioleyl-sn-glycerol-3-phosphoethanolamine (DOPE), 1,2-

distearoyl-sn-glycero-3-phospho-(1'-rac-glycerol) (DSPG), 1,2-dioleyl-3-trimethylammonium-

propane (DOTAP)) and cholesterol.

In water, amphiphilic lipids tend to form bilayers since they are poorly water soluble

with a critical micelle concentration (CMC) of 10-12

to 10-8

M. Spontaneously formed

multilamellar vesicles (MLV) are very heterogenous in lamellarity and in size, ranging

from 500 to 5000 nm. More sophisticated, small unilamellar vesicles (SUV, <100 nm) and

large unilamellar vesicles (LUV, 100–800 nm) can be prepared by sonication or extrusion

(Ulrich 2002, Torchilin 2005).

Hydrophobic drug molecules can be entrapped passively in the liposome bilayer during

the preparation of the liposomes using the aforementioned methods. Hydrophilic drugs are

encapsulated in the aqueous core of the liposome (or in the aqueous phase between

bilayers in the case of MLVs) using passive loading procedures, such as reverse phase

evaporation (Szoka Jr., Papahadjopoulos 1978), dehydration-rehydration method (Shew,

Deamer 1985), or active loading involving pH-gradient across the liposome membrane

(Mayer, Bally & Cullis 1986, Hwang et al. 1999). Remote loading of doxorubicin into

preformed liposomes using ammonium sulfate gradient as a driving force results in the

efficient and stable entrapment of the drug (Bolotin et al. 1994). Some liposomal cancer

drugs that are currently employed clinically, utilise remote loading; including Caelyx® and

Myocet® loaded with doxorubicin, and Daunoxome

® with daunorubicin.

Cationic liposomes can be used for complexation of negatively charged DNA or RNA

molecules. The formed complexes are called lipoplexes. The size of the highly cationic

Page 15: In vitro, in vivo, and in silico investigations of polymer

15

lipoplexes varies typically between 100 and 450 nm, whereas the lipoplexes carrying a

charge close to neutral are more heterogenic, varying from 350 to 1200 nm in diameter. In

lipoplexes, two types of structures have been observed; multilamellar structure (Lα), where

DNA is located as a monolayer between cationic membranes (Radler et al. 1997), or

inverted hexagonal structure (HII), where DNA is encapsulated within cationic lipid

monolayer tubes (Koltover et al. 1998) (Figure 3). To enhance gene delivery, so called

helper lipids, such as 1,2-dioleyl-sn-glycerol-3-phosphoethanolamine (DOPE), are often

mixed with cationic lipids to promote conversion of the lamellar phase into a hexagonal

structure (Hafez, Cullis 2000).

Figure 3 Schematic structures of lamellar (A) and inverted hexagonal phase (B) in the

cationic lipid/DNA complexes. Modified from Morille et al. (2008).

2.1.2 Polymeric carriers

Polymer based carriers can be divided into different categories by their structure; 1)

polymeric nanoparticles have a structure of a capsule or matrix, 2) polymeric micelles of

amphiphilic polymers with core-shell structure are spontaneously formed in water, 3)

polymersomes are polymeric vesicles, membrane bilayer constructed from amphiphilic

polymers, and 4) dendrimers are hyperbranched structures, composed of multiple

branched monomers emerging radially from the core (Cho et al. 2008, Brinkhuis, Rutjes &

Van Hest 2011). Drug is usually either linked covalently to a polymer or physically

entrapped into the polymer capsule or matrix (Rawat et al. 2006). Even though natural

polymers such as albumin, chitosan, and heparin have been used for the delivery of drugs

and genetic material, the synthetic polymers may be preferable because they can be

designed and synthesized to achieve required properties. Among various polymers tested

N-(2-hydroxypropyl)-methacrylamide copolymer (HPMA), polyethylene glycol (PEG),

and poly-L-glutamic acid (PGA) are examples of synthetic polymers used for drug

delivery (Cho et al. 2008). Albumin-bound paclitaxel (Abraxane®) is an example of a

nanoparticle formulation in clinical use for the treatment of metastatic breast cancer.

Cationic polymers are able to bind and condense DNA into polyplexes which are even

smaller than lipoplexes formed after condensation of DNA with cationic liposomes

(Dunlap et al. 1997). Poly(ethylene imine) (PEI), poly-L-lysine (PLL), and poly(2-

(dimethylamino)ethyl methacrylate) (PDMAEMA) are well known polymers for gene

delivery (Figure 4). Charge ratio between positive nitrogen atoms of polymer and negative

Page 16: In vitro, in vivo, and in silico investigations of polymer

16

phosphate groups of nucleic acid (n/p ratio) is an important factor in mediation of

transfection and toxicity. An excess of positive charges, n/p ratio 2.5 in the case of PEI, is

needed for total DNA condensation (Boeckle et al. 2006). Increasing the n/p ratio of

PEI/DNA complexes from 2 to 20 results in a decrease in a particle size from 1000 nm to

100-200 nm and a simultaneous reduction in a polydispersity (Erbacher et al. 1999). High

cationic charge of polyplexes leads to enhanced transfection efficiency, but as a drawback,

to higher toxicity because of the free polymer in solution (Hanzlíková et al. 2011).

It has been demonstrated that polymer architecture has an impact on the DNA

condensation and gene transfection properties of the polyplexes. Männistö et al. (2002)

demonstrated lower transfection activity for dendritic PLL compared to linear PLL. They

reasoned it might be due to unfavourable shape and orientation of dendritic amines for

DNA binding. However, in the case of PDMAEMA having a star-shaped architecture,

which mimics the structure of dendrimer, the polymer showed enhanced transfection

efficacy and reduced cytotoxicity compared to linear PDMAEMA (Xu et al. 2009). It has

been shown that the molecular weight of PEI strongly influences on the transfection

efficiency (Godbey, Wu & Mikos 1999). Choosakoonkriang et al. (2003) showed that both

branched and linear PEI (25 kDa) mediated higher transgene expression than smaller,

branched PEI 2 kDa. Linear PEI 22 kDa (ExGen 500) has proven to be more effective

than branched PEI 25 kDa in mediating transfection in lung epithelial cells both in vitro

and in vivo (Wiseman et al. 2003). However, linear PEI forms large unstable aggregates in

salt-containing medium that might explain its high gene delivery ability in vivo

(Wightman et al. 2001).

Figure 4 Chemical structures of some essential polymeric DNA carriers. Branched

poly(ethylene imine) (PEI), poly-L-lysine (PLL), poly(2-(dimethylamino)ethyl methacrylate)

(PDMAEMA), and polyethylene glycol (PEG).

Page 17: In vitro, in vivo, and in silico investigations of polymer

17

2.1.3 Hybrid particles

To design an optimal drug or gene delivery vehicle, a combination of lipids, polymers, and

proteins, forming different types of nanostructures, could be used. The materials can be

linked together covalently or mixed as a physical mixture. Probably the most well-known

modification of nanoparticles is the steric stabilization of the particle surface with a

hydrophilic polymer, most commonly PEG. PEG is a bio-compatible, water-soluble, and

chemically inert synthetic polymer. It generates a stealth effect on the surface of the

particles thus reducing aggregation, enhancing the stability, and prolonging the circulation

time in the body (Allen et al. 2002). The surface coverage of the particles is determined by

the molecular weight of the polymer as well as the graft density. PEG molecular weight of

2000–5000 has been shown to be the most effective in prolonging circulation half-life of

liposomes (Allen et al. 1991). It has been proposed that if the graft density of PEG2000 on

liposomes is below 5%, it takes the shape of a “mushroom” or a half sphere, and if the

density is high (> 5%), it takes an extended “brush” shape (Allen et al. 2002) (Figure 5).

The inclusion of 3-10% of PEG in liposomes has been shown to prolong their circulation

times (Allen et al. 1991).

The surface of the nanoparticles can be functionalized with targeting ligands, such as

antibodies, antibody fragments, or peptides that are able to bind to the specific cell types.

For the purpose of diagnostics and imaging, different kinds of markers such as radio

ligands or fluorescent markers can be used (Figure 5). Membrane-active, cationic peptides

on the surface of the particles have proven to enhance intracellular delivery (Kale,

Torchilin 2007, Ye et al. 2010). When the targeting ligands or membrane active peptides

are added to the surface of PEGylated particles, these ligands are usually coupled to the

end of the PEG chains rather than straight onto the particle surface (Hansen et al. 1995).

This minimizes the interference of the PEG shield thus enabling the interaction between

the ligand and the target cell or antigen (Shiokawa et al. 2005). The correct orientation of

the targeting moieties is important in order to achieve efficient interaction with the

receptors.

Figure 5 Functionalization of a liposome. Steric stabilization with PEG2000 at < 5 mol %

results in “a mushroom” shape of the PEG molecules (a), if the graft density is > 5 mol % PEG

takes“a brush” shape (b). Targeting antibody (c) and cell penetrating peptide (d) coupled to the

distal end of a PEG chain. Hydrophilic drugs or imaging agents can be encapsulated in the core

of the liposome (e) and lipophilic ones into the liposome bilayer (f).

Page 18: In vitro, in vivo, and in silico investigations of polymer

18

2.2 Challenges in efficient drug and gene delivery

Despite the great potential of drug and gene carriers, they face multiple challenges on their

way from the vial to the site of action (Figure 6). The vector has to remain stable during

storage, and also in physiological conditions. It should not be cleared too fast from the

blood circulation or cause immunological reactions. Intravenously administered carrier

must pass from the vasculature to the target tissue, bind, and internalize to the target cells.

The drug or gene should be released from the carrier at the target site and, in certain cases,

be imported into the nucleus (Mastrobattista, Koning & Storm 1999, Wang, Upponi &

Torchilin 2011). These issues should be taken into account during the development of new

nanocarrier systems.

Figure 6 Critical steps for efficient drug and gene delivery. Storage stability (1), stability and

half-life in blood circulation (2), extravasation from the blood stream (3), specific binding and

internalization into the target cells (4), escape from the endosomes and intracellular trafficking

(5), and nuclear localization (6).

2.2.1 Stability of the vector

In order to have a good nanoparticle formulation for clinical use, it should first be stable

during storage. Uncoated particles are prone to aggregation, which results from many

factors, such as ionic strength and pH of the solution, the initial size distribution of the

particles and storage temperature (Lee, Mount & Ayazi Shamlou 2001). Charge-neutral

complexes or complexes formed at low n/p ratios tend to aggregate because of

hydrophobic interactions or van der Waals forces. Whereas higher surface charge reduces

aggregation because of electrostatic repulsion (Tros de Ilarduya, Sun & Düzgüneş 2010).

In addition to physical stability, chemical stability also has to be taken into account; for

example, the lipids may be hydrolysed, resulting in lysolipids, and especially unsaturated

lipids can be oxidized easily. Hydrolysis and oxidation finally lead to degradation of

lipidic carriers. To enhance chemical stability, antioxidants can be added to the

Page 19: In vitro, in vivo, and in silico investigations of polymer

19

preparation, or liposomes can be stored as lyophilized powders, but the size distribution,

morphology, and entrapped cargo must be examined after reconstitution (Ulrich 2002).

Reserving the stability is even more difficult in physiological conditions. In blood

circulation there are serum proteins, mainly albumin, but also lipoproteins (high- and low

density lipoproteins, HDL and LDL) and many other proteins which may interact with

polymeric and lipidic particles. They can alter the complex diameter and zeta potential,

especially in the case of cationic complexes, and lead to premature release of encapsulated

material (Zelphati et al. 1998). Extracellular space and cell surface contain negatively

charged glycosaminoglycans (GAGs), components of connective tissues that are often

covalently linked to protein in the form of proteoglycans. Sulphated GAGs, such as

chondroitin sulphate and heparan sulphate are able to block the transfection of cationic

polyplexes and lipoplexes (Ruponen, Ylä-Herttuala & Urtti 1999, Ruponen et al. 2004).

To prevent aggregation and premature disruption in vivo, the carrier system should be

neutral in charge, and the particle size and shape should be optimal (Tao et al. 2011).

Interestingly, polymeric nanoparticles, having shapes like long cylinders (Geng et al.

2007) or elliptical discs (Muro et al. 2008), have demonstrated longer blood circulation

times than their spherical counterparts. Most of the current nanocarriers, however, are

cylindrical in shape, probably because of ease of manufacture. PEG-shield on the surface

of the particles can mask the possible charges and form a hydrated steric barrier against

aggregation (Tirosh et al. 1998, Erbacher et al. 1999).

2.2.2 Tissue distribution and elimination

To be able to find their targets in the body, therapeutic particles should remain long

enough in the blood circulation. Still, only a small fraction of the dose can reach the

tumor. In mice, 24 h from intravenous injection of PEGylated liposomes, roughly, only

0.5–5% of the injected dose has internalized the tumor xenograft, 10–20% still remains in

the blood circulation, 10–20% is up taken by the liver, and 2–5% is up taken by the spleen

(Chang et al. 2007, Chow et al. 2009, Lee et al. 2010). The defence mechanisms of the

body react rapidly against foreign material. The mononuclear phagocyte system (MPS)

consists of phagocytic cells in spleen, liver (Kuppfer cells), lungs, and lymph nodes.

Especially cationic or hydrophobic particles can interact with serum proteins, be

opsonised, and removed from the blood circulation by MPS (Allen et al. 1991, Dash et al.

1999). This can be seen as high accumulation of carrier systems in liver, spleen, and lungs.

Aggregation of the complexes may also lead to embolization in the lungs, which is

obviously life threatening (Morille et al. 2008). Opsonization can also activate the

complement system that induces phagocytosis and initiates inflammatory responses

against the foreign particles (Müller-Eberhard 1988). Vauthier et al. (2011) reported

binding of bovine serum albumin (BSA), fibrinogen, and a complement activating protein,

C3, on the nanoparticles consisting of poly(isobutylcyanoacrylate)-dextran co-polymers.

Adsorption of BSA on the nanoparticles following C3 protein binding activated the

complement cascade, while fibrinogen induced aggregation of the particles.

There are several approaches to reduce MPS recognition of the particles after i.v.

administration, referred in Harasym, Bally & Tardi (1998). The first method is to modify

Page 20: In vitro, in vivo, and in silico investigations of polymer

20

the particle surface properties, e.g. by incorporation of hydrophilic polymers and size

(favourably 50–150 nm) to inhibit the recognition of the immune system. In the second

approach, the phagocytic cells are saturated by increasing the dose of the particles beyond

the level that is required for therapy (>100 mg lipid/kg in murine studies). Certain drugs,

such as liposomal doxorubicin, can also be used to block the phagocytic system. From

these methods, the first option is more relevant because of its safety. Coating the liposome

surface with inert, hydrophilic polymers (e.g. PEG) provides steric stabilization against

interactions with opsonic factors. PEGylated liposomes show longer circulation times and

reduced uptake by MPS compared to conventional, non-PEGylated ones (Allen et al.

1991, Lu et al. 2004). PEG forms a highly hydrated shield around the liposome that has

been thought to sterically inhibit both electrostatic and hydrophobic interactions with the

serum proteins (Mastrobattista, Koning & Storm 1999, Ogris et al. 1999). However, there

is increasing evidence suggesting that PEG does not inhibit plasma protein binding on the

liposome surface (Moghimi, Szebeni 2003, Dos Santos et al. 2007). In fact, PEG may

even enhance complement activation via binding of immunoglobulin M (IgM) and IgG

(Moghimi, Szebeni 2003). Instead, the mechanism for prolonged circulation time provided

by PEG could be prevention of aggregation of the liposomes (Dos Santos et al. 2007).

Attachment of targeting ligands to the nanoparticle surface may lower the stability

and alter the pharmacokinetics of the carrier. Especially antibody-coupled carriers are

rapidly recognized by the immune system and cleared from the blood circulation. The

higher the antibody density on the particles, the faster the clearance (Aragnol, Leserman

1986). Harding et al.´s (1997) pharmacokinetic study with repeated injections of antibody-

coupled liposomes showed even more rapid clearance after second and third injection

compared to initial administration, evidencing immunogenicity of the formulation.

Interestingly, they also demonstrated that antibodies coupled to liposomes are more

immunogenic than free antibodies. Using smaller antibody fragments (Fab´, scFv) instead

of the whole antibody molecule, the half-life can be prolonged to almost the same level

with PEGylated, non-targeted liposomes (Maruyama et al. 1997, Pastorino et al. 2003b). A

prolonged circulation time is prerequisite for efficient accumulation of the particles to the

target site. The mechanism of the liposomal accumulation from blood circulation into the

tumors is discussed in section 2.3.1.

2.2.3 Cellular uptake

When the nanocarrier reaches the target tissue, for example tumor, it is facing the next

barrier, the cell membrane. There are two main routes for cellular uptake: endocytic and

non-endocytic. Endocytic cell uptake can occur via several pathways: clathrin-mediated

endocytosis, caveolae-mediated endocytosis, macropinocytosis, or phagocytosis

(Hillaireau, Couvreur 2009). Fusion and penetration of the nanocarriers through the

cellular membrane are examples of non-endocytic pathways (Xiang et al. 2012).

It has been shown that endocytosis (Figure 7) is the predominant route for

internalization of polymeric and lipidic nanoparticles (Wang, Upponi & Torchilin 2011).

Cationic particles trigger endocytosis by interacting non-specifically with the negatively

charged cell surface via cell membrane associated proteoglycans (Mislick, Baldeschwieler

Page 21: In vitro, in vivo, and in silico investigations of polymer

21

1996, Mounkes et al. 1998) or other anionic components on the cellular membrane. On the

other hand, it has been shown that cell-surface glycosaminoglycans can also inhibit

cellular uptake and gene expression of cationic DNA complexes (Ruponen et al. 2004).

Figure 7 Cellular uptake mechanisms of drug-loaded nanoparticles. Non-specific adsorption

and internalization via endocytosis (a), target-specific binding followed by receptor-mediated

endocytosis (b). If drug carrier binds to non-internalizing receptor, drug can be released outside

of the cell (c). Lipidic carriers can fuse with the cell membrane (d) or exchange the lipid

components with the cell membrane (e), leading to drug release inside the cell. Modified from

Torchilin (2005).

Receptor-mediated endocytosis can take place after specific binding of the targeting

ligand to its receptor. This is a very important mechanism by which the cells take up

nutrients and regulatory proteins, and it is nowadays also utilized in drug and gene

delivery. Binding to the receptor does not automatically mean rapid internalization into the

cell. In the case that the drug carrier is targeted to a non-internalizing receptor, the carrier

should release the drug outside the cell after which the free drug is taken up by the host

cell and also by the neighbouring cells (Figure 7). This kind of “bystander effect” might be

preferable in solid tumors where diffusion of large carrier systems is limited or all of the

cancer cells do not express the targeted antigens (Mastrobattista, Koning & Storm 1999,

Sapra, Allen 2003). However, liposomal drug carriers endocytosed via receptor binding

have been shown to have enhanced antitumoral efficacy over the carriers bound to non-

internalized receptors (Chuang et al. 2010).

Non-endocytic pathways are preferable for non-viral gene delivery because the

destructive effect of lysosomes is then usually avoided (Morille et al. 2008). To enhance

the internalization of drug and gene carriers, cationic membrane active peptides can be

coupled to the particle surface. These cell penetrating peptides (CPP), for example trans-

acting activator of transcription (TAT) peptide from HIV-1, can mediate intracellular

Page 22: In vitro, in vivo, and in silico investigations of polymer

22

delivery via endocytic, and according to some studies, also via non-endocytic pathways

(Torchilin et al. 2001, Bolhassani 2011). Nonetheless, the non-endocytic mechanism of

cellular penetration of CPPs can be considered to be ambiguous. In the study of Subrizi et

al. (2012), only the endocytic uptake mechanism for CPPs could be seen.

Lipidic carriers are able to fuse with the cellular membrane and directly release the

contents to the cytoplasm before entering the endocytic pathways. However, the main

uptake route for lipoplexes is the endocytic pathway, whilst fusion plays an important role

in releasing the DNA in endosomes (Hafez, Maurer & Cullis 2001, Xiang et al. 2012).

2.2.4 Intracellular distribution and cargo release

Following internalization via endocytic pathway, endosome capture and subsequent

lysosomal degradation are the major obstacles to efficient gene delivery. To be effective,

the vector, or at least its contents must be released from the endosome before its

maturation into the lysosome. DNA and RNA degrade easily in the lysosomal

compartments by hydrolytic enzymes. The endosomal release should happen rather fast

since after endocytosis, the endosomal vesicles mature into lysosomes in 10–20 min

(Simões et al. 2004).

For polymer-based vectors, two possible escape mechanisms have been proposed. The

first one, physical disruption of the negatively charged endosomal membrane via

interaction with cationic polymers has been suggested by Zhang, Smith (2000). They

noticed that high generation poly(amidoamine) (PAMAM) dendrimers were much more

effective than PLL in inducing lipid mixing and leakage of the contents. This escape

mechanism seems to depend also on the composition of the cellular membrane (cell type).

The other, better known mechanism, “proton sponge” effect can be applied by PEI,

PDMAEMA and PAMAM which contain protonable secondary and tertiary amines

having pKa-value of 5–7, near to endosomal pH (Boussif et al. 1995). The proton-sponge

hypothesis is based on high buffering capacity of the polymers. Increase in the endosomal

pH causes transportation of protons into the endosome that then results in an influx of

counter ions (Cl-). This promotes osmotic swelling and finally rupture of the endosomal

membrane (Boussif et al. 1995, Sonawane, Szoka & Verkman 2003).

Cationic lipid-based carriers are able to destabilize the anionic endosomal membrane

via electrostatic interactions. So called flip-flop-mechanism has been described by Xu,

Szoka (1996) and Zelphati, Szoka (1996). After endocytosis, the cationic complex

destabilizes endosomal membrane resulting in flip-flop of anionic lipids. The anionic

lipids diffuse into the complex, forming a charge neutral ion pair with cationic lipids. As a

consequence, entrapped nucleic acids dissociate from the complex and are released into

the cytoplasm. Destabilization of the endosomal membrane can also occur after lipid

phase transition. Helper lipid, DOPE, as discussed earlier, is able to acquire an inverted

hexagonal phase (HII) which is unstable and rapidly fuses and releases DNA or drug upon

adhering to endosomal vesicles (Koltover et al. 1998, Mönkkönen, Urtti 1998).

To release the DNA or the drug in a controlled manner at the desired target site,

delivery systems that are sensitive to a certain signal have been developed. The pH inside

the endosomes is 5–6, which is more acidic compared to its environment. The low pH can

Page 23: In vitro, in vivo, and in silico investigations of polymer

23

trigger the release of cargo from pH-sensitive carriers. A combination of

DOPE/cholesteryl hemisuccinate (CHEMS) is widely used in pH-sensitive formulations

(Kirchmeier et al. 2001, Simões et al. 2001, Shi et al. 2002b, Fattal, Couvreur & Dubernet

2004). In an acidic environment, anionic CHEMS becomes protonated, and this neutral

form induces formation of fusogenic hexagonal phase with DOPE. Whereas, at neutral or

alkaline pH, CHEMS stabilizes DOPE into a more stable lamellar phase (Hafez, Cullis

2000).

In addition to pH, increased temperature and enzymatic activity have been utilized to

trigger drug release. In tumors, the temperature is slightly higher than in healthy tissues,

but in practice the temperature difference is so small that it makes the controlled release

challenging. Drug release from thermosensitive carriers can be triggered by using

localized external heating. Kullberg, Mann & Owens (2009) used external heating up to

42 °C to trigger calcein release from temperature-sensitive 1,2-dipalmitoyl-sn-glycero-3-

phosphocholine (DPPC)-based immunoliposomes. The drug release is based on a sharp

gel to liquid crystalline phase transition of thermosensitive lipids at a certain temperature.

Paasonen et al. (2007) created a liposomal system where the contents of the liposomes

were released by UV-light induced heating of the gold nanoparticles incorporated into the

liposome bilayer. Local heating of the gold nanoparticles resulted in leakage of

thermosensitive liposomes. Enzymatically active carriers, for example human serum

albumin (HSA) nanoparticles, have shown degradation and drug release in the presence of

physically existing enzymes: trypsin, proteinase K, protease, pepsin, and intracellular

enzyme cathepsin B (Langer et al. 2008).

2.2.5 Diffusion in cytoplasm and nuclear import

Some drugs can act in the cytoplasm, but others, such as plasmid DNA and many

cytostatic drugs, must enter the nucleus to reach the site of action (Table 1). After escape

from the endosomes, the drug then faces the challenges of intracellular trafficking and

nuclear localization. For large molecular weight DNA in particular, these are difficult

barriers to overcome. Mobility of DNA in cytoplasm is slow because of the tight network

of cytoskeletal filaments, the presence of cell organelles, and high protein concentration

(Lechardeur, Verkman & Lukacs 2005). The diffusion rate of DNA depends strongly on

the size of the molecule. Plasmid DNA, containing 1 000–10 000 base pairs (bp), diffuses

much slower than DNA or RNA molecules under 250 bp (Dauty, Verkman 2005). Slow

diffusion makes DNA an easy target for cytoplasmic nucleases (Lechardeur, Verkman &

Lukacs 2005).

Page 24: In vitro, in vivo, and in silico investigations of polymer

24

Table 1 Examples of target sites inside the cell of different therapeutic agents.

If DNA remains complexed in the cytosol, the resistance against nucleases can be

increased (Lechardeur et al. 1999). Pollard et al. (1998) showed that 1% of cytoplasmic

DNA/PEI complexes entered the nucleus, which was 10 times more than uptake of free

plasmid DNA. Since passive diffusion into the nucleus via nuclear pore complexes is

limited for particles less than 10 nm in diameter, the nuclear entry of plasmid DNA can

occur mainly during cell division when the nuclear envelope is reformed (Görlich, Mattaj

1996). Toropainen et al. (2007) demonstrated substantially higher transgene expression in

dividing human corneal epithelial (HCE) cells compared to differentiated HCE cells after

transfection with PEI/DNA and DOTAP/DOPE/DNA complexes. More specifically,

higher nuclear accumulation has been seen in the cells which are close to mitosis phase

compared to the cells in post-mitotic phase (Gap 1) (Männistö et al. 2007). Männistö et al.

(2007) also demonstrated that despite the high amount of imported transgene in the

nucleus, only 10-6

to 10-4

parts were totally released from the carrier and thus available for

transcription. Release of DNA from the carrier is thus a critical step since premature

disassembly can lead to DNA degradation while incomplete release impairs gene

expression.

Therapeutic agent Site of action in the cell

Genetic drugs

plasmid DNA nucleus

oligonucleotides (siRNA, miRNA) mRNA, cytoplasm/nucleus

Small molecular anticancer drugs

doxorubicin DNA, nucleus

paclitaxel microtubules, cytoplasm

camptothecin DNA enzyme topoisomerase I, nucleus

Page 25: In vitro, in vivo, and in silico investigations of polymer

25

2.3 Targeted cancer therapy

Anti-cancer drugs are usually toxic for the cell, which is why they are undesirable in

healthy tissues. With targeted nanocarrier systems, the drug concentration could be

increased at the tumor site (ElBayoumi, Torchilin 2009) and thus the spreading of the

harmful drug to the normal tissues may be reduced. Targeting also improves

internalization of the drug into cancer cells (Mamot et al. 2003, Dubey et al. 2004). Tumor

targeting can be divided into two types; passive and active (Figure 8).

Figure 8 Passive targeting and active targeting of nanoparticles. A. Passive targeting is

based on leaky vasculature of the tumor and long circulation time of nanocarriers. B. In active

targeting, nanocarriers can be targeted to the receptors overexpressed either on tumor cells (1) or

on angiogenic endothelial cells (2). Modified from Danhier, Feron & Preat (2010).

2.3.1 Passive targeting

When tumor volume reaches 1–2 mm3, it starts to form new blood vessels in order to bring

oxygen and nutrients to the growing cells (Feron 2004). This blood vessel formation is

called angiogenesis. The morphology of tumor vasculature differs from the normal

vessels. The tumor vessel endothelium is malformed and leaky; having 100–600 nm gaps

between the endothelial cells, whereas normal endothelial cells form a continuous,

uniform monolayer (Yuan et al. 1995, Hashizume et al. 2000). Pericytes, the cells

surrounding the endothelial cells, are also malformed in angiogenic tumor vessels

(Morikawa et al. 2002). Thus, small 50–200 nm particles can enter the tumor. Moreover,

due to a non-functional, or absent, lymphatic drainage system, nanoparticles can be also

retained in the tumor interstitium. This phenomenon is called the “Enhanced permeability

and retention” (EPR) effect and is utilized in passive targeting of nanoparticles into tumor

(Maruyama 2011). Because of the EPR effect, it is possible to achieve even 10–50 fold

local concentrations of nanoparticles in tumor compared to normal tissues (Iyer et al.

2006).

The properties of the nanocarriers can influence on the EPR effect. The carriers should

have a long half-life in blood in order to have enough time for efficient tumor

Page 26: In vitro, in vivo, and in silico investigations of polymer

26

accumulation. Optimally, the size of the particle should be more than 10 nm to avoid

filtration through the kidneys, but under 100 nm to avoid a capture by the liver (Danhier,

Feron & Preat 2010). In addition, they should be sterically stabilized to avoid aggregation

and rapid recognition by the MPS system.

The tumor environment also provides some barriers for successful therapy, such as

heterogenous blood flow, increased interstitial fluid pressure, and large transport distances

in the tumor interstitium (Jain 1990, Harrington et al. 2000). Harrington et al. (2000)

demonstrated the influence of tumor size on the uptake of PEGylated liposomes. In large

tumors, the uptake of liposomes is probably reduced due to higher osmotic pressure and

because of a relatively low vascular volume, reflecting areas of poor perfusion or even

necrotic areas.

2.3.2 Active targeting

To mediate active tumor targeting, cancer cell specific targeting ligands are attached to the

surface of the nanocarrier. The chosen ligand actively binds to the receptors that are either

selectively expressed or overexpressed in cancerous cells compared to normal cells (Sapra,

Allen 2003). The most commonly used ligands for liposome and nanoparticle targeting

are: monoclonal antibodies (mAb) (ElBayoumi, Torchilin 2009), fragments of the

antibodies (Fab or svFc) (Pastorino et al. 2003b, Iyer et al. 2011), growth factors (Lee et

al. 2010), peptides (Moreira et al. 2001, Temming et al. 2005, Xiong et al. 2005), small

molecule ligands (such as folate and transferrin) (Gabizon et al. 1999, Voinea et al. 2002,

Riviere et al. 2011), sugars (such as galactosamine, lactose, and trivalent galactose) (David

et al. 2004), and aptamers (Tong et al. 2010) (Table 2).

To attach the ligands on the sterically stabilized liposomes, the ligands are preferably

coupled to the termini of the PEG chains. When the ligands are attached on the bilayer of

the liposome, the PEG may serve as a steric hindrance for both ligand coupling and later

on for binding to the receptors, especially in the case of small molecular weight ligands

(Sapra, Allen 2003). The end group of the PEG-spacer can be functionalized for the

chemical ligand coupling, for example with maleimide (Kirpotin et al. 1997) or N-(3´-

(pyridyldithio)propionoylamino (PDP) (Allen et al. 1995) for the thiol-containing ligands,

or with biotin for avidin-coupled ligands (Loughrey, Bally & Cullis 1987). The amount of

targeting ligands attached on the liposomes is crucial since excessive ligand density leads

to rapid clearance of the liposomes, while insufficient ligand density fails to facilitate

satisfactory targeting efficiency. Only 10–20 molecules of whole targeting antibody or

Fab´ fragments per liposome are required for sufficient internalization to the target cell

(Park et al. 1997, Iden, Allen 2001). For antibody density in excess of 35

molecules/liposome, an increased rate of clearance has been reported (Allen et al. 1995).

In the case of small peptides, even 200–500 peptide molecules/liposome did not cause

highly elevated blood clearance compared to PEGylated, non-targeted, liposomes

(Zalipsky et al. 1995).

Active tumor targeting could be achieved by direct targeting, where the targeting

ligands are coupled straight on the drug carrier, or by a pre-targeting (multistep) approach.

In the pre-targeting method, ligands are not covalently linked to the carrier system;

Page 27: In vitro, in vivo, and in silico investigations of polymer

27

instead, the target-specific ligand is administered as a first step. Once ligand has bound to

the target receptor, the ligand-binding drug-containing nanoparticles are administered.

Pre-targeting is commonly based either on biotin-avidin binding, that shows extremely

high affinity of Kd ~ 1015

M-1

, (Weber et al. 1989, Lesch et al. 2010) or on bispecific

antibodies (Sharkey et al. 2003). Pre-targeting has been utilized in targeting of polymeric

nanoparticles (Nobs et al. 2006, Pulkkinen et al. 2008) and liposomes (Xiao et al. 2002,

Pan et al. 2008).

2.3.2.1 Cancer cell targeting in solid tumors

To reach the cancer cells throughout the solid tumor, nanocarriers should extravasate from

blood circulation to tumor interstitial space and diffuse evenly around. Moreover, the

target receptors should be expressed homogenously on all targeted cells. The most

targeted receptors in solid tumors are: 1) human epidermal growth factor receptors (EGFR

and HER-2), overexpressed in many tumor types, e.g. in breast, colon, ovarian, pancreatic,

head and neck cancer, and non-small cell lung cancer; 2) transferrin receptor, which

participates in iron transfer, expressed in tumor cells 100-fold more than in normal cells;

and 3) folate receptor, which takes care of folic acid intake and is also overexpressed in

many human cancer types (Danhier, Feron & Preat 2010).

Even though a high affinity to the target receptor is desirable, it can also limit the

tumor penetration properties of the nanocarriers. Adams et al. (2001) showed that with

low affinity (Kd = 3.2 x 10-7

M), anti HER2/neu scFv exhibited broad diffusion from the

vasculature to the tumor, whereas the high affinity scFv (Kd = 1.5 x 10-11

M) failed to

traverse more than 2-3 cell diameters. This study was done with labelled scFv, (molecular

weight 27 kDa) without nanocarrier. After coupling these high affinity ligands onto

nanoparticles or liposomes (molecular weight of millions), even more restricted spreading

to the tumor would be expected. Interestingly, Sugahara et al. (2010) showed that co-

administration of non-conjugated tumor-penetrating peptide (iRGD) improved tumor

tissue penetration and therapeutic efficacy of free doxorubicin, nanoparticles

(Abraxane®), doxorubicin liposomes, and antibody trastuzumab. The mechanism of tumor

penetration for iRGD is distinct from the passive EPR-effect, since it is receptor-mediated

and energy-dependent.

2.3.2.2 Targeting to the tumor vasculature

Tumor vasculature targeting aims to obstruct the blood supply of the tumor. That leads to

a lack of nutrients and oxygen, in turn causing tumor cell starvation and death. When

compared to tumor cell targeting, vascular targeting has some advantages: 1) direct

accessibility to the endothelial cells from blood circulation avoiding the problems related

to poor extravasation and tumor tissue penetration; 2) high efficacy, since one tumor

capillary supplies hundreds of tumor cells; 3) avoidance of drug resistance, because

endothelial cells are genetically stable compared to tumor cells that may become resistant

Page 28: In vitro, in vivo, and in silico investigations of polymer

28

to the therapy; 4) broad applicability, since most solid tumors are dependent on

neovascularization (Mastrobattista, Koning & Storm 1999, Feron 2004).

Besides during tumor growth, angiogenic vessels are formed also in some non-

malignant conditions such as in atherosclerosis, wound healing, psoriasis, and in certain

eye diseases, e.g. in the wet form of age-related macular degeneration and

neovascularisation of the cornea. Angiogenic vessels express markers such as vascular

endothelial growth factor receptor (VEGFR) and integrins (αvβ3 and αvβ5) that are not

present in the resting blood vessels of normal tissues (Ruoslahti 2002). The integrins are

also upregulated in different tumor cells, including metastic melanoma cells (Conforti et

al. 1992, Seftor, Seftor & Hendrix 1999). Integrins can be specifically recognized by

RGD-peptide, consisting of arginine, glycine, and aspartic acid. RGD-peptide was found

by screening of phage display peptide libraries (Pasqualini, Koivunen & Ruoslahti 1997)

and it is one of the most studied tumor vasculature homing peptides.

Page 29: In vitro, in vivo, and in silico investigations of polymer

29

Table 2 Nanocarrier-based targeted therapeutics in clinical development and examples of

preclinical studies.

Targeting ligand Formulation Target Study

phase

Reference

Clinical trials

stomach cancer

specific GAH mAb

PEGylated liposomal

doxorubicin (MCC-

465)

tumor antigen in

stomach cancer

I Reviewed in

(Cheng et al.

2012)

anti-transferrin

receptor scFv

liposomal p53 plasmid

DNA (SGT-53)

transferrin receptor I

human transferrin liposomal oxaliplatin

MBP-426)

transferrin receptor I/II

human transferrin siRNA loaded

nanoparticles

(CALAA-01)

transferrin receptor I

peptide docetaxel-loaded

polymeric

nanoparticles (BIND-

014)

prostate specific

antigen

I

Preclinical studies

cancer cell specific

mAb 2C5

PEGylated liposomal

doxorubicin

cancer cell surface bound

nucleosomes

(ElBayoumi,

Torchilin 2009)

Fab´ fragment of

cetuximab

PEGylated liposomal

doxorubin/vinorelbin

epidermal growth factor

receptor

(Mamot et al.

2005)

mesothelioma

targeting scFv

(M1)

PEGylated 111

In-

labeled liposomes

surface antigens on human

mesothelioma tumor cells

(Iyer et al. 2011)

epidermal growth

factor

111In-labeled

polymeric micelles

epidermal growth factor

receptor

(Lee et al. 2010)

A10 aptamer polymer-paclitaxel

conjugates

prostate-specific membrane

antigen

(Tong et al.

2010)

folate PEGylated liposomal

doxorubicin

folate receptor (Riviere et al.

2011)

cyclic RGD-

peptide

PEGylated liposomal

5-fluorouracil

αvβ3 integrins (Dubey et al.

2004)

NGR-peptide PEGylated liposomal

doxorubicin

angiogenic endothelial cell

marker aminopeptidase N

(Pastorino et al.

2003a)

Fab´ fragment of

anti-VEGFR-2

PEGylated liposomal

doxorubicin

vascular endothelial growth

factor receptor

(Roth et al.

2007)

mAb anti-

disialoganglioside

and NGR-peptide

PEGylated liposomal

doxorubicin

disialoganglioside receptor and

aminopeptidase N

(Pastorino et al.

2006)

Page 30: In vitro, in vivo, and in silico investigations of polymer

30

3 Aims of the study

The general objective of this study was to develop and evaluate lipid and polymer based

nanocarriers for targeted drug and gene delivery using in vitro, in vivo, and in silico

methods. The specific aims were:

1. To investigate the effects of the architecture and flexibility of cationic amphiphilic

star and linear PDMAEMA-based block copolymers on DNA complex formation,

in vitro transfection efficiency, and cytotoxicity.

2. To determine the DNA binding ability, in vitro transfection efficiency, and

cytotoxicity of novel BSA- and hydrophobin (HFBI)-dendron conjugates.

3. To develop an extracellularly stable gene delivery vector that can release its

contents at the acidic endosomal pH.

4. To investigate a targeted liposomal drug delivery system when novel activated

endothelium targeted peptide (AETP) is used as a targeting ligand.

5. To explore an epidermal growth factor receptor (EGFR) targeted liposomes using

direct targeting and pre-targeting approaches.

Page 31: In vitro, in vivo, and in silico investigations of polymer

31

4 Overview of the methods

The materials and prepared formulations for gene delivery (I – III) and for targeted

liposomal drug delivery (IV, V) are summarized in Table 3. The cell lines used in the

original publications are shown in Table 4. General methods of physicochemical and

biological studies are shown in Tables 5 and 6, respectively. In addition, a computational

study combining molecular dynamics simulation and ligand-protein docking was

performed (IV). Materials and methods are described in detail in the publications.

Table 3 Summary of materials and prepared formulations in the publications.

Material/Formulation Source/Preparation method Publication

plasmid DNA (pCMVβ) encoding β-

galactosidase

amplification in E.coli, purification by

column separation

I - III

Polymeric materials

PDMAEMA block copolymers:

(PS-PDMAEMA)6,

(PBuA-PDMAEMA)6,

PDMAEMA-PS-PDMAEMA,

PDMAEMA-PBuA-PDMAEMA

synthesized, purified and characterized

in the Department of Chemistry,

Laboratory of Polymer Chemistry,

University of Helsinki

I

protein-polyamine dendron

conjugates:

HFBI, HFBI-G1, HFBI-G2,

BSA, BSA-G1, BSA-G2

synthesized, purified and characterized

in the Department of Engineering,

Physics, and Mathematics, and Center

for New Materials, Helsinki University

of Technology

II

other polymers:

PEI 25K, PLL

commercially available I - III

Lipidic materials

DOPE, CHEMS, Egg PC, Egg SM,

DMPG, cholesterol

commercially available III

HSPC, cholesterol, DSPE-PEG2000,

DSPE-PEG2000-maleimide,

DSPE-PEG2000-biotin

commercially available IV, V

Targeting ligands

AETP phage display and peptide synthesis

using Fmoc chemistry in Karyon Ltd.,

Finland

IV

cetuximab commercially available V

Formulations

cationic polyplexes self-assembling I - III

lipid-coated polyplexes detergent removal III

calcein containing liposomes reverse-phase evaporation III

PEGylated liposomes extrusion IV, V

Page 32: In vitro, in vivo, and in silico investigations of polymer

32

AETP-targeted liposomes extrusion + post-insertion method of

AETP

IV

cetuximab-targeted liposomes extrusion + biotin-avidin-biotin binding

of cetuximab

V

doxorubicin- liposomes extrusion + remote-loading of

doxorubicin

IV, V

CMV = cytomegalovirus, PDAMEMA = poly(2-(dimethylamino)ethyl methacrylate), PS = polystyrene , PBuA = poly(n-butyl acrylate), HFBI = hydrophobin, BSA = bovine serum albumin, G1 = first generation, G2 = second generation, PEI = poly(ethylene imine), PLL = poly-L-lysine, DOPE = 1,2-dioleyl-sn-glycerol-3-phosphoethanolamine, CHEMS = cholesteryl hemisuccinate, Egg PC = egg phosphatidyl choline, Egg SM = egg sphingomyelin, DMPG = 1,2-dimyristoyl-sn-glycero-3-phospho-(1'-rac-glycerol), HSPC = fully hydrogenated phosphatidyl choline, DSPE-PEG2000 = 1,2-distearoyl-sn-glycero-3-phosphoethanolamine-N-[amino(polyethylene glycol)-2000], AETP = activated endothelium targeting peptide, Fmoc = 9-fluorenylmethyloxycarbonyl

Table 4 The cells used in the publications.

Table 5 Physicochemical characterization methods used in the publications.

Study objective Method Probe Publication

particle size dynamic light scattering - I, III - V

DNA-complex integrity ethidium bromide intercalation

assay

ethidium

bromide

I-III

pH-sensitivity calcein-dequenching assay calcein III

phospholipid (liposome)

concentration

fluorescence fluorescein-PE IV, V

liposomal drug

concentration

absorbance doxorubicin IV, V

ligand coupling efficiency fluorescence tryptophan IV, V

Cell line Type Species Publication

ARPE-19 retinal pigment epithelial cell line human I

C8161 melanoma cell line human IV

CV1 kidney fibroblast cell line monkey I-III, V

primary

HUVEC

umbilical vein endothelial cells human IV

KS1767 kaposi’s sarcoma cell line human IV

SKOV-3 ovarian adenocarcinoma cell line human V

SKOV3.ip1 ovarian adenocarcinoma cell line human V

SVEC4-10 lymph node endothelial cell line mouse IV

Page 33: In vitro, in vivo, and in silico investigations of polymer

33

Table 6 Biological methods used in the publications.

Study objective Method Probe Publication

in vitro studies

cytotoxicity/therapeutic activity

of DNA-complexes and

liposomes

MTT-assay formazan II - IV

Alamar Blue®-assay resorufin I, V

transfection efficacy ONPG-assay ONPG I - III

cellular affinity/uptake of the

formulations

FACS-analysis EMA-DNA III

fluorescein IV - V

in vivo studies

pharmacokinetics: half-life of

liposomes

TRF Europium IV

biodistribution of liposomes TRF Europium IV

confocal microscopy Rho/FITC IV

LC-MS analysis doxorubicin V

SPECT-CT and

gammacounting

99mTechnetium V

MTT = 3-(4,5-dimethylthiazol-2-yl)-2,5-diphenyltetrazolium bromide, ONPG = ortho-nitrophenyl-β-D-galactopyranoside, FACS = fluorescence-activated cell sorting, EMA-DNA = ethidiummonoazide, TRF = time-resolved fluorescence, Rho = rhodamin, FITC = fluorescein isothiocyanate, LC-MS = liquid chromatography - mass spectrometry, SPECT-CT = single-photon emission computed tomography - computed tomography

Page 34: In vitro, in vivo, and in silico investigations of polymer

34

10 Summary of the main results

The main results of the experimental data and molecular modeling are presented in

Table 7.

Table 7 Summary of the main results.

Gene delivery vectors Publi-

cation

DNA-binding

properties

All PDMAEMA block co-polymers bound DNA. Only (PS-

PDMAEMA)6 was not able to condense DNA totally.

I

HFBI-G1, HFBI-G2 and BSA-G2 showed the highest binding

capacity towards DNA.

II

LCDCs condensed DNA completely. After treatment with

anionic dextran sulphate, LCDCs prepared by detergent dialysis

method were stable at least for 24 h, while LCDCs prepared by

ethanol injection method had released ~50% of DNA at 24 h

time point.

III

pH-sensitivity LCDCs were stable at neutral pH, but they were able to fuse with

calcein-containing endosome-mimicking liposomes at endosomal pH

(5-6).

III

Transfection

efficacy in

vitro

Linear PDMAEMA-PBuA-PDMAEMA was able to transfect ARPE-

19 cells with efficacy comparable to PEI, and CV-1 cells with

efficacy of 1/10 of that of PEI. The other carriers showed some

transfection efficacy in the following order: PDMAEMA-PS-

PDMAEMA ≈ (PBuA-PDMAEMA)6 > (PS-PDMAEMA)6.

I

Of the protein-dendron conjugates, only HFBI-G2 was able to

transfect CV-1 cells. Transfection efficacy was only 1/20 of that of

PEI.

II

Transfection efficiency of LCDCs was ~80% of that of PEI/DNA

polyplexes in the absence of serum. In the presence 10% serum, the

transgene expression level decreased dramatically.

III

Cytotoxicity

in vitro

All PDMAEMA block co-polymers showed over 80% viability in

ARPE-19 and CV-1 cells at low n/p ratios (0.5–4). With increasing

n/p ratio, the cell viability decreased gradually.

I

All protein-dendron conjugates showed over 80% viability in CV-1

cells at n/p low ratios (0.125–4). HFBI-G1 and HFBI-G2 were

observed to be slightly cytotoxic at higher n/p ratios.

II

Cancer-targeted liposomes

Cellular

affinity in

vitro

AETP did not increase the cellular affinity of liposomes in HUVEC-

cells.

IV

EGFR-targeted liposomes had 22–38 times higher affinity towards

SKOV-3 cells and 13–17 times higher affinity towards CV-1 cells

compared to affinity of non-targeted liposomes. Competition with

V

Page 35: In vitro, in vivo, and in silico investigations of polymer

35

free cetuximab decreased the affinity of the targeted liposomes to the

level of the non-targeted liposomes. Pre-targeting showed lower

cellular association than direct targeting.

Cytotoxic

activity in

vitro

AETP did not increase the cytotoxic efficacy of doxorubicin-loaded

liposomes in HUVEC or SVEC4-10 cells.

IV

EGFR-targeted doxorubicin-loaded liposomes showed slightly higher

cytotoxicity than non-targeted liposomes in SKOV-3 cells. However,

free DXR was significantly more cytotoxic than both of the liposome

types.

V

Pharmacokine

tics in vivo

AETP-targeted liposomes showed similar elimination half-life (~7–

10 h) to non-targeted liposomes (~6–7 h).

IV

Cellular

affinity in

vivo

Significant differences in co-localization of liposomes and

endothelial cells could not be seen between the AETP-targeted and

non-targeted liposomes in confocal microscopy.

IV

Uptake in

tumor

No significant difference in tumoral uptake between AETP-targeted

and non-targeted liposomes could be seen.

IV

After direct targeting or pre-targeting, no significant difference in

uptake in tumor between EGFR-targeted and non-targeted liposomes

could be seen at 24 h time point. However, intraperitoneal

administration of the liposomes, targeted or not, led to faster and

higher tumoral accumulation of the liposomes than intravenous

administration.

V

Orientation of

the targeting

ligands in

silico

Computational modeling revealed that both AETP and RGD peptides

located in the PEG region of the PEGylated liposomes. However,

AETP was more covered by the PEG chains, while RGD was more

exposed to the solvent.

IV

Affinity

towards

serum

albumin in

silico

Protein-ligand docking showed HSA having 12 times stronger

binding affinity to PEG and 14 times stronger affinity to AETP than

to RGD peptide.

IV

Page 36: In vitro, in vivo, and in silico investigations of polymer

36

11 General discussion

11.1 Structure-activity relationship of polymeric DNA carriers on DNA-complex formation, transfection efficacy, and toxicity

To be able to condense the DNA, polymer should have a sufficient density of positive

charges. On the other hand, high cationic charge density is often related to increased

toxicity. Electrostatic interactions with the plasma membrane are known to damage the

cell (Lv et al. 2006). An additional critical step of gene delivery is escape from the

endosomes. Because cationic polymers lack hydrophobic parts, they cannot destabilize the

endosomal membrane. However, some polymers do protonate at endosomal pH (e.g. PEI,

PAMAM or PDMAEMA) and they can act as swelling proton sponges. To modify the

properties of the cationic polymers, they can be synthesized in different lengths, with

different architecture (linear or branched), with different substitutions or additions of

functional groups (Tros de Ilarduya, Sun & Düzgünes 2010). Nonetheless, it seems that

there is no universal rule regarding how the architecture of a polymer effects its properties

as a gene delivery vehicle.

To investigate the effect of polymer composition and architecture on the ability to

condense DNA and mediate transgene expression, two different star block copolymers and

corresponding linear triblock copolymers were synthesized. The hydrophobic core of star-

like polymers was composed of either glassy polystyrene (PS) or rubbery poly(n-

butylacrylate) (PBuA). The outer block consisted of hydrophilic and cationic PDMAEMA.

All of the synthesized polymers were able to form spherical core-shell micelles in aqueous

solutions. The linear polymers and star-like (PBuA-PDMAEMA)6 could condense plasmid

DNA completely, while glassy (PS-PDMAEMA)6 formed loose DNA complexes. The

gene transfection efficiency was highest for linear PDMAEMA-PBuA-PDMAEMA at n/p

ratio 2 and 4, and lowest for star-like (PS-PDMAEMA)6. Cytotoxicity was increased at

charge ratios of n/p 8 and higher, (PS-DMAEMA)6 being least toxic.

The poor transfection efficiency of star-like (PS-PDMAEMA)6 may arise from glassy

polystyrene in the core of the star polymers which stiffens the structure of the polymer and

may restrict the interaction with plasmid DNA. On the contrary, rubbery pBuA core may

not limit the contact of PDMAEMA arms and the DNA. Linear polymer chain is more

flexible making the structure more favorable for DNA condensation and transfection

compared to star-like structures. In addition, the amine groups are more available for DNA

binding in the linear form of the polymer. Linear architecture and rubbery pBuA core

seem to be favorable features for mediating transfection. The positive effect of linear

architecture was also shown in the case of poly-L-lysine polymers (Männistö et al. 2002).

This is the first known investigation into the effects of rubbery and glassy states on gene

transfer.

This study has revealed evidence that polymer architecture and composition effect

gene packing ability and transfection efficiency of the studied block copolymers.

Accordingly, linear PDMAEMA-PBuA-PDMAEMA showed transfection efficiency close

to PEI 25 kDa in a retinal pigment epithelial cell line in vitro. However, similar to all

Page 37: In vitro, in vivo, and in silico investigations of polymer

37

cationic polyplexes, PDMAEMA-based polyplexes are also liable to deactivation by

negatively charged serum proteins and other polyanions in vivo.

In another study, protein-dendron conjugates, consisting of either BSA or hydrophobin

(HFBI) protein and first-generation (G1) or second-generation (G2) polyamine dendrons,

were synthesized. BSA is a rather large (66.4 kDa) biocompatible protein that exhibits

long circulation times, while HFBI is a small (8.7 kDa) surface-active protein that is also

well known for its safety. Since both of the proteins lack DNA binding motifs, cationic

polyamine dendrons were conjugated to them to achieve sufficient DNA binding. All of

the dendron-conjugated proteins showed enhanced DNA binding over the naked proteins.

HFBI-conjugates expressed stronger affinity to DNA than corresponding BSA-conjugates,

especially BSA-G1. This is not surprising because the dendron (~1 kDa) is relatively small

compared to the size of BSA and plasmid DNA.

HFBI-G2 was the only conjugate with enhanced transfection activity, even though

BSA-G2 could bind to DNA with a similar affinity. This could be due to the amphiphilic

nature of HFBI that may lead to a favorable interaction with cell membrane structures.

Despite the successful use of BSA as a drug nanocarrier, it was not able to mediate

transfection when conjugated with dendrons. It has been shown earlier that the small

polyamine dendrons are relatively inefficient in mediating transfection on their own,

without endosomal disrupting agent, chloroquine (Hardy et al. 2006). Nonetheless, in

combination with hydrophobic HFBI, enhanced transgene activity without increased

toxicity could be achieved. The transfection efficiency of HFBI-G2 was, however, fairly

low compared to PEI 25 kDa. It has been shown that by increasing the size of the

PAMAM dendrimers, higher transfection could be achieved (Haensler, Szoka 1993). This

may also be the case for HFBI-dendrons. The endosomal escape of HFBI-dendrons would,

however, most likely persist as a bottleneck because of deficiency of protonable tertiary

amines in the structure of polyamine dendrons.

The concept of protein-dendron conjugates worked in the sense of efficient DNA

condensation, but the transfection efficiency was very modest compared to PEI 25 kDa.

The protein-dendron conjugates are not suitable for gene delivery as such; the concept

should be further developed.

11.2 Lipid-coated DNA-complexes as stable gene delivery vectors

Cationic lipoplexes and polyplexes are able to transfect cells in vitro, but in vivo they are

susceptible to deactivation by several polyanionic proteins and polysaccharides, such as

GAGs. Extracellular GAGs can bind to cationic complexes and are able to alter both

cellular uptake and intracellular behavior thus decreasing the gene transfer (Ruponen et al.

2004). Optimal gene delivery vectors would be stable in extracellular space but are

capable of releasing the cargo inside the cell. Coating of cationic DNA polyplexes with a

stabilizing layer of anionic lipids results in envelope-type particles (Guo, Lee 2000,

Mastrobattista et al. 2001, Nahde et al. 2001, Guo, Gosselin & Lee 2002, Khalil et al.

2007). The stability of such particles against GAGs has not been evaluated, however.

Page 38: In vitro, in vivo, and in silico investigations of polymer

38

In this study, a method to coat the PEI/DNA complexes by anionic and fusogenic lipid

mixture was developed. In this detergent removal coating procedure, the polyplexes are

slowly coated by a mixed micellar solution of DOPE/CHEMS/octyl glucoside. The

resulting mixture is subsequently diluted with buffer that leads to extraction of the

surfactant from the mixed micelles and to the formation of a lipid bilayer. The lipid-coated

DNA complexes (LCDC) showed a good stability against a model GAG, dextran sulphate,

while uncoated polyplexes disintegrated rapidly in the presence of the GAGs. LCDCs

prepared by the detergent removal coating method showed superior stability compared to

the LCDCs prepared by ethanol injection procedure, in which the polyplexes were coated

with preformed liposomes. The addition of mixed micellar solution to the polyplex surface

apparently forms a more uniform coating compared to the addition of preformed

liposomes.

The coated complexes showed pH-sensitivity at low pH (5-6) by fusion and

aggregation. This means that the complexes are able to fuse with endosomes and release

their contents, which is important for efficient gene delivery. Because of the negative

surface charge, the cellular uptake was lower than the uptake of non-coated cationic

PEI/DNA complexes. This may explain why the transfection efficacy also remained lower

compared to PEI/DNA polyplexes. Cellular uptake of LCDCs was improved by

neutralizing the negative charges with PEGylation, but this modification reduced the pH-

sensitivity and lowered the level of transfection. PEG is ideal for preventing aggregation

and thus prolonging the time in blood circulation, but it becomes unnecessary after particle

internalization into the cells. To avoid the obstructing effects of PEG inside the cells,

cleavable, acid labile PEG chains have been generated (Guo, Szoka 2001, Romberg,

Hennink & Storm 2008).

Local DNA delivery, for example in the eye or in some surgical situations, may not

require PEGylation because the particles are not removed by the reticuloendothelial

system in such cases. The anionic lipid coat or PEG coating may be useful however, in

preventing the interactions with local anions, like hyaluronic acid in the vitreous body of

the eye. Kurosaki et al. (2013) showed higher transfection activity in the retina of rabbits

for PEI/DNA polyplexes coated with anionic polymer-coating than for uncoated cationic

PEI/DNA polyplexes. Because of the aggregation, cationic polyplexes were immobilized

in the vitreous and only a small amount could reach the target cells in the posterior eye. To

improve the cellular uptake of LCDCs, cell penetrating agents could be employed. This

kind of envelope-type structure, with a functionalized surface, could be one step towards a

virus-mimicking gene delivery vector.

In conclusion, LCDCs were stable against GAGs, but they were able to fuse with the

endosomal membranes at low pH. The overall cellular uptake of the LCDCs remained

low, however. By enhancing the cellular uptake with targeting peptides, the LCDCs might

reach higher cellular uptake and become suitable for local gene delivery, to locations such

as the eye.

Page 39: In vitro, in vivo, and in silico investigations of polymer

39

11.3 Hindering effect of liposomal PEG on the targeting efficiency of a small hydrophobic peptide, AETP

Steric stabilization of therapeutic nanoparticles with PEG is important for maintaining

their stability and long blood circulation times, but, as discussed earlier, it can hamper

complex internalization and endosomal membrane diffusion (Holland et al. 1996, Shi et al.

2002a). In addition, PEG shielding may also impair the interaction between the targeting

moiety of the particle and the target antigen on the cell surface, as was hypothesized in the

study of activated endothelium targeting peptide (AETP)-targeted liposomes.

AETP was first discovered by phage display screening of activated human umbilical

vein endothelial cells (HUVEC) and then of human Kaposi’s sarcoma xenograft in mice. It

was thus expected to be an efficient targeting moiety towards neovascular endothelium.

Despite successful phage display screening, the AETP was not able to act as a targeting

moiety when conjugated to drugs (Bergman, M., personal communication). This was due

to the very low water solubility of the peptide. Therefore, it was assumed that AETP could

function as a targeting ligand on PEGylated liposomes. However, neither the cellular

affinity nor the cytotoxic efficacy of AETP-targeted doxorubicin-containing liposomes

was enhanced compared to the non-targeted ones. Pharmacokinetic investigation

demonstrated that AETP-targeted liposomes had a similar or longer blood half-life

compared to non-targeted, PEGylated liposomes. This was a positive result, since the

peptide did not disturb the stealth effect of the PEG on the liposomes, but it could also

mean that the peptide was covered with the PEG shield. AETP-targeted liposomes did not

show any enhanced accumulation in the tumor tissue or enhanced specific binding to the

endothelial cells in vivo when compared to non-targeted liposomes.

The interaction between AETP moieties and liposomal PEG was demonstrated by

molecular modeling. A well-known RGD-peptide that has been found to be an effective

targeting ligand was used as a comparison in simulations. As a significantly more

hydrophobic molecule, AETP was found to be covered by PEG while the surface of the

more hydrophilic RGD was more exposed to water. In addition, AETP was also located

deeper in the PEG shield than RGD peptide, rendering it unavailable to bind the receptors.

An obscuring effect of PEG on targeting efficiency of PEG2000-liposomes targeted with

PEG2000-folate has also been observed by Gabizon et al. (1999) and Shiokawa et al.

(2005). When folate was coupled to the end of a clearly longer PEG arm, PEG3500 instead

of PEG2000, clearly extruding from the PEG2000 shield, the cellular association of the

liposomes increased. Recently, Stefanick et al. (2013) revealed the importance of linker

length effect on cellular uptake of peptide-targeted PEGylated liposomes. However, using

longer PEG-spacers may not work in the case of hydrophobic AETP because it tends to

escape from the solvent and interact with the PEG molecules. PEG also has hydrophobic

properties, even though it is described to form a hydrophilic sheath around the liposome.

Hydrophobic interactions between AETP and PEG may explain the unfavorable

orientation of AETP molecules for targeting purposes.

Ligand docking studies showed a high binding affinity of human serum albumin

(HSA) to both AETP and PEG. Although the effect of protein binding on AETP targeting-

ability warrants further studies, molecular modeling shows very strong evidence that PEG

coating hinders target binding of AETP. Protein binding to PEG has been previously

Page 40: In vitro, in vivo, and in silico investigations of polymer

40

revealed by others experimentally (Price, Cornelius & Brash 2001, Dos Santos et al.

2007).

Even though PEG presents many advantages and it is accepted for clinical use, it may

not be suitable for all cases of drug delivery. In the case of having hydrophobic targeting

ligands, PEG-coating on the liposomes could be replaced by more hydrophilic polymer to

avoid the interactions between the ligand and the steric coating. More rigid polymer may

also work, given that it can force the ligand out from the polymer cloud. But whether the

stealth effect is then lost due to a firmer polymer structure should be assessed. Some

possible alternatives for PEG have been suggested by Knop et al. (2010). When

PEGylated nanosystems are used it might be beneficial to choose a hydrophilic targeting

ligand.

In a summary, AETP, a promising targeting candidate, failed to show any targeting

efficiency in these studies. This was most likely due to the interference of liposomal PEG

shield that may have prevented the interaction of the peptide with the target receptors.

Nevertheless, cancer targeting is a complicated task, and many other factors may influence

the efficacy of drug targeting.

11.4 Pre-targeting and local administration of liposomes as potential approaches in tumor targeting

Tumor targeting with immunoliposomes has been studied in numerous institutes over the

past 30 years. Targeting efficiency over the non-targeted liposomes was somewhat

successful, but accelerated clearance from the blood circulation proved problematic,

particularly when whole antibodies were used as targeting moieties (Aragnol, Leserman

1986). Shorter half-life usually compromises the benefit that has been achieved by

targeting. When smaller fragments of antibodies have been used, half-lives comparable to

non-targeted liposomes have been reached (Maruyama et al. 1997). Another option to

prolong the residence time in blood circulation could be pre-targeting technology, since it

has been noticed by Harding et al. (1997) that separate injections of cetuximab and

PEGylated liposomes did not cause immune response, while immunogenicity was

potentiated by antibody-coupled liposomes.

In the current study, both direct targeting and pre-targeting (Figure 9) approaches were

used to target PEGylated liposomes to ovarian adenocarcinoma cells (SKOV-3) in vitro

and intraperitoneal xenografts in mice in vivo. Biotin-neutravidin technology was utilized

in both cases to link the endothelial growth factor receptor (EGFR) antibody (cetuximab)

to the liposomes. Direct targeting of the liposomes to SKOV-3 cells was receptor-specific

and efficient in vitro, whilst the pre-targeting was not as efficient, possibly due to

premature internalization of the cetuximab-receptor-complex. Antibody-neutravidin-

complex should remain available for interaction with biotinylated liposomes. The other

explanation could be the hindrance effect of PEG as discussed earlier. As a small

molecule, biotin could be partly covered by the PEG shield and be hindered from the

interaction with neutravidin that is bound to the receptor.

Page 41: In vitro, in vivo, and in silico investigations of polymer

41

Figure 9 Schematic presention of pre-targeting and direct targeting approaches. In pre-

targeting method (A), antibody-linked neutravidin is administered first (step 1). Once antibody

complex has found its target, biotinylated liposomes are administered (step 2). In direct targeting

approach (B), antibodies are coupled on the surface of the liposomes and the formed

immunoliposomes are administered as a single dose.

In animal studies, accumulation of the targeted liposomes in the tumor was not higher

than that of non-targeted liposomes with either targeting approach. However,

intraperitoneally (i.p.) injected biotinylated liposomes, regardless of targeting,

accumulated faster and reached a higher concentration in tumors compared to

intravenously (i.v.) administered liposomes. This was comparable to the observation of

Lin et al. (2009), who noticed rapid tumor accumulation of PEGylated liposomes after i.p.

injection. I.p. administered liposomes can thus accomplish a fast local effect on

intraperitoneal tumors, followed by systemic drug delivery after passing into the blood

circulation. Because of the tendency of ovarian cancer to spread to the abdominal cavity,

i.p. administration of cytotoxic drugs may be beneficial to reach both the primary tumor

and the metastases.

No significantly enhanced accumulation of targeted liposomes to the tumor was

evident, when compared to non-targeted, PEGylated liposomes, even though higher

uptake was seen in cell culture. This is somewhat expected, because due to the EPR effect,

the uptake in the solid tumor tissue may not be further enhanced by targeting. Without

higher accumulation at the tumor site, enhanced cancer cell-specific internalization

(Kirpotin et al. 2006) and increased therapeutic efficacy (Mamot et al. 2005) for targeted

liposomes over non-targeted ones have been observed.

The pre-targeting method described here still requires optimization: the choice of the

antibody should be reconsidered, but also the timing between the antibody administration

and the liposome injections needs optimization. In principle, a pre-targeting approach has

potential because of separate injections of targeting agents and drug formulation; different

antigens can be targeted simultaneously with the same drug formulation. This flexibility

enables the use of the same nanoformulation in different types of cancers. When antibody

is not chemically attached to the liposome, the stability during storage could be improved,

Page 42: In vitro, in vivo, and in silico investigations of polymer

42

and also the developmental costs may not rise as high as the costs for more complicated

directly targeted formulation.

Pre-targeting proved to be less efficient, compared to direct targeting, regarding uptake

in cancer cells in vitro. Because of fast tumor accumulation in vivo, however, the concept

of local tumor pre-targeting might be beneficial after further development.

Page 43: In vitro, in vivo, and in silico investigations of polymer

43

12 Conclusions

1. Polymer architecture and composition had a clear effect on DNA complexation

and gene transfer efficacy. Linear structure and rubbery poly-n-butyl acrylate

block in the middle of PDMAEMA-chain improved DNA complexation and

transfection efficiency, while star-shaped and glassy polystyrene core hindered

DNA condensation as well as transfection.

2. Hydrophobin and BSA-conjugated polyamino dendrons bound to DNA with high

efficacy and were biocompatible in vitro; but transfection was observed only for

amphiphilic hydrophobin-dendron conjugates (second generation) with a low

transgene expression.

3. Lipid-coated DNA complexes with a cationic PEI/DNA core covered with

negatively charged DOPE/CHEMS mixture were successfully produced by

detergent removal method. These complexes were resistant against extracellular

glycosaminoglycans, and were able fuse with endosomal membrane at acidic pH

and mediate transfection.

4. Molecular modeling revealed that hydrophobic AETP targeting moieties were

located deep in the PEG layer of the liposomes, and thus might have been

prevented from interacting with target receptors. In addition, peptide binding to

serum proteins may further inhibit target binding. These findings may be useful in

the development of targeted nanocarriers.

5. Direct targeting with EGFR-antibody liposomes was superior to non-targeted and

pre-targeted liposomes in the cell studies, but in mice the accumulation in the

tumor remained low. Tumoral accumulation after intraperitoneal administration of

pre-targeted and non-targeted liposomes was faster and greater compared to

intravenous administration. Local tumoral pre-targeting method warrants further

studies as a potential approach in cancer therapy.

Page 44: In vitro, in vivo, and in silico investigations of polymer

44

13 Future prospects

Nanoparticles – drugs of the future?

Nanoparticle technology in pharmaceutical research has provided great promise for more

efficient and safe drug therapy. Targeted nanomedicines in particular have garnered

growing enthusiasm among researchers reflected by a massively increasing number of

publications. Despite decades of developmental work, there are no targeted nanoparticle

formulations in clinical use and only a few are currently in clinical trials. When a drug

formulation becomes more complicated, the risks in the development phase most

definitely increase.

Somewhat surprisingly, most of the nanoparticle systems are developed for the

treatment of cancer. Development of systemically administered targeted nanoparticles

faces many technical challenges as a risk of too short half-life, poor tumoral perfusion, and

diffusion barriers at the binding site. The leakiness of the tumor vessels varies between the

tumor types, in some cases the tumor core may be not well perfused (Chauhan et al. 2011).

In addition, the expression levels of the target receptors can vary between different cancers

(Perez-Soler 2004) and the receptor density may also decrease dramatically during the

treatment (Jiang et al. 2008, Cheng et al. 2012). This brings challenges to nanoparticle

delivery to tumors. Possibly, some other disease states in closed systems, such as in the

eye, or the surgical situations, might be more beneficial delivery targets for nanomedicine

development, because of the possibility of local administration. In this case, the

nanomedicine would provide advantages, such as protection of the drug from enzymatic

degradation, prolonged activity, cell specificity, and optimization of intracellular

distribution. Non-malignant diseases also lack rapid mutation of the target cells which

makes the targeting easier.

The approval process for a nanoformulation is much more difficult than for parent

drug. This likely reflects why big pharmaceutical companies do not want to invest money

and time for a small increase in performance that might be achieved by reformulating a

currently approved drug inside a nanocarrier (Venditto, Szoka Jr. 2013). For example,

liposomal cancer therapeutics Caelyx and Ambisome are able to reduce the toxicity of the

parent drug, but improving the efficacy is still modest. Even if reduced toxicity is very

important for the patients, the minimal improvement in the efficacy may become an issue

for nanotherapeutics (Juliano 2013).

Although nanotherapeutics has not yet realised its promises, it will certainly find its

place in the pharmaceutical field. Increasing knowledge of the strengths and the

weaknesses of nanomedicines will be useful in the developmental phase. In addition,

utilization of computational modeling may be helpful in screening the best nanocandidates

before carrying out the costly preclinical and clinical studies.

Page 45: In vitro, in vivo, and in silico investigations of polymer

45

References

Adams, G.P., Schier, R., McCall, A.M., Simmons, H.H., Horak, E.M., Alpaugh, R.K.,

Marks, J.D. & Weiner, L.M. 2001, "High affinity restricts the localization and tumor

penetration of single-chain fv antibody molecules", Cancer research, vol. 61, no. 12,

pp. 4750-4755.

Allen, C., Dos Santos, N., Gallagher, R., Chiu, G.N., Shu, Y., Li, W.M., Johnstone, S.A.,

Janoff, A.S., Mayer, L.D., Webb, M.S. & Bally, M.B. 2002, "Controlling the physical

behavior and biological performance of liposome formulations through use of surface

grafted poly(ethylene glycol)", Bioscience reports, vol. 22, no. 2, pp. 225-250.

Allen, T.M., Hansen, C., Martin, F., Redemann, C. & Yau-Young, A. 1991, "Liposomes

containing synthetic lipid derivatives of poly(ethylene glycol) show prolonged

circulation half-lives in vivo", Biochimica et biophysica acta, vol. 1066, no. 1, pp. 29-

36.

Allen, T.M., Brandeis, E., Hansen, C.B., Kao, G.Y. & Zalipsky, S. 1995, "A new strategy

for attachment of antibodies to sterically stabilized liposomes resulting in efficient

targeting to cancer cells", Biochimica et Biophysica Acta (BBA) - Biomembranes, vol.

1237, no. 2, pp. 99-108.

Aragnol, D. & Leserman, L.D. 1986, "Immune clearance of liposomes inhibited by an

anti-Fc receptor antibody in vivo", Proceedings of the National Academy of Sciences

of the United States of America, vol. 83, no. 8, pp. 2699-2703.

Boeckle, S., Fahrmeir, J., Roedl, W., Ogris, M. & Wagner, E. 2006, "Melittin analogs with

high lytic activity at endosomal pH enhance transfection with purified targeted PEI

polyplexes", Journal of Controlled Release, vol. 112, no. 2, pp. 240-248.

Bolhassani, A. 2011, "Potential efficacy of cell-penetrating peptides for nucleic acid and

drug delivery in cancer", Biochimica et biophysica acta, vol. 1816, no. 2, pp. 232-

246.

Bolotin, E.M., Cohen, R., Bar, L.K., Emanuel, N., Ninio, S., Lasic, D.D. & Barenholz, Y.

1994, "Ammonium sulfate gradients for efficient and stable remote loading of

amphiphilic weak bases into liposomes and ligandoliposomes", Journal of Liposome

Research, vol. 4, no. 1, pp. 455-479.

Boussif, O., Lezoualc'h, F., Zanta, M.A., Mergny, M.D., Scherman, D., Demeneix, B. &

Behr, J.P. 1995, "A versatile vector for gene and oligonucleotide transfer into cells in

culture and in vivo: polyethylenimine", Proceedings of the National Academy of

Sciences of the United States of America, vol. 92, no. 16, pp. 7297-7301.

Brinkhuis, R.P., Rutjes, F.P.J.T. & Van Hest, J.C.M. 2011, "Polymeric vesicles in

biomedical applications", Polymer Chemistry, vol. 2, no. 7, pp. 1449-1462.

Page 46: In vitro, in vivo, and in silico investigations of polymer

46

Chang, Y.J., Chang, C.H., Chang, T.J., Yu, C.Y., Chen, L.C., Jan, M.L., Luo, T.Y., Lee,

T.W. & Ting, G. 2007, "Biodistribution, pharmacokinetics and microSPECT/CT

imaging of 188Re-bMEDA-liposome in a C26 murine colon carcinoma solid tumor

animal model", Anticancer Research, vol. 27, no. 4B, pp. 2217-2225.

Chauhan, V.P., Stylianopoulos, T., Boucher, Y. & Jain, R.K. 2011, "Delivery of molecular

and nanoscale medicine to tumors: transport barriers and strategies", Annual review of

chemical and biomolecular engineering, vol. 2, pp. 281-298.

Cheng, Z., Al Zaki, A., Hui, J.Z., Muzykantov, V.R. & Tsourkas, A. 2012,

"Multifunctional nanoparticles: cost versus benefit of adding targeting and imaging

capabilities", Science, vol. 338, no. 6109, pp. 903-910.

Cho, K., Wang, X., Nie, S., Chen, Z.G. & Shin, D.M. 2008, "Therapeutic nanoparticles for

drug delivery in cancer", Clinical cancer research, vol. 14, no. 5, pp. 1310-1316.

Choosakoonkriang, S., Lobo, B.A., Koe, G.S., Koe, J.G. & Middaugh, C.R. 2003,

"Biophysical characterization of PEI/DNA complexes", Journal of pharmaceutical

sciences, vol. 92, no. 8, pp. 1710-1722.

Chow, T.H., Lin, Y.Y., Hwang, J.J., Wang, H.E., Tseng, Y.L., Wang, S.J., Liu, R.S., Lin,

W.J., Yang, C.S. & Ting, G. 2009, "Improvement of biodistribution and therapeutic

index via increase of polyethylene glycol on drug-carrying liposomes in an HT-29/luc

xenografted mouse model", Anticancer Research, vol. 29, no. 6, pp. 2111-2120.

Chuang, K.H., Wang, H.E., Chen, F.M., Tzou, S.C., Cheng, C.M., Chang, Y.C., Tseng,

W.L., Shiea, J., Lin, S.R., Wang, J.Y., Chen, B.M., Roffler, S.R. & Cheng, T.L. 2010,

"Endocytosis of PEGylated agents enhances cancer imaging and anticancer efficacy",

Molecular cancer therapeutics, vol. 9, no. 6, pp. 1903-1912.

Conforti, G., Dominguez-Jimenez, C., Zanetti, A., Gimbrone, M.A.,Jr, Cremona, O.,

Marchisio, P.C. & Dejana, E. 1992, "Human endothelial cells express integrin

receptors on the luminal aspect of their membrane", Blood, vol. 80, no. 2, pp. 437-

446.

Danhier, F., Feron, O. & Preat, V. 2010, "To exploit the tumor microenvironment: Passive

and active tumor targeting of nanocarriers for anti-cancer drug delivery", Journal of

controlled release, vol. 148, no. 2, pp. 135-146.

Dash, P.R., Read, M.L., Barrett, L.B., Wolfert, M.A. & Seymour, L.W. 1999, "Factors

affecting blood clearance and in vivo distribution of polyelectrolyte complexes for

gene delivery", Gene therapy, vol. 6, no. 4, pp. 643-650.

Dauty, E. & Verkman, A.S. 2005, "Actin cytoskeleton as the principal determinant of size-

dependent DNA mobility in cytoplasm: a new barrier for non-viral gene delivery",

The Journal of biological chemistry, vol. 280, no. 9, pp. 7823-7828.

David, A., Kopečková, P., Minko, T., Rubinstein, A. & Kopeček, J. 2004, "Design of a

multivalent galactoside ligand for selective targeting of HPMA copolymer–

Page 47: In vitro, in vivo, and in silico investigations of polymer

47

doxorubicin conjugates to human colon cancer cells", European journal of cancer,

vol. 40, no. 1, pp. 148-157.

Dos Santos, N., Allen, C., Doppen, A.M., Anantha, M., Cox, K.A., Gallagher, R.C.,

Karlsson, G., Edwards, K., Kenner, G., Samuels, L., Webb, M.S. & Bally, M.B. 2007,

"Influence of poly(ethylene glycol) grafting density and polymer length on

liposomes: relating plasma circulation lifetimes to protein binding", Biochimica et

biophysica acta, vol. 1768, no. 6, pp. 1367-1377.

Dubey, P.K., Mishra, V., Jain, S., Mahor, S. & Vyas, S.P. 2004, "Liposomes modified

with cyclic RGD peptide for tumor targeting", Journal of drug targeting, vol. 12, no.

5, pp. 257-264.

Duncan, R. & Gaspar, R. 2011, "Nanomedicine(s) under the microscope", Molecular

pharmaceutics, vol. 8, no. 6, pp. 2101-2141.

Dunlap, D.D., Maggi, A., Soria, M.R. & Monaco, L. 1997, "Nanoscopic structure of DNA

condensed for gene delivery", Nucleic acids research, vol. 25, no. 15, pp. 3095-3101.

ElBayoumi, T.A. & Torchilin, V.P. 2009, "Tumor-targeted nanomedicines: enhanced

antitumor efficacy in vivo of doxorubicin-loaded, long-circulating liposomes

modified with cancer-specific monoclonal antibody", Clinical cancer research, vol.

15, no. 6, pp. 1973-1980.

Erbacher, P., Bettinger, T., Belguise-Valladier, P., Zou, S., Coll, J., Behr, J. & Remy, J.

1999, "Transfection and physical properties of various saccharide, poly(ethylene

glycol), and antibody-derivatized polyethylenimines (PEI)", The journal of gene

medicine, vol. 1, no. 3, pp. 210-222.

Fattal, E., Couvreur, P. & Dubernet, C. 2004, ""Smart" delivery of antisense

oligonucleotides by anionic pH-sensitive liposomes", Advanced Drug Delivery

Reviews, vol. 56, no. 7, pp. 931-946.

Feron, O. 2004, "Targeting the tumor vascular compartment to improve conventional

cancer therapy", Trends in pharmacological sciences, vol. 25, no. 10, pp. 536-542.

Gabizon, A., Horowitz, A.T., Goren, D., Tzemach, D., Mandelbaum-Shavit, F., Qazen,

M.M. & Zalipsky, S. 1999, "Targeting folate receptor with folate linked to extremities

of poly(ethylene glycol)-grafted liposomes: in vitro studies", Bioconjugate chemistry,

vol. 10, no. 2, pp. 289-298.

Geng, Y., Dalhaimer, P., Cai, S., Tsai, R., Tewari, M., Minko, T. & Discher, D.E. 2007,

"Shape effects of filaments versus spherical particles in flow and drug delivery",

Nature nanotechnology, vol. 2, no. 4, pp. 249-255.

Giacca, M. & Zacchigna, S. 2012, "Virus-mediated gene delivery for human gene

therapy", Journal of Controlled Release, vol. 161, no. 2, pp. 377-388.

Page 48: In vitro, in vivo, and in silico investigations of polymer

48

Godbey, W.T., Wu, K.K. & Mikos, A.G. 1999, "Size matters: molecular weight affects the

efficiency of poly(ethylenimine) as a gene delivery vehicle", Journal of Biomedical

Materials Research, vol. 45, no. 3, pp. 268-275.

Görlich, D. & Mattaj, I.W. 1996, "Nucleoplasmic transport", Science, vol. 271, pp. 1513-

1518.

Guo, W., Gosselin, M.A. & Lee, R.J. 2002, "Characterization of a novel diolein-based

LPDII vector for gene delivery", Journal of controlled release, vol. 83, no. 1, pp.

121-132.

Guo, W. & Lee, R.J. 2000, "Efficient gene delivery using anionic liposome-complexed

polyplexes (LPDII)", Bioscience reports, vol. 20, no. 5, pp. 419-432.

Guo, X. & Szoka, F.C.,Jr 2001, "Steric stabilization of fusogenic liposomes by a low-pH

sensitive PEG-diortho ester-lipid conjugate", Bioconjugate chemistry, vol. 12, no. 2,

pp. 291-300.

Haensler, J. & Szoka, F.C.,Jr 1993, "Polyamidoamine cascade polymers mediate efficient

transfection of cells in culture", Bioconjugate chemistry, vol. 4, no. 5, pp. 372-379.

Hafez, I.M. & Cullis, P.R. 2000, "Cholesteryl hemisuccinate exhibits pH sensitive

polymorphic phase behavior", Biochimica et biophysica acta, vol. 1463, no. 1, pp.

107-114.

Hafez, I.M., Maurer, N. & Cullis, P.R. 2001, "On the mechanism whereby cationic lipids

promote intracellular delivery of polynucleic acids", Gene therapy, vol. 8, no. 15, pp.

1188-1196.

Hansen, C.B., Kao, G.Y., Moase, E.H., Zalipsky, S. & Allen, T.M. 1995, "Attachment of

antibodies to sterically stabilized liposomes: evaluation, comparison and optimization

of coupling procedures", Biochimica et biophysica acta, vol. 1239, no. 2, pp. 133-

144.

Hanzlíková, M., Ruponen, M., Galli, E., Raasmaja, A., Aseyev, V., Tenhu, H., Urtti, A. &

Yliperttula, M. 2011, "Mechanisms of polyethylenimine-mediated DNA delivery:

Free carrier helps to overcome the barrier of cell-surface glycosaminoglycans",

Journal of Gene Medicine, vol. 13, no. 7-8, pp. 402-409.

Harasym, T.O., Bally, M.B. & Tardi, P. 1998, "Clearance properties of liposomes

involving conjugated proteins for targeting", Advanced Drug Delivery Reviews, vol.

32, no. 1-2, pp. 99-118.

Harding, J.A., Engbers, C.M., Newman, M.S., Goldstein, N.I. & Zalipsky, S. 1997,

"Immunogenicity and pharmacokinetic attributes of poly(ethylene glycol)-grafted

immunoliposomes", Biochimica et biophysica acta, vol. 1327, no. 2, pp. 181-192.

Page 49: In vitro, in vivo, and in silico investigations of polymer

49

Hardy, J.G., Kostiainen, M.A., Smith, D.K., Gabrielson, N.P. & Pack, D.W. 2006,

"Dendrons with spermine surface groups as potential building blocks for nonviral

vectors in gene therapy", Bioconjugate chemistry, vol. 17, no. 1, pp. 172-178.

Harrington, K.J., Rowlinson-Busza, G., Syrigos, K.N., Abra, R.M., Uster, P.S., Peters,

A.M. & Stewart, J.S. 2000, "Influence of tumour size on uptake of(111)ln-DTPA-

labelled pegylated liposomes in a human tumour xenograft model", British journal of

cancer, vol. 83, no. 5, pp. 684-688.

Hashizume, H., Baluk, P., Morikawa, S., McLean, J.W., Thurston, G., Roberge, S., Jain,

R.K. & McDonald, D.M. 2000, "Openings between defective endothelial cells

explain tumor vessel leakiness", The American journal of pathology, vol. 156, no. 4,

pp. 1363-1380.

Hillaireau, H. & Couvreur, P. 2009, "Nanocarriers' entry into the cell: relevance to drug

delivery", Cellular and molecular life sciences, vol. 66, no. 17, pp. 2873-2896.

Holland, J.W., Hui, C., Cullis, P.R. & Madden, T.D. 1996, "Poly(ethylene glycol)--lipid

conjugates regulate the calcium-induced fusion of liposomes composed of

phosphatidylethanolamine and phosphatidylserine", Biochemistry, vol. 35, no. 8, pp.

2618-2624.

Hwang, S., Maitani, Y., Qi, X., Takayama, K. & Nagai, T. 1999, "Remote loading of

diclofenac, insulin and fluorescein isothiocyanate labeled insulin into liposomes by

pH and acetate gradient methods", International journal of pharmaceutics, vol. 179,

no. 1, pp. 85-95.

Iden, D.L. & Allen, T.M. 2001, "In vitro and in vivo comparison of immunoliposomes

made by conventional coupling techniques with those made by a new post-insertion

approach", Biochimica et biophysica acta, vol. 1513, no. 2, pp. 207-216.

Iyer, A.K., Su, Y., Feng, J., Lan, X., Zhu, X., Liu, Y., Gao, D., Seo, Y., Vanbrocklin, H.F.,

Courtney Broaddus, V., Liu, B. & He, J. 2011, "The effect of internalizing human

single chain antibody fragment on liposome targeting to epithelioid and sarcomatoid

mesothelioma", Biomaterials, vol. 32, no. 10, pp. 2605-2613.

Iyer, A.K., Khaled, G., Fang, J. & Maeda, H. 2006, "Exploiting the enhanced permeability

and retention effect for tumor targeting", Drug discovery today, vol. 11, no. 17–18,

pp. 812-818.

Jain, R.K. 1990, "Physiological barriers to delivery of monoclonal antibodies and other

macromolecules in tumors", Cancer research, vol. 50, no. 3 Suppl, pp. 814s-819s.

Jiang, W., Kim, B.Y., Rutka, J.T. & Chan, W.C. 2008, "Nanoparticle-mediated cellular

response is size-dependent", Nature nanotechnology, vol. 3, no. 3, pp. 145-150.

Juliano, R. 2013, "Nanomedicine: is the wave cresting?", Nature reviews. Drug discovery,

vol. 12, no. 3, pp. 171-172.

Page 50: In vitro, in vivo, and in silico investigations of polymer

50

Kale, A.A. & Torchilin, V.P. 2007, ""Smart" drug carriers: PEGylated TATp-modified

pH-sensitive liposomes", Journal of Liposome Research, vol. 17, no. 3-4, pp. 197-

203.

Khalil, I.A., Kogure, K., Futaki, S., Hama, S., Akita, H., Ueno, M., Kishida, H., Kudoh,

M., Mishina, Y., Kataoka, K., Yamada, M. & Harashima, H. 2007, "Octa-arginine-

modified multifunctional envelope-type nanoparticles for gene delivery", Gene

therapy, vol. 14, no. 8, pp. 682-689.

Kirchmeier, M.J., Ishida, T., Chevrette, J. & Allen, T.M. 2001, "Correlations between the

rate of intracellular release of endocytosed liposomal Doxorubicin and cytotoxicity as

determined by a new assay", Journal of Liposome Research, vol. 11, no. 1, pp. 15-29.

Kirpotin, D., Park, J.W., Hong, K., Zalipsky, S., Li, W.L., Carter, P., Benz, C.C. &

Papahadjopoulos, D. 1997, "Sterically stabilized anti-HER2 immunoliposomes:

design and targeting to human breast cancer cells in vitro", Biochemistry, vol. 36, no.

1, pp. 66-75.

Kirpotin, D.B., Drummond, D.C., Shao, Y., Shalaby, M.R., Hong, K., Nielsen, U.B.,

Marks, J.D., Benz, C.C. & Park, J.W. 2006, "Antibody targeting of long-circulating

lipidic nanoparticles does not increase tumor localization but does increase

internalization in animal models", Cancer research, vol. 66, no. 13, pp. 6732-6740.

Knop, K., Hoogenboom, R., Fischer, D. & Schubert, U.S. 2010, "Poly(ethylene glycol) in

drug delivery: pros and cons as well as potential alternatives", Angewandte Chemie,

vol. 49, no. 36, pp. 6288-6308.

Koltover, I., Salditt, T., Radler, J.O. & Safinya, C.R. 1998, "An inverted hexagonal phase

of cationic liposome-DNA complexes related to DNA release and delivery", Science,

vol. 281, no. 5373, pp. 78-81.

Kreiss, P., Cameron, B., Rangara, R., Mailhe, P., Aguerre-Charriol, O., Airiau, M.,

Scherman, D., Crouzet, J. & Pitard, B. 1999, "Plasmid DNA size does not affect the

physicochemical properties of lipoplexes but modulates gene transfer efficiency",

Nucleic acids research, vol. 27, no. 19, pp. 3792-3798.

Kullberg, M., Mann, K. & Owens, J.L. 2009, "A two-component drug delivery system

using Her-2-targeting thermosensitive liposomes", Journal of drug targeting, vol. 17,

no. 2, pp. 98-107.

Kurosaki, T., Uematsu, M., Shimoda, K., Suzuma, K., Nakai, M., Nakamura, T., Kitahara,

T., Kitaoka, T. & Sasaki, H. 2013, "Ocular gene delivery systems using ternary

complexes of plasmid DNA, polyethylenimine, and anionic polymers", Biological &

pharmaceutical bulletin, vol. 36, no. 1, pp. 96-101.

Langer, K., Anhorn, M.G., Steinhauser, I., Dreis, S., Celebi, D., Schrickel, N., Faust, S. &

Vogel, V. 2008, "Human serum albumin (HSA) nanoparticles: Reproducibility of

preparation process and kinetics of enzymatic degradation", International journal of

pharmaceutics, vol. 347, no. 1–2, pp. 109-117.

Page 51: In vitro, in vivo, and in silico investigations of polymer

51

Lechardeur, D., Sohn, K.J., Haardt, M., Joshi, P.B., Monck, M., Graham, R.W., Beatty,

B., Squire, J., O'Brodovich, H. & Lukacs, G.L. 1999, "Metabolic instability of

plasmid DNA in the cytosol: a potential barrier to gene transfer", Gene therapy, vol.

6, no. 4, pp. 482-497.

Lechardeur, D., Verkman, A.S. & Lukacs, G.L. 2005, "Intracellular routing of plasmid

DNA during non-viral gene transfer", Advanced Drug Delivery Reviews, vol. 57, no.

5, pp. 755-767.

Lee, H., Hoang, B., Fonge, H., Reilly, R.M. & Allen, C. 2010, "In vivo distribution of

polymeric nanoparticles at the whole-body, tumor, and cellular levels",

Pharmaceutical research, vol. 27, no. 11, pp. 2343-2355.

Lee, L.K., Mount, C.N. & Ayazi Shamlou, P. 2001, "Characterisation of the physical

stability of colloidal polycation-DNA complexes for gene therapy and DNA

vaccines", Chemical Engineering Science, vol. 56, no. 10, pp. 3163-3172.

Lesch, H.P., Kaikkonen, M.U., Pikkarainen, J.T. & Ylä-Herttuala, S. 2010, "Avidin-biotin

technology in targeted therapy", Expert opinion on drug delivery, vol. 7, no. 5, pp.

551-564.

Lin, Y.Y., Li, J.J., Chang, C.H., Lu, Y.C., Hwang, J.J., Tseng, Y.L., Lin, W.J., Ting, G. &

Wang, H.E. 2009, "Evaluation of pharmacokinetics of 111In-labeled VNB-

PEGylated liposomes after intraperitoneal and intravenous administration in a

tumor/ascites mouse model", Cancer biotherapy & radiopharmaceuticals, vol. 24,

no. 4, pp. 453-460.

Loughrey, H., Bally, M.B. & Cullis, P.R. 1987, "A non-covalent method of attaching

antibodies to liposomes", Biochimica et Biophysica Acta (BBA) - Biomembranes, vol.

901, no. 1, pp. 157-160.

Lu, W.L., Qi, X.R., Zhang, Q., Li, R.Y., Wang, G.L., Zhang, R.J. & Wei, S.L. 2004, "A

pegylated liposomal platform: pharmacokinetics, pharmacodynamics, and toxicity in

mice using doxorubicin as a model drug", Journal of pharmacological sciences, vol.

95, no. 3, pp. 381-389.

Lv, H., Zhang, S., Wang, B., Cui, S. & Yan, J. 2006, "Toxicity of cationic lipids and

cationic polymers in gene delivery", Journal of Controlled Release, vol. 114, no. 1,

pp. 100-109.

Mamot, C., Drummond, D.C., Greiser, U., Hong, K., Kirpotin, D.B., Marks, J.D. & Park,

J.W. 2003, "Epidermal growth factor receptor (EGFR)-targeted immunoliposomes

mediate specific and efficient drug delivery to EGFR- and EGFRvIII-overexpressing

tumor cells", Cancer research, vol. 63, no. 12, pp. 3154-3161.

Mamot, C., Drummond, D.C., Noble, C.O., Kallab, V., Guo, Z., Hong, K., Kirpotin, D.B.

& Park, J.W. 2005, "Epidermal growth factor receptor-targeted immunoliposomes

significantly enhance the efficacy of multiple anticancer drugs in vivo", Cancer

research, vol. 65, no. 24, pp. 11631-11638.

Page 52: In vitro, in vivo, and in silico investigations of polymer

52

Marshall, E. 1999, "Gene therapy death prompts review of adenovirus vector", Science,

vol. 286, no. 5448, pp. 2244-2245.

Maruyama, K. 2011, "Intracellular targeting delivery of liposomal drugs to solid tumors

based on EPR effects", Advanced Drug Delivery Reviews, vol. 63, no. 3, pp. 161-169.

Maruyama, K., Takahashi, N., Tagawa, T., Nagaike, K. & Iwatsuru, M. 1997,

"Immunoliposomes bearing polyethyleneglycol-coupled Fab′ fragment show

prolonged circulation time and high extravasation into targeted solid tumors in vivo",

FEBS letters, vol. 413, no. 1, pp. 177-180.

Mastrobattista, E., Kapel, R.H.G., Eggenhuisen, M.H., Roholl, P.J.M., Crommelin, D.J.A.,

Hennink, W.E. & Storm, G. 2001, "Lipid-coated polyplexes for targeted gene

delivery to ovarian carcinoma cells", Cancer gene therapy, vol. 8, no. 6, pp. 405-413.

Mastrobattista, E., Koning, G.A. & Storm, G. 1999, "Immunoliposomes for the targeted

delivery of antitumor drugs", Advanced Drug Delivery Reviews, vol. 40, no. 1-2, pp.

103-127.

Mayer, L.D., Bally, M.B. & Cullis, P.R. 1986, "Uptake of adriamycin into large

unilamellar vesicles in response to a pH gradient", Biochimica et Biophysica Acta -

Biomembranes, vol. 857, pp. 123-126.

Mislick, K.A. & Baldeschwieler, J.D. 1996, "Evidence for the role of proteoglycans in

cation-mediated gene transfer", Proceedings of the National Academy of Sciences of

the United States of America, vol. 93, no. 22, pp. 12349-12354.

Moghimi, S.M. & Szebeni, J. 2003, "Stealth liposomes and long circulating nanoparticles:

Critical issues in pharmacokinetics, opsonization and protein-binding properties",

Progress in lipid research, vol. 42, no. 6, pp. 463-478.

Moreira, J.N., Hansen, C.B., Gaspar, R. & Allen, T.M. 2001, "A growth factor antagonist

as a targeting agent for sterically stabilized liposomes in human small cell lung

cancer", Biochimica et biophysica acta, vol. 1514, no. 2, pp. 303-317.

Morikawa, S., Baluk, P., Kaidoh, T., Haskell, A., Jain, R.K. & McDonald, D.M. 2002,

"Abnormalities in pericytes on blood vessels and endothelial sprouts in tumors", The

American journal of pathology, vol. 160, no. 3, pp. 985-1000.

Morille, M., Passirani, C., Vonarbourg, A., Clavreul, A. & Benoit, J. 2008, "Progress in

developing cationic vectors for non-viral systemic gene therapy against cancer",

Biomaterials, vol. 29, no. 24-25, pp. 3477-3496.

Mounkes, L.C., Zhong, W., Cipres-Palacin, G., Heath, T.D. & Debs, R.J. 1998,

"Proteoglycans mediate cationic liposome-DNA complex-based gene delivery in vitro

and in vivo", The Journal of biological chemistry, vol. 273, no. 40, pp. 26164-26170.

Muro, S., Garnacho, C., Champion, J.A., Leferovich, J., Gajewski, C., Schuchman, E.H.,

Mitragotri, S. & Muzykantov, V.R. 2008, "Control of endothelial targeting and

Page 53: In vitro, in vivo, and in silico investigations of polymer

53

intracellular delivery of therapeutic enzymes by modulating the size and shape of

ICAM-1-targeted carriers", Molecular therapy, vol. 16, no. 8, pp. 1450-1458.

Müller-Eberhard, H.J. 1988, "Molecular Organization and Function of the Complement

System", Annual Review of Biochemistry, vol. 57, pp. 321-347.

Männistö, M., Reinisalo, M., Ruponen, M., Honkakoski, P., Tammi, M. & Urtti, A. 2007,

"Polyplex-mediated gene transfer and cell cycle: effect of carrier on cellular uptake

and intracellular kinetics, and significance of glycosaminoglycans", The journal of

gene medicine, vol. 9, no. 6, pp. 479-487.

Männistö, M., Vanderkerken, S., Toncheva, V., Elomaa, M., Ruponen, M., Schacht, E. &

Urtti, A. 2002, "Structure-activity relationships of poly(L-lysines): effects of

pegylation and molecular shape on physicochemical and biological properties in gene

delivery", Journal of controlled release, vol. 83, no. 1, pp. 169-182.

Mönkkönen, J. & Urtti, A. 1998, "Lipid fusion in oligonucleotide and gene delivery with

cationic lipids", Advanced Drug Delivery Reviews, vol. 34, no. 1, pp. 37-49.

Nahde, T., Müller, K., Fahr, A., Müller, R. & Brüsselbach, S. 2001, "Combined

transductional and transcriptional targeting of melanoma cells by artificial virus-like

particles", The journal of gene medicine, vol. 3, no. 4, pp. 353-361.

Nobs, L., Buchegger, F., Gurny, R. & Allemann, E. 2006, "Biodegradable nanoparticles

for direct or two-step tumor immunotargeting", Bioconjugate chemistry, vol. 17, no.

1, pp. 139-145.

O'Brien, M.E., Wigler, N., Inbar, M., Rosso, R., Grischke, E., Santoro, A., Catane, R.,

Kieback, D.G., Tomczak, P., Ackland, S.P., Orlandi, F., Mellars, L., Alland, L.,

Tendler, C. & CAELYX Breast Cancer Study Group 2004, "Reduced cardiotoxicity

and comparable efficacy in a phase III trial of pegylated liposomal doxorubicin HCl

(CAELYX/Doxil) versus conventional doxorubicin for first-line treatment of

metastatic breast cancer", Annals of Oncology, vol. 15, no. 3, pp. 440-449.

Ogris, M., Brunner, S., Schüller, S., Kircheis, R. & Wagner, E. 1999, "PEGylated

DNA/transferrin-PEI complexes: reduced interaction with blood components,

extended circulation in blood and potential for systemic gene delivery", Gene

therapy, vol. 6, no. 4, pp. 595-605.

Paasonen, L., Laaksonen, T., Johans, C., Yliperttula, M., Kontturi, K. & Urtti, A. 2007,

"Gold nanoparticles enable selective light-induced contents release from liposomes",

Journal of controlled release, vol. 122, no. 1, pp. 86-93.

Pan, H., Han, L., Chen, W., Yao, M. & Lu, W. 2008, "Targeting to tumor necrotic regions

with biotinylated antibody and streptavidin modified liposomes", Journal of

controlled release, vol. 125, no. 3, pp. 228-235.

Page 54: In vitro, in vivo, and in silico investigations of polymer

54

Park, J.W., Hong, K., Kirpotin, D.B., Meyer, O., Papahadjopoulos, D. & Benz, C.C. 1997,

"Anti-HER2 immunoliposomes for targeted therapy of human tumors", Cancer

letters, vol. 118, no. 2, pp. 153-160.

Pasqualini, R., Koivunen, E. & Ruoslahti, E. 1997, "Alpha v integrins as receptors for

tumor targeting by circulating ligands", Nature biotechnology, vol. 15, no. 6, pp. 542-

546.

Pastorino, F., Brignole, C., Di Paolo, D., Nico, B., Pezzolo, A., Marimpietri, D., Pagnan,

G., Piccardi, F., Cilli, M., Longhi, R., Ribatti, D., Corti, A., Allen, T.M. & Ponzoni,

M. 2006, "Targeting liposomal chemotherapy via both tumor cell-specific and tumor

vasculature-specific ligands potentiates therapeutic efficacy", Cancer research, vol.

66, no. 20, pp. 10073-10082.

Pastorino, F., Brignole, C., Marimpietri, D., Cilli, M., Gambini, C., Ribatti, D., Longhi,

R., Allen, T.M., Corti, A. & Ponzoni, M. 2003a, "Vascular damage and anti-

angiogenic effects of tumor vessel-targeted liposomal chemotherapy", Cancer

research, vol. 63, no. 21, pp. 7400-7409.

Pastorino, F., Brignole, C., Marimpietri, D., Sapra, P., Moase, E.H., Allen, T.M. &

Ponzoni, M. 2003b, "Doxorubicin-loaded Fab' fragments of anti-disialoganglioside

immunoliposomes selectively inhibit the growth and dissemination of human

neuroblastoma in nude mice", Cancer research, vol. 63, no. 1, pp. 86-92.

Perez-Soler, R. 2004, "HER1/EGFR targeting: refining the strategy", The oncologist, vol.

9, no. 1, pp. 58-67.

Pollard, H., Remy, J.S., Loussouarn, G., Demolombe, S., Behr, J.P. & Escande, D. 1998,

"Polyethylenimine but not cationic lipids promotes transgene delivery to the nucleus

in mammalian cells", The Journal of biological chemistry, vol. 273, no. 13, pp. 7507-

7511.

Price, M.E,, Cornelius, R.M. & Brash, J.L. 2001, "Protein adsorption to polyethylene

glycol modified liposomes from fibrinogen solution and from plasma", Biochimica et

Biophysica Acta, vol. 1512, no. 2, pp. 191-205.

Pulkkinen, M., Pikkarainen, J., Wirth, T., Tarvainen, T., Haapa-aho, V., Korhonen, H.,

Seppälä, J. & Järvinen, K. 2008, "Three-step tumor targeting of paclitaxel using

biotinylated PLA-PEG nanoparticles and avidin-biotin technology: Formulation

development and in vitro anticancer activity", European journal of pharmaceutics

and biopharmaceutics, vol. 70, no. 1, pp. 66-74.

Radler, J.O., Koltover, I., Salditt, T. & Safinya, C.R. 1997, "Structure of DNA-cationic

liposome complexes: DNA intercalation in multilamellar membranes in distinct

interhelical packing regimes", Science, vol. 275, no. 5301, pp. 810-814.

Rawat, M., Singh, D., Saraf, S. & Saraf, S. 2006, "Nanocarriers: promising vehicle for

bioactive drugs", Biological & pharmaceutical bulletin, vol. 29, no. 9, pp. 1790-1798.

Page 55: In vitro, in vivo, and in silico investigations of polymer

55

Riviere, K., Huang, Z., Jerger, K., Macaraeg, N. & Szoka, F.C.,Jr 2011, "Antitumor effect

of folate-targeted liposomal doxorubicin in KB tumor-bearing mice after intravenous

administration", Journal of drug targeting, vol. 19, no. 1, pp. 14-24.

Romberg, B., Hennink, W.E. & Storm, G. 2008, "Sheddable coatings for long-circulating

nanoparticles", Pharmaceutical research, vol. 25, no. 1, pp. 55-71.

Roth, P., Hammer, C., Piguet, A., Ledermann, M., Dufour, J. & Waelti, E. 2007, "Effects

on hepatocellular carcinoma of doxorubicin- loaded immunoliposomes designed to

target the VEGFR-2", Journal of drug targeting, vol. 15, no. 9, pp. 623-631.

Ruoslahti, E. 2002, "Specialization of tumour vasculature", Nature reviews. Cancer, vol.

2, no. 2, pp. 83-90.

Ruponen, M., Honkakoski, P., Tammi, M. & Urtti, A. 2004, "Cell-surface

glycosaminoglycans inhibit cation-mediated gene transfer", Journal of Gene

Medicine, vol. 6, no. 4, pp. 405-414.

Ruponen, M., Ylä-Herttuala, S. & Urtti, A. 1999, "Interactions of polymeric and liposomal

gene delivery systems with extracellular glycosaminoglycans: physicochemical and

transfection studies", Biochimica et biophysica acta, vol. 1415, no. 2, pp. 331-341.

Sapra, P. & Allen, T.M. 2003, "Ligand-targeted liposomal anticancer drugs", Progress in

lipid research, vol. 42, no. 5, pp. 439-462.

Seftor, R.E., Seftor, E.A. & Hendrix, M.J. 1999, "Molecular role(s) for integrins in human

melanoma invasion", Cancer metastasis reviews, vol. 18, no. 3, pp. 359-375.

Sharkey, R.M., McBride, W.J., Karacay, H., Chang, K., Griffiths, G.L., Hansen, H.J. &

Goldenberg, D.M. 2003, "A universal pretargeting system for cancer detection and

therapy using bispecific antibody", Cancer research, vol. 63, no. 2, pp. 354-363.

Shew, R.L. & Deamer, D.W. 1985, "A novel method for encapsulation of macromolecules

in liposomes", BBA - Biomembranes, vol. 816, no. 1, pp. 1-8.

Shi, J., Votruba, A.R., Farokhzad, O.C. & Langer, R. 2010, "Nanotechnology in Drug

Delivery and Tissue Engineering: From Discovery to Applications", Nano Letters,

vol. 10, pp. 3223-3230.

Shi, F., Wasungu, L., Nomden, A., Stuart, M.C., Polushkin, E., Engberts, J.B. & Hoekstra,

D. 2002a, "Interference of poly(ethylene glycol)-lipid analogues with cationic-lipid-

mediated delivery of oligonucleotides; role of lipid exchangeability and non-lamellar

transitions", The Biochemical journal, vol. 366, no. Pt 1, pp. 333-341.

Shi, G., Guo, W., Stephenson, S.M. & Lee, R.J. 2002b, "Efficient intracellular drug and

gene delivery using folate receptor-targeted pH-sensitive liposomes composed of

cationic/anionic lipid combinations", Journal of controlled release, vol. 80, no. 1-3,

pp. 309-319.

Page 56: In vitro, in vivo, and in silico investigations of polymer

56

Shiokawa, T., Hattori, Y., Kawano, K., Ohguchi, Y., Kawakami, H., Toma, K. & Maitani,

Y. 2005, "Effect of polyethylene glycol linker chain length of folate-linked

microemulsions loading aclacinomycin A on targeting ability and antitumor effect in

vitro and in vivo", Clinical cancer research, vol. 11, no. 5, pp. 2018-2025.

Simões, S., Moreira, J.N., Fonseca, C., Düzgünes, N. & Pedroso de Lima, M.C. 2004, "On

the formulation of pH-sensitive liposomes with long circulation times", Advanced

Drug Delivery Reviews, vol. 56, no. 7, pp. 947-965.

Simões, S., Slepushkin, V., Düzgünes, N. & Pedroso de Lima, M.C. 2001, "On the

mechanisms of internalization and intracellular delivery mediated by pH-sensitive

liposomes", Biochimica et.biophysica acta, vol. 1515, no. 1, pp. 23-37.

Sonawane, N.D., Szoka, F.C.,Jr & Verkman, A.S. 2003, "Chloride accumulation and

swelling in endosomes enhances DNA transfer by polyamine-DNA polyplexes", The

Journal of biological chemistry, vol. 278, no. 45, pp. 44826-44831.

Stefanick, J.F., Ashley, J.D., Kiziltepe, T. & Bilgicer, B. 2013, "A Systematic Analysis of

Peptide Linker Lenght and Liposomal Polyethylene Glycol Coating on Cellular

Uptake of Peptide-Targeted Liposomes", ACS Nano, vol 7, no. 4, pp. 2935-2947.

Subrizi, A., Tuominen, E., Bunker, A., Róg, T., Antopolsky, M. & Urtti, A. 2012, "Tat(48-

60) peptide amino acid sequence is not unique in its cell penetrating properties and

cell-surface glycosaminoglycans inhibit its cellular uptake", Journal of Controlled

Release, vol. 158, no. 2, pp. 277-285.

Sugahara, K.N., Teesalu, T., Karmali, P.P., Kotamraju, V.R., Agemy, L., Greenwald, D.R.

& Ruoslahti, E. 2010, "Coadministration of a tumor-penetrating peptide enhances the

efficacy of cancer drugs", Science, vol. 328, no. 5981, pp. 1031-1035.

Szoka Jr., F. & Papahadjopoulos, D. 1978, "Procedure for preparation of liposomes with

large internal aqueous space and high capture by reverse-phase evaporation",

Proceedings of the National Academy of Sciences of the United States of America,

vol. 75, no. 9, pp. 4194-4198.

Tao, L., Hu, W., Liu, Y., Huang, G., Sumer, B.D. & Gao, J. 2011, "Shape-specific

polymeric nanomedicine: emerging opportunities and challenges", Experimental

biology and medicine, vol. 236, no. 1, pp. 20-29.

Temming, K., Schiffelers, R.M., Molema, G. & Kok, R.J. 2005, "RGD-based strategies

for selective delivery of therapeutics and imaging agents to the tumour vasculature",

Drug resistance updates, vol. 8, no. 6, pp. 381-402.

Tirosh, O., Barenholz, Y., Katzhendler, J. & Priev, A. 1998, "Hydration of polyethylene

glycol-grafted liposomes", Biophysical journal, vol. 74, no. 3, pp. 1371-1379.

Tong, R., Yala, L., Fan, T.M. & Cheng, J. 2010, "The formulation of aptamer-coated

paclitaxel–polylactide nanoconjugates and their targeting to cancer cells",

Biomaterials, vol. 31, no. 11, pp. 3043-3053.

Page 57: In vitro, in vivo, and in silico investigations of polymer

57

Torchilin, V.P. 2005, "Recent advances with liposomes as pharmaceutical carriers",

Nature reviews. Drug discovery, vol. 4, no. 2, pp. 145-160.

Torchilin, V.P., Rammohan, R., Weissig, V. & Levchenko, T.S. 2001, "TAT peptide on

the surface of liposomes affords their efficient intracellular delivery even at low

temperature and in the presence of metabolic inhibitors", Proceedings of the National

Academy of Sciences of the United States of America, vol. 98, no. 15, pp. 8786-8791.

Toropainen, E., Hornof, M., Kaarniranta, K., Johansson, P. & Urtti, A. 2007, "Corneal

epithelium as a platform for secretion of transgene products after transfection with

liposomal gene eyedrops", The journal of gene medicine, vol. 9, no. 3, pp. 208-216.

Tros de Ilarduya, C., Sun, Y. & Düzgünes, N. 2010, "Gene delivery by lipoplexes and

polyplexes", European journal of pharmaceutical sciences, vol. 40, no. 3, pp. 159-

170.

Ulrich, A.S. 2002, "Biophysical aspects of using liposomes as delivery vehicles",

Bioscience reports, vol. 22, no. 2, pp. 129-150.

Vauthier, C., Persson, B., Lindner, P. & Cabane, B. 2011, "Protein adsorption and

complement activation for di-block copolymer nanoparticles", Biomaterials, vol. 32,

no. 6, pp. 1646-1656.

Venditto, V.J. & Szoka Jr., F.C. 2013, "Cancer nanomedicines: So many papers and so

few drugs!", Advanced Drug Delivery Reviews, vol. 65, no. 1, pp. 80-88.

Voinea, M., Dragomir, E., Manduteanu, I. & Simionescu, M. 2002, "Binding and uptake

of transferrin-bound liposomes targeted to transferrin receptors of endothelial cells",

Vascular Pharmacology, vol. 39, no. 1–2, pp. 13-20.

Wang, T., Upponi, J.R. & Torchilin, V.P. 2011, "Design of multifunctional non-viral gene

vectors to overcome physiological barriers: Dilemmas and strategies", International

journal of pharmaceutics, vol. 427, no. 1, pp. 3-20.

Weber, P.C., Ohlendorf, D.H., Wendoloski, J.J. & Salemme, F.R. 1989, "Structural

origins of high-affinity biotin binding to streptavidin", Science, vol. 243, pp. 85-88.

Wightman, L., Kircheis, R., Rossler, V., Carotta, S., Ruzicka, R., Kursa, M. & Wagner, E.

2001, "Different behavior of branched and linear polyethylenimine for gene delivery

in vitro and in vivo", The journal of gene medicine, vol. 3, no. 4, pp. 362-372.

Wiseman, J.W., Goddard, C.A., McLelland, D. & Colledge, W.H. 2003, "A comparison of

linear and branched polyethylenimine (PEI) with DCChol/DOPE liposomes for gene

delivery to epithelial cells in vitro and in vivo", Gene therapy, vol. 10, no. 19, pp.

1654-1662.

Xiang, S., Tong, H., Shi, Q., Fernandes, J.C., Jin, T., Dai, K. & Zhang, X. 2012, "Uptake

mechanisms of non-viral gene delivery", Journal of Controlled Release, vol. 158, no.

3, pp. 371-378.

Page 58: In vitro, in vivo, and in silico investigations of polymer

58

Xiao, Z., McQuarrie, S.A., Suresh, M.R., Mercer, J.R., Gupta, S. & Miller, G.G. 2002, "A

three-step strategy for targeting drug carriers to human ovarian carcinoma cells in

vitro", Journal of Biotechnology, vol. 94, no. 2, pp. 171-184.

Xiong, X.B., Huang, Y., Lu, W.L., Zhang, H., Zhang, X. & Zhang, Q. 2005, "Enhanced

intracellular uptake of sterically stabilized liposomal Doxorubicin in vitro resulting in

improved antitumor activity in vivo", Pharmaceutical research, vol. 22, no. 6, pp.

933-939.

Xu, F.J., Zhang, Z.X., Ping, Y., Li, J., Kang, E.T. & Neoh, K.G. 2009, "Star-shaped

cationic polymers by atom transfer radical polymerization from beta-cyclodextrin

cores for nonviral gene delivery", Biomacromolecules, vol. 10, no. 2, pp. 285-293.

Xu, Y. & Szoka, F.C.,Jr 1996, "Mechanism of DNA release from cationic liposome/DNA

complexes used in cell transfection", Biochemistry, vol. 35, no. 18, pp. 5616-5623.

Ye, G., Gupta, A., DeLuca, R., Parang, K. & Bothun, G.D. 2010, "Bilayer disruption and

liposome restructuring by a homologous series of small Arg-rich synthetic peptides",

Colloids and surfaces.B, Biointerfaces, vol. 76, no. 1, pp. 76-81.

Yuan, F., Dellian, M., Fukumura, D., Leunig, M., Berk, D.A., Torchilin, V.P. & Jain, R.K.

1995, "Vascular permeability in a human tumor xenograft: molecular size dependence

and cutoff size", Cancer research, vol. 55, no. 17, pp. 3752-3756.

Zalipsky, S., Puntambekar, B., Boulikas, P., Engbers, C.M. & Woodle, M.C. 1995,

"Peptide attachment to extremities of liposomal surface grafted PEG chains:

preparation of the long-circulating form of laminin pentapeptide, YIGSR",

Bioconjugate chemistry, vol. 6, no. 6, pp. 705-708.

Zelphati, O. & Szoka, F.C.,Jr 1996, "Mechanism of oligonucleotide release from cationic

liposomes", Proceedings of the National Academy of Sciences of the United States of

America, vol. 93, no. 21, pp. 11493-11498.

Zelphati, O., Uyechi, L.S., Barron, L.G. & Szoka Jr., F.C. 1998, "Effect of serum

components on the physico-chemical properties of cationic lipid/oligonucleotide

complexes and on their interactions with cells", Biochimica et Biophysica Acta (BBA)

- Lipids and Lipid Metabolism, vol. 1390, no. 2, pp. 119-133.

Zhang, L., Gu, F.X., Chan, J.M., Wang, A.Z., Langer, R.S. & Farokhzad, O.C. 2008,

"Nanoparticles in Medicines: Therapeutic Applications and Developments", Clinical

Pharmacology & Therapeutics, vol. 83, no. 5, pp. 761-769.

Zhang, Z.Y. & Smith, B.D. 2000, "High-generation polycationic dendrimers are unusually

effective at disrupting anionic vesicles: membrane bending model", Bioconjugate

chemistry, vol. 11, no. 6, pp. 805-814.