tailored hollow mesoporous silica nanoplatforms with biological...

193
I Tailored hollow mesoporous silica nanoplatforms with biological labels for colon targeted drug delivery systems by Xiaodong She (Masters of Engineering) Submitted in fulfilment of the requirements for the degree of Doctor of Philosophy Deakin University November, 2014

Upload: others

Post on 25-Jun-2020

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

I

Tailored hollow mesoporous silica nanoplatforms with biological labels for colon targeted drug

delivery systems

by

Xiaodong She

(Masters of Engineering)

Submitted in fulfilment of the requirements for the degree of

Doctor of Philosophy

Deakin University

November, 2014

Page 2: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms
sfol
Retracted Stamp
Page 3: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms
sfol
Retracted Stamp
Page 4: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

IV

Acknowledgement

Every PhD project involves a new journey that starts with an idea and travels its own

distinct path until a whole thesis is created. I love my PhD journey, with its unexpected

turns and twists, surprises, joy and excitement. But more than anything else, when I look

back at the journey I have taken, I would like to gratefully acknowledge a number of

people who played a crucial role in helping me with my project.

Herein I would like to acknowledge my principle supervisor Prof Lingxue Kong, who

made it possible for you to read this thesis. Thanks to him for his trust and consistent

support, for giving his precious time, and for seeing the potential of this project when it

was just an idea.

I would also like to make this opportunity to express my sincere thanks to Prof Wei Duan,

Dr. Sarah Shigdar in the School of Medicine, Dr. Leonora Velleman, Dr. Fenghua She

Dr. Weimin Gao and Dr Ludovic Dumee for their wonderful assistances, encouragement

and invaluable help. Special thanks to Lijue Chen, Chengpeng Li and Tao Wang whom I

had the immense honour and privilege to work with. Without their help, this doctoral

work would not have been accomplished. Marion Wright, Dr John Denman in Ian Wark

Research Institute, Dr Cara Doherty in CSIRO Materials Science & Engineering, and Dr

Haijin Zhu in Centre of Excellence for Electromaterials Science have provided a lot of

help regarding my experimental work, and hereby acknowledged.

Personally, I would like to thank my husband Canzhong HE for his encouragement and

company during my PHD journey, where he offered support and understanding in all

wethers. I would also like to thank my parents, sisters and other family members for their

consistent support.

Page 5: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

V

Abstract

Nanoparticle based targeted drug delivery systems hold great promise for cancer therapy

because of their targeting functions, sustained drug release profiles, reduced side-effects

and ability to overcome multidrug resistance (MDR). Nevertheless, inadequate specificity

and stability, insufficient intratumor drug concentration, and high toxicity related to

carriers are still not well solved and thus restrict clinical translation of such systems. To

overcome these shortcomings, hollow mesoporous silica nanoparticles (HMSNs) were

chosen as carriers for engineering a novel drug delivery system with specifically targeting,

high drug loading, sustained release behaviour, and negligible toxicity.

HMSNs are one of the most promising carriers for effective drug delivery due to their

large surface area, high volume for drug loading and excellent biocompatibility. However

the non-ionic surfactant templated HMSNs often have a broad size distribution, a

defective mesoporous framework or a missed hollow structure because of the difficulties

involved in controlling the formation and organization of micelles for the growth of silica

framework. In this study, a novel “Eudragit assisted” strategy has been developed to

fabricate HMSNs by utilising the Eudrgit nanoparticles as cores and to assist the self-

assembly of micelle organisation. Highly dispersed mesoporous silica spheres with intact

hollow interiors and through pores on the shell were achieved after the removal of

Eudragit core and surfactant micelles. The HMSNs have shown a high surface area (670

m2/g), small diameter (120 nm) and uniform pore diameter (2.5 nm). Eudragit S100 was

proven to assist in the spontaneous organization of the Triton X-100 (TX100) micelles

into structurally ordered nanoscale architecture via hydrogen bonding. The HMSNs have

demonstrated non-cytotoxicity to colorectal cancer cells SW480. The high quality of

HMSNs and their excellent biocompatibility demonstrate that the Eudragit assisted

HMSNs can be potential intracellular vehicles for targeted drug delivery.

Besides biocompatibility, targeting functions are essential for drug delivery systems as

they have the potential to precisely target and kill cancerous cells while leaving normal

cells unharmed. Given that human epidermal growth-factor receptor (EGFR) is over

expressed by tumour cells and their natural ligand, epidermal growth factor (EGF) has

Page 6: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

VI

emerged as an attractive targeting molecule labelled onto nanoparticles for cancer therapy.

To generate targeting function of the synthesized HMSNs, EGF was grafted to HMSNs.

HMSNs with and without EGF bio-ligands were covalently linked to a fluorescent dye

(FITC), making them detectable by fluorescence microscopy and flow cytometry. In vitro

studies with colorectal cancer cells SW 480 overexpressing EGFR demonstrated that

HMSNs with EGF bioconjugates (HMSN-EGF) achieved nearly 25 times higher cellular

internalization than the unfunctionalised control at a concentration of 10 μg/mL and an

incubation time of 30 mins. In addition, this superior cellular uptake of HMSN-EGF was

inhibited when EGFR in cancer cells was knocked down by free EGF. Similarly, the

targeting function of HMSN-EGF was very limited when they were applied to cells

without EGFR expression (SW 620), revealing the very high specificity of HMSN-EGF

to EGFR positive cells and the contribution of EGF-EGFR interaction.

On the other hand, the properties of grafted targeting ligands, in terms of effective

immobilization, stability and binding efficiency, significantly affect the subsequent

cellular interactions of HMSNs because the performance of biomaterials in a biological

environment is largely influenced by their surface properties. Although the utilization of

EGF for targeting delivery is well established, a precise control of the density of EGF

attachments on nanoparticles is limited. It is found that a small change in the quantity of

EGF labels on HMSNs may make a significant difference in the subsequent cell activities

of HMSN-EGF. To address this issue, hollow mesoporous silica nanoparticles

functionalized with amine group were conjugated with EGF using carbodiimide

chemistry. The time of flight secondary ion mass spectrometry (ToF-SIMS), a very

surface specific technique (penetration depth < 1.5 nm), was employed to study the

binding efficiency of the EGF to the nanoparticles. Principal component analysis (PCA)

was employed to track the relative surface concentrations of EGF on the surface of

HMSNs. Effective control of the quality and density of the resulting grafted EGF on

nanoparticles was achieved which would provide opportunities for engineering targeted

drug delivery systems with tailored targeting efficacy in the future.

Furthermore, the capacity to encapsulate drugs and perform a sustained release is one of

the most important requirements for a drug delivery system. To evaluate the drug loading

and releasing effectiveness of the HMSNs, the model drug, 5-fluorouracil (5-FU), was

Page 7: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

VII

loaded in the resultant HMSNs and achieved a sustained release profile. However, for

drug delivery systems, an adequate intratumor concentration of targeted drug is essential

for effective cancer therapy, which requires a high drug loading capacity of drug carriers.

Moreover, it was reported that, in addition to sustained release, specific drug release rates

are always pursued for personal treatment as different stages of disease evolution need

different drug release rates. In this regard, smart delivery systems with a higher drug

encapsulation and more effective control of drug release rate were of great importance.

To meet these requirements, ways to enhance the loading capacity of HMSNs specifically

to 5-FU were investigated and the results showed that 5-FU loading capacity of HMSNs

can be tuned by introducing chemical groups onto HMSNs, varying pH values for loading

process and generating electrostatic attractions between 5-FU and HMSNs. The highest

5-FU encapsulation ability of HMSNs was achieved when these nanocariers conjugated

with amine groups (-NH2). The highest loading capacity of 288.9 mg(5-FU)/g(HMSNs) was

observed at pH 8.0 which is 2 times of that reported in a similar study. On the other hand,

the release rate of 5-FU from HMSNs was controlled by changing the structure parameter

of HMSNs in terms of pore size, shell thickness and particle diameter. Experimental

results revealed that a larger pore size resulted in a quicker release rate while a thicker

shell decreased the amount of drug released from HMSNs. Moreover, particle diameter

is not as important as pore size and shell thickness in determining the release profile of

HMSNs.

In addition to excellent loading and release, overcoming multidrug resistance (MDR) is

a new requirement set for modern drug delivery systems. MDR referring to the ability of

cancer cells to adapt and survive from chemotherapy is one of the most complex and

challenging problems in cancer treatments. Although nanopariticle based drug delivery

systems are often used to prevent MDR in cancer treatment by sidestepping drug

resistance mechanisms, the development of drug resistant genes by cancer cells ultimately

leads to unsatisfied outcomes. To overcome the 5-FU resistance, HMSNs were modified

to deliver siRNA to silence the gene expression involved in MDR. It is found that rather

than surface charge, pore size is a key parameter of HMSNs for effective siRNA

encapsulation. In particular, when the pore size of HMSNs was expanded form 2.5 nm to

4.3 nm, an obvious increase in siRNA loading was found. Almost all siRNA can be

encapsulated into HMSNs when HMSNs have a pore size larger than 4.3 nm and the

Page 8: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

VIII

particle to siRNA ratio is 100.

In summary, high quality HMSNs were synthesized via a newly developed strategy

assisted by Eudragit. These HMSNs were proven to be non-cytotoxic to colorectal cancer

cells SW480. The HMSNs have a high drug and siRNA loading capacity, sustained and

controlled release, and efficient and selective cell internalization when bioconjugated

with Epidermal Growth Factor (EGF). The promising in vitro cell tests have revealed the

great potential of the HMSNs to be used for colorectal cancer therapy.

Page 9: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

IX

CContents

Chapter 1 Introduction ................................................................................................... 1

1.1. Research aims and objects ................................................................................... 1

1.2. Outline of the thesis ............................................................................................. 3

Chapter 2 Literature review ........................................................................................ 5

2.1. Background ................................................................................................................. 5

2.2. Controlled and slowed drug delivery systems .......................................................... 7

2.2.1. Controlled and slowed drug delivery ..................................................................... 7

2.2.2. Nanoparticle based drug delivery systems ............................................................. 9

2.2.3. Active targeting drug delivery ............................................................................. 10

2.2.4. Platforms for nanoparticles based drug delivery systems .................................... 12

2.3. Preparation and modification of HMSNs ............................................................... 19

2.3.1. Architectural features of MSNs ........................................................................... 19

2.3.2. Fabrication of HMSNs ......................................................................................... 20

2.3.3. Modification of MSNs ......................................................................................... 24

2.3.4. Qualitative characterization of HMSNs ............................................................... 28

2.3.5. MSNs’ pharmacokinetics ..................................................................................... 33

2.4. Drug loading and release .......................................................................................... 34

2.5. Drug resistant phenomenon ..................................................................................... 37

Chapter 3 Materials and methodology ...................................................................... 41

3.1 Introduction ............................................................................................................... 41

3.2 Materials .................................................................................................................... 41

3.3 Preparation of versatile HMSNs ............................................................................. 42

3.3.1 Formation of Eudragit S-100 nanoparticles ......................................................... 42

3.3.2 Fabrication of HMSNs ......................................................................................... 42

3.3.3 Control of particle size ......................................................................................... 43

3.3.4 Control of shell thickness .................................................................................... 43

3.3.5 Pore expansion of HMSNs .................................................................................. 44

3.4 Modifications of HMSNs .......................................................................................... 44

3.4.1 Amine-, methyl- and cyano-functionalization ..................................................... 44

3.4.2 Carboxyl-modification (HMSN-COOH) ............................................................. 45

3.4.3 FITC-modification (FITC-HMSNs) .................................................................... 45

3.4.4 EGF labelling (HMSN-EGF) ............................................................................... 45

3.5 5-FU loading and release study ................................................................................ 46

3.5.1 Measurement of 5-FU oil : water partition coefficients ....................................... 46

Page 10: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

X

3.5.2 5-FU loading ........................................................................................................ 46

3.5.3 5-FU release ......................................................................................................... 47

3.6 Cell experiments ........................................................................................................ 48

3.6.1 Cell culture ........................................................................................................... 48

3.6.2 Cellular uptake of HMSNs and EGF-HMSNs in colorectal cancer cells ............ 48

3.7 Characterizations ...................................................................................................... 49

3.7.1 Electron microscopy ............................................................................................ 49

3.7.2 Nitrogen absorption and desorption experiments ................................................ 49

3.7.3 Small angle X-Ray diffraction (SAXRD) ............................................................ 49

3.7.4 Thermo-gravimetric analysis (TGA) ................................................................... 49

3.7.5 Fourier transform infrared spectrometer (FTIR) .................................................. 50

3.7.6 Time of flight secondary ion mass spectrometry (ToF-SIMS) ............................ 50

3.7.7 Nuclear magnetic resonance (NMR) ................................................................... 51

3.7.8 Dynamic light scattering (DLS) ........................................................................... 51

3.7.9 Confocal microscopy ........................................................................................... 51

Chapter 4 Synthesis of tailored HMSNs for nanotheranostics ................................. 52

4.1 Introduction ............................................................................................................... 52

4.2 Synthesis of HMSNs with designed properties for drug delivery systems ........... 54

4.2.1 Preparation and characterization of HMSNs ....................................................... 54

4.2.2 Innovations of Eudragit-assisted HMSNs fabrication ......................................... 60

4.2.3 Mechanism of Eudragit-assisted HMSNs fabrication .......................................... 63

4.2.4 5-FU encapsulation and the in vitro release ......................................................... 69

4.3 Pore size tuning for controlled release .................................................................... 71

4.3.1 BET analysis of pore size .................................................................................... 73

4.3.2 PALS analysis of pore size ................................................................................. 74

4.3.3 Controlled and sustained release of HMSNs with different pore size ................ 76

4.4 Cytotoxicity and non-specific cellular uptake of HMSNs ..................................... 80

4.5 Conclusions ................................................................................................................ 83

Chapter 5 Modification of HMSNs for targeting delivery ...................................... 84

5.1 Introduction ............................................................................................................... 84

5.2 A facile, one step strategy for EGF labelling .......................................................... 85

5.2.1 Synthesis of amine functionalized HMSNs (HMSNs-NH2) ............................... 86

5.2.2 Preparation of EGF grafted HMSNs and characterization of EGF labelling ........... ............................................................................................................................. 92

5.2.3 The effect of EGF concentration on surface chemistries of the HMSNs ............. 95

5.2.4 Analysis on immonium ions distinguishing samples with different surface chemistry .............................................................................................................. 99

Page 11: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

XI

5.3 Comparison of newly developed grafting method with other methods.............. 101

5.3.1 Superior EGF grafting efficiency by the new grafting method ......................... 101

5.3.2 Increase EGF labels on HMSNs for enhanced targeting efficiency .................. 103

5.4 Effectively target colorectal cancer cells by EGF grafted HMSNs ..................... 106

5.4.1 Effective targeting effect of HMSN-EGF to EGFR positive CRC cells ............ 106

5.4.2 Reduced targeting effect of HMSN-EGF by pretreating SW480 cells with free EGF .................................................................................................................... 109

5.4.3 Non-targeting effect of HMSN-EGF to EGFR negative cell line SW620 ......... 110

5.5 Conclusions .............................................................................................................. 111

Chapter 6 Functionalization of HMSNs for favourite 5-FU loading and siRNA encapsulation .......................................................................................... 113

6.1 Introduction ............................................................................................................. 113

6.2 Enhanced 5-FU loading capacity of HMSNs by precise functionalization ........ 114

6.2.1 Study of 5-FU partition coefficient .................................................................... 114

6.2.2 Preparation and characterization of functionalized HMSNs .............................. 117

6.2.3 Regulate 5-FU loading capacity of HMSNs by modifications .......................... 123

6.3 Control of 5-FU loading and release ..................................................................... 124

6.3.1 The effect of pH on drug loading and release .................................................... 124

6.3.2 The effect of particle size and shell thickness on drug loading and release ...... 126

6.3.3 The effect of EGF on HMSNs’ drug loading and release behaviours ............... 129

6.4 SiRNA encapsulation based on HMSNs ................................................................ 131

6.4.1 Increased surface charge of HMSNs for SiRNA loading .................................. 131

6.4.2 Pore size expansion of HMSNs for efficient SiRNA encapsulation .................. 135

6.5 Conclusions .............................................................................................................. 142

Chapter 7 Conclusions and future works ................................................................. 144

7.1 Overview of current work ...................................................................................... 144

7.2 Limitations and future work .................................................................................. 148

7.2.1 Targeting efficiency of EGF-HMSNs ................................................................ 148

7.2.2 siRNA release and knockdown efficiency ......................................................... 149

7.2.3 Preparation multi-anticancer drug loaded HMSNs ............................................ 149

7.2.4 Animal experiences............................................................................................ 151

Bibliography ................................................................................................................. 152

Publication List ............................................................................................................. 175

Page 12: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

XII

List of Table Table 2. 1 Methods employed to introduce functional groups ....................................... 25

Table 4. 1 Proton diffusion coefficients (D) of Eudragit and Eudragit/TX100 mixture . 67

Table 4. 2 Physicochemical Properties of HMSNs ......................................................... 74

Table 4. 3 PALS data for HMSNs .................................................................................. 75

Table 4. 4 Loading capacity and other physical parameters of HMSNs ......................... 77

Table 4. 5 Values of diffusional exponent and corresponding drug release mechanism 79

Table 5. 1 Positive fragment ions (immonium ions) used in PCA ................................. 97

Table 6. 1 Partition coefficient of 5-FU and relevant parameters ................................. 115

Table 6. 2 Structure parameters of functionalized and non-functionalized HMSNs .... 120

Table 6. 3 Loading capacity of functionalized and non-functionalized HMSNs .......... 123

Table 6. 4 Parameters of HMSNs and HMSN-NH2 ...................................................... 132

Table 6. 5 siRNA encapsulation efficiency................................................................... 135

Table 6. 6 Structure parameter of HMSNs-S, HMSN-M and HMSNs-L ..................... 138

Page 13: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

XIII

List of Figure

Figure 2. 1 Tine line showing advancements made in CRC treatment ............................. 5

Figure 2. 2 History of adjuvant therapy of colon cancer................................................... 6

Figure 2. 3 Drug’s pharmacokinetic fates of different administration modes .................. 7

Figure 2. 4 Schematic structure of EGFR ....................................................................... 10

Figure 2. 5 EGFR signalling pathway ............................................................................. 11

Figure 2. 6 Multifunctional carbon nanotubes in cancer therapy................................... 14

Figure 2. 7 Preparation of modified SWCNTs............................................................... 15

Figure 2. 8 Mesoporous silica nanoparticles as a platform for drug delivery ................. 16

Figure 2. 9 Transmission electron microscopy of mesoporous silica nanoparticles ....... 16

Figure 2. 10 Schematic representation of the drug delivery system by using mesoporous

silica nanoparticles .......................................................................................................... 18

Figure 2. 11 Different particle morphologies of MSN. ................................................... 19

Figure 2. 12 The synthesis process of HMSNs ............................................................... 21

Figure 2. 13 Schematic illustration of the formation process of HMSs .......................... 22

Figure 2. 14 Scheme of the synthetic procedure of HMSs ............................................. 23

Figure 2. 15 Formation (left) and microscopic structure (right) of hollow/rattle-type

MSNs .............................................................................................................................. 23

Figure 2. 16 Introduction of functional groups in different regions of MSN. ................ 24

Figure 2. 17 Functionalized hollow mesoporous silica nanoparticles ............................ 26

Figure 2. 18 TEM morphology of HMSNs ..................................................................... 28

Figure 2. 19 XRD patterns of mesoporous silica nanoparticles synthesized at different

pH .................................................................................................................................... 29

Page 14: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

XIV

Figure 2. 20 Six categories of gas adsorption isotherms for porous materials with

different pore size. ........................................................................................................... 30

Figure 2. 21 Adsorption isotherms of nitrogen in mesoporous silica nanoparticles of

different pore sizes.. ........................................................................................................ 31

Figure 2. 22 Pore geometry ............................................................................................. 32

Figure 2. 23 Effect of pore geometry on shapes of hysteresis loop ................................ 32

Figure 2. 24 TEM images of MSNs (black dots) endocytosed by human cervical cancer

cells ................................................................................................................................. 34

Figure 2. 25 HMSNs based drug delivery system for simultaneous joint lubrication and

treatment .......................................................................................................................... 35

Figure 2. 26 HMSNs based drug delivery system for high-intensity focused ultrasound

(HIFU)-mediated intravenous drug delivery. .................................................................. 36

Figure 2. 27 Rational design of a multifunctional micellar nanomedicine for targeted co-

delivery of siRNA and DOX to overcome multidrug resistance .................................... 38

Figure 2. 28 System for delivery of doxorubicin (DOX)/siRNA combination based on

quantum dots (QDs) modified with β-cyclodextrin ........................................................ 40

Figure 4. 1 Fabrication possess of HMSNs ..................................................................... 54

Figure 4. 2 a, b & c) SEM images of HMSNs and d, e & f) TEM image of HMSNs

displaying hollow interior ............................................................................................... 56

Figure 4. 3 DLS curve of HMSNs .................................................................................. 57

Figure 4. 4 Small angle XRD pattern of HMSNs ........................................................... 58

Figure 4. 5 Nitrogen adsorption-desorption isotherm (A) and the BJH pore size

distribution of HMSNs (B) ............................................................................................. 59

Page 15: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

XV

Figure 4. 6 TEM images (a) and SEM images (b & c) of HMSNs-S and TEM images

(d) of HMSNs-C and SEM images (e & f) ..................................................................... 61

Figure 4. 7 Zeta potential of HMSNs-S and HMSNs-N ................................................. 62

Figure 4. 8 Stability study of HMSNs-C and HMSNs-S ................................................ 62

Figure 4. 9 Formation schematics and of TX100/Eudragit S100 composite micelles (a)

and proposed reactions between TX100 and Eudragit S100 .......................................... 64

Figure 4. 10 DLS of Eudragit S100 nanoparticles, TX100 and their mixture ................ 65

Figure 4. 11 FTIR spectra of Eudragit S100 nanoparticles, TX100 and the mixture of

Eudragit and TX100 ........................................................................................................ 66

Figure 4. 12 1H NMR spectra of Eudragit S100 nanoparticles (upper) and

Eudragit/TX100 mixture (bottom); Inserts are TX100 molecular structure. .................. 67

Figure 4. 13 SEM images (a) and TEM image (b) of non-Eudragit assisted MSNs....... 68

Figure 4. 14 (a):5-FU loading profile and (b): 5-FU release profile of HMSNs ............ 70

Figure 4. 15 Three stages of the slow release kinetic model .......................................... 71

Figure 4. 16 Representative schematic illustration of controlling drug release rate by

tuning the pore size of HMSNs ....................................................................................... 72

Figure 4. 17 Adsorption-desorption isotherms (A) and corresponding pore size

distribution (B) of HMSNs ............................................................................................. 73

Figure 4. 18 The relationship between porosity and τn ................................................... 75

Figure 4. 19 In vitro release profiles of HMSNs with different pore size in PBS with a

pH of 5.0 ......................................................................................................................... 77

Figure 4. 20 In vitro release kinetics of HMSNs with different pore size in PBS with a

pH of 5.0 ......................................................................................................................... 79

Figure 4. 21 Cytotoxicity of HMSNs on SW 480 cells at different concentrations....... 80

Page 16: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

XVI

Figure 4. 22 Cell uptake of FITC-HMSNs with varying time and concentration by

SW480 cells. ................................................................................................................... 82

Figure 5. 1. a, b & c) SEM images of HMSNs and d, e & f) TEM image of HMSNs

displaying hollow interior ............................................................................................... 88

Figure 5. 2 A) Nitrogen adsorption-desorption isotherm; B) Small angle XRD pattern of

HMSNs; C) BJH pore size distribution. .......................................................................... 89

Figure 5. 3 FTIR spectra of HMSNs-NH2 and HMSNs ................................................. 90

Figure 5. 4 TGA thermograms of HMSNs & HMSNs-NH2 (a) and 29Si single pulse

excitation (SPE) spectrum of HMSNs-NH2 .................................................................... 91

Figure 5. 5 Schematic diagram showing the process of EGF grafting onto HMSNs ..... 92

Figure 5. 6 Positive survey mass spectra for: (A) HMSNs surface; (B) HMSNs-NH2

surfaces; (C) EGF; (D) HMSNs-NH2-EGF..................................................................... 94

Figure 5. 7 Positive survey mass spectra of HMSN-NH2-EGF surface using different

EGF concentration .......................................................................................................... 95

Figure 5. 8 Plot of ratios of the intensities of the 28 m/z (from HMSNs) to 59 m/z, 70

m/z and 58 m/z (from EGF) ............................................................................................ 96

Figure 5. 9 Scores on PC1 and PC2 (a) and corresponding loading plots on PC1 (b) and

PC2 (c) of immonium ions from static positive spectra of HMSNs-NH2-EGF prepared

with EGF concentration .................................................................................................. 98

Figure 5. 10 Comparison of characteristic immonium ions intensities between samples

....................................................................................................................................... 100

Figure 5. 11 Routes of EGF lebelling based on carbodiimide chemistry ..................... 102

Figure 5. 12 SEM images of HMSN-COOH-EGF (a) and HMSN-NH2-EGF (b)........ 103

Figure 5. 13 Positive survey mass spectra .................................................................. 105

Page 17: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

XVII

Figure 5. 14 Cellular uptake of EGF grafted and plain HMSNs in SW480 cells ......... 106

Figure 5. 15 Confocal microscopic images of SW480 cells after incubating with FITC

labelled HMSN-EGF (a & b) or HMSNs (c & d) with a concentration of 5mg/ml and an

incubation time of 2 hours ............................................................................................ 108

Figure 5. 16. The cellular uptake rate of HMSN-EGF and HMSNs in SW 480 cells with

or without 5 μM of free EGF pretreatment ................................................................... 109

Figure 5. 17. Confocal microscope images of EGF-HMSNs and HMSNs in SW 480

cells. .............................................................................................................................. 110

Figure 5. 18. The mean fluorescence intensity in the SW 620 cells treated with different

concentrations of HMSNs or EGF-HMSNs for 2 hours. .............................................. 111

Figure 6. 1 percentages of 5-FU dissolved in each phase with varying octanol to water

ratio ............................................................................................................................... 116

Figure 6. 2 TEM images of HMSNs (a), HMSN-NH2 (b), HMSN-COOH (c), HMSN-

CN (d) and HMSN-CH3 (e)........................................................................................... 117

Figure 6. 3 N2 adsorption-desorption isotherms (a) and the corresponding pore size

distributions (b) of plain and functionalized HMSNs ................................................... 119

Figure 6. 4 FTIR spectra of plain and functionalized HMSNs and plain HMSNs ....... 120

Figure 6. 5 TGA curves of functionalized HMSNs and plain HMSNs ........................ 122

Figure 6. 6 5-FU loading capacity of HMSNs and HMSN-NH2 at different pH values

(a) and UV spectra of remaining 5-FU after 24 hours loading by HMSN-NH2 at

different pH values (b) .................................................................................................. 125

Figure 6. 7 Two possible 5-FU anions (AN1 and AN3) and dianion ........................... 126

Figure 6. 8 SEM images of HMSNs with different particle size: 100 nm (a & d); 200 nm

(b & e) and 300 nm (c & f) ........................................................................................... 127

Page 18: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

XVIII

Figure 6. 9 Loading capacities (a) and release profiles (b) at pH 5.5 of HMSNs with

varying particle size ...................................................................................................... 127

Figure 6. 10 SEM images of HMSNs with different shell thickness: 10 nm (a & d); 15

nm (b & e) and 30 nm (c & f) ....................................................................................... 128

Figure 6. 11 Loading capacities (a) and release profiles (b) at pH 5.5 of HMSNs with

varying shell thickness .................................................................................................. 129

Figure 6. 12 SEM images (a & b) and TEM image (c) of HMSN-EGF ....................... 130

Figure 6. 13 Loading (a, at pH 8.0) and release (b, at pH 5.5) profiles of HMSNs with

and without EGF attachments ....................................................................................... 130

Figure 6. 14 SEM image (a) and zeta potential (b) of HMSN-NH2 .............................. 132

Figure 6. 15 siRNA encapsulation efficiency of HMSN-NH2. Various weight ratio of

particles/siRNA from 5 to 200 were applied. ............................................................... 133

Figure 6. 16 Zeta potential of HMSNs, HMSN-NH2-1, HMSN-NH2-2 ....................... 134

Figure 6. 17 siRNA encapsulation efficiency of HMSN-NH2-2 (a) and HMSN-NH2-3

(b). Various weight ratio of particles/siRNA from 5 to 200 were applied. ................... 134

Figure 6. 18 SEM and TEM images of HMSNs-S (a & b), HMSNs-M (c & d) and

HMSNs-L (e & f) .......................................................................................................... 137

Figure 6. 19 Zeta potential (a) and TGA (b) of aminated HMSNs, HMSNs-S, ........... 139

Figure 6. 20 siRNA encapsulation efficiency of HMSNs-S, HMSNs-M and HMSNs-L.

Various weight ratio of particles/siRNA from 5 to 200 were applied. ......................... 141

Figure 6. 21 Proposed siRNA encapsulation of HMSNs with different pore sizes ...... 142

Page 19: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

1

Chapter 1 Introduction

1.1. Research aims and objects

Colorectal cancer (CRC) remains a significant health burden in the world. For its

treatment, the anticancer drug of fluorinated pyrimidine 5-fluorouracil (5-FU) as the

backbone stands alone in CRC therapy [1]. Nearly every chemotherapy regimen for CRC

includes 5-FU. However, owing to its inadequate intratumor concentrations and severe

side effects, CRC patients pay great prices when receive the treatments. Obviously, the

therapeutic utility of 5-FU on CRC treatment has been hindered by lacking a safe delivery

system. The challenge is to extend its circulating half-life and to control its distribution

in tissues.

One way to address this challenge is to engineer a system where a constant dose of 5-FU

is delivered directly to colorectal cancer cells over an extended period [2, 3]. After

reviewing the relationship between drug’s pharmacokinetic fates and administration

modes, an optimally efficacious approach toward achieving the goal of continuous,

nontoxic bioactivity would involve controlled and sustained drug delivery systems based

on nanocarriers labelled with cell targeted molecules.

When compared to polymer and carbon nanotubes which have been widely investigated

as carriers in drug delivery systems, hollow mesoporous silica nanoparticles (HMSNs)

have the potential to serve as a versatile drug nanocarrier for smart drug delivery. For

example, HMSNs can level off the drug concentration at the targeted area as the drug

molecules released from the ordered pores by a tunable diffusion. This will reduce drug

dosage and side effects. Zero premature release before reaching the targeted sites can be

attained when using functionalized HMSNs as carriers. Furthermore, HMSNs have large

pore volume and surface area and are an ideal platform for cancer treatment because of

their higher drug loading [4].

Page 20: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

2

However, owning to the lack of structure and morphology controlling, applications of

HMSNs for drug delivery are still limited. Thus, technology to fabricate, functionalize

and modify HMSNs with expected shapes, sizes and defined properties is vitally essential.

Additionally, as pointed out by Argyo [5] in a very recent review, cell type specificity is

a challenge that still needs to be overcome by using these types of materials as

nanoplatforms for drug delivery systems. This is because extensive studies on

nanoparticle based targeted drug delivery systems were relied on the enhanced-

permeation end-retention effect (EPR) [6, 7] of cancer cells. However, this effect is

limited just to vascularized tumours and is usually not sufficient for a complete

eradication of the cancer [8]. Base on the comprehensive literature review, we

hypothesized that active targeting of cancer cells by epidermal growth factor (EGF) on

the surface of the nanoparticles [9-11] could be an efficient strategy to provide an

enhanced accumulation of the nanoparticles in targeted cell population (CRC cells) and

thus represent a powerful tool to improve the efficacy of the cancer treatment.

Therefore the main objectives of this project are:

1. Engineering new ways to prepare HMSNs with designed properties and tailored

structure for targeted drug delivery and investigating the biosafety of these

synthesized HMSNs.

2. Detail studies on chemical modifications of HMSNs to achieve high 5-FU loading

and sustained release and the fabrication of HMSNs with controllable drug release

rate. On the other hand, the 5-FU behaviours at different pHs were different, which

is another factor that needs to be investigated to obtain a high 5-FU encapsulation

rate by HMSNs.

3. Grafting of targeting ligands on HMSNs to attain a high specificity of HMSNs for

further application. The relationship between grafted biological attachments and

properties as well as biological behaviours need to be fully studied.

4. In vitro studies with colorectal cancer cells to assess the biological activities of

Page 21: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

3

HMSNs in biological environment are require to demonstrate HMSNs as an effective

vehicle for successful intracellular delivery of 5-FU for colorectal cancer therapy.

1.2. Outline of the thesis

This thesis consists of 7 chapters as follows:

Chapter 2 provides a comprehensive review of the challenges existing in colorectal cancer

treatments, the background information and recent studies in the areas of nanoparticle

based drug delivery systems, nanoplatforms used for sustained release, and biomarkers

identified for targeted drug delivery. Throughout this chapter, gaps in 5-FU delivery

specifically for colorectal cancer therapy are identified. In addition, feasible methods

proposed to fill these gaps are introduced and discussed.

Chapter 3 outlines all the materials and experimental procedures used in this thesis.

Techniques for the fabrication and modification of HMSNs as well as the strategies

applied for testing the drug loading and release properties, for introducing biological

targeting labels and for in vitro experiments are illustrated. Methods for the

characterization of HMSNs such as SEM, TEM, XRD, FTIR and TGA etc. are described

in details.

Chapter 4 focuses on the synthesis of HMSNs and the fabrication mechanism.

Characterizations of structure, morphology and biocompatibility are also studied in this

chapter.

Chapter 5 aims at the study of the targeting function of current drug delivery systems.

Extremely high specificity of HMSNs for targeting delivery is generated by introducing

a bio-ligand, EGF, onto HMSNs. Especially, the characterization of EGF attachments and

EGF concentration on the surface of nanoplatforms were fully investigated in this chapter.

Page 22: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

4

Chapter 6 explores ways to effectively increase 5-FU and siRNA loading capacity of

HMSNs. In addition, strategies regarding better control of drug release kinetics from

HMSNs are investigated.

Chapter 7 summarizes the major findings and contributions of this project and identifies

some limitations in the current study. Future research directions in this area are also

proposed in this chapter.

Page 23: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

5

Chapter 2 Literature review

2.1. Background

Colorectal cancer (CRC) manifesting as cancerous growths in the colon and rectum, is

the fourth most commonly diagnosed malignancy in the world. Over the past 25 years,

significant advances have been made in the treatment of CRC (Figure 2.1), leading to a

steady improvement of the treatment of this illness [12]. However, it still remains a

significant health burden in the world with over 1,000,000 cases and a disease-specific

mortality of nearly 33% even in developed countries [1]. Obviously, the treatment of CRC

is still full of challenges and many of them are still unsolved, such as toxic effects of

chemotherapy [13] and tumours developing resistance to drugs [14].

Figure 2. 1 Tine line showing advancements made in CRC treatment [12]

Among drugs for CRC treatment, the fluorinated pyrimidine 5-fluorouracil (5-FU) as the

backbone stands alone in CRC therapy (Figure 2.2). Fluorouracil inhibits thymidylate

synthase which is the rate-limiting enzyme for pyrimidine nucleotide synthesis [15].

Nearly all chemotherapy regimens shown to be effective for CRC treatments incorporate

5-FU [1, 16], highlighting its crucial and first-line role in treating colorectal cancer.

Page 24: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

6

Figure 2. 2 History of adjuvant therapy of colon cancer [17]

Many 5-FU intravenous administration schedules are in clinical use, but the efficacy is

limited, because intravenous 5-FU injection is hard to reach optimal dose effectiveness at

targeted sites due to 5-FU’s very short half-life and quick metabolism in plasma

circulation [18]. Moreover, intravenous administration of 5-FU has been shown to cause

severe side effects [19] because it induces high drug concentration above the maximum

tolerable drug concentration in the plasma [20, 21]. Obviously, the therapeutic effect of

5-FU on cancer treatment has been hindered by the lack of a safe delivery system.

In summary, 5-FU plays an important role in CRC chemotherapy. However, the more

extensive use of 5-FU is limited by its inadequate intratumor concentrations and severe

side effects which may result from its short half-life and non-specific systemic organ

distribution. Therefore, new strategies for effective and safe 5-FU delivery is urgently

needed for colorectal cancer therapy [22].

Page 25: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

7

2.2. Controlled and slowed drug delivery systems

2.2.1. Controlled and slowed drug delivery

One way to overcome the limitations associated with 5-FU for CRC chemotherapy is to

engineer a systems where a constant dose of 5-FU is delivered directly to colorectal

cancer cells over an extended period. With the development of pharmaceutics, a

medicinal agent’s therapeutic effect regarding its pharmacokinetic fate in vitro and in vivo

can be optimized by the utility of precise drug administrations [23]. Figure 2.3 depicts the

relationship between drug’s serum concentration and administration modes. Obviously,

an optimally efficacious approach toward achieving the goal of continuous, nontoxic

bioactivity would involve controlled and sustained administrations which provide a

considerably extended therapeutic delivery of the drugs.

Figure 2. 3 Drug’s pharmacokinetic fates of different administration modes [23]

Page 26: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

8

Engineering sequential release dosage forms is beneficial for optimal therapy in terms of

safety, efficacy and patient compliance. Ideally, a controlled release dosage form will

maintain a therapeutic concentration of the drug in the blood throughout the dosing

interval [24, 25]. With this aim, techniques for the preparation of sustained release

formulations were investigated including preparation of prodrug [26, 27], incorporation

of drugs into carriers and coating with special materials [28, 29]. In addition, liquisolid

technique, referring to the control of the dissolution rate of drugs, is one of the best and

most successful methods because of its simplicity and low cost [30]. Particularly, a

“liquisolid system” involved in blending drug solutions with selected coating materials

or carriers before drying to form compactible powder mixtures [31]. It has been used to

produce sustained release formulations of water-soluble drugs such as propranolol

hydrochloride [24] and the release of drug from these formulations had been proven to

follow zero-order release kinetics [30-32].

At present, 5-FU sequential release pellets have been prepared [22, 33, 34] and evidenced

to exhibit inhibition for tumour growth and proliferation to some extent [35]. This

augmentative therapeutic efficiency is due to the extension of 5-FU plasma half-life by

the sequential release [34]. However, when compared to free drug, these pellets showed

limited advantages in minimising side effects of 5-FU because only limited control over

the release kinetics is possible [36], where an initially rapid release is presented and

followed by an inability to maintain a sufficient drug concentration for an extended period

of time [37]. This limitation is due, in part, to the poor water-soluble nature of 5-FU and

its low molecular weight (130.08). An alternative approach has been reported to control

the release of 5-FU by attaching the drug covalently to the network of selected materials

or carriers [38] via degradable linkages. Nevertheless, the dependence on temperature,

pressure or pH [39] and susceptibility to degradation by esterases [40] often present a

challenge, which finally leads to deviations from predicted release kinetics.

While sustain release formulations offer promising opportunities to enhance anti-tumour

efficacy of 5-FU with low systemic toxicity, the unpredictable tumour environment and

susceptibility of drugs to collapse on the way to tumour site make it difficult to precisely

control the drug release kinetics. Therefore, an alternative or complementary strategy is

Page 27: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

9

required to be developed to overcome these shortcomings, where a biomaterial platform

is applied to enable sustainable, predictable, and tunable 5-FU release.

2.2.2. Nanoparticle based drug delivery systems

Recently, nanomedicine has been demonstrated to significantly enhance the efficacy of

cancer therapy because nanoparticles maintain a prolonged effective drug concentration

which renders cancer cells expose sustainedly to the drug [20]. The feasibility of

nanoparticle-based anticancer therapy has been proven in clinical use [6]. Owning to their

small size and the characteristic features of tumour biology, like the leaky vasculature

and the impair lymphatic system, nanoparticles can reach certain solid tumours and

release biologically active cargos by diffusion into the tumour tissues. This method is

called passive targeting approach [6]. Previously, passive targeting nanoparticles had

been used as platforms to successfully control and sustain drug release for therapy. For

example, it was shown that selective anticancer effect had been enhanced by passive

targeting self-assembled nanoparticles [41]. Similarly, silica nanoparticles as drug

carriers have effectively demonstrated heart-targeting profile by passive targeting

approach [42].

These inspired studies could help to engineer a 5-FU delivery system to the colon basing

on nanoparticles with small enough size to improve the transportation efficiency of 5-FU

to cancer cells. Although there have been some preliminary reports [43], the application

of this passive targeting approach in colon targeted drug delivery systems is still hindered.

Because of the random nature of this approach, it is hard to control the drug release

kinetics and this in turn decreases the efficacy of localizing nanomedicines at cancer cells

in vivo. As we know, our body is so perfectly structured that some physiological drug

barriers in our body are existing to protect organs from external toxins, like the

gastrointestinal barrier [44]. Moreover, cancer cells are smart enough that their survival

is achieved by presenting barriers to fight with anti-cancer drugs, such as drug resistant

genes and transport proteins for expelling drugs out of cancer cells. Unfortunately, as

colon is the distal segment of the gastrointestinal tract, when delivering 5-FU to the colon,

nanocarriers will encounter barriers motioned above on their route to the targeted site.

Page 28: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

10

Therefore, it would be desirable to combine the rational design of nanocarriers with novel

nanoscale targeting molecules.

2.2.3. Active targeting drug delivery

To address this challenge, rather than passive delivery, it is essential to find an actively

targeting region on the surface of colorectal cancer cells with higher specificity and

binding affinity. Therefore, understanding the pathology of colorectal cancer becomes

imperative.

In recent years, overexpression of growth factors and receptors has been implicated in

tumorigenesis and tumour progression of CRC [45, 46]. Dysfunctions in signalling

pathways have also been demonstrated to lead to proliferation, survival and migration of

cancers [47, 48]. The overexpression of epidermal growth factor receptor (EGFR) has

been documented in 30 – 90% of cases of advanced CRC, which has shown to involve in

numerous cellular responses including proliferation and metastatic spread [49]. It thus

becomes vital to understand the EGFR signalling pathway in CRC during neoplastic

transformation as it may enable us to programme the nanocarriers so they can actively

bind to colorectal cancer cells.

Figure 2. 4 Schematic structure of EGFR

EGFR is an endogenous cell surface glycoprotein with a molecular weight of 170 KDa

[11]. It comprises three domains that include a 622 amino acid extracellular domain, a 23

Page 29: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

11

amino acid hydrophobic transmembrane domain, and an intracellular tyrosine Kinase

domain with several phosphorylation sites (Figure 2.4) [49].

The EGFR signalling pathway consists of 4 sub-pathways: PI3K-mTOR pathway, JAK-

STAT pathway, PLC signalling pathway and RAS-ERK signalling pathway. On ligand

binding (signal activation), EGFR homo-dimerizes activate the intracellular kinase

domain. Then, importins (adopter proteins, like PI3K, JAK and PLC) bind with

transcription factors (like STAT and ERK), leading to the nucleus accumulation of EGFR.

As a result, tumour proliferation, metastasis and survival happen (Figure 2.5) [1].

Figure 2. 5 EGFR signalling pathway

(JAK, Janus kinase; PLC, phospholipase C; STAT, signal transducer and activator of transcription.)

With its association with tumour survival and proliferation, EGFR has the potential to

serve as an actively targeting region of colon targeted drug delivery system. Epidermal

growth factor (EGF), hepatin-binding EGF (HB-EGF), betacellulin, transforming

growth-factor-a (TGF-a), epiregulin and amphiregulin are six known endogenous ligands

of EGFR. Although any of these ligands can be used as a targeting agent for EGFR, EGF

with a good compromise between the binding affinity and the size of the molecule (EGF,

6.1 kDa) is one of the most commonly detected factors in humans [8]. It is hypothesized

Page 30: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

12

that introducing EGF as a ligand to EGFR on the surface of nanocarriers will exhibit

specific delivery of 5-FU into colorectal cancer cells, and thus enhance the resultant anti-

cancer activity. To the best of our knowledge, at present there is not EGFR targeting 5-

FU delivery system being exploited specifically for colorectal cancer.

2.2.4. Platforms for nanoparticles based drug delivery systems

Before designing a colon targeted drug delivery system, several questions should be

addressed. How to achieve a tight association between the targeting molecules and the

surface of nanocarrier is one of key questions. Also, as for colon targeted drug delivery

system, biodegradability and toxicological harmlessness is another point that should be

considered beforehand because the materials to be used for drug delivery may accumulate

in the human body on prolonged use. As we mentioned above, nanoparticles exhibit slow-

release profile, but different functions can be observed even for the same drug delivery

system when using different nanocarriers. In this regard, it would therefore be desirable

to study materials used for drug delivery and find out the one that can achieve the expected

functions greatly.

In order to achieve an ideal colon targeted drug delivery system, a variety of materials

have been exploited to meet different therapeutic needs, many with great potential for

topical delivery. Reports on materials with controlled release properties for topical

delivery include: cellobiose-derived monomer [50], carbon nanotubes [51], metals [52,

53], chitosan [54, 55], pectins [54, 56], guar gum [57, 58], inulin, and dextran. These

materials can be divided into steroids, e.g. polymer, carbon nanotubes and inorganic

nanoparticles.

2.2.4.1.Polymer

Polymers are a kind of widely used materials for colon targeted drug delivery system due

to their inherent properties. For example, excellent stability and safety, less toxicity and

gel forming make the utilisation of polymers for colon specific delivery wide and

intensive. Moreover they are more cost-effective, as they are abundant in various

Page 31: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

13

structures and exhibit diversified properties.

A large number of studies have been done in exploiting polymers as potential carriers for

colon specific delivery including chitosan [59], pectin [60, 61], sugar gum [58] and

dendrimer [62]. The effects of formulation parameters on drug release pattern are studied

by Das, Chaudhury et al. [54]. Elkhodairy, Barakat et al. [63] used low molecular weight

chitosan to prepare a new guar-based colon delivery formulation. Hiorth, Skoien et al.

[64] revealed that a high degree of swelling of chitosan-containing pellets may boost the

drug release. pH-responsive fluorinated dendrimer-based particulates were prepared by

Criscione, Le et al. [65] which exhibit controlled release of agents. Moreover, these dense

particulates are self-assembled and possess a high density of fluorine spins that can be

easily detected in vivo.

However, the issue associated with the utilisation of polymers is their high water

solubility. Polymers are easily assimilated by microorganisms or degraded by enzymes

on its way to the targeting site such as the human colon, resulting in breaking down of the

molecules back bone [66]. As a consequence, the molecular weight reduces as the

polymer undergoes a further degradation leading to the loss of mechanical strength,

unable to hold the targeting molecules (such as EGF) and drug (like 5-FU) any longer.

This is how the burst release and premature release happen.

2.2.4.2.Carbon nanotube

Carbon nanotubes (CNTs), regarded as one of the most promising nanomaterials, possess

the ability to deliver drug into living systems [67]. Owning to their unique

physicochemical properties like their inherent stability, novel structure, high aspect ratio

as well as high electrical and thermal conductivity, CNTs can not only detect the

cancerous cell, but also transport biologically active cargo into these living cells [68].

During the last several years, CNTs have been widely investigated in the field of cancer

treatment modality [69], containing gene [70] and photodynamic therapy [71], lymphatic

targeted chemotherapy [72], drug delivery [73] and thermal therapy [74] (Figure 2.6).

Research has found that CNTs present great potential to meet future drug delivery

Page 32: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

14

challenges [75]. For instance, when the titanium alloys were implanted, the self-organized

TiO2 nanotubes are found to be a promising drug delivery system [76]. Again,

molecularly imprinted nanotubes are found to possess the potential of controlled drug

release [77].

Figure 2. 6 Multifunctional carbon nanotubes in cancer therapy [51]

Novel SWNT-based tumor-targeted drug delivery systems which are mainly made up of

functionalized SWNTs, tumor-targeting ligands, and anticancer drugs have gained

considerable interests [51]. By linking precise peptides or ligands on their surface,

functionalized CNTs can recognize cancer-specific receptors on the cell surface and cross

the mammalian cell membrane by endocytosis or other mechanisms to recognize cancer-

specific receptors on the cell surfaces. Finally, therapeutic drugs are released more safely

and effectively than untargeted ones for the purpose of targeted killing of cancer cells

[78]. A pH-triggered targeted drug delivery system based on modified single wall carbon

nanotubes is prepared by Zhang, Meng et al. (Figure 2.7). The loading efficiency and

release rate of the anticancer drug doxorubicin (DOX) can be controlled [79].

Page 33: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

15

Figure 2. 7 Preparation of modified SWCNTs [79]

(SWCNTs: single wall carbon nanotubes; ALG: sodium alginate; CHI: chitosan; DOX: doxorubicin)

Inspired by these studies, it seems that CNTs could serve as a nanoplatform with specific

colon targeting function. However, the application of CNTs in biomedicine has been

concerned over their side effects and toxicity. Studies performed has found that the CNTs

toxicity not rely on the presence of metal impurities and their length [80] but the CNTs

themselves like physical form, concentration and functionalization degrees. In addition,

multi-walled carbon nanotubes could result in oxidative stress and cell death as they may

cause incomplete phagocytosis [81]. What is more, in the research performed by Ji, Liu

et al., correlation between toxicity of CNTs and the route of administration is suggested

[51]. Therefore, CNTs is not the most suitable nanoplatform due to its high toxicity.

2.2.4.3.Mesoporous silica nanoparticles

The synthesis and application of mesoporous silica nanoparticles (MSN) has attracted

great attention in the last two decades since the first ordered mesoporous silica molecular

sieves (the type of MCM-41) was synthesized by Kresge, Leonowicz et al. in 1992 [82].

The tunability of the structure, morphology, modification, functionalization, surface

properties, particle size and pore diameter are drawing increasing attention for the purpose

of their applications in the sectors of pharmaceutical [83], optics [84], environmental

protection [85] and catalysis [86]. Particularly, owning to their unique properties, MSNs

possess great potential to be applied in controlled release (Figure 2.8). In addition, in

order to develop the application of MSNs in biomedicine, research on the

Page 34: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

16

biocompatibility [87], cellular uptake [88] and cytotoxicity [89] of MSNs has been

performed [90].

Figure 2. 8 Mesoporous silica nanoparticles as a platform for drug delivery

A=gatekeepers B=drugs C=stimuli-responsive linkers K=functional groups G=complexation with plasmid DNA [88]

Figure 2. 9 Transmission electron microscopy of mesoporous silica nanoparticles

The honeycomb-like structure of the nanoparticles is visualized with the parallel stripes (black arrow) and

the hexagonally packed light dots (red arrow) [88]

Page 35: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

17

MSNs are synthesized inorganic nanoparticles with a honeycomb-like porous structure

(Figure 2.9), which is originated from the idea of extending the application of zeolites

[89]. When compared to most general organic materials like polysaccharides, liposomes

and micelles, MSNs possess many unique properties such as high surface area and pore

volume, ordered pores and sizes, tunable morphology [90]. Firstly, MSNs are widely used

in molecular sieving, ion exchange, adsorption [91] and separation. Soon, the well-

ordered sizes and pores extend their application to new fields. Specifically, MSNs were

used in sewage circulation purification by functionalization with mercaptopropyl groups.

Moreover MSNs are also used as templates for conductive polymers, or for controlled

polymerization processes [92]. There is also potential for MSNs to be applied for

macromolecule and micromolecule immobilization.

In particular, as there are mesopores (channels) on MSNs which are able to encapsulate

pharmaceutical cargoes, they are exploited to be used as carriers in order to attain the

controlled release in recent years. More importantly, more and more research proves that

MSNs with the diameter range of 100nm to 200nm can be taken by cells through the

endocytosis [93]. In addition, MSNs have significant advantages to be used in targeting

drug delivery systems [90]. For example, MSNs possess the capacity to level off the drug

concentration at the targeted area as the drug molecules are released from the ordered

pores or channels of MSNs by a tuneable diffusion. This gives rise to a reduction in

dosage and the preventions of side effects. Zero premature release before reaching the

targeted sites can be attained when using functioned MSNs as carriers in targeting drug

delivery systems. Furthermore, the large pore volume and large surface areas lead to a

higher drug loading [4]. Choi, Jaworski et al. functionalized MSNs for the purpose of

controlling drug release, and developed a drug delivery system using MSNs as carriers

based on the host-guest concept (Figure 2.10) [83] . Mamaeva and co-workers [94]

proposed MSNs as a platform for cancer treatments. In their study, MSNs based drug

delivery system can block the Notch signalling in cancer by controlling the delivery of γ-

secretase inhibitors to cancer cells.

Page 36: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

18

Figure 2. 10 Schematic representation of the drug delivery system by using mesoporous

silica nanoparticles: (a) preparation of mesoporous silica nanoparticle, (b) immobilization

of 1 as gate molecule, (c) encapsulation of curcumin (2), (d) complex formation of 18-

crown-6 moiety with Cs+ ion and (e) release of curcumin (2) upon addition of K+ ion.

[83]

Most importantly, recent studies pointed out that MSNs are biocompatible, and particles

are degraded and excreted mainly through urine and faeces without being retained within

the organs [94]. Lu and co-workers [87] find that 95% degraded products of MSNs could

be safely eliminated out of body by renal excretion.

To summarize, great advancements are made in the synthesis of versatile drug

nanocarriers to date, including polymers, carbon nanotubes and MSNs [66]. However,

due to high toxicity of the nanotubes and premature leakages of the polymeric carriers

[61], the development of colon targeted drug delivery systems that can reduce the toxic

effect while augment therapeutic efficacy of 5-FU remains a significant challenge. From

the literature review, it can be found that MSN have been highlighted as compelling

carriers for anticancer drugs owning to their unique properties, such as high surface area,

ordered porous structure, and easily functionalized surface. More importantly, its

biocompatibility and physical and chemical stability make MSNs the potential drug

delivery vehicles that target colon cancer cells.

Page 37: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

19

2.3. Preparation and modification of HMSNs

2.3.1. Architectural features of MSNs

During the MSN synthesis process, the silicate polymerization directed by templates can

be controlled in order to achieve different designed architectural features [95]. The

morphology of MSNs can be controlled in different ways, including varying templates,

controlling the synthesis pH and choosing additives to control the magnitude of the

interactions between the growing silica polymer and the assembled templates. Moreover,

the control of the rates of the silica source hydrolysis and condensation is important to

attain certain architectural features of MSNs. Sphere, rod, shuttle and lamel are the main

structures of MSNs (Figure 2.11).

Figure 2. 11 Different particle morphologies of MSN: (a) rod, (b) sphere, (c) lamel, (d)

shuttles, (e) core-shell and (d) hollow particles [4, 96].

Among the different MSNs, it is important to choose the right one that can achieve the

expected functions. For drug delivery, it was reported that the nanospheres exhibit a

higher efficiency of cell internalization of when compared with nanorods. This is based

on the positive relationship between the particle surface areas and the speed of cell

membrane enclosing a particle [88, 97]. Moreover, with the development of

nanomedicine, demands for nanoplatforms with specific performance such as large drug

Page 38: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

20

loading gradually increase. To meet such new requirements, more and more new methods

are under investigation for developing new drug carriers. Recently, hollow MSNs

(HMSNs) have been studied as a new-generation drug platform for delivery systems

owning to their extraordinarily high loading capacity. Li and co-workers [98] exploited

HMSNs as drug carriers to increase the capacity of storing aspirin.

Based on these studies, it is proposed that hollow mesoporous silica nanospheres is an

ideal nanocarrier for colon targeting 5-FU delivery system as it exhibits spherical

morphology to enhance cell internalization and hollow cave in the middle to increase drug

loading. However, up to date, fabricating hollow MSNs with good stability and

dispersibility remains a main challenge owning to the complex effects of physicochemical

parameters on synthesis reactions. Thereby, developing new methods to fabricate hollow

MSNs is important for the further development of efficient drug delivery systems.

2.3.2. Fabrication of HMSNs

The formation of meso-structure materials is usually attained by the synthesis of liquid-

crystal-like arrays at low-temperature. The liquid-crystal-like arrays consist of precursors

and templating agents, typically an amphiphilic organic surfactant [99]. For hollow MSN,

the synthesis is usually achieved by using a dual template method. More specifically, a

hard or soft template (cave template) which is used for the generation of the interior cave

is dispersed in a neutral or charged surfactant (pore template). Then pore template

micelles are formed around the surface of cave templating agents. After that, silica matrix

coating on the core template is generated by adding silica source agents. Finally, both

core and cave templates are removed by thermal calcination or solvent extraction to obtain

hollow MSNs with hollow interiors and mesoporous shells [100] (Figure 2.12). Now,

soft-templating, hard-templating and select-etching approaches are investigated for

successfully fabricating hollow MSNs.

Page 39: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

21

Figure 2. 12 The synthesis process of HMSNs

2.3.2.1.Soft-templating method

Soft-templating method refers to using surfactants as cave templates to build a hollow

interior cave after the removal of surfactants. For example, Feng and co-workers [101]

synthesized hollow MSNs through the aggregation of micelles. In their work, (-)-N-

dodecyl-N-methylephedrinium bromide (DMEB) was employed as a cave and pore

template and carboxyethylsilanetriol sodium salt (CSS) was used to stabilize the DMEB

micelles. Importantly, the particles size and mesoporous shell thickness could be changed

by altering the pH values during the synthesis possess (Figure 2.13). It has been suggested

that DMEB, CSS and pH values determine the morphology of hollow MSNs. Other

surfactants using for hollow MSNs fabrication through soft-templating route include

tetrapropylammonium hydroxide [98], hexadecane [102], Pluronic F127 [103] and

polystyrene latex [104].

Page 40: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

22

Figure 2. 13 Schematic illustration of the formation process of HMSs [101]

Soft-templating method has been proved to be a simple and effective synthesis approach

of hollow MSNs. However, hollow MSNs obtain through this method tend to collapse in

the surfactant removing process. Moreover, it is hard to completely remove the cave

template through solvent extraction and this in turn will lead to severe side effects when

hollow MSNs are used to deliver drugs to human [100].

2.3.2.2.Hard-templating approach

Another strategy for the synthesising of HMSNs relies on employing some solid particles

like silica and hematite as hard templates to build a hollow interior cave after template

removal process (Figure 2.14). This fabrication route is made up of four steps: (a) the

formation of solid cave templates; (b) core template dispersion, sometimes

functionalization is needed to ensure core surfactants coat the cave templates favourably;

(c) silica mesoporous shell generation; and (d) the removal of solid cave. The diameter of

the spheres and the caves can be easily tuned by changing the solid core size. Zhang and

co-workers [105] prepared hollow MSNs with tunable shell porosity by the

transformation of solid silica particles to hollow structure in NaBH4 solution. Other hard

templates being used to create hollow structures are nanospheres of carbon, metals and

metal oxides [106]. In general, the challenge of this approach is to fabricate hollow MSNs

with small diameters (under 100nm).

Page 41: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

23

Figure 2. 14 Scheme of the synthetic procedure of HMSs [107]

2.3.2.3.Select-etching technique

Figure 2. 15 Formation (left) and microscopic structure (right) of hollow/rattle-type

MSNs [108]

Recently, the fabrication of hollow MSNs through select-etching method was developed.

Hollow interiors are generated by making use of etching agents to etch the inner cores.

This strategy is based on the structural differences or compositional differences between

the core and shell structures [108]. Chen and co-workers [96, 108] prepared hollow/rattle-

type MSNs by using sodium carbonate solution to etch solid silica nanoparticles, resulting

in the building of a hollow cave in the middle (Figure 2.15). Gao and co-workers [109]

synthesized hollow MSNs with three different pore sizes by a structural difference based

selective etching methods. In their work, hollow MSNs exhibited the function of

controlling drug release rate which is achieved by changing the pore sizes. Similarly,

Page 42: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

24

Zhang and co-workers [110] described a surface protected approach to prepare hollow

MSNs. The interior of solid silica spheres is selectively etched by NaOH when their near

surface layer is protected by poly (vinyl pyrrolidone).

2.3.3. Modification of MSNs

Figure 2. 16 Introduction of functional groups in different regions of MSN (a) at the

external surface, (b) at the pore entrances, or (c) within the walls [4].

After the preparation of MSNs, it is essential that these as synthesized nanocarriers

possess designed properties for specific drug delivery systems, such as high cellular

uptake, effective cell internalization, higher binding affinity with targeting molecules. In

this regard, modification of MSNs is always needed to make hollow MSNs smart carriers

for targeting.

As there are many hydroxyl groups existing on the surfaces, MSNs can be modified very

easily in different regions including pore walls, the interior/exterior surfaces and the pore

entrance (Figure 2.16). The introduction of functional groups results in more versatile

MSNs. For example, creating hydrophobic and hydrophilic surfaces, positively and

negatively charged surfaces, improving biocompatibility and targeting functions,

regulating drug release rates can be attained by the functionalization of MSNs. He and

co-worker [111] changed the surface charge of SBA-15-structured mesoporous silica

nanoparticles from -31.1 mV to 29.6 mV by grafting rhodamine B onto the surfaces. The

resulting MSNs showed an outstanding sustained-release property when used for drug

delivery. Again, Lu and co-workers [112] improved the dispersion of MSNs by modifying

Page 43: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

25

the surfaces with trihydroxysilylpropyl methylphosphonate.

Methods used for the surface modification of MSNs include post synthesis grafting,

molecular imprinting, and organosiloxane/siloxane co-condensation. Some functional

groups have been widely introduced onto the surface of MSNs through one or more of

these methods (Table 2.1).

Table 2. 1 Methods employed to introduce functional groups onto

the MSNs surface [89]

Functional group Method

Ureidoalkyl Co-condensation

Mercaptoalkyl Co-condensation, grafting

Cyanoalkyl Co-condensation, grafting

Aminoalkyl Co-condensation, grafting

Allyl Co-condensation, grafting

Isicyanatoalkyl Grafting

Epoxyalkyl Grafting

Phosphonatoalkyl Grafting

Metal-aminoalkyl Molecular imprinting

2.3.3.1.Co-condensation

Inorganic–organic hybrid networks of MSNs are usually attained using co-condensation

of a tetraalkoxysilane and one or more organo-substituted trialkoxysilanes. This method

has the advantages of the homogeneous coverage of functional groups. However, the

selective functionalization of the external or internal surfaces of MSN is hard to achieve

through this method. The degree of organic functionalization can be controlled through

studying the electrostatic matching effects between substituent groups and the surfactant

[113, 114]. It is found that the anions which are less hydrated than Br- can displace it

from the cetyltrimethylammonium bromide (CTAB) micelle and bind tightly to the

Page 44: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

26

cationic CTAB. The anionic lyotropic series reported by Larsen for interactions with

CTAB is: citrate < CO32- < SO4

2- < CH3CO2- < F- < OH- < HCO2

- < Cl- < NO3- < Br- <

CH3C6H4SO3- [113] .

2.3.3.2.Grafting

Grafting is a method commonly used for the post-synthesis modification of MSNs.

Introduction of some functional groups on the external or the internal surfaces of MSNs

is usually achieved by silylation on free silanol groups ( Si–OH) and germinal silanol

groups (-Si(OH)2). When compared to the co-condensed method, grafted materials

exhibit better hydrothermal stability. However, the problem that this method encountered

is the inhomogeneous surface coverage of functional groups. More specifically,

functional groups grafted onto the surfaces of MSN are mainly located on the external

surface and in close proximity to the pore entrances [84].

2.3.3.3.HMSNs modified with chemical groups for certain applications

Figure 2. 17 Functionalized hollow mesoporous silica nanoparticles [115]

Page 45: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

27

HMSNs with a large cavity inside each original MSN, have recently been highlighted to

have great potential for biomedical applications due to their greatly enhanced drug

loading ability [115]. Additionally, with the availability of well-established techniques

for surface functionalization as illustrated above, HMSNs possessing biocompatibility

can be selectively modified for various applications, such as specific targeting and

imaging (Figure 2.17).

Recently, HMSNs have been modified and successively employed as highly attractive

multifunctional nanoplatforms for increasing surface hydrophobicity [116], enhancing

cellular uptake [117], cell targeting [106], bioimaging [115], and controlled drug release

[109]. For example, by integrating different types of chemical groups on the internal or

external surface, HMSNs can be used as a stationary phase in HPLC to separate both

biomolecules and small aromatic molecules, such as proteins [118]. By changing the type

(and length) of the employed alkyl chains, the accessible pore sizes of HMSNs can be

tailored for controlling the release kinetics of loaded cargos [119, 120]. Gao [109] and Jia

et al. [121] progressively increased the pore dimensions of HMSNs by extending the

etching time of HMSNs with the Na2CO3 and NaBH4 solutions, respectively and showed

sustained and controlled release of these modified HMSNs.

Nevertheless, although modifications have made HMSNs more versatile for applications,

they also have disadvantages that negatively affect HMSNs when they are exploited as

drug carriers. Firstly, if not well controlled, modifications would cause structure collapse

[102] of HMSNs. Further, the introduction of chemical groups on the surface of HMSNs

may bring additional toxicity which makes HMSNs unsuitable as carriers for drug

delivery systems. Moreover, in many cases, it is difficult to couple several functional

groups in sufficient concentration, since the number of attachment sites on the particle

surface is limited. In addition, each functionalization step might negatively affect the

suspension stability of the particulate system, depending on the physicochemical

properties of the added function. Therefore, all these issues should be considered when a

drug delivery system is engineered based on modified HMSNs and the development of

effective methods to deal with these problems are of great importance.

Page 46: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

28

2.3.4. Qualitative characterization of HMSNs

2.3.4.1. Morphology

HMSNs as a kind of mesoporuous materials possess a porous shell with well-defined

cylindrical/hexagonal mesopores and a spherical hollow cavity. The characterization of

HMSNs is similar to other mesoporous materials. In general, a variety of techniques

including transmission electron microscopy (TEM), scanning electron microscopy (SEM),

electron diffraction (ED), X-ray diffraction (XRD), small-angle X-ray/neutron scattering

(SAXS/SANS), gas adsorption measurements and nuclear magnetic resonance (NMR)

were applied to study the morphology and microstructure of HMSNs [122]. Particularly,

the hollow structure can be clearly seen under TEM, where a contrast in colour between

the cavity and the shell was observed due the reduced density in the cavity (Figure 2.18).

Moreover, the existence of mesoporous shell of HMSNs is often confirmed by the

presence of low angel peaks (2θ < 5˚) in HMSNs’ XRD or SAXS patterns (Figure 2.19)

[123-125].

Figure 2. 18 TEM morphology of HMSNs [104, 126]

Conventionally, sorption isotherms represent the most widely used method to provide

detailed information for determining porous structure, pore geometry and size as well as

Page 47: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

29

pore distribution of HMSNs [127, 128]. According to the IUPAC classification, there are

six categories of gas adsorption isotherms, Types I-V and IVc (Figure 2.20) [129], which

is a source to study structural information of porous materials due to their different

characteristic shapes.

Figure 2. 19 XRD patterns of mesoporous silica nanoparticles synthesized at different

pH [125]

On the basis of the diameter, pores are classified as micropores (< 2 nm), mesopores (2 -

50 nm), and macropores (> 50 nm). In detail, Type I isotherm exhibiting prominent

adsorption at low relative pressures and then levelling off, is usually related to adsorption

in micropores. However, it may also be considered to be indicative of mesoporous

materials with pore sizes close to the micropore range. Type II and III isotherms are often

observed for macroporous materials while Type IV in general, and IVc in particular, is

typical for many materials with accessible mesopores.

Adsorption on mesoporous solids is a multilayer adsorption process followed by capillary

condensation in mesopores, which give rise to Type IV isotherms [129]. It should be

noted that some porous materials may exhibit a combination of the six types of isotherms

as a result of the presence of several different types of pores in the structure.

Page 48: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

30

Figure 2. 20 Six categories of gas adsorption isotherms for porous materials with different

pore size [129].

Page 49: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

31

Based on the above illustration, nitrogen adsorption-desorption isotherm of HMSNs

should follow the type IV as illustrated in Figure 2.21, with a high pore volume and a

large surface area.

Figure 2. 21 Adsorption isotherms of nitrogen in mesoporous silica nanoparticles of

different pore sizes. Closed symbols, adsorption; open symbols, desorption [122].

2.3.4.1. Pore geometry and pore size

Pore geometry is an essential part of investigating HMSNs as pore types and shapes play

important roles in the applications of HMSNs. According to the types of pores that define

the networks (Figure 2.22), MSN are divided into several types, such as MCM-41, SBA-

15 and FDU-1 [130].

Page 50: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

32

Figure 2. 22 Pore geometry [129]

Nitrogen Absorption and Desorption (BET/BJH) is commonly used to characterize the

pore geometry and pore size of MSNs. Isotherm linear plots (Figure 2.23, type H1-H4)

obtained from BET/BJH adsorption and desorption equipment provide information of

pore geometry [128, 129, 131]. Owning to different pore structure, the way of gas

evaporation differs from one another, which results in the change in shape of the

hysteresis loops on nitrogen adsorption-desorption isotherms (Figure 2.23). Therefore the

shape of a gas adsorption and desorption isotherm is one of the sources that provide

important qualitative structural information. By studying shapes of the corresponding

isotherms, the information about pore geometry can be obtained.

Figure 2. 23 Effect of pore geometry on shapes of hysteresis loop

Page 51: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

33

For an ideal 5-FU delivery system, it is essential to make sure that the drug is easily loaded

and released in a designed profile. The pore connection would therefore be a vital element

that should be taken into consideration, as it plays a key role in assessing the feasibility

of hollow MSNs nanocarriers for 5-FU delivery. The accessibility of mesopores as well

as the pore size can be acquired through this method. However, this characterization

method is limited to pores with a size of 2-50 nm.

Positron annihilation lifetime spectroscopy (PALS) is another technique used to identify

the porosity of mesoporous materials. PALS is a powerful characterization giving useful

information on hierarchical porosity, especially on micropores and molecular scale free

volume elements that is information hardly available from other characterisation methods,

such as BET and XRD.

2.3.5. MSNs’ pharmacokinetics

The pharmacokinetics of MSNs is fundamentally significant to understand, interpret and

assess the effectiveness of particles as it is closely connected to the toxicity and

biocompatibility of MSNs as drug carriers. Thus, the study of MSNs’ pharmacokinetics

is vitally essential before the utilization of MSNs for delivery systems. The current

investigations on MSNs’ pharmacokinetics include blood circulation [132], retention,

biodegradation, excretion and bio-distribution.

MSNs have found to be biocompatible and well tolerated when injected in mice with a

dosage of 50 mg kg-1. This dosage is proved to be an adequate concentration to provide

effective cancer therapy. On the other hand, higher dosage of MSNs (1.5 g kg-1) would

cause death to mice. The biocompatibility of MSNs can be improved by coating

biocompatible polymers.

Page 52: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

34

Figure 2. 24 TEM images of MSNs (black dots) endocytosed by human cervical cancer

cells [88]

The pathways of HMSNs for entry into cells are endocytosis but the surface

functionalization and change in structural properties of HMSNs can lead to different

pathways (Figure 2.24) [88]. Lu [87] found that MSNs were rapidly and completely

cleared from animal bodies through urine and feces. The biodegradation of HMSNs

follows a three-stage degradation, involving a fast bulk degradation stage in the first hour,

a decelerated degradation stage hindered by calcium/magnesium silicate layer and finally,

a slow diffusion stage on day-scale [133, 134]. The bio-distribution of HMSNs is mainly

in the liver and spleen [87, 135]. Moreover, the influence of surface potential on the

pharmacokinetics of HMSNs is significant. Studies conducted by Mou and co-workers

[136] reveal that HMSNs with positive charge on the surface undergo faster transportation

and excretion.

2.4. Drug loading and release

As for drug delivery system, high drug loading and controlled or sustained drug release

are highly desired. In this case, hollow MSNs possessing both nano-scaled particle size

and large pore and cave diameter are one of the promising candidates being pursued

(Figure 2.25). Size, surface potential and hydrophilicity are factors to be controlled to

make MSNs as ideal drug carriers with extraordinarily high load capacity as well as

controlled release profile [137].

Page 53: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

35

Figure 2. 25 HMSNs based drug delivery system for simultaneous joint lubrication and treatment [138]

Hollow MSNs have open pores for drug molecules to be encapsulated and have hollow

interiors for a high drug storage (Figure 2.26) which makes drug loading and release quite

different from polymeric carriers [139, 140]. The hollow cave and pore size would have

significant effect on drug loading and release capacity. It is reported that selective

adsorption of drug greatly depends on the pore size diameter [100]. Moreover, surface

potential and hydrophilicity also play important roles on drug loading capacity. In general,

hollow MSNs with both the opposite potential and similar hydrophilicity of drug

molecules are more likely to achieve higher drug adsorption. Therefore, such parameters

should be taken into account while designing nanocarriers for a given drug molecule. He,

Zhang et al. [111, 141] developed a drug delivery system by fabricating MSNs with

positive charge on their surface to load a negative charged drug for anti- hepatic fibrosis.

It is found that the positively charged MSNs present better drug loading capacity and

controlled release profiles than negatively charged ones. Gao and co-workers [109]

regulated the release rate of doxorubicin by tuning the pore diameter of hollow MSNs.

Page 54: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

36

Figure 2. 26 HMSNs based drug delivery system for high-intensity focused ultrasound

(HIFU)-mediated intravenous drug delivery [142].(a) The schematic illustration of the

synthesis route of HMSN-LM and its applications in multi-drug co-delivery, temperature-

responsive release and HIFU treatment enhancement; TEM images of HMSNs (b) and

drug loaded HMSNs (c), and the insets are their corresponding schematic illustration.

Transmembrane transport of hollow MSNs is another element that affects drug release

kinetics in vivo. It is well established that the internalization of nanocarriers via

endocytosis are size and cell dependent. MSNs can be prepared with a wide range of

particle size (30-300nm) and tailored shapes (spheres, plates and rods). However, only

particular size and shape can exhibit good transmerbrane transport. Lu and co-workers

[143] found out MSNs with 50 nm in diameter exhibit the best cellular uptake efficiency.

This is also proved by Tang et al [100] who pointed out that the order of cellular uptake

efficacy of MSNs is 50 nm > 30 nm > 110 nm > 280 nm > 170 nm. Importantly, it is

suggested that the most effective size range of MSN in drug delivery is 50-150 nm from

this standing point of cellular uptake efficacy. In addition, for transmembrane transport,

the feature of the net negative surface charge of cell membrane attracts attention from

researchers [88, 144]. Nanocarriers are modified to possess positive charge on their

surfaces leading to high efficiency in the endocytic trafficking in mammalian cells [136,

Page 55: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

37

145]. He et al [146] pointed out that the higher surface hydrophobicity and larger particle

size will prohibit the transmembrane delivery.

Inspired by these studies, as for 5-FU, in order to attain high adsorption, it is essential to

fabricate hollow MSNs with a particle size within the range of 50-150 nm. Also, hollow

MSN carriers possessing hydrophilic and positively charged groups on its mesopore walls

and surfaces would enhance their 5-FU loading capacity.

2.5. Drug resistant phenomenon

Systemic toxicity of anti-cancer drugs as well as the emergence of multidrug resistance

(MDR) by cancer cells have greatly limited the success of chemotherapy, making cancer

remains one of the major causes of mortality and morbidity in the world [147-150].

The systemic toxicity of 5-FU mainly resulted from its nonspecific distribution in a

biological system, leading to the death of normal cells [35, 151]. Nanocarriers such as

HMSNs have been developed that can offer targeting delivery of chemotherapy drugs to

cancer cells by taking the advantage of enhanced permeation and retention (EPR) e ect

[152, 153] in the tumour site. Further functionalization of the surface of HMSNs allows

for enhanced recognition and internalization of drug loaded HMSNs by target cell

population, thereby protecting normal tissues from exposing to the toxic effects of anti-

cancer drugs [154-157].

However, even though the toxicity of 5-FU would be overcome by a well-designed

HMSNs based targeted drug delivery system, the effective cancer treatment of 5-FU

cannot be maintained for a relative long period for a complete killing of the cancel cells.

Instead, the therapeutic effect of 5-FU decreased along the time, leading to treatment

failure and recurrence or relapse of tumours [158, 159]. This is mainly due to the cancer

cells developing resistance to the drug 5-FU, which is another obstacle to the successful

cancer treatment [160, 161].

Page 56: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

38

Figure 2. 27 Rational design of a multifunctional micellar nanomedicine for targeted co-delivery of siRNA and DOX to overcome multidrug resistance [162]

(a) Chemical structure of functionalized copolymers with ligands at the end of the PEO block and

conjugated moieties on the PCL block. (b) Assembly of multifunctional micelles with DOX and siRNA in

the micellar core and RGD and/or TAT on the micellar shell. (c) Schematic diagram showing the proposed

model for the intracellular processing of targeted micelles in MDR cancer cells after receptor-mediated

endocytosis, leading to cytoplasmic siRNA delivery followed by P-gp down-regulation on the cellular and

nuclear membrane and endosomal DOX release, followed by DOX nuclear accumulation

Multidrug resistance (MDR) involves cancer cells overexpressing drug transporter

proteins, like P-glycoprotein (P-gp) to pump the drugs out of the cells [163, 164]. To

overcome this problem, combination delivery of chemotherapy drugs [165, 166] is

emerged as a strategy and is proved to be effective for some short-term treatments.

However, in many case, long-terms therapy is required for a full cure of cancers, where

co-delivery of chemotherapy drugs shows inability [167]. This is mainly due to the fact

that after a particular type of cancer has developed resistance to a given drug, the cancer

cells develop resistance to other drugs as well. Fortunately, the mechanism and genes

behind MDR are well explored and the inhabitation of P-gp functions was achieved by

specific small biological molecules [162]. Despite this effort, because of their nonspecific

action that caused intolerable toxicity, no P-gp inhibitor has been approved for clinic use.

Therefore, new strategies to inhibit the expression (rather than the function) of P-gp is of

great importance.

Page 57: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

39

In this regard, the use of small interfering RNA (siRNA) for a potent post-transcriptional

gene silencing was developed [163]. However, the delivery of siRNA through systemic

administration is often obsessed with poor stability, low cellular uptake, and rapid

clearance of siRNA from the circulation [162]. Similarly, nanoplatforms can provide

efficient delivery of siRNA with high specificity and low off-target effect. A better

therapeutic effect in MDR tumours was observed by co-delivery of drugs and siRNA to

the same cancer cells [162, 167]. For example, as shown in Figure 2.27, siRNA was

successfully delivered by a multifunctional micellar system to overcome multidrug

resistance. Mao et al. [168] used functionalized quantum dots for simultaneous delivery

of siRNA and anticancer drug DOX, which showed effectively induced apoptosis and

suppressed tumour growth (Figure 2.28).

Inspired by the research illustrated above, we hypothesised that co-delivery of 5-FU and

siRNA using HMSNs as nanocarriers would simultaneously deal with the problems

regarding systemic toxicity and drug resistance. Nevertheless, co-delivery of drug and

siRNA is a challenging task due to their huge different properties [169]. Thus, when

engineering a co-delivery system, multiple factors need to be considered beforehand.

Firstly, mutual chemical compatibility should be ensured. The capacity to load and

release drug-siRNA combination then needs to be determined [170]. More importantly,

the differences in hydrophobicity, metabolic stability and molecular weight between

drugs and siRNA are very critical for nanocarriers [167, 171].

Based on this literature review, we propose to develop a co-delivery system through a

more feasible option. HMSNs will be loaded with 5-FU and siRNA separately. Then by

combining 5-FU loaded HMSNs and si-RNA loaded HMSNs with different ratio will be

delivered to colorectal cancer cells. This proposed strategy take the advantage of avoiding

co-encapsulation of 5-FU and si-RNA, which will provide a more-potent, independent

control and less-toxic means for cancer treatment.

Page 58: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

40

Figure 2. 28 System for delivery of doxorubicin (DOX)/siRNA combination based on quantum dots (QDs) modified with β-cyclodextrin [168]

Page 59: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

41

Chapter 3 Materials and methodology

3.1 Introduction

This chapter outlines the chemical reagents, HMSNs fabrication and modification

techniques, in vitro experimental methods as well as characterization technologies

applied in this thesis. Firstly, all the materials including chemicals and auxiliary reagents

are specified, followed by a full illustration of preparation and chemical functionalization

technologies for HMSNs. Additionally, strategies used for 5-FU loading and release study

are described. Furthermore, detailed in vitro experimental methods for assessing the

biological activity of HMSNs as drug carriers for targeted drug delivery system are

discussed. Finally, technologies utilised to characterize the morphology and properties of

samples in terms of particle size, pore size, surface area, modification efficiency, drug

loading capacity, release profile, biocompatibility and targeting ability are fully

introduced.

3.2 Materials

Tetraethyl orthosilicate (TEOS, Si(OCH2CH3)4), Triton X-100 (TX100), Poly(ethylene

glycol)20-block-poly(propylene glycol)70-block-poly(ethylene glycol)20 (P123),

Poly(ethylene glycol)106-block-poly(propylene glycol)70-block-poly(ethylene glycol)106

(F127), Tweens 20, ammonia solution (28-30%), Sulphuric acid, Sodium carbonate

(Na2CO3), Octadecyltrimethoxysilane (OTMS), N-Hydroxysuccinimide (NHS), 1-ethyl-

3-(3-dimethylaminopropyl) carbodiimide (EDC), (3-Aminopropyl)triethoxysilane

(APTES), 3-Cyanopropyltriethoxysilane (CPTES), Fluorescein isothiocyanate (FITC),

fluorinated pyrimidine 5-fluorouracil (5-FU), deuterium oxide, Epidermal growth factor

(EGF, human), 100× penicillin/streptomycin solution, Trypsin−EDTA Solution (0.25%

trypsin, 0.1% EDTA), paraformaldehyde (PFA), 4',6-Diamidino-2-Phenylindole (DAPI)

were all purchased from Sigma Aldrich (Sydney, Australia). Dulbecco's Modified Eagle

Medium (DMEM), Fetal Bovine Serum (FBS), Phosphate-Buffered Saline (PBS) and

Page 60: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

42

GlutaMAX™ were from life technologies. Eudragit S100 which is made in Germany was

purchased from Shenzhen Youpuhui Pharmaceutical Co. Ltd (Shenzhen, China).

3.3 Preparation of versatile HMSNs

3.3.1 Formation of Eudragit S-100 nanoparticles

The synthesis of the Eudragit core particles were achieved by nanoprecipitation [172].

Specifically, 20 mL acetone solution of Eudragit S-100 (4 mg/mL) was added dropwise

to 100 mL deionised water under stirring. The suspension was left to proceed for 5 h

under agitation at room temperature before being used for HMSNs fabrication.

3.3.2 Fabrication of HMSNs

For the formation of the mesoporous silica shell, 120 mL Eudragit S-100 nanoparticle

solution (0.75 mg/mL) was ultrasonically dispersed for 3 min (100W), followed by

addition of 1.0 g TX100 under stirring. After TX100 was completely dissolved, 2.3 mL

TEOS was added and this solution was left to proceed for 24 h before being transferred

into a sealed container and hydrothermally treated overnight under 100 C. The resulting

nanoparticles (as-synthesized HMSNs) were collected by centrifugation (10000 rpm for

20 mins) and redispersed in Acetone for the removal of Eudragit core and TX100

templates. The particles were then washed 3 times with ethanol and deionized water

respectively before finally freeze-dried (deionized water was used as the solvent for all

freeze-drying procedures).

For stability study, paralleled samples were prepared using the same method illustrated

above to obtain as-synthesized HMSNs. Then these as-synthesized HMSNs were freeze-

dried before undergoing a calcination process, where 550 ˚C was applied for 6 h to

remove the Eudragit core and TX100 templates.

Page 61: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

43

3.3.3 Control of particle size

The particle size of HMSNs was controlled by synchronously changing the amount of

TEOS and the concentration of Eudragit S-100 solution applied in HMSNs fabrication.

In detail, 20 mL acetone solution of Eudragit S-100 (4 mg/ml, 8 mg/ml and 16 mg/ml)

was added dropwise to 100 ml deionised water under stirring. The suspension was left to

proceed for 5 h under agitation at room temperature. Then the resulting Eudragit S-100

nanoparticles suspensions were ultrasonically dispersed for 3 min, followed by addition

of 1.0 g TX100 under stirring. After completely dissolving TX100, 2.3 mL, 3.5 mL and

8.0 mL TEOS were added, respectively and these solutions were left to proceed for 24 h

before transferred into sealed containers and hydrothermally treated overnight under 100

˚C. The resulting HMSNs with different size were collected by centrifugation and

redispersed in Acetone to remove Eudragit core and TX100 templates. Finally all the

particles were collected and freeze-dried.

3.3.4 Control of shell thickness

HMSNs with varying shell thickness (10 nm, 15 nm and 30 nm) were fabricated by using

different types of nanoparticles as cores. Typically, HMSNs with 10 nm shell thickness

was fabricated according to the method illustrated in 3.3.2 while those with 15 nm and 30

nm shell thickness were synthesised with the following procedures.

Six milliliter TEOS was added dropwise to a mixture containing 100 mL ethanol and 6

mL ammonia solution, which was hydrolysed and condensed for 0.5 and 1 h at 30 ˚C to

obtain silica nanoparticles with different diameters as cores for further HMSNs

fabrications. After that, 2 mL OTMS and 5 mL TEOS were firstly mixed and then added

to the above solutions, which was left to react for another 1 and 4 h at 30 ˚C under

magnetic stirring. Finally, the yield particles were collected by centrifugations and

redispersed in Na2CO3 solutions (0.6 M) for 1 h at 80 ˚C before calcinations and freeze-

drying.

Page 62: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

44

3.3.5 Pore expansion of HMSNs

The control of the pore size of then HMSNs were achieved by varying the molecular

weight of non-ionic surfactants used for HMSNs fabrications. Larger pores were

developed by surfactants with longer molecular chain which forms bigger micelles that

finally evolved to pores of HMSNs. F127, P123 and Tweens 20 were used instead of

TX100 for HMSNs synthesis described in 3.3.2.

Another method for pore expansion is to apply different calcination time to HMSNs,

where small pores were collapsed and merged into larger pores. In detail, 6 mL TEOS

was added dropwise to a mixture containing 100 mL ethanol and 6 mL ammonia solution,

which was hydrolysed and condensed for 1 h at 30 ˚C. Then, 2 mL OTMS and 5 mL

TEOS were mixed first and added to the above solutions, which was left to react for

another 1 h at 30 ˚C under magnetic stirring. After that, the yield particles were collected

by centrifugations and redispersed in Na2CO3 solutions (0.6 M) for 1 h at 80 ˚C. The

resultant particles were divided into 3 groups which underwent calcinations for 3, 6 and

12 h at 550 ˚C before freeze-drying. HMSNs with pore size of 2.5, 4.3 and 20.1 nm were

acquired.

3.4 Modifications of HMSNs

3.4.1 Amine-, methyl- and cyano-functionalization

For amine-functionalization (HMSN-NH2), 50 mg HMSNs were ultrasonically dispersed

in 50 mL ethanol, 25 μL APTES were added and stirred for 24 h before centrifugations

and washed 3 times with ethanol then deionized water respectively. Finally these yielded

HMSN-NH2 were collected and freeze dried.

The experimental procedure for methyl- and cyano-functionalization (HMSN-CH3 and

HMSN-CN) is similar to amine-functionalization. The only difference is to use OTMS

Page 63: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

45

and CPTES instead of APTES for chemical modification.

3.4.2 Carboxyl-modification (HMSN-COOH)

Fifty milligrams of HMSNs were ultrasonically dispersed in 50 mL ethanol, 25 μL

APTES were added and stirred for 24 h. The resulting nanoparticles were collected by

centrifugations and redispersed in sulphuric acid (50% v/v), heated at 150 ˚C for 4 h. The

final products were re-collected, washed and freeze dried.

3.4.3 FITC-modification (FITC-HMSNs)

Twenty milligrams of HMSNs were suspended in 20 mL ethanol, followed by addition

of 10 mL APTES, then this mixture was stirred for 24 h at room temperature. The

resulting amino functionalized HMSNs were collected by centrifugation and then

redispersed in 10 mL FITC ethanol solution (1 mg/mL) and stirred for another 24h. The

final FITC modified HMSNs (FITC-HMSNs) were collected for freeze-drying.

3.4.4 EGF labelling (HMSN-EGF)

Sixteen milligrams of NH2-HMSNs were ultrasonically dispersed in 1 mL deionised

water, to which 2.1 mg NHS and 17.3 mg EDC were added. After that, 0.4mL of EGF

solution (0.2 mg/mL in PBS solution) and 0.8 μL of triethyl amine were added. The

mixture was left to proceed for 1 h under stirring at room temperature. The EGF grafted

hollow mesoporous silica nanoparticles (EGF-HMSNs) were collected by centrifugation

and washed in deionised water. These EGF-HMSNs were redispersed in 8 mL FITC

ethanol solution (1 mg/mL) and stirred for 24h. The final FITC modified EGF-HMSNs

(EGF-HMSN-FITC) were collected by centrifugation and washed with ethanol then

deionized water. Finally, the product was freeze-dried.

Page 64: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

46

3.5 5-FU loading and release study

3.5.1 Measurement of 5-FU oil : water partition coefficients

Measurement of 5-FU oil : water partition coefficients was conducted via shake flash

method according to OECD guideline 107 [173]. Firstly, n-octanol and Milli-Q water

were saturated, respectively prior to the experiments and herein their saturated solutions

were referred to as octanol and water. Then, 5-FU was dissolved in water to obtain a

concentration of 266.7 μg/mL. A combination of octanol and water with an octanol:water

ratio of 1:1, 1:2 and 2:1 were added to centrifuge tubes and a total volume of 40 mL was

maintained. Each ratio was carried out in triplicate. 1mL 5-FU was added to the centrifuge

tubes and the phases were mixed by vortex for 10 min before centrifuging at 1000 rpm

for a further 10 min. Both 5-FU concentrations in water and octanol phases were

determined via UV-Vis spectroscopy (Ocean Optics system, USA).

The partition coefficient Pow is calculated from the data of each run using the following

equation (Equ. 3.1):

3.1

3.5.2 5-FU loading

The loading of 5-FU was carried out by soaking HMSNs in a concentrated drug solution

at room temperature for 24 h while stirring to ensure the diffusion of the drug molecules

through the mesopores. Firstly, 5-FU was dissolved in deionized water to a concentration

of 3 mg/ml. 150 mg of HMSNs was ultrasonically dispersed in 50 mL of the 5-FU solution.

The mixture was stirred at room temperature. At timed intervals of 0, 2, 4, 6, 8, 12 and

24 h, 0.5 mL of the solution was extracted, centrifuged to collect the supernatant for UV-

Vis analysis at a wavelength of 266 nm while the 5-FU -loaded HMSNs precipitate was

placed back into the drug solution. At intervals of 0, 2, 4, 6, 8 and 12h, the HMSNs only

loaded a certain amount of 5-FU and they still have the capacity for further encapsulation

Page 65: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

47

of the drug. Therefore, the 5-FU loaded HMSNs precipitate collected at intervals were

placed back to the mixture for further loading. After the loading, the 5-FU loaded HMSNs

were washed with deionized water before freezer-drying. The loading amount of 5-FU in

the HMSNs was calculated by subtracting the amount of 5-FU in the supernatant from

the amount in the original drug loading solution.

The drug loading capacity was calculated using the following equation:

3.5.3 5-FU release

To measure the drug release profile from 5-Fu loaded HMSNs, 30 mg of 5-FU-loaded

HMSNs were dispersed in 150 mL of phosphate buffer saline (PBS) solutions at pH 5.5

or pH 8.0 and the mixtures were gently stirred at 37 C. At timed intervals, 1 mL of the

solution was removed and centrifuged (13000 rpm for 5 mins) for UV-Vis analysis and 1

mL of fresh PBS was added to the drug release solution to keep the volume constant. For

drug release, it is important to make sure that a certain amount of the drug is released into

a constant volume. A changing volume can affect the release rate and release kinetics to

some extent. The method used in this work for measuring the drug release behavior of

HMSNs is a standard method which has been widely used [174-177].

The release percentage at each time point was analysed by UV-Vis at a wavelength of

266 nm and calculated using the following equation:

×100%

where Ct is the drug concentration at a time t, Ct-1, Ct-2 are the drug concentrations detected

at different time points previous to t (C0 =0). V1 is the total volume for drug release (150

mL), V2 is the volume that is taken for UV measurement (1 mL), W1 is the weight of drug

loaded HMSNs (0.03 g) and L is the drug loading capacity of HMSNs (194.5 mg(5-

FU)/g(HMSNs)). The 5-Fu release experiment was conducted in triplicate and the mean

Page 66: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

48

results were reported.

3.6 Cell experiments

3.6.1 Cell culture

The colorectal cancer cell line, SW 480 and SW 620, with and without EGFR

overexpression, were cultured in DMEM containing 10% foetal bovine serum and 1%

penicillin/streptomycin supplemented with 1% GlutaMAX™. The culture was

maintained at 37 C in a humidified atmosphere containing 5% CO2.

3.6.2 Cellular uptake of HMSNs and EGF-HMSNs in colorectal cancer cells

SW480 and SW 620 were seeded into 24-well plates with 1×105 cells per well in 0.5 mL

of complete growth medium. After 24 hours of cell attachment at 37 in 5% CO2, the

cells were treated with 2, 5 or 10 μg/mL of EGF-HMSN-FITC or FITC-HMSNs for 30

min, 2 hours or 4 hours. The cells were washed with PBS three times and harvested by

trypsinization. Then the cells were resuspended in PBS containing 10% of FBS for flow

cytometric analysis. The green fluorescently-labelled live cells were counted as positive

cells which are associated with nanoparticles and the mean fluorescence intensity (MFI)

was calculated. The cells incubated with DMEM only were used as a negative control.

To visualize the uptake of HMSNs via confocal microscope, SW 480 and SW 620 cells

were seeded into 8-well chamber slides at 5 ×104 per well for 24 hours before incubation

with 10 μg/mL of EGF-HMSN-FITC or FITC-HMSNs for 30 min, 2 hours or 4 hours.

Cells were fixed with 4% paraformaldehyde and stained with DAPI for confocal

microscopy imaging. Confocal imaging of the cells was carried out on a laser scanning

confocal microscope.

Page 67: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

49

3.7 Characterizations

3.7.1 Electron microscopy

The morphology of HMSNs was studied using scanning electron microscopy (SEM, Carl

Zeiss AG SEM Supra 55VP) and transmission electron microscopy (TEM, LaB6 JEOL

JEM-2100). Before imaging under SEM, all the samples were coated with carbon through

a BAL-TEC SCD 050 sputter coater (Leica Microsystems, Australia). TEM imaging were

conducted under 200 kV.

3.7.2 Nitrogen absorption and desorption experiments

Nitrogen absorption and desorption isotherms of HMSNs were conducted on a

Micromeritics Tristar 3000 (Particle & Surface Science, UK) equipment at 77K. Before

measurements, samples were degassed at 523 K for 1 h under nitrogen. The specific

surface areas of HMSNs were calculated using the Brunauer-Emmett-Teller (BET)

method. The Barrett-Joyner-Halenda (BJH) model was utilized to obtain the pore size

distributions from the desorption branch of isotherms.

3.7.3 Small angle X-Ray diffraction (SAXRD)

The mesoscopic ordering of the nanoparticles was confirmed by small angle X-ray

diffraction (SAXRD) using a Bruker D8 Advance (Bruker, Germany) X-ray

diffractometer with Cu-Kα radiation (λ=0.1542 nm). The instruments were operated at 40

kV and 40 mA current.

3.7.4 Thermo-gravimetric analysis (TGA)

TGA of HMSNs were conducted using a TG Simultaneous Thermal Analyser (STA 449C,

ZETZSCH, Germany). Samples with a weight of ~5 mg were placed into aluminium

oxide crucibles and heated under nitrogen condition at a heating rate of 10 ˚C/min from

room temperature to 600 ˚C.

Page 68: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

50

3.7.5 Fourier transform infrared spectrometer (FTIR)

Fourier transform infrared (FTIR) spectra were measured with a Vertex 70 spectrometer

(Bruker, Germany). All the samples were compressed into KBr pellets and recorded at 64

scans from 4000cm-1 to 400cm-1 with a resolution of 4 cm-1.

3.7.6 Time of flight secondary ion mass spectrometry (ToF-SIMS)

ToF-SIMS experiments were performed using a Physical Electronics Inc. PHI TRIFT V

nanoTOF instrument (Physical Electronics Inc., Chanhassen, MN, USA) equipped with

a pulsed liquid metal 79+Au primary ion gun (LMIG), operating at 30 kV energy. Dual

charge neutralisation was provided by an electron flood gun (10 eV electrons) and 10 eV

Ar+ ions. Experiments were performed under a vacuum of 5x10-6Pa or better. “Bunched”

Au1 instrumental settings were used to optimise mass resolution for the collection of

spectra. Positive SIMS spectra were collected from areas of 100×100 micron, with an

acquisition time of 2 minutes.

In order to meet the static SIMS regime, the corresponding total primary ion dose of less

than 1 × 1012 ions cm-2 was typically achieved [178]. A mass resolution at nominal m/z

= 27 amu (C2H3+) was m/ m of > 5000. Ten positive ion mass spectra were collected

from areas that did not overlap for each sample. Sample spectra were processed and

interrogated using WincadenceN software (Physical Electronics Inc., Chanhassen, MN,

USA). Integrations were performed on all recognisable and unobscured immonium ions

in the 2-160 amu mass range. The peaks were normalised to the total intensity of all

selected peaks [179].

PCA was performed using Statistica software (StatSoft Pacific Pty Ltd, North Melbourne,

VIC, Australia).

Page 69: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

51

3.7.7 Nuclear magnetic resonance (NMR)

The pulse-field gradient NMR (PFG-NMR) diffusion experiments were carried out on a

Bruker Avance III 500 MHz wide bore spectrometer (with proton Larmor frequency of

500.07 MHz) equipped with a 5 mm diff50 pulse-field gradient probe. Each sample was

packed in a 5 mm Schott E NMR tube to a height of 50 mm. The pulse-field gradient

stimulated echo (PFG-STE) pulse sequence [180] was used to obtain diffusion

coefficients and the 1H NMR signal was used for the determination of diffusion

coefficients of different molecules. The maximum gradient strength of the amplifier is

29.454 T/m.

3.7.8 Dynamic light scattering (DLS)

The particle size of Eudragit and micelle sizes of TX100 were investigated by dynamic

light scattering (Malvern ZetaSizer Nano). Each sample was tested in triplicate and the

mean result was reported.

Zeta potential of HMSNs were analysed on this instrument as well. Samples were

dispersed in ethanol with a concentration of 1 mg/ml. Measurements of each sample were

conducted in triplicate and the average value was considered.

3.7.9 Confocal microscopy

Confocal laser scanning microscope (CLSM) images of cells were took via a confocal

microscopy (Leica TCS SP5, Germany) using 40× objective, 488 nm/405 excitation.

Page 70: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

52

Chapter 4 Synthesis of tailored HMSNs for nanotheranostics

4.1 Introduction

Hollow mesoporous silica nanoparticles (HMSNs) with large hollow interiors and

penetrating pores in the shell are exceptionally promising platforms in many current and

emerging fields such as pharmaceutics [1], optics [2], environmental protection [3] and

catalysis [4]. In particular, they were exploited as nanocarriers for drug delivery [5]

because their huge cavities provide large drug loading capacities [6] and the open

mesopores allows transportation of drugs [7]. In addition, owing to their hollow structure,

HMSNs also possess other characteristics such as a decreased particle density and

increased active area for drug loading. In this regard, HMSNs are superior to non-hollow

mesoporous silica nanoparticles (MSNs) and other corresponding mesoporous silica

nanoparticle composites [8]. However, owing to the complex effects of physicochemical

parameters on synthesis reactions, fabricating HMSNs with good stability and dispersion

remains a main challenge [5] which limits their further application in drug delivery.

Thereby, developing new, reliable and facile methodologies for synthesising discrete and

dispersible HMSNs is important and could contribute to the further development of drug

delivery systems.

One of the typical methods to produce mesoporous silica materials is self-assembly which

is a well-known synthesis strategy and occurs when molecules interact with one another

through a balance of attractive and repulsive interactions [9, 10]. For mesoporous silica

materials, self-assembly is achieved via a surfactant templating route which involves the

spontaneous organization of templates into structurally ordered micelles through van der

Waals and Coulomb interactions [11] or hydrogen bonding. As shown in previous studies,

non-ionic surfactants such as tri-block copolymers with certain hydrophilic/hydrophobic

ratios and neutral surfactants like Triton X-100 (TX100) have been demonstrated to be

extremely suitable for assembling ordered mesoporous silicates [12, 13]. However, due

to the difficulty in adjusting micelle structure and the poor understanding of the

Page 71: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

53

mechanisms behind the self-assembly formation of mesoporous structure in sol–gel

chemistry, the controllable tailored design of non-ionic surfactant templated HMSNs is

still far from satisfactory. Therefore, there is immense value in the development of new

methodologies for regulating the structure of non-ionic surfactant micelles in order to

tailor HMSNs for certain applications.

Current efforts are ongoing to optimize the parameters and yield of mesoporous structures

by changing material ratios or fabrication parameters, such pH [14, 15] and synthesis

temperature [16]. From a technical point of view, an insight into the micelle organization

mechanism can contribute to the improvement of the silica structure including stability,

pore size and volume. In addition, customisation of the silica structure will allow

facilitation of the host-guest specific interaction for the intended purpose.

In this chapter Eudragit S100 was introduced as a core particle to bind and facilitate the

self-assembly of the non-ionic surfactant into stable composite micelles. The reactions

between surfactant molecules and Eudragit S100 was investigated by Fourier Transform

Infrared (FTIR), Dynamic Light Scattering (DLS) and Pulse Field Gradient NMR (PFG-

NMR) analyses. The addition of Eudragit S100 were inferred to result in the redistribution

of surfactant micelles by hydrogen-bonding. Uniform and highly ordered HMSNs were

achieved, demonstrating the efficiency of using Eudragit to assist in the morphological

formation of surfactant micelles and consequently, the silica shell structure.

The efficacy of Eudragit-assisted HMSNs as drug carriers was ascertained by examining

the loading and release behaviour of the first line anticancer drug, 5-FU. Additionally, in

vitro cytotoxicity and cellular uptake of these HMSNs by colorectal cancer cells SW480

were investigated.

Page 72: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

54

4.2 Synthesis of HMSNs with designed properties for drug delivery

systems

4.2.1 Preparation and characterization of HMSNs

For HMSNs preparation, typically a hard or soft sphere is used to create the hollow cavity

in HMSNs while surfactants are employed to develop pores in the shell and this in turn,

involves a template-eliminated process [106, 107, 116, 181] . One of the typical methods

to remove the core template and the pore template is calcination, where the as-synthesized

HMSNs (with templates) subject to a very high temperature (>500 ˚C). However, during

high-temperature, irreversible particle aggregation occurs, leading to the formation of

nondispersible and conglutinated materials. Current efforts have been made in creating

the intrinsic space in HMSNs, but little attention has been paid to solve the problem of

particle dispersibility in solvents [182].

To obtain dispersible HMSNs, a new method was developed to synthesise hollow

mesoporous silica nanoparticles (HMSNs) by employing Eudragit S100 nanoparticles as

the dissolvable core templates. Both the Eudragit S100 core and the structure-directing

surfactant can be easily and synchronously removed through solvent extraction in acetone

rather than calcination.

Figure 4. 1 Fabrication possess of HMSNs

Page 73: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

55

As a working model, the designed protocol as shown in Figure 4.1 depicts the fabrication

possess where Eudragit S 100 nanoparticles were employed as a core template and the

non-ionic surfactant Triton X-100 (TX100) was used as the pore template. Firstly, TX100

micelles were assembled on the surface of Eudragit S 100 nanoparticles. Subsequently,

TEOS was added as a silica source to form the mesoporous silica shell with embedded

TX100 micelles. TX100 micelles and Eudragit S 100 nanoparticles can be easily removed

by solvent extraction at the same time resulting in HMSNs with open mesopores and a

hollow cavity.

The microstructure of synthesized nanoparticles which are made up of a huge hollow

cavity and a mesopourous shell are proved by SEM and TEM images (Figure 4.2). The

spherical hollow interior can clearly be seen from the TEM images in Figure 4.2c-f, with

a shell thickness of 10 nm (Figure 4.2e). The synthesized particles were spherical in shape

with a mean diameter of 100nm. The rough surface of these hollow mesoporous silica

nanoparticles shown in Figure 4.2b further suggested their porous structures [183, 184].

Page 74: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

56

Figure 4. 2 a, b & c) SEM images of HMSNs and d, e & f) TEM image of HMSNs displaying hollow interior

Page 75: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

57

1 10 100 10000

5

10

15

20

25

Inte

nsity

(%)

Size (nm)

PDI: 0.048

Figure 4. 3 DLS curve of HMSNs

The particle distribution collected by dynamic light scattering (DLS) in Figure 4.3

definitely displays a narrow size distribution centred at 164 nm. The polydispersity index

(PDI) of these HMSNs is approximately 0.048, suggesting a high size uniformity in these

products [185]. It is found that the diameters showed here are slightly bigger than the size

was measured by TEM (~100nm). This is presumably due to the fact that the DLS

technique measured the particle size in solution where a hydrate layer exists on the

HMSNs, thereby producing a larger diameter than that in its dried form observed from

TEM or SEM.

The small angle X-ray diffraction (SAXRD) pattern of the synthesized HMSNs was

presented in Figure 4.4. Only one distinct broad Bragg diffraction peak was presented at

2θ=2-3 , which was the characteristic XRD pattern of a mesophase pore system [146,

186, 187]. The characteristic peak of mesoporous materials was centred at 2θ=2.1˚. This

result was in good agreement with other mesoporous silica nanoparticles synthesized by

nonionic surfactant templating [101, 118, 135, 188].

Page 76: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

58

2 4 6 8

0

1100

2200

3300

4400

Inte

nsity

/(a.u

.)

2θ/degree

Figure 4. 4 Small angle XRD pattern of HMSNs

According to IUPAC classification, HMSNs exhibited a classical IV type N2 adsorption-

desorption isotherm with well-defined steps at relative pressures (P/P0) of 0.1-0.3 and 0.9-

1.0 corresponding to capillary condensation and desorption in open mesopores and

interstitial pores respectively [189]. The N2 adsorption-desorption isotherm for HMSNs

(Figure 4.5a) displayed two hysteresis loops at P/P0=0.3-0.5 and P/P0=0.9-0.96,

respectively, which further supports the XRD results that suggested HMSNs have a

mesoporous structure. The HMSNs were calculated to have a narrow pore size

distribution centred at 2.5 nm (Figure 4.5b). The specific surface area and pore volume

of HMSNs were 760m2/g and 0.45cm3/g, respectively.

Page 77: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

59

0.0 0.2 0.4 0.6 0.8 1.0

650

700

750

800

850

P/P0=0.9

P/P0=0.96P/P0=0.5

P/P0=0.3

Vad

s/cm

3 g-1

P/P0

(A)

P/P0=0.1

5 10 15 20 25

0.0

0.5

1.0

1.5

2.0

dV/d

logD

(cm

3 /g)

pore size (nm)

(B)2.5nm

Figure 4. 5 Nitrogen adsorption-desorption isotherm (A) and the BJH pore size

distribution of HMSNs (B)

Page 78: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

60

4.2.2 Innovations of Eudragit-assisted HMSNs fabrication

When compared to other conventional synthesis strategies of HMSNs, the use of Eudragit

as template is novel. Eudragit S100 is a co-polymer of methacrylic acid and methyl

methacrylate with the ratio of 1:2. It is stable under acidic pH (< 7) but becomes

dissolvable in base condition (pH >7) or in Acetone [190, 191]. Because of this property,

Eudragit S100 is often used as an enteric polymers for drug delivery [192-194]. By

applying Eudragit S 100 nanoparticles as a core template, the yield products HMSNs

don’t have to undergo the calcination process for the removal of the core template to

create a hollow cavity in HMSNs. Instead, Eudragit core can be easily removed by

soaking the as-synthesized HMSNs (HMSNs with Eudragit cores) in base solution or in

Acetone under room temperature. This solvent extraction process to clean the Eudragit

cores is facile and gentle, which can remarkably prevent the HMSNs from further damage

or aggregation caused by subjecting to a very high temperate. To our knowledge, this is

the first time to report the use of Eudragit S100 nanoparticles as a core template for

HMSNs synthesis. More importantly, Eudragit S100 is non-toxic, which is also superbly

useful for the fabrication of HMSNs as drug carriers.

To further demonstrate the above illustration, as-synthesized HMSNs (with Eudragit

cores) underwent solvent extraction in Acetone and calcination at 550 ˚C for 6 h

separately. Both treatments achieved the same goal, to remove the Eudragit cores, and the

yield products were named as HMSNs-S and HMSNs-C, respectively. The EM images

in Figure 4.6 definitely displayed the superiority of solvent extraction. It can be seen that

HMSNs-C were aggregated and some of them were damaged due to the high temperature.

In contrast, HMSNs-S exhibited an excellent dispersion revealing that the solvent

extraction method plays a crucial role in preparation of discrete HMSNs.

Page 79: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

61

Figure 4. 6 TEM images (a) and SEM images (b & c) of HMSNs-S and TEM images (d) of HMSNs-C and SEM images (e & f)

Page 80: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

62

-150 -100 -50 0 50 100 150

0

20000

40000

60000

80000

100000

120000-4.8 mV

Tota

l Cou

nts

Zeta potential (mV)

HMSNs-S HMSNs-C

-25.6 mV

Figure 4. 7 Zeta potential of HMSNs-S and HMSNs-C

In addition, the aqueous stability of HMSNs was thoroughly studied. The zeta potential

of HMSNs-S and HMSNs-C were -25.6 mV and -4.8 mV at pH 7.4, respectively (Figure

4.7). The higher surface charge of HMSNs-S was due to the abundant Si–OH groups on

the surface [195]. In comparison, when HMSNs-C were subject to calcination at 550 ˚C

for 6 h, some of the Si–OH groups condense with one another resulting in decreased Si–

OH groups on the surface and thus a decreased surface charge (-5.89 mV) was observed.

Figure 4. 8 Stability study of HMSNs-C and HMSNs-S

Page 81: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

63

As shown in Figure 4.8, no obvious precipitation tendency of HMSNs-S was observed

over one week indicating its excellent stable dispersion in water. On the other hand,

HMSNs-C with a low surface charge exhibited decreased stability in aqueous solution,

which was proved by the presence of sediment at the bottom of the sample vial after 2

days of storage.

As such, it is demonstrated that the new Eudragit-assisted strategies for the fabrication of

HMSNs is superior to conventional methods used to fabricate HMSNs in preparation of

highly uniform, discrete and stable HMSNs, given that particle stability is one of the

critical considerations in the biomedical use of mesoporous silica nanoparticles.

Therefore, the aggregation state of HMSNs beginning at the synthesis process should be

taken into account before administration.

4.2.3 Mechanism of Eudragit-assisted HMSNs fabrication

Considering the successful results observed for Eudragit-assisted HMSNs fabrication, it

is important to further elucidate the fabrication mechanism. Generally, Eudragit S100

nanoparticles could be readily synthesized by nanoprecipitation [172] and TX 100

micelles are formed in water at a given concentration and temperature [196]. Hydrogen

bonds may be formed between Eudragit S100 nanoparticles and TX100 micelles due to

the existence of carboxyl groups in Eudragit molecules and hydroxyl groups in TX100

molecules.

If Eudragit S100 nanoparticles and TX 100 micelles are combined to form a core/shell

structure, hydrogen-bonding may be significantly complementary to the Coulombic

interaction between Eudragit S 100 nanoparticles and TX100 micelles. The interactions

between TX100 micelles and Eudragit S100 nanoparticles are proposed in Figure 4.9,

demonstrating the formation of TX100/Eudragit composite micelles. The hydroxyl

groups of surfactant micelles react with the carboxyl groups of Eudragit S100 particles to

form hydrogen bonds. These are critical properties for the formation of TX100 micelles

on the surface of Eudragit nanoparticles which finally form mesopores of HMSNs.

Page 82: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

64

Figure 4. 9 Formation schematics and of TX100/Eudragit S100 composite micelles (a) and proposed reactions between TX100 and Eudragit S100 (equations 1-3)

To demonstrate this hypothesis, Eudragit S100 nanoparticles and TX 100 micelles were

prepared and their reactions were characterized by dynamic light scattering (DLS), FTIR

and PFG-NMR techniques. As shown in Figure 4.10, the formation of composite micelles

was tracked by DLS. Firstly, TX100 micelles with a diameter of approximately 8.7 nm

were formed in water. Subsequently, after adding Eudragit nanoparticles (diameter ~ 68

nm), the TX100/Eudragit composite micelles were formed with a diameter of ~ 92 nm

indicating the coupling between Eudragit nanoparticles and TX100 micelles.

From the FTIR results in Figure 4.11, the investigated composite micelles

(TX100/Eudragit S100) were found to display a broad and intense absorption (continuum)

in the 1500 – 800 cm-1 region, typical for short hydrogen bonds [197-201]. Broad

Page 83: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

65

absorption arise from the bending and stretching vibrations of the O-H group involved in

the short OH….O hydrogen bonds. In addition, the red shift of the O-H stretch, which

varies from 3545 to 3500 cm-1, provides unambiguous information about the formation

of hydrogen bond [202, 203]. Thus, the extrinsic interaction between surfactant and

Eudragit led to the nanocoating of surfactant micelles around Eudragit nanoparticles

through self-assembly.

1 10 100 1000

0

5

10

15

20

25

Inte

nsity

(%)

size (nm)

Eudragit TX100 Eudragit/TX100

Figure 4. 10 DLS of Eudragit S100 nanoparticles, TX100 and their mixture

To provide more direct evidence of the interactions between Eudragit S100 nanoparticles

and TX100 micelles, 1H NMR spectra of Eudragit S100 nanoparticles and the mixture of

TX100/ Eudragit S100 were recorded in Figure 4.12 and the diffusion coefficients of

different species in both samples were listed in Table 4.1

Page 84: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

66

0 500 1000 1500 2000 2500 3000 3500 4000 4500

Wavenumber(cm)

1730

1732 3500

Eudragit

TX100

Eudragit & TX100

3545

Figure 4. 11 FTIR spectra of Eudragit S100 nanoparticles, TX100 and the mixture of

Eudragit and TX100

As shown in Figure 4.12, the 1H NMR spectra of Eudragit S100 nanoparticles presented

two obvious peaks at 4.81 ppm (peak I) and 2.25ppm (peak II), which are assigned to the

residual water and acetone in the D2O solution, respectively. The three peaks labelled as

‘III’ are attributed to the Eudragit S100 molecules. The low intensity of the Eudragit peaks

is due to its poor solubility in D2O at room temperature. In Eudragit S100 nanoparticle

samples, the Eudragit S100 and the acetone molecules show similar diffusion coefficients,

which are largely different from the diffusion coefficient of H2O molecules. This result

suggests that the Eudragit molecules exist in the acetone phase, which is separated from

the bulk water phase. Remarkably, upon the addition of the TX100, the diffusion

coefficient of Eudragit S100 dropped by 75%, from 6.5×10-10 m2/s to 1.635×10-10 m2/s,

whereas the acetone molecules maintain a similar diffusion coefficient to that in the

system of pure Eudragit S100 nanoparticles. This is a strong indication of interactions

between the Eudragit and the TX100 molecules which have a much lower diffusion

coefficient (4.51 ×10-11 m2/s).

Page 85: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

67

Figure 4. 12 1H NMR spectra of Eudragit S100 nanoparticles (upper) and Eudragit/TX100 mixture (bottom); Inserts are TX100 molecular structure.

Table 4. 1 Proton diffusion coefficients (D) of Eudragit and Eudragit/TX100 mixture

a The detailed assignment of each peak is shown in the inset of Figure 4.12. b The diffusion coefficient of TX100 molecules was obtained by fitting the sum of

the integrals of all the corresponding peaks to the diffusion equation.

Samples Peaks Assignment D (m2/s) SDb

Eudragit I H2O (in D2O) 1.22×10-9 1.22×10-3

II Acetone (in D2O) 7.25×10-10 4.65×10-3

III Eudragit 6.5×10-10 4.25×10-3

Eudragit/TX100 I' H2O (in D2O) 1.25×10-9 1.017×10-3

II' Acetone (in D2O) 7.74×10-10 5.32×10-3

III' Eudragit 1.635×10-10 7.4×10-3

1,2,3,4,5,6a TX100 4.51×10-11 b 6.1×10-3

Page 86: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

68

The measured NMR diffusion coefficient of the Eudgragit in Eudragit/TX100 samples is

lower than that of the free Eudragit molecules (in the pure Eudragit sample), yet higher

than that of the TX100 molecules. This can be explained by the fast exchange process

between the free and associated Eudragit molecules. In the time scale of NMR

measurements (10 ms), the Eudragit molecules exhibit a very fast exchange between the

free and associated states. The measured diffusion coefficient (also termed as apparent

diffusion coefficient) is then a population-weighted value between the free and associated

values. Assuming that the associated Eudragit molecules are of the same diffusion

coefficient as the TX100 molecules (4.51 ×10-11 m2/s), it can be easily calculated that

about 19% of the Eudragit molecules are free and 81% of the Eudragit molecules are

associated with TX100 molecules in the equilibrium state.

Figure 4. 13 SEM images (a) and TEM image (b) of non-Eudragit assisted MSNs

To further confirm the effect of the coordination between Eudragit S-100 nanoparticles

and TX-100 micelles on the formation of the hollow mesoporous silica nanoparticles,

Eudragit S-100 was omitted during HMSNs fabrication while keeping other experimental

conditions the same. As seen from the SEM and TEM results (Figure 4.13), without

Eudragit S-100 nanoparticles, the mesoporous silica nanoparticles obtained did not

contain a hollow cavity. More importantly, the dispersibility, yield, size distribution and

overall quality of the structure of these nanopartilces were significantly inferior to the

HMSNs fabricated with the assistance from Eudragit. Thus, the role of Eudragit S-100 is

crucial in the formation of high quality mesoporous nanoparticles. More than being a

Page 87: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

69

solid core particle to develop the hollow cavity, Eudragit S100 nanoparticles also assisted

in the cooperative assembly of TX100 micelles, which contributed to the excellent

morphology of the HMSNs.

4.2.4 5-FU encapsulation and the in vitro release

To assess the potential application of HMSNs in drug delivery, the drug loading and

release behaviour of HMSNs were investigated. 5-fluorouracil (5-FU) standing alone in

colorectal cancer therapy was chosen as a model drug for this study. The 5-FU loading

process predominantly relied on a physical adsorption mechanism in mesopores. The

results shown in Figure 4.14a indicated that HMSNs have a loading capacity of 194.5

mg(5-FU)/g(HMSNs) at 3 mg/mL of 5-FU loading solution. It has been found that a maximum

loading content can be obtained after 8 h.

Figure 4.14b presents the release of 5-FU from HMSNs in pH 5.5 and 8.0 PBS solutions.

The results demonstrated that the release pattern of 5-FU in both pHs is similar. There

were two stages during the release of 5-FU from HMSNs. The first stage of release was

initially rapid, and within this stage, nearly 23.38% and 26.58% of the 5-FU was released

into the aqueous media at pH 8.0 and 5.5, respectively. This may be due to the rapid

diffusion of the 5-FU molecules which were adsorbed on the surface and in pore entrances

of the HMSNs. When 5-FU loaded HMSNs were initially immersed into the PBS solution,

there is a concentration gradient of 5-FU between the PBS buffer and the HMSNs. Thus,

5-FU diffuses more quickly from the pores to the outside media driven by the diffusion

effect. After 1h, the second stage of release was relatively slow and the drug concentration

levelled off in 25 h which is an excellent release profile for cancer treatments.

Page 88: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

70

0 5 10 15 20 25

0

10

20

Cum

ulat

ive

load

ed (%

)

Time (h)

(A)

0 1 2 3 4 15 20 25

0

10

20

30

40

50(B)

Time (hours)

Cum

ulat

ive

rele

ase

(%)

PH 8.0 PH 5.5

Figure 4. 14 (a):5-FU loading profile and (b): 5-FU release profile of HMSNs

Page 89: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

71

On the other hand, when the pH was decreased from 8.0 to 5.5, the release rate was

accelerated and the total amount of 5-FU released increased. Near the isoelectric point of

5-FU (pKa = 8.0), the net charge of this drug molecule was decreased. Thus, electrostatic

attractions between the inorganic carrier and drug molecules were minimal [204].

However, at lower pH (pH =5.5) the elevated electrostatic repulsion would lead to a

higher transport of 5-FU through the mesoporous silica shell and thus a greater release

capacity [205]. The amount of 5-FU released from HMSNs in pH 8.0 and 5.5 PBS

solutions was ~37.98% and ~43.99% at 25h, respectively. These results illustrated that

these novel HMSNs exhibited great sustained release due to their mesoporous structure.

4.3 Pore size tuning for controlled release

An ideal delivery system is the one that can provide a considerably extended period of

effective drug delivery. The ideal slow release kinetic model [23] has three stages (Figure

4.15). The release rate in stage I is initially rapid and as a results, the drug concentration

in plasma can reach its therapeutic concentration rapidly. In stage II, the therapeutic

concentration is maintained for a period of time then drop off gradually to the minimum

effective concentration in stage III. The HMSNs based drug delivery system can exhibit

such a sustained release profile as demonstrated above.

Figure 4. 15 Three stages of the slow release kinetic model

I II III

therapeutic zone

Page 90: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

72

However, with the further development of the cancer treatments, more requirements were

introduced to drug delivery systems. It was reported that, in addition to sustained release,

specific drug release rates are always pursued for personal treatment as different stages

of disease evolution need different drug release rates [206-208]. In this regard, smart

delivery systems with more effective control of drug release rate were of great importance.

Fortunately, the release rate of HMSNs based drug delivery system can be further

controlled and adjusted by changing the pore diameter (Figure 4.16). By varying the pore

size in the shell, the release rate of drug would be tuned to a specific rate for personal

treatments. It is anticipated that the release rate will increase along with the increase of

pore size because the space provided by small pores is limited for the diffusion of drug

molecules while large ones can supply bigger room for more drug molecules to go

through at the same time, which results in a quicker release speed.

Figure 4. 16 Representative schematic illustration of controlling drug release rate by

tuning the pore size of HMSNs

Three HMSNs with different pore size were synthesized using the same fabrication

system. The pore size of HMSNs was successfully tuned by using different surfactants

with different molecular weight (poly ethylene oxide block copolymers F127, P123 and

low molecular weight nonionic surfactant Tweens 20). The higher the molecular weight

Page 91: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

73

of surfactant, the bigger the micelles would be form which were finally evolved to

mesopores of HMSNs.

The pore size of HMSNs can be investigated by a suite of characterisation techniques,

such as BET, XRD and PALS. These techniques provide different information whilst are

complementary to each other. BET and PALS were used to study the pore size distribution

of these three HMSNs (HMSN-1, HMSN-2 and HMSN-3). Combining these methods

would allow a complementary pore size analysis of materials.

4.3.1 BET analysis of pore size

0.0 0.2 0.4 0.6 0.8 1.0

Vad

s/cm

3 g-1

P/P0

HMSNs-2

HMSNs-3

HMSNs-1(a)

0 10 20 30 40 50 600

1

2

Por

e vo

lum

e/(c

m3 g

-1)

Pore size/nm

HMSNs-2 HMSNs-1 HMSNs-3

9.0nm

5.6nm(b)

Figure 4. 17 Adsorption-desorption isotherms (A) and corresponding pore size distribution (B) of HMSNs

Adsorption-desorption isotherms of HMSNs-1, HMSNs-2 and HMSNs-3 were recorded

in Figure 4.17a and the corresponding pore size distribution was shown in Figure 4.17b.

In addition, some physicochemical properties of HMSNs were listed in Table 4.2. The

adsorption-desorption isotherms (Figure 4.17a) of all HMSNs exhibit typical type IV

curves with well-defined capillary condensation steps. This suggested that all HMSNs

had mesoporous structure which was further confirmed by their high surface areas of 687,

589, 760 m2g-1and pore volumes of 0.73, 0.96, 0.45 cm3g-1 for HMSNs-1, HMSNs-2 and

HMSNs-3, respectively (Table 4.2). The pore size of HMSNs-1, HMSNs-2 and HMSNs-

3 was calculated and presented in Figure 4.17b. The appearance of sharp peaks in Figure

4.17b indicated that HMSNs-1 and HMSNs-2 had a majority of mesopores with the size

of 9.0nm and 5.6nm. However, for HMSNs-3, no peaks could be seen from Figure 4.17b.

Page 92: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

74

The reason of this phenomenon may be that the average pore diameter of HMSNs-3 is

relatively small beyond the limit of this characterization. Compared with F127 and P123,

the molecular weight of Tweens 20 and TX100 is much smaller, which produces smaller

micelles during fabrication process and therefore results in smaller pore size of

nanoparticles. As pore size, surface area and pore volume of the final products highly

depend on the kind of surfactant used for fabrication, HMSNs synthesized with

surfactants having similar properties may have similar or even the same physicochemical

properties, including the surface area [209, 210].

Table 4. 2 Physicochemical Properties of HMSNs

Sample

name

surfactants BET

surface

area (m2/g)

Pore size

(nm)

Pore

volume

(cm3/g)

HMSNs-1 F127 589 9.0 0.96

HMSNs-2 P123 687 5.6 0.73

HMSNs-3 Tweens 20 760 - 0.45

4.3.2 PALS analysis of pore size

The PALS is another technique used to identify the porosity of mesoporous materials.

The results gained from PALS are presented in Table 4.3 which illustrates the multiple

scales of porosity readily detectable with positrons. It is known that for mesoporous silica

nanoparticles, τ3 is associated with micro-porosity in the amorphous silica network on

molecular scale (D=0.2-0.4nm) while τ4 provides information on bridging micropore

elements that connect mesopores (D≈1 nm). This information is not easily attainable from

BET results. τ5 is related to the mesopores (D=5-20nm) [211]. Figure 4.18 depicts the

relationship between porosity and τn.

Page 93: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

75

Figure 4. 18 The relationship between porosity and τn [211]

Table 4. 3 PALS data for HMSNs

Sample τ3 (ns) D3

(nm)

I3 (%) τ4 (ns) D4

(nm)

I4 (%) τ5 (ns) D5

(nm)

I5 (%)

HMS-1

0.63

0.23

35.00

5.19

1.07

1.14

79.65

7.8

19.02

HMS-2

0.68

0.26

25.51

6.46

1.18

1.19

72.55

6.6

24.87

HMS-3

0.60

0.21

37.05

7.18

1.24

13.42

72.45

6.5

2.29

τn: o-Ps lifetimes; D: average free path of corresponding pore length; I: corresponding

intensity

Page 94: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

76

As shown in Table 4.3, the pore size calculations were fitted to 3 other ortho-positronium

components that are associated with the pore size (τ3, τ4 and τ5). Each of the components

has an associated intensity value (I3, I4 and I5, relative number of pores) and a calculated

pore size (D3, D4, and D5) using the RTE spherical model. τ3 representing very small

micropores are very uniform between the samples, indicating that it is inherent to these

silica products. τ4 is also very small for pores with the diameter of around 1.20nm. It was

noticed that I4 of HMSNs-3 was much higher than that of other samples. This is a strong

indication of the fact that there is a large proportion of micropores existing in HMSNs-3

and this might be the reason for the absence of pore size distribution peak in Figure 4.17b.

Given that the pore size obtained from BET is calculated by BJH method whose

application is limited to mesopores (2 - 50nm) [212-214], micropores cannot be detected

via this method. The largest pore size D5 (calculated from Tau 5) is related to the pore

size from the BET data. When compared to BET results, the pore size D5 of HMSNs-1

and HMSNs-2 fluctuated to some extent. However only a range of pore sizes rather than

a uniform pore size can be seen from PALS [211]. In addition, it is noteworthy that I5 of

HMSNs-3 is only 2.29, so it is not a significant proportion of pores in this range for this

sample.

In conclusion, BET technique using BJH for pore size calculation is only effective to

evaluate the size of mesopores such as HMSNs-1 and HMSNs-2. When refers to material

with a large proportion of micropores like HMSNs-3, other methods (PALS et al.) should

be applied in the measurement of pore size. Further, the pore size of HMSNs can be

controlled by changing the surfactants used for synthesis. Under the reaction conditions

used here, Tweens 20 resulted in the formation of micropores (< 2nm), while P123 and

F127 tend to yield mesopores. The possible reason for this phenomenon would be

attributed to the surfactant behaviour during fabrication process.

4.3.3 Controlled and sustained release of HMSNs with different pore size

In order to investigate the release behaviour of HMSNs with different pore size, 5-FU had

been loaded into HMSNs-1, HMSNs-2 and HMSNs-3 to obtain 5-FU loaded HMSNs.

The loading capacity and some physical parameters are listed in Table 4.4. Three HMSNs

Page 95: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

77

exhibit a similar loading capacity of around 10%. However, HMSNs are expected to

achieve a higher drug loading (20% ~ 50%) as they have the advantage of the high surface

area and high pore volume. The low drug encapsulation ability of HMSNs might be due

to the electrostatic repulsions between negatively charged silica nanoparticles and

negatively charged 5-FU molecules. Therefore the surface of the HMSNs needs to be

modified to improve 5-FU loading which will be specifically discussed in Chapter 6.

Table 4. 4 Loading capacity and other physical parameters of HMSNs

0 5 10 15 20 25

0

10

20

30

40

50

Cum

ulat

ive

rele

ase

(%)

Time (h)

HMSNs-3 HMSNs-2 HMSNs-1

Figure 4. 19 In vitro release profiles of HMSNs with different pore size in PBS with a pH of 5.0

The release of 5-FU from HMSNs with three pore sizes was examined in phosphate-

buffered solution (PBS) with a pH of 7.4. As evidenced in Figure 4.19, all HMSNs

displayed a sustained release and the release rate is related to the pore size of HMSNs. In

the first 2 hours, three HMSNs demonstrated a similar release rate because at the first

Samples Pore size (nm) Loading

capacity (%)

Zata potentials

(mV)

HMSNs-1 9.0 10.8 -5.32

HMSNs-2 5.6 9.3 -18.6

HMSNs-3 < 2 13.6 -27.1

Page 96: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

78

stage of drug release (stage I presented in Figure 4.15), drug encapsulated on the surface

or at the entrances of pores are released. 5-FU loaded HMSNs have been washed at the

end of the drug loading procedure before freeze-drying. Therefore, some of the drugs

attached on the surface of HMSNs were washed away. However, the release of the drug

remained on the surface, in addition to those encapsulated in the pore entrance and near

the pore entrance leads to the rapid release behavior of HMSNs. At this release stage,

drug molecules were easily released from the HMSNs and the effect of pore size on drug

release is not significant. In addition, the concentration difference of the drug between

the environment (PBS solution) and inside the HMSNs is huge at this release stage, which

could account for the rapid release as well. As the size of the pores is much bigger than

that of the drug molecule (~0.5 nm), which provides enough room for the free

transportation of the drug, the drug can be encapsulated into the pores and the cavity of

HMSNs, not just attached on the surface. When the release proceeds further, drug stored

in the cavity of HMSNs passes through the pores to get released. As a result, pore size

becomes a more influential factor, determining the release speed of 5-FU. As expected,

HMSNs with larger pores present a faster release rate in addition to a higher cumulative

release rate. Within 24 hours, about 42.5%, 37.4% and 33.6% of 5-FU were released from

HMSNs-1, HMSNs-2 and HMSNs-3, respectively.

In the attempt to investigate the release kinetics of HMSNs, an empirical equation (Eq. 1)

was used to describe their release behaviour. Eq. 1 was introduced by Ritger and Peppas

in 1987 for description of solute release from non-swellable and swellable devices in

controlled release systems [215, 216].

(1)

The nature logarithm form of Eq. 1 is:

(2)

Page 97: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

79

where Mt/M is the accumulative release percentage of 5-FU at time t; k is a constant and

n is the diffusional exponent indicating the transport mechanism of the drug from their

carriers. Table 4.5 lists the diffusional exponent values (n) and their corresponding drug

release mechanism based on literature [217, 218].

Table 4. 5 Values of diffusional exponent and corresponding drug release mechanism

Figure 4. 20 In vitro release kinetics of HMSNs with different pore size in PBS with a pH of 5.0

2.4

2.6

2.8

3

3.2

3.4

3.6

3.8

-1 0 1 2 3 4

ln (M

t/M∞)

lnt (hours)

HMSNs-1 HMSNs-2 HMSNs-3

Diffusional exponent, n

Drug release

mechanism Thin film Cylindrical

samples

Spherical samples

< 0.50

< 0.45

< 0.43

Quasi-Fickian

diffusion

0.50 0.45 0.43 Fickian diffusion

0.50< n < 1.00

0.45 < n < 1.00

0.43 < n < 1.00

Non-Fickian

diffusion

(Anomalous

transport)

1.00 1.00 1.00 Zero-order release

Page 98: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

80

To obtain n values for three HMSNs, data extracted from drug release studies (Figure

4.19) were calculated according to Eq. 2. n of HMSNs (the slope) were graphically

determined by plotting versus ln t (Figure 4.20). The linearity of the regression

line was evaluated by the relevance coefficient (R2). As shown in Figure 4.20, the release

data of three HMSNs fit well to the Ritger-Peppas model which was evidenced by a high

R2 value (0.986, 0.981 and 0.943 for HMSNs-1, HMSNs-2 and HMSNs-3, respectively).

n is under 0.43 for all three HMSNs, demonstrating that three HMSNs follow the Quasi-

Fickian diffusion mechanism in which 5-FU molecules transport through the pores of

porous silica carriers without any relevant deformation and stresses during the whole

release process. Further, although the pore size regulates the speed of 5-FU molecules

passing through the pores, it shows no significant effect on their release mechanism. In

other words, the release rate of drug loaded HMSNs can be adjusted to a certain rate for

specific application while maintains their sustained release mechanism.

4.4 Cytotoxicity and non-specific cellular uptake of HMSNs

Figure 4. 21 Cytotoxicity of HMSNs on SW 480 cells at different concentrations

As nanocarriers for drug delivery systems, HMSNs are expected to be biocompatible with

100.00107.72 110.10

100.59

0

30

60

90

120

150

control 20 μg/ml 50 μg/ml 100 μg/ml

Cell

viab

ility

(%)

Page 99: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

81

no toxic function on cancer cells. The biocompatibility of HMSNs was studied by an

MTT assay. Colorectal cancer cells SW 480 cells were cultured with HMSNs of varying

concentration for 24 hours. The cytotoxicity of HMSNs was expressed by the percentage

of cell viability compared to the control group which was not exposed to HMSNs (Figure

4.21). The viability of cells treated with HMSNs in the concentration range of 20 ~ 100

g/mL was similar to cells that were not exposed to HMSNs, demonstrating that HMSNs

are non-toxic and biocompatible with SW 480 cells. Therefore, HMSNs are a potential

nanocarrier for drug delivery systems to colorectal cancer cells.

To evaluate the natural ability of HMSNs to enter cancer cells, the cell uptake of FITC

modified HMSNs by SW480 cancer cells was investigated in vitro by flow cytometry and

fluorescence microscopy. After cell attachment in a 24 well culture plate at 37 C in 5%

CO2 for 24 hours, the SW480 cells (100000/well) were treated with FITC-HMSNs of

different concentration (2, 5, and 10 μg/mL) for various incubation times (30 min, 2 h

and 4 h).

As shown in Figure 4.22a, with increased HMSNs concentration and incubation time,

more particles were found to be internalised into SW480 cells, indicating that cell

internalisation of HMSNs are concentration and time dependent. The confocal

microscopy images and the corresponding 3D microscope projections showed in Figure

4.22b-i proved that the FITC-HMSNs can penetrate the plasma membrane of SW480 cells

and translocate into the cytoplasm. However, the uptake of HMSNs by SW480 cells is

relatively low, reaching a maximum of 50.8 % cells associated with HMSNs at 10 μg/mL

and 4 h incubation time (Figure 4.22a). Thus it is important to develop targeted drug

delivery systems that will improve the internalisation rate.

Page 100: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

82

Figure 4. 22 Cell uptake of FITC-HMSNs with varying time and concentration by SW480 cells (A), confocal laser scanning microscope (CLSM) images of SW480 cells cultivated with FITC-HMSNs (10 μg/mL) for 4 hours (B-G) and corresponding 3D microscope projections (H-I). Blue: the nucleus (B & E); green: FITC (C & F); merging of image B and C (D); merging of image E and F (G); Scale bars are 40 μm for images B, C and D and 10 μm for images E-I.

1.7 2.21 3.393.05

9.29

28.5

7.99

21.3

50.8

0

10

20

30

40

50

60

2 5 10

% c

ells

asso

ciat

ed w

ith H

MSN

s

Concentration (μg/mL)

30 min

2 h

4h

B C D

E F G

A

Page 101: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

83

4.5 Conclusions

Hollow mesoporous silica nanoparticles with large internal cavities were synthesized

through a novel and facile method by using Eudragit S100 nanoparticles as both a core

template and an assistant in the self-organization of surfactant micelles. The aggregation

problem has been effectively depressed through this “Eugragit assisted” strategy. The

HMSNs were found to exhibit a high drug loading and sustained release of chemotherapy

drug 5-FU. The HMSNs followed the Quasi-Fickian diffusion mechanism in which 5-FU

molecules transport through the pores of porous silica carriers without any relevant

deformation and stresses during the whole release process. Further, pore size is an

important parameter to regulate the release speed of 5-FU molecules from HMSNs. More

importantly, the synthesized HMSNs were found to exhibit nontoxic to colorectal cancer

cells and can enter colorectal cancer cells with a concentration and time dependent

manner. The successful synthesis of HMSNs in addition to their excellent

biocompatibility demonstrate the potential of HMSNs as effective drug carriers for

building versatile cancer-targeted drug delivery systems.

Page 102: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

84

Chapter 5 Modification of HMSNs for targeting delivery

5.1 Introduction

Nanoparticles have been extensively studied as drug carriers for cancer treatments [6, 20,

219], because they can be passively accumulated in tumour tissues due to a phenomenon

called the enhanced-permeation end-retention effect (EPR) [6, 7]. However, this effect is

limited just to vascularized tumours and is usually not sufficient for a complete

eradication of the cancer [8]. Numerous research groups are therefore trying to develop

an efficient system for the active targeting of cancer cells by binding various ligands on

the surface of the nanoparticles [9-11]. Such systems could provide a specific

accumulation of the nanoparticles in targeted cells and thus represent a powerful tool to

improve the efficacy of the cancer treatment.

The over expression of certain receptors on the surface of tumour cells can be exploited

for targeting by drug carriers. EGFR, as one of receptors, has shown to involve in cellular

responses including proliferation and metastatic spread [49]. As a result, EGFR is one of

the most heavily investigated tumour targeting biomarkers for constructing nanoparticles

for targeted cancer therapeutic applications [7, 11, 220]. EGF, hepatin-binding EGF (HB-

EGF), betacellulin, transforming growth-factor-a (TGF-a), epiregulin and amphiregulin

are six known endogenous ligands of EGFR. Although any of these ligands can be used

as a targeting agent for EGFR, EGF with a good compromise between the binding affinity

and the size of the molecule (EGF, 6.1 kDa) is one of the most commonly detected factors

in humans [8]. It is hypothesized that EGF as a ligand to EGFR on the surface of

nanocarriers will exhibit specific delivery of anticancer drug, 5-FU, into colorectal cancer

cells, and thus enhance the resultant anti-cancer activity and reduce the side effects of

anticancer drugs.

Page 103: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

85

Several surface modification methods have been implemented to covalently couple EGF

to the surface of nanoparticles [221] and carbodiimide chemistry has been found to be a

preferential method for the binding of EGF to silica nanoparticles [75, 221, 222]. For a

HMSNs based drug delivery system, the challenge lies in successfully maintaining an

intact structure while combining the targeting ligands and drugs into nanocarriers. In

many cases, HMSNs based targeted delivery systems end up with collapsed HMSNs

because too many functionalization steps were applied and each of them might involve

several times of centrifugations and re-dispersions which negatively affect the stability

of HMSNs. Therefore, facile method with less steps for EGF immobilization is of great

importance.

In order to protect the structure of HMSNs, a facile, single step strategy for EGF grafting

will be introduced, where EGF is covalently linked to the surface of amine functionalized

HMSNs by using aqueous 1-(3-(dimethylamino)propyl)-3- ethylcarbodiimide

hydrochloride (EDC) and N-Hydroxysuccinimide (NHS) as promoters. The covalent

attachment of the EGF to the particle surface was first confirmed by time of flight

secondary ion mass spectrometry (ToF-SIMs). Such functionalized particles are proved

by cell experiments to be suitable candidates as drug carriers for targeting delivery due

to the combination of the large loading capacity and the targeting function. Importantly,

the control of the quantity of EGF attachments on the surface of HMSNs was achieved

by changing the EGF concentration during the grafting. An increase in EGF on HMSNs

will increase the targeting efficiency to colorectal cancer cells.

5.2 A facile, one step strategy for EGF labelling

The covalent binding of proteins, such as EGF, to nanoparticles is highly challenging

since their active conformation must be preserved during the binding process, which

limits the choice of the reaction conditions. Several surface modification methods have

been implemented to covalently couple EGF to the surface of nanoparticles [221] and

carbodiimide chemistry is found to be the most effective method for the binding of EGF

to silica nanoparticles [75, 221, 222].

Page 104: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

86

Carbodiimide chemistry is to generate crosslink between carboxylic groups and primary

amines using EDC and NHS and is a versatile and powerful tool for preparing

biomolecular probes, crosslinking peptides and proteins, and immobilizing

macromolecules for use in numerous cell biology detection and protein analysis. EGF as

a protein, contains both carboxylic acids and primary amines (C-and N-termini, and also

in amino acids presenting in side-chains of EGF). Thus, HMSNs bearing either amine or

carboxyl groups can be easily crosslinked to EGF by the utility of EDC and NHS.

However, most previous research only focused on generating carboxyl functionalized

silica nanoparticles for the binding of EGF [75, 221, 222] and the comparatively simple

route of preparing amine functionalized ones were not reported. In particular, HMSNs

with a hollow cavity are often obsessed with the problem of collapse when undergoing

too many times of centrifugation or modification steps. Compared to amine group

functionalization, carboxyl group functionalization of silica nanoparticles is relatively

complicated as it involves more than one step [223, 224]. For example, Kralj et al

achieved carboxyl-functionalized nanoparticles using a follow-up reaction of amine-

functionalized surfaces [225] and Yiu et al firstly modified silica nanoparticles with

cyano groups, then oxidated them to carboxyl groups [226]. As a result, carboxyl

functionalization of silica nanoparticles is more time consuming and might negatively

affect the physicochemical properties (such as suspension stability) of the products. To

overcome these shortcomings, amine modification was employed to produce HMSNs

with functional groups while maintaining an intact structure for EGF labelling.

5.2.1 Synthesis of amine functionalized HMSNs (HMSNs-NH2)

HMSNs-NH2 were synthesized by co-condensation method in which NH2 groups were

introduced during the fabrication process. This results in a more homogeneous coverage

of functional groups onto HMSNs. As shown in TEM and SEM images (Figure 5.1), all

morphologic features in plain HMSNs are remained in HMSNs-NH2, such as the hollow

structure (Figure 5.1c-f), the particle diameter (~150 nm), and a good dispersion, making

them very suitable for targeted drug delivery systems. The deliberately broken HMSNs-

NH2 shown in Figure 5.1c and f further proved their hollow structure [102]. It is

Page 105: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

87

noteworthy that most HMSNs-NH2 have intact structure (Figure 5.1a and d). Similar to

HMSNs’ characterization, XRD and BET were performed to confirm the mesophase of

HMSNs-NH2. A broad Bragg diffraction peak within the range of 2θ=0-3˚ was found in

Figure 5.2b, illustrating that HMSNs-NH2 have a mesoporous structure [146, 186, 187].

BET results were collected and presented in Figure 5.2a, in addition to a classical IV type

isotherm, a hysteresis loop at relative pressures (P/P0) of 0.45-0.9 was observed which

further proved the mesoporous structure of HMSNs [189]. The pores of HMSNs-NH2

have a narrower size distribution centred at 3.0 nm (Figure 5.2c). The specific surface

area and pore volume of HMSNs-NH2 are 544m2/g and 0.57cm3/g, respectively.

It is noteworthy that, compared to Figure 4.4, the Bragg diffraction peak of HMSNs-NH2

(Figure 5.2b) shifted to a lower angle, which could be due to the grafting of –NH2

functional groups. According to the TEM, SEM and XRD results, the introduction of –

NH2 onto HMSNs didn’t have significant influence on the porous structure of HMSNs

which is used for the drug loading and controlled release. However, due to the coverage

of –NH2, slight decrease in the surface area has been identified via BET results.

Page 106: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

88

Figure 5. 1. a, b & c) SEM images of HMSNs-NH2 and d, e & f) TEM image of HMSNs displaying hollow interior; c & f are broken HMSNs-NH2 showing their hollow structure

Page 107: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

89

0.0 0.2 0.4 0.6 0.8 1.00

10

20

30

40

50

V ads/c

m3 g-1

P/P0

A

Figure 5. 2 A) Nitrogen adsorption-desorption isotherm; B) Small angle XRD pattern of HMSNs-NH2; C) BJH pore size distribution.

0 2 4 6 8 10

0

1000

2000

3000

4000

5000

Inte

nsity

/(a.u

.)

2θ/degree

B

0 5 10 15 200.0

0.5

1.0

1.5

2.0

dV/d

logD

Pore size (nm)

C

Page 108: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

90

NH2 groups were qualified via FTIR and quantified by thermogravimetric analysis (TGA)

and NMR techniques. FTIR spectra of HMSNs-NH2 and the control sample (HMSNs)

are shown in Figure 5.3. Both samples have strong bands located at around 460 cm-1 (δO–

Si–O), 1100 (ʋasym. Si–O–Si), and 3300-3500 cm-1 (ʋSi–OH), confirming the presence of the

SiO2 inorganic phase. In particular, the bands with the highest intensity, centred at 1089

and 1073 of HMSNs and HMSNs-NH2, respectively are associated with the SiO2

stretching vibration [227]. The NH2 functionalization is clearly evidenced by the bands

detected at around 1500 cm-1, which is unambiguously attributed to the NH bending

vibration [227-230]. Results obtained by FTIR prove the effective NH2 functionalization

of mesoporous silica nanoparticles.

0 1000 2000 3000 4000

3311

1089

465

Wavenumber(cm-1)

HMSN HMSN-NH2

NH2

460

1073

3445

Figure 5. 3 FTIR spectra of HMSNs-NH2 and HMSNs

The TAG thermograms and 29Si single pulse excitation (SPE) spectrum were given in

Figure 5.4 for the evaluation of NH2 quantity. A weight loss of 14.6% was found in

HMSNs while HMSNs-NH2 presented an increased weight loss percentage of 21.5%

(Figure 5.4a). The higher weight loss of HMSNs-NH2 is due to the degradation of grafted

functional groups, inferring that around 5.9% (w/w) of HMSNs-NH2 was related to NH2

immobilization. The 29Si SPE of HMSNs-NH2 spectrum is shown in Figure 5.4b in which

Qm (m=2-4 indicating 2, 1 or 0 –OH groups are linked to Si) and T signals provide

Page 109: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

91

information of silanol groups and grafted functional groups (-NH2), respectively [228,

231]. The peaks at -110, -100 and -90 ppm are associated to Q4 [(SiO)4Si], Q3

[(SiO)3SiOH] and Q2 [(SiO)2Si(OH)2], given that –OH groups of the HMSNs react with

amine silane (APTES) for –NH2 immobilization which results in a decrease in the number

of –OH groups on HMSNs. Thus, the presence of signals with decreased amount of –OH

groups (Q4 & Q3) in addition to T signal can reveal the successful grafting of –NH2 [231-

235]. The relative concentration of -NH2 in HMSNs (15%) was calculated by taking the

signal summation (T + Q2 + Q3 + Q4) as 100%.

100 200 300 400 500 600 7000

20

40

60

80

100

Wei

ght (

%)

Temperature (0C)

HMSNs HMSNs-NH2

weight loss:HMSNs 14.6%HMSNs-NH2 21.5%

(A)

100 50 0 -50 -100 -150 -200

T

Q4

Q3

Q2

46%

34%

5%

Chemical Shift (ppm)

15%

(B)

Figure 5. 4 TGA thermograms of HMSNs & HMSNs-NH2 (a) and 29Si single pulse excitation (SPE) spectrum of HMSNs-NH2

Page 110: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

92

5.2.2 Preparation of EGF grafted HMSNs and characterization of EGF labelling

Figure 5. 5 Schematic diagram showing the process of EGF grafting onto HMSNs

A facile, single-step process was developed to graft EGF onto HMSNs (Figure 5.5).

After grafting, the characterization of EGF attachment is a crucial subsequent step.

However, since the amount of EGF on HMSNs is very low, the determination of EGF

immobilization emerges to be a big challenge as most characterization methods, such as

thermo-gravimetric analysis (TGA), energy-dispersive X-ray spectroscopy (EDX),

electron energy loss spectroscopy (ELLS) and even X-ray photoelectron spectroscopy

(XPS) have their limit. As EGF is used as a targeting ligand to target cancer cells [75],

cell and/or even animal experiments were carried out by researchers to evaluate the

effectiveness of EGF and thus validate their success and effectiveness of the grafting

[236-238]. For example, comparison between EGF and non-EGF grafted samples in

cellular uptake by cancer cells was largely used in the illustration of effective EGF

grafting. Even though these passive methods are useful, they cannot provide direct

information on stability and density of targeting ligands on the surface of nanoparticles

and thus fail to control the quantity of targeting molecules on nanoparticles for specific

application.

Time of flight secondary ion mass spectrometry (ToF-SIMS) as a highly specific

analytical tool in the study of proteins, has been shown to be a powerful technique in the

detection of extremely low amount of proteins which were undetectable by XPS [239].

Inspired by the high chemical selectivity and surface sensitivity of ToF-SIMS [239-242],

it was hypothesized that ToF-SIMS would allow for the observation of successful EGF

attachment to the HMSNs in addition to the quantity of the EGF immobilization.

Page 111: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

93

In this regard, ToF-SIMS analysis was performed on HMSN surfaces before and after

EGF grafting. Amino acids presenting in the backbone of proteins such as EGF

are cracked into smaller immonium ions (amino acid mass fragments) during the

measurement of TOF-SIMS. Identification of characteristic amino acid mass fragments

by ToF-SIMS lead to its extremely high sensitivity (better than XPS) for the detection of

proteins [240].

Figure 5.6 displays the positive survey ToF-SIMS spectra (m/z 0 - 200) for the

unmodified HMSNs, amine group grafted HMSNs (HMSNs-NH2), EGF and HMSNs

labelled with EGF (HMSN-NH2-EGF). For the HMSN surfaces (Figure 5.6a) the intense

ion signals at nominal m/z 28, 29 and 45 amu can be attributed mainly to Si+, SiH+ and

SiHO+ fragments [239]. Some peaks raised by the amine groups on HMSNs-NH2 surface

are shown in Figure 5.6b. When compared to HMSNs, HMSNs-NH2 have new peaks at

15, 41, 43, 55, 57 and 69 amu which correspond to HN+, C3H5+, C2H5N+, C3H5N+,

C3H7N+ and C3H3NO+ fragments, confirming the successful grafting of NH2.

A control spectrum of the EGF is presented in Figure 5.6c. The main peaks at nominal

m/z 15, 27, 41, 43 and 55 happen to overlap with those of HMSNs-NH2. However in

addition to these peaks, intense peaks at nominal m/z 30, 70, 72, 73, 84, 86, 110 and 130

are observed which correspond to CH4N+, C4H8N+, C4H10N+, C2H7N3+, C5H10N+,

C5H12N+, C5H8N3+ and C9H8N+ and can be attributed to fragments from the EGF

backbone [242].

Figure 5.6d presents the ToF-SIMS spectra for EGF grafted HMSN surfaces. Ion signals

of the major fragments for both HMSNs and EGF were detected. For example, ion signals

of unmodified HMSNs reported above at nominal m/z 28, 29 and 45 amu can be seen. In

addition, the ion signals at nominal m/z 30, 70, 72, 73, 84 and 86 amu which were

assigned to fragments of EGF from the spectra of pure EGF (Figure 5.6c) are present

thereby suggesting successful EGF attachment. Interestingly, besides these characteristic

ion signals of HMSNs and EGF, new intense ion signals at nominal m/z 58 and 129 were

detected which are related to C3H8N+ and C5H13N4+ fragments. After grafting onto the

HMSNs surface, the conformation and orientation of EGF might change which can

subsequently lead to different distributions of the amino acid fragments detected in a

Page 112: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

94

static ToF-SIMS experiment [241]. Therefore, these new peaks could be seen as

characteristic peaks of EGF grafting using the methods we illustrated above. These ToF-

SIMS results demonstrate the successful EGF labelling of HMSNs.

Figure 5. 6 Positive survey mass spectra for: (A) HMSNs surface; (B) HMSNs-NH2 surfaces; (C) EGF; (D) HMSNs-NH2-EGF.

Page 113: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

95

5.2.3 The effect of EGF concentration on surface chemistries of the HMSNs

The stability and density of EGF on the surface of HMSNs may play a very crucial role

in targeting the cancer cells, though only one EGF molecule is needed to bind to the EGFR

on the cell to initiate internalisation. However, the more EGFs grafted onto the HMSNs,

the higher the probability of HMSNs-NH2-EGF targeting the cancer cells. It is

consequently important to quantify the EGF on the surface of HMSNs. In this regard,

EGF grafting was explored by varying the concentration of EGF during the attachment

procedure from 0.1 to 0.8 mg/ml and analysing the surfaces with ToF-SIMS (Figure 5.7).

Figure 5. 7 Positive survey mass spectra of HMSN-NH2-EGF surface using different EGF concentration: 0.1,0.2,0.4 and 0.8 mg/ml.

Page 114: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

96

Figure 5. 8 Plot of ratios of the intensities of the 28 m/z (from HMSNs) to 59 m/z, 70 m/z and 58 m/z (from EGF)

It appears that the EGF concentration has a limited effect on the resulting surface as no

significant differences can be indentified between samples prepared with 0.1, 0.2, 0.4 and

0.8 mg/ml of EGF (Figure 5.7). Differences, however, emerge upon close inspection of

the ratios of 28/59, 28/70 and 28/58 mass fragments. 28 m/z is an ion signal specific for

HMSNs while 59 m/z, 70 m/z and 58 m/z are specific for EGF, thus the ratios of 28/58,

28/70 and 28/58 track the relative amount of EGF grafted on the surface of HMSNs.

Figure 5.8 displays that the ratio of 28/59 decreases with an increase in EGF concentration,

therefore suggesting that an increase in EGF concentration can lead to a greater amount

of EGF to be attached onto the surface of HMSNs. As ToF-SIMS is a very surface specific

technique (penetration depth < 1.5 nm) [241], an increase in EGF surface coverage would

result in less silica but more protein being detected, thus the ratio of 28/59 decreased.

Moreover, the same trend can be seen for the 28/70 and 28/58 ratios.

The differences between the samples shown in Figure 5.7 are developed from comparing

only a few peaks selected from ToF-SIMs spectra and the information contained in

unselected peaks is disregarded. Although this univariate analysis is useful, it is not

enough to properly evaluate the EGF attachment of these surfaces, because the remaining

peaks can provide important information for fully classifying the differences of surface

chemistries between samples. Thereby, PCA was employed to extract the most influential

factors from the complex mass spectra and to help in data interpretation. In the attempt to

0

2

4

6

8

10

12

14

16

0 0.2 0.4 0.6 0.8 1

28/59 28/70 28/58

Page 115: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

97

classify the TOF-SIMS data by PCA, 35 peaks listed in Table 5.1 were used. The selection

of these peaks is based on the assignments taken from literature [243-246]. Only the

characteristic nitrogen containing peaks were studied which arise from amino acid

fragments ions so they are indicators of EGF immobilization.

Table 5. 1 Positive fragment ions (immonium ions) used in PCA

NO. Fragment m/z No Fragment m/z

1 CH4N+ 30 18 C4H10N+ 72

2 C2H4N+ 42 19 C2H7N3+ 73

3 CH3N2+ 43 20 C3H8NO+ 74

4 CH4N2+ 44 21 C4H6NO+ 84

5 C2H6N+ 44 22 C5H10N+ 84

6 C2H3S+ 47 23 C5H12N+ 86

7 C3H4N+ 54 24 C3H7N2O+ 87

8 C3H6N+ 56 25 C3H8NO2+ 88

9 C2H4NO+ 58 26 C4H10N3+ 100

10 C3H8N+ 58 27 C4H11N3+ 101

11 CH5N3+ 59 28 C4H8NO2+ 102

12 C2H5S+ 61 29 C7H7O+ 107

13 C4H6N+ 68 30 C5H8N3+ 110

14 C4H5O+ 69 31 C8H10N+ 120

15 C3H4NO+ 70 32 C5H11N4+ 127

16 C4H8N+ 70 33 C5H13N4+ 129

17 C3H3O2+ 71 34 C9H8N+ 130

35 C9H8O+ 132

Page 116: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

98

Figure 5.9 displays the scores plot of positive fragment ions on PC1 and PC2 for the EGF

grafted HMSN samples and their corresponding loading plots show the influence of each

amino acid peak on PC1 and PC2. Each sample was characterised by 35 groups of positive

fragment ions derived from replicated positive ion mass spectra. About 91 % of the total

variances were captured by PC1 and PC2 revealing that these two-PC’s models retain the

most of the original information.

-0.08 -0.04 0.00 0.04 0.08-0.12

-0.08

-0.04

0.00

0.04

0.08

EGF-0.1 EGF-0.2 EGF-0.4 EGF-0.8

PC2

(34.

4%)

PC1 (56.7%)

(a)

-0.015

-0.010

-0.005

0.000

0.005

0.010

0.015

0.020

0.025

C4H8N+

CH5N3+

C3H8N+C2H4N

+

Load

ing

on P

C1

(56.

72%

) CH4N+ (b)

-0.025

-0.020

-0.015

-0.010

-0.005

0.000

0.005

0.010

0.015

0.020

0.025

C3H8N+

Load

ing

on P

C2

(34.

44%

)

CH4N+

(c)

Figure 5. 9 Scores on PC1 and PC2 (a) and corresponding loading plots on PC1 (b) and PC2 (c) of immonium ions from static positive spectra of HMSNs-NH2-EGF prepared

with EGF concentration of 0.1, 0.2, 0.4 and 0.8 mg/ml

Page 117: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

99

As shown in Figure 5.9a, with an increase in EGF concentration, a decrease of PC1 scores

was observed. The EGF concentration of 0.1 and 0.2 mg/ml groups having the highest

PC1 scores overlaps with both PC’s, which reflects their similar surface chemistries in

terms of nitrogen species. Marked differences are observed when the EGF concentration

is increased further. When the concentration increases to 0.4 mg/ml, the samples have

negative scores on both PC1 and PC2 and that are separated from low concentration ones

(0.1 and 0.2 mg/ml). In contrast, the samples with 0.8 mg/ml EGF concentration have

negative scores on PC1 but positive scores on PC2 and that are well separated from other

three samples, signifying that they have quite different surfaces. This may be due to more

EGF proteins being attached onto the HMSNs surfaces. Figure 5.9b and c indicate a high

negative loading of C4H8N+ (at m/z 70 amu) on PC1 and C3H8N+ (at m/z 58 amu) on PC2,

which reflects these two fragments contribute the most to the differences between samples

with different amount of EGF attachments.

In summary, PCA on the ToF-SIMS data demonstrates EGF has been attached onto the

NH2 modified HMSNs surface and different concentration of EGF grafting leads to

different surface chemistries of the final products.

5.2.4 Analysis on immonium ions distinguishing samples with different surface chemistry

The loadings of immonium ions on PC1 and PC2 point at the peaks (CH4N+, C2H4N+,

C3H8N+, CH5N3+ and C4H8N+) that discriminate the groups (Figure 5.9 b and c). Therefore,

the relative intensities of these ions, which have the most impact on distinguishing

samples with different surface chemistry, are closely related to the amount of EGF that

has been grafted onto the nanoparticles. The detailed effect of EGF concentration on the

amount of EGF attachments can be observed by comparing the relative intensities of these

main immonium ions between different samples. In order to gain an acute insight into the

quantity of EGF on samples, Si+ and SiOH+ fragments are included in the analysis. Thus,

each sample was characterised by 7 groups of immonium ions derived from replicated

positive static secondary ion mass spectra.

Page 118: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

100

Figure 5. 10 Comparison of characteristic immonium ions intensities between samples

Shown in Figure 5.10, unmodified HMSNs had significantly high Si+ and SiOH+

intensities but low immonium ions intensities (almost zero), consistent with the fact that

no EGF protein is present on the surface of this sample. NH2 functionalized HMSNs

present a similar trend to unmodified HMSNs, however these samples have relatively

high CH4N+ and C2H4N+ signals due to the presence of amine functional groups on the

surface. On the contrary, all EGF grafted samples had enhanced immonium ion signals

and decreased silica intensities. Further, with the increase in EGF concentration, the Si+

and SiOH+ intensities decreased inferring that EGF concentration had a positive effect on

facilitating the EGF grafting.

Significant differences can be seen by close inspecting the changes of immonium ions

signals between samples. Firstly, CH5N3+ and C4H8N+ fragments are the EGF indicators

since HMSNs-NH2 didn’t show such fragments. The intensities of these two immonium

ions increased along with an increase in EGF concentration revealing more EGF has been

grafted onto the surface of HMSNs, which is consistent with the gradual decrease of silica

signals. If the intensities of these two fragments of pure EGF were set as 100%, at the

highest EGF concentration (0.8 mg/ml), the detected intensities of these two fragments

0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

CH4N C2H4N C3H8N CH5N3 C4H8N Si SiOH

Nor

mal

ised

cou

nts

+SIMS Species

EGF ProteinPlain HMSNsNH2 functionalised HMSNsNH2 +EGF 0.1NH2 +EGF 0.2NH2 +EGF 0.4NH2 +EGF 0.8

0

0.01

0.02

CH5N3 C4H8N

Page 119: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

101

of HMSNs-NH2-EGF are 66.7 % and 78.6%, respectively revealing more than 50% of

EGFs was grafted on HMSNs. Secondly, the CH4N+ and C2H4N+ fragments presented in

EGF and the amine group grafted samples varied differently. Along with an increase in

EGF concentration, although a gradually growing trend was not observed with these two

ions signal intensities, four EGF grafted HMSNs showed improved signals compared to

pure EGF and HMSNs-NH2 samples. These enhanced intensities were attributed to the

synergy of joint signals of CH4N+ and C2H4N+ generated by EGF and free amine groups.

Lastly, C3H8N+ fragment was uniquely shown in EGF grafted samples.

ToF-SIMS combined with PCA were proved in this study to be able to evaluate the

binding efficiency between EGF and silica nanoparticles. This method could be extended

as a general approach to analyse protein biomarkers in targeted drug delivery systems.

5.3 Comparison of newly developed grafting method with other

methods

5.3.1 Superior EGF grafting efficiency by the new grafting method

As mentioned before, HMSNs with a cavity inside tend to form a defect structure as too

many functional steps or harsh experimental procedures (such as high temperature) are

involved in the process which will cause the collapse of HMSNs, a main issue that need

to be considered. Therefore, in order to protect HMSNs from damage, a more facile

process with fewer steps to graft EGF onto HMSNs was developed. To compare, EGF

grafted HMSNs were prepared via two routes and final products were named as HMSN-

COOH-EGF (produced from commonly use route) and HMSN-NH2-EGF (fabricated

from the newly developed route) and characterized through ToF-SIMS (Figure 5.11). As

shown from ToF-SIMS results, ion signals of major fragments of both silica and EGF

were detected in both HMSN-COOH-EGF and HMSN-NH2-EGF. The intense ion signals

at nominal m/z 28 can be attributed to Si+, while peaks at nominal m/z 58, 70, 71, and 72

amu correspond to EGF. Importantly, although EGF can be attached to HMSNs through

either of the routes, it is worth noticing that HMSN-NH2-EGF have demonstrated

enhanced EGF signals and decreased silica signals inferring more EGF were introduced

Page 120: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

102

to covered the surface of HMSNs. Therefore, less silica signals were detected in this

sample. This result proved that new EGF immobilization method has a superior EGF

grafting efficiency when compared to the commonly used ones.

Figure 5. 11 Routes of EGF lebelling based on carbodiimide chemistry: (A) A commonly used route; (B) Newly developed route with reduced steps to protect HMSNs from damage and Positive survey mass spectra for: (C) HMSN-COOH-EGF surface; Insets: High mass resolution of static ToF-SIMs in the 70-72 amu region. (D) HMSNs-NH2-EGF surfaces;

Further, SEM images were taken to study the morphology of EGF grafted HMSNs

through different routes (Figure 5.12). After EGF grafting, HMSN-NH2-EGF remain the

intact structure while HMSN-COOH-EGF show broken particles. This is due to more

modification steps applied and thus more centrifugation rounds were involved in route 1

(Figure 5.11).

Page 121: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

103

Figure 5. 12 SEM images of HMSN-COOH-EGF (a) and HMSN-NH2-EGF (b)

5.3.2 Increase EGF labels on HMSNs for enhanced targeting efficiency

The use of TOF-SIMs provides opportunities to determine relative EGF concentration on

HMSNs. Therefore, the amount of EGF grafted onto HMSNs was compared by changing

the EGF concentration during modification process. In 5.2.3 and 5.2.4, it was reported

that increasing EGF concentration can result in higher EGF attachments on nanoparticles

via route 2 illustrated in Figure 5.11. Again, 4 different EGF concentrations for EGF

grafting (0.1, 0.2, 0.4 and 0.8 mg/ml) were compared based on carboxyl group

functionalized HMSNs (route 1 in Figure 5.11). Unfortunately, no differences can be

found from these four spectra (Figure 5.13a). Moreover, as shown in Figure 5.13b,

although different EGF concentration were employed, all samples were overlapped along

both PC’s indicating their similar surface chemistries (a comparison to Figure 5.9). From

Figure 5.13a and b, 4 important ion signals were selected: 28 m/z (specific for HMSNs),

30 m/z, 44 m/z and 58 m/z (correspond to EGF). The ratio of silica to protein: 28/30,

28/44 and 28/58 of different EGF grafted samples was calculated and presented in Figure

5.13c (a comparison to Figure 5.8). When the EGF concentration was increased gradually

from 0.1-0.8 mg/ml, rather than a regular trend, the ratio of 28/58, 28/44 and 28/30

fluctuated. Thereby an effective control of EGF quantity on HMSNs is failed with the

commonly used route for EGF immobilization.

Page 122: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

104

A new grafting method was developed specific to improve the EGF grafting efficiency of

HMSNs in 5.3.1. This new strategy for EGF labelling can tailor make HMSN-NH2-EGF

with a certain amount of EGF attachments on HMSNs (illustrated in 5.2.3). This will

contribute to more versatile applications of EGF grafting. A precise control of EGF on

HMSNs could lead to an effective regulation of nanoparticles entering the cells and

thereby controlling the drug concentration in targeted cells. It is known that the cancer

development is a multistep process where different drug concentration and exposure time

were required at different stages.

Page 123: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

105

-0.15 -0.10 -0.05 0.00 0.05 0.10 0.15-0.04

-0.03

-0.02

-0.01

0.00

0.01

0.02

0.03

0.04

EGF-0.1 EGF-0.2 EGF-0.4 EGF-0.8

PC

2 (7

.9%

)

PC1 (82.6%)

(B)

Figure 5. 13 Positive survey mass spectra (a) and scores on PC1 and PC2 (b) of HMSN-COOH-EGF surface using different EGF concentration: 0.1,0.2,0.4 and 0.8 mg/ml; and plot of ratios of the intensities of the 28 m/z (from HMSNs) to 30 m/z, 44 m/z and 58 m/z (from EGF) (c)

0

2

4

6

8

10

0 0.2 0.4 0.6 0.8 1

Inte

nsity

EGF Concentration (mg/ml)

(C)

28/30

28/44

28/58

Page 124: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

106

5.4 Effectively target colorectal cancer cells by EGF grafted HMSNs

5.4.1 Effective targeting effect of HMSN-EGF to EGFR positive CRC cells

Figure 5. 14 Cellular uptake of both EGF grafted and plain HMSNs in SW480 cells

Targeting is essential for drug delivery systems as it has the potential to precisely target

and kill cancerous cells while leaving normal cells unharmed. To configure HMSNs for

targeted drug delivery, EGF was grafted onto HMSNs (HMSN-EGF) to target the EGFR

positive colorectal cancer cells (SW 480 cells). HMSNs with and without EGF attachment

were covalently linked to a fluorescent dye (FITC), making them visible by fluorescence

microscopy and flow cytometry. SW 480 cells were treated with varying doses (1, 2, 5,

10 and 20 μg/ml) of HMSN-EGF for 2 hours in addition to non-EGF grafted HMSNs for

control.

With increased nanoparticle concentration, more HMSNs and EGF-HMSNs are found to

be associated with SW480 cells (Figure 5.14), indicating that the uptake of both HMSNs

and EGF-HMSNs by the cells are concentration dependent. It was found for nanoparticle

concentrations below 20 μg/ml, the cellular uptake of HMSN-EGF was higher than

unmodified HMSNs (Figure 5.14). In order for nanoparticles to internalise into > 90 %

of cells only 5 μg/mL of EGF-HMSNs was required while for unmodified HMSNs 20

0.00

20.00

40.00

60.00

80.00

100.00

120.00

0 5 10 15 20 25

% C

ells

ass

ocia

ted

with

HM

SNs

Concentration (μg/ml)

HMSNs

HMSN-EGF

Page 125: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

107

μg/mL was necessary to achieve the same result. This improved internalisation was due

to the ability of the targeting molecules, EGF, to attach to their receptors, EGFR, which

were found to be overexpressed on cancer cells. Therefore EGF attachment can facilitate

the cellular uptake of nanoparticles and the EGF-EGFR reaction could account for this

observation.

On the other hand, at a concentration of 20 μg/mL both HMSNs and HMSN-EGF were

found to be taken up by ~ 100% of cells. However it is more preferable to deliver a smaller

quantity of HMSNs to minimise side effects to patients. Additionally, due to the targeting

function of EGF it will be possible to reduce systemic elimination of HMSNs from the

body and the problematic delivery of chemotherapeutic drugs to normal cells.

Figure 5.15 a-f displays confocal microscopic images of SW480 cell incubated with FITC

labelled HMSN-EGF and HMSNs (10 μg/mL) for 0.5, 2 and 4 hours. Green dots are

nanoparticles and cell nucleus were visualized in blue colour. As observed in Figure 5.15

a-f, with increased time, more HMSNs and EGF-HMSNs are taken up by SW480 cells,

demonstrating that the uptake of both HMSNs and EGF-HMSNs by the cells are time

dependent. Also, the uptake of HMSN-EGF into target cells was significantly higher than

that of HMSNs when the same incubation time was applied, consistent with the fact that

EGF attachment enhances the capacity of nanoparticle entering the targeted cells (EGFR

positive cells).

To further prove that the particles are actually penetrating into the cells, the cytoplasm of

the cells was stained and 3D projections of the cellular uptake of EGF-HMSNs were

presented in Figure 5.15 g-h. The cells were sliced with a slice thickness of 1 μm and all

the slices were combined and presented as a 3D projection. Green dots were seen within

the 3D projections indicating the EGF-HMSNs have been internalized into the cells.

Page 126: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

108

Figure 5. 15 (a-c): the confocal laser scanning microscope (CLSM) images of SW480 cells cultivated with 10 μg/mL FITC-HMSNs; (d-f): CLSM images of SW480 cells cultivated with 10 μg/mL FITC-EGF-HMSNs; and (g-h): 3D microscope projections of SW 480 cells cultivated with 10 μg/mL FITC-EGF-HMSNs for 4h.Green dots are nanoparticles while cell nucleus were visualized in blue colour.

Page 127: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

109

5.4.2 Reduced targeting effect of HMSN-EGF by pretreating SW480 cells with free EGF

Figure 5. 16. The cellular uptake rate (mean fluorescence intensity) of HMSN-EGF and HMSNs in SW 480 cells with or without 5 μM of free EGF pretreatment

As HMSN-EGF had a higher uptake in SW 480 cells, to validate that this enhanced uptake

rate is facilitated specifically by the EGF targeting bioconjugates, SW480 cells were

treated with 5 μM of free EGF before incubation with 5 μg/mL of HMSN-EGF and plain

HMSNs. As seen in Figure 5.16, when cells were pretreated with free EGF, the ability to

internalize HMSN-EGF decreased while the capacity to internalize plain HMSNs

remained the same. After pretreating with free EGF, the mean fluorescence intensity (MFI)

value of cells treated with HMSN-EGF decreased about 25%. However, no obvious

differences can be seen in the uptake of plain HMSNs between cells with and without

free EGF pretreatments. This demonstrated that part of the EGF receptors on the cell

surface can be occupied by free EGF which may inhibit the binding between HMSN-EGF

and EGFR. This result can be further confirmed by the confocal microscope images, as

seen in Figure 5.17, the uptake of the nanoparticles into target cells was visualized by

luminescence of green emitting FITC (green dots) and the cell nucleus was shown in blue

colour. After 2 hours of treatment with HMSNs, only a few of the cells were seen

associated with HMSNs while the uptake of HMSN-EGF was much higher. After

0.0

500.0

1000.0

1500.0

2000.0

2500.0

3000.0

3500.0

HMSNs HMSN-EGF

Cells without freeEGF pretreating

Cells with free EGFpretreating

25%

Page 128: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

110

pretreating with free EGF, cells associated with HMSN-EGF decreased indicating the

uptake of HMSN-EGF was inhibited.

HMSNs HMSN-EGF HMSN-EGF + free EGF

Figure 5. 17. Confocal microscope images of EGF-HMSNs and HMSNs in SW 480 cells.

5.4.3 Non-targeting effect of HMSN-EGF to EGFR negative cell line SW620

The cellular uptake studies on the SW 480 cells have demonstrated that HMSN-EGF

exhibited a higher uptake rate than HMSNs and the cellular uptake can be inhibited by

pre-treating the cells with 5 μM of free EGF. To further confirm that the higher uptake

of HMSN-EGF was mediated specifically via EGFR, cellular uptake experiments were

conducted on SW 620 cells which do not express EGFR.

As shown in Figure 5.18, the SW 620 cells were treated with different concentrations of

HMSNs or HMSN-EGF for 2 hours and the MFI were measured by flow cytometer.

Similar to the cellular uptake in SW 480 cells, the cell uptake of both HMSN-EGF and

HMSNs were concentration dependant: the mean fluorescence intensity increased as the

concentration of nanoparticles increased. However, the difference between HMSN-EGF

and HMSNs is not obvious in this cell line as the MFI of cells treated with HMSNs and

HMSN-EGF on each concentration point was close to each other.

Page 129: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

111

Figure 5. 18. The mean fluorescence intensity in the SW 620 cells treated with different concentrations of HMSNs or EGF-HMSNs for 2 hours.

5.5 Conclusions

To precisely kill colorectal cancer cells, HMSNs were successfully synthesized and

bioconjugated with specific receptor ligand EGF. The construction of a targeting

molecule (EGF) to the surface of HMSNs was achieved through a newly developed

strategy, where the carboxyl groups of EGF were conjugated to an amine terminated

mesoporous silica surface. This method to graft EGF on silica surface proved be facile

and effective and can protect HMSNs from damage during modification. The accurate

control of the number of EGF attachments on HMSNs was achieved via this method,

proving its superiority to commonly used ones.

The characterization of EGF attachments was carried out using ToF-SIMS. The ToF-

SIMS data revealed the successful grafting of EGF attachments. PCA was implemented

to further extract and clarify the information from the complex ToF-SIMs data and

illustrated that EGF concentration played a crucial role in the binding efficiency of the

EGF to the nanoparticles. In all cases, the combination of ToF-SIMS and PCA proved

0

200

400

600

800

1000

1200

1400

0 5 10 15 20 25

MFI

Particle concentration

HMSNs-FITC

EGF-HMSNs-FITC

Page 130: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

112

the effectiveness of the method in classifying relative surface concentration of EGF

attachments.

In vitro experiments present direct evidences that HMSN-EGF can efficiently and

selectively enter the EGFR positive colorectal cancer cells (SW480). However, the

targeting function of the current drug delivery system had no effect on EGFR negative

cancer cell lines SW620. More importantly, cellular uptake results on SW 480 and SW

620 cells suggested that the EGF-EGFR endocytosis pathway accounted for the improved

internalisation of HMSN-EGF. Therefore, by utilizing the EGF-EGFR interactions, the

specific delivery of drugs in colorectal cancer cells can be achieved, which could

guarantee a more precise killing of cancer cells without damaging the normal cells.

Page 131: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

113

Chapter 6 Functionalization of HMSNs for favourite 5-FU loading and siRNA

encapsulation

6.1 Introduction

Inadequate intracellular concentration of anti-cancer drug is always associated to the

failure of cancer therapy as drug concentration is a key parameter for successful treatment.

The excellent targeting function of current drug delivery system was demonstrated based

on HMSNs in last chapter, which can boost the number of nanoparticles to be internalized

by targeted cell population and thus increasing the amount of drug transported and

accumulated into cancer cells. However, even though with high surface area, high volume

and large cavity, HMSNs showed a relative low 5-FU loading capacity (~ 10%) which

may be due to the electrostatic repulsion between the drug and the carriers. In this case,

compared to those carriers with a higher drug loading capacity, more HMSNs was

required to be taken up by cancer cells to achieve an effective cure concentration for

cancer therapy. From a technical point of view, it is more preferable to deliver a smaller

quantity of HMSNs to minimise patient side effects [247]. Therefore, increasing the drug

loading capacity of HMSNs is essential for further optimizing the current drug delivery

system. The encapsulation of drugs in nanoparticles can be improved by functionalization

[248] or by optimizing the drug encapsulation conditions. Precisely controlling the

grafting of chemical groups on HMSNs would result in an enhancement in drug loading.

On the other hand, drug resistance is another issue associated with cancer therapy. Even

though a very high 5-FU loading and excellent targeting behaviour of HMSNs could be

achieved, the therapeutic effect of 5-FU is still limited due to the appearance of drug

resistance in cancer cells. To overcome this problem, siRNA encapsulation appeared to

be another potential application for HMSNs. Therefore, in this chapter, the studies focus

on the modifications of HMSNs for enhanced 5-FU loading and effective siRNA

Page 132: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

114

encapsulation. Different functionalizations of HMSNs are introduced in this chapter to

make HMSNs more versatile for particular applications and ways of improving siRNA

encapsulation of HMSNs are explored.

6.2 Enhanced 5-FU loading capacity of HMSNs by precise functionalization

6.2.1 Study of 5-FU partition coefficient

The partition coefficient (Pow) is defined as the quotient of two concentrations of a test

substance (like 5-FU) in a two-phase system, where two largely immiscible solvents (n-

octanol and water) were mixed with a fixed volume ratio. The logarithm of Pow between

n-octanol and water (log Pow) has often been used as a key parameter to describe the

lipophilicity of a molecule. In the pharmaceutic field, logP is utilized to estimate the

biological activities of the drugs, such as solubility [249], biodegradation rate, bio-

accumulation, membrane permeability [250], and drug absorption and toxicity

predictions. Therefore, the investigation of the Pow of 5-FU is essential to predict its

transport and activity, thereby providing information for the subsequent modification of

HMSNs specific for high 5-FU loading.

The shake flask method is one of the common and standard experimental procedures

adopted for logPow study. This method is applied to determine the solubility and

hydrophobicity of compounds with logP values ranging from –2 to 4. The determination

is based on the principle that hydrophobic substances soluble in the lipid phase show a

positive logPow (log Pow > 0), while negative logPow values (log Pow < 0) typifies polar

compounds soluble in the water phase.

According to OECD guideline 107 [173], an effective partition coefficient study refers to

all obtained log Pow values falling within a range of -2 ~ 4 with a offset of ± 0.3 units. As

shown in Table 6.1, the log Pow values of 5-FU with all ratios (octanol : water) are within

the range mentioned above and are consistently negative, indicating that 5-FU is

hydrophilic. Specifically, when the volume of water decreased from 26.67 mL to 13.33

mL, the 5-FU concentration in water phase increased significantly from 7.747 μg/mL to

Page 133: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

115

13.17 μg/mL. This is due to the fact that most 5-FU molecules prefer congregating in the

water phase, thus with a decrease in the water volume, a more concentrated 5-FU solution

is observed. In contrast, no obvious difference was found in 5-FU concentration in the

octanol phase. With varying octanol volume, from 26.67 mL to 13.33 mL, 5-FU

concentration remained at around 2.2 ~ 2.6 μg/mL, which can be explained by the

hydrophilic property of 5-FU.

Table 6. 1 Partition coefficient of 5-FU and relevant parameters

1:2 (O:W)

1:1 (O:W)

2:1 (O:W)

Water volume (mL) 26.67 20 13.33

Oil volume (mL) 13.33 20 26.67

C water (μg/mL)

7.747

9.925 13.17

C octanol (μg/mL) 2.279 2.380

2.629

Pow 0.2942 0.2398 0.1996

log Pow -0.5314 -0.6201 -0.6997

As shown in Figure 6.1, even though the volume of water and octanol changes, it appears

that in all ratios, around 75% of 5-FU are dissolved in the water phase while the other

~25% is found in the octanol phase. Therefore, a more hydrophilic surface of HMSNs

would increase the loading of 5-FU.

Page 134: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

116

Figure 6. 1 percentages of 5-FU dissolved in each phase with varying octanol to water ratio

77.48% 74.44% 75.82%

22.52% 25.56% 24.18%

0%

20%

40%

60%

80%

100%

120%

1:2 (O:W) 1:1 (O:W) 2:1 (O:W)

Oil

Water

Page 135: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

117

6.2.2 Preparation and characterization of functionalized HMSNs

Figure 6. 2 TEM images of HMSNs (a), HMSN-NH2 (b), HMSN-COOH (c), HMSN-CN (d) and HMSN-CH3 (e)

Page 136: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

118

Surface-functionalized HMSNs were synthesized by post grafting method where

chemical groups were introduced to the surface of as-synthesized HMSNs using

appropriate silanes. CPTES, APTES and TEMS were used to prepare HMSN-CN,

HMSN-NH2 and HMSN-CH3 while HMSN-COOH was obtain by oxidizing HMSN-CN.

In all cases, plain HMSNs without any modification was used as a control. The

morphology of modified and plain HMSNs were characterized by TEM and BET. As

shown in Figure 6.2, all modified HMSNs (Figure 6.2 b-e) had a similar morphology to

that of plain HMSNs (Figure 6.2 a) revealing that the structure of these nanoparticles can

be well preserved during all functionalization processes. In particular, the surface-

functionalized HMSNs maintained a hollow structure and a very spherical shape with a

mean diameter of ~120 nm and a shell thickness of ~10 nm. In addition, all HMSNs

displayed a highly uniform size, shell thickness and more importantly, a mono-dispersion

which is one of the main requirements for drug delivery systems.

The presence of mesoporous structure in all HMSNs was confirmed by BET (Figure 6.3).

The nitrogen adsorption/desorption experiments demonstrated that both modified and

plain HMSNs showed characteristic isotherm of mesoporous materials, a type IV

isotherm, indicating the existence of mesopores. Well-defined steps occurred in the

relative pressure of 0.35-0.55 and 0.9-1.0 were related to capillary condensation and

desorption in open mesopores and interstitial pores, respectively [189]. As seen in Figure

6.3b, narrow pore size distributions are presented in HMSNs before and after surface

modifications. The average pore size of all HMSNs was calculated using the desorption

branch of the isotherm by the typical Barrett-Joyner-Halenda (BJH) method, according

to which HMSNs, HMSN-NH2, HMSN-CN, HMSN-COOH and HMSN-CH3 have a

pore size of 2.78 nm, 2.47 nm, 2.28 nm, 2.61 nm and 2.51 nm, respectively.

Page 137: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

119

0.0 0.2 0.4 0.6 0.8 1.0

1.000.90

0.55

Vol

ume

abso

rbed

Relative pressure (P/P0)

HMSNs HMSN-CH3

HMSN-CN HMSN-COOH HMSN-NH2

0.35

(A)

0 2 4 6 8 10 12 14 16 18 20

HMSNs HMSN-CH3

HMSN-NH2

HMSN-COOH HMSN-CN

pore size (nm)

2.78nm (B)

2.51nm

2.47nm

2.61nm

2.28nm

Figure 6. 3 N2 adsorption-desorption isotherms (a) and the corresponding pore size distributions (b) of plain and functionalized HMSNs

Page 138: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

120

Table 6. 2 Structure parameters of functionalized and non-functionalized HMSNs

0 500 1000 1500 2000 2500 3000 3500 4000 4500

NH2

CH3

COOH

Wavenumber(cm-1)

HMSNs HMSN-COOH HMSN-CN HMSN-CH3

HMSN-NH2

CN

δO–Si–O

νO–Si–O

νSi–OH

Figure 6. 4 FTIR spectra of plain and functionalized HMSNs and plain HMSNs

HMSNs HMSN-CH3

HMSN-CN

HMSN-NH2

HMSN-COOH

Plain HMSNs

Pore size (nm) 2.51 2.28 2.47 2.61 2.78

Pore volume (cm3/g)

0.43 0.50 0.43 0.47 0.53

Surface area (m2/g)

434 514 413 498 820

Page 139: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

121

Structure parameters of the HMSNs with and without modifications were summarized in

Table 6.2. According to Table 6.2, not significant differences can be found in the pore

size and pore volume between all kinds of nanoparticles, proving that surface

functionalization had a limited effect on the two parameters of final products. The

maintaining of pore size and pore volume guarantees large spaces in functionalized

HMSNs for drug loading that ideally should be same as non-functionalized ones.

However, due to the coverage of functional groups, the surface area of functionalized

samples decreased when compared to plain HMSNs. On the other hand, due to being

different batches, the surface area of the plain HMSNs is slightly different from that

presented in Section 4.2.1 even though the same fabrication procedures were applied.

Compared with F127 and P123, the molecular weight of Tweens 20 and TX100 is much

smaller, which produces smaller micelles during the fabrication and therefore results in

smaller pore size of nanoparticles. As pore size, surface area and pore volume of the final

products highly depend on the kind of surfactant used for fabrication, HMSNs

synthesized with surfactants having similar properties may have similar or even the same

physicochemical properties, including the surface area. These results were consistent

with similar studies [228].

Different chemical groups of -NH2, -CN, -COOH and –CH3 were grafted onto the surface

of HMSNs separately and the successful grafting was evidenced by FTIR technique.

FTIR spectra of modified HMSNs and the control sample (HMSNs) were shown in Figure

6.4. All samples have strong bands located at around 460 (δO–Si–O), 1100 (ʋasym. Si–

O–Si), and 3300-3500 cm-1 (ʋSi–OH), confirming the presence of the SiO2 inorganic phase

[227]. The -NH2 functionalization is clearly evidenced by the bands detected at around

1500 cm-1 and 3000 cm-1, which are unambiguously attributed to the NH bending

vibration and C-N stretching vibration, respectively [227-230]. Similarly, -CN and –

COOH functionalizations were proved by the observation of relative characteristic peaks.

The peak at 2200cm-1 corresponding to cyano groups (CN) was observed in the spectrum

of HMSN-CN while HMSN-COOH was shown at a peak of 1730cm-1 relating to C=O

vibration of COOH. The bands appeared at around 3000 cm-1 in the spectrum of HMSN-

CH3 are assigned to C-H vibration of methyl groups (–CH3) confirming the existence of

–CH3 on the surface of HMSN-CH3. Results obtained by FTIR proved the effective

surface functionalization of HMSNs, thereby confirming the presentence of -NH2, -CN,

Page 140: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

122

-COOH and –CH3 on the surface of HMSN-NH2, HMSN-CN, HMSN-COOH and

HMSN-CH3, respectively.

Thermogravimetric analysis (TGA) was carried out to further confirm the effective

surface modification of functionalized HMSNs. All samples were tested under nitrogen

condition at the temperature ranging from room temperature up to 650 C. The weight

loss of samples obtained from TGA arose from the decomposition of chemical groups

thereby a difference in the weight loss between samples can indicate the relative amount

of chemical groups grafted onto particles. As seen from Figure 6.5, the total weight loss

of plain HMSNs was ~8.0% due to the dehydroxylation of silanol groups (Si-OH). In

contrast, all functionalized HMSNs showed a total weight loss of ~20% which was

mainly from thermal cracking of grafted functional groups. Therefore, the success of the

grafting has been proven by TGA results which was in accordance with the FTIR results

illustrated above and the relative amount of functional groups of modified HMSNs was

~12% (w/w).

100 200 300 400 500 60060

80

100

Temperature (0C)

Wei

ght (

%)

HMSNs HMSN-CN HMSN-NH2

HMSN-CH3

HMSN-COOH

Figure 6. 5 TGA curves of functionalized HMSNs and plain HMSNs

Page 141: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

123

6.2.3 Regulate 5-FU loading capacity of HMSNs by modifications

Table 6. 3 Loading capacity of functionalized and non-functionalized HMSNs

The 5-FU loading capacity of functionalized and non-functionalized HMSNs has been

monitored using a UV spectrophotometer. An initial drug concentration of 3 mg/ml was

applied and the concentration change after 24 hours of loading was tracked. The loading

capacity of all HMSNs were calculated and listed in Table 6.3. As expected, a difference

in the amount of 5-FU loading after surface functionalization was observed, which

resulted from the different hydrophobicity and hydrophilicity of the functional groups

[228]. The plain HMSNs presented a loading percentage of 18.34% which increase to

28.89% (the best) after amine functionalization. It was reported that by using amine

functionalized mesoporous silica nanoparticles (MSNs, non-hollow structure)

specifically for 5-FU encapsulation, a 5-FU loading of 124 mg 5-FU/g MSNs was

achieved [251]. Therefore, by designing and fabricating MSNs with different structure

such as the hollow MSNs, the 5-FU loading can be significantly improved. This

enhanced loading capacity could be attributed to the hollow structure of HMSNs and the

high efficiency of functionalization.

Furthermore, -CH3 functionalization decreases the loading capacity of HMSNs to some

extent (from 18.34% to 12.73%). 5-FU is a water-soluble drug therefore it is harder to

approach HMSNs functionalized with methyl groups which is slightly more hydrophobic.

On the other hand, -CN and -COOH functionalized HMSNs showed moderate loading

ability with a loading percentage of 22.54% and 20.73%, respectively. Although HMSN-

COOH is hydrophilic which is similar to 5-FU, its loading capacity is inferior to HMSN-

NH2. This can be explained by the different surface charge of HMSN-NH2 and HMSN-

COOH. Given that 5-FU is negatively charged, it is much easier for 5-FU to get inside

HMSNs HMSN-CH3

HMSN-CN

HMSN-NH2

HMSN-COOH

Plain HMSNs

Loading capacity (%) 12.73 22.54 28.89 20.73 18.34

Loading capacity (mg 5-FU/g HMSNs)

127.3 225.4 288.9 207.3 183.4

Page 142: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

124

the positively charged HMSN-NH2 via electrostatic attractions. However, the existence

of negative charge of –COOH groups may repel 5-FU molecules and decrease the 5-FU

loading capacity of HMSN-COOH.

In summary, the loading capacity of HMSNs specific to 5-FU can be improved via

surface functionalization. The presence of –NH2 groups on the surface of nanoparticles

resulted in a highest 5-FU loading capacity. A combination of a similarity in

hydrophilicity and a presence of reverse charge of the drug gives rise to this highest

loading capacity of HMSN-NH2. This finding in addition to the higher EGF

immobilization by HMSN-NH2 illustrated in Chapter 5 will definitely render the current

drug delivery system more versatile for various applications.

6.3 Control of 5-FU loading and release

6.3.1 The effect of pH on drug loading and release

The above study illustrated that HMSN-NH2 can significantly increase the amount of 5-

FU encapsulated into nanoparticles due to the similar hydrophilicity to drug and the

existence of positive charge provided by –NH2 groups. Therefore the control of the

intensity of electrostatic force between nanoparticles and 5-FU may further regulate 5-

FU loading capacity of particles.

In this regard, the loading of 5-FU by HMSN-NH2 was carried out in 4 different pHs, 4.0,

5.5, 7.4 and 8.5. As shown in Figure 6.6, as pH increases, the loading of 5-FU capacity

of HMSN-NH2 increases and attains the maximum loading capacity of 27.52% at pH of

8.5. HMSN-NH2 is positively charged at a wider pH range because of the protonation of

the amine groups while under different pH values, the behaviour of 5-FU in terms of

surface charge is different [252, 253]. As shown in Figure 6.7, the deprotonation of 5-FU

can occur at N1 or N3 or both to form the monoanions AN1, AN3 and dianion under

alkaline condision. It is reported that at pH 7-10, AN1, AN3 and dianion of 5-FU coexist

as a mixture in aqueous medium [252, 253]. Therefore, at acidic pHs (4.0 and 5.5), 5-FU

becomes less deprotonated or non-deprotonated resulting in a very low surface charge of

Page 143: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

125

5-FU. Therefore, lower drug loading capacity of HMSN-NH2 (18.56% and 19.03%) was

observed at these two pHs. Nevertheless, when pH increased to 7.4 and 8.5, the

deprotonation of 5-FU enhanced thereby enhancing electrostatic attractions between

HMSN-NH2 and 5-FU. Instead of being a neutral molecule in the medium, most 5-FU

molecules transferred to monoanions and dianion, which generated negative charges and

led to more 5-FU being encapsulated into the nanoparticles.

18.56

20.92

27.03 27.52

19.0320.45

19.4218.47

0

5

10

15

20

25

30

8.57.45.5

Load

ing

capa

city

(%)

pH values

HMSN-NH2

HMSNs

4.0

A

200 300 400 5000.0

0.5

1.0

Abs

orba

nce

(a.u

.)

Wavelength (nm)

pH 8.5 pH 5.5 pH 7.4 pH 4.0

B

Figure 6. 6 5-FU loading capacity of HMSNs and HMSN-NH2 at different pH values

(a) and UV spectra of remaining 5-FU after 24 hours loading by HMSN-NH2 at

different pH values (b)

Page 144: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

126

Figure 6. 7 Two possible 5-FU anions (AN1 and AN3) and dianion

On the other hand, no obvious differences can be seen in the loading capacity of HMSNs

at various pHs. Due to the presence of negatively charged silanolate groups on the surface,

the loading capacity of HMSNs was inferior to HMSN-NH2 at all pHs. The encapsulation

of 5-FU by HMNSs might be mainly based on hydrogen bonding rather than electrostatic

force. In addition, under alkaline pHs where 5-FU possesses more negative charges, a

slightly decrease in drug loading capacity of HMSNs was found. This could be due to

the repulsive interaction between carriers and the drug.

6.3.2 The effect of particle size and shell thickness on drug loading and release

To study the effect of particle size on 5-FU loading and release behaviours, HMSNs with

different particle size (100, 200 and 300 nm) were synthesized and their morphology

were shown by SEM images (Figure 6.8). As seen from Figure 6.8, all HMSNs exhibit

very spherical shape with different diameters. In addition, the size of each kind of

AN3

1 2

3 34

5

6

AN1

Dianion

+H+

-H+

+H+

-H+ 1 3

3

1

Page 145: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

127

HMSNs is uniform and is at the target size 100, 200 and 300 nm. A mono-dispersion is

found consistently for all particles.

Figure 6. 8 SEM images of HMSNs with different particle size: 100 nm (a & d); 200 nm (b & e) and 300 nm (c & f)

20.3621.45

19.23

100 nm 200 nm 300 nm0

10

20

5-Fu

load

ing

capa

city

(%)

Particle size (nm)

A

0 300 600 900 1200 1500

0

10

20

30

40

50

C

umul

ativ

e re

leas

e (%

)

Release time (mins)

100 nm 200 nm 300 nm

B

Figure 6. 9 Loading capacities (a) and release profiles (b) at pH 5.5 of HMSNs with varying particle size

Drug loading and release experiments were performed on HMSNs with varying particle

size and their loading capacities and release behaviours are shown in Figure 6.9. HMSNs

with particle size of 100, 200 and 300 nm obtained similar 5-FU loading capacities of

~20% and cumulative release percentages of ~40% (Figure 6.9), indicating particle size

Page 146: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

128

is not an important factor of HMSNs in determining 5-FU loading and release ability. On

the other hand, three kinds of HMSNs displayed sustained release profiles with a

maximum release percentage approached at 5 hours and maintained for another 20 hours.

Similarly, HMSNs with a varying shell thickness were prepared for testing the effect of

shell thickness on drug loading and release performances. As presented in Figure 6.10,

the shell thickness of HMSNs can be changed from ~10 nm to ~30 nm while the particle

size remains the same as ~120 nm. HMSNs with different shell thickness showed

spherical appearance and hollow structure. Along with the increase in shell thickness, a

decrease in cavity size was found, which resulted from the fact that the size of the

nanoparticles used to develop cavities in HMSNs were different.

Figure 6. 10 SEM images of HMSNs with different shell thickness: 10 nm (a & d); 15

nm (b & e) and 30 nm (c & f)

The loading and release performance of HMSNs with different shell thickness were

summarized in Figure 6.11. Firstly, the drug loading capacity of HMSNs with a shell

thickness of 10, 15 and 30 nm were 18.42%, 16.93% and 12.16%, respectively. The

ability to encapsulate 5-FU decreases with an increase in shell thickness and HMSNs

with the maximum shell thickness (30 nm) shows an obvious decrease in 5-FU loading

capacity. This might be due to the reduced size of cavities in this kind of HMSNs and

therefore less space in HMSNs available for drug storage. Secondly, the release

Page 147: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

129

behaviour of HMSNs with shell thickness of 10 and 15 nm generates a similar release

behaviour with a final released drug proportion of ~ 40%. However, even though a

sustained release profile was presented for HMSNs with 30 nm shell, only ~ 15% of

loaded drug can be release from the carriers. In addition, the release rate was slower than

the other two samples, which could be due to the thicker shell that provides a longer way

for drug molecules to transport.

18.4216.93

12.16

10 nm 15 nm 30 nm0

4

8

12

16

20

5-Fu

load

ing

capa

city

(%)

Shell thickness (nm)

A

-60 0 60 120 180 240 1200 1800

0

10

20

30

40

50

Cum

ulat

ive

rele

ase

(%)

Release time (mins)

10nm 15nm 30nm

B

Figure 6. 11 Loading capacities (a) and release profiles (b) at pH 5.5 of HMSNs with

varying shell thickness

6.3.3 The effect of EGF on HMSNs’ drug loading and release behaviours

EGF attachment had been proved to be an excellent targeting ligand in chapter 5 which

enables HMSNs to precisely target cancer cells. After targeting, an effective killing of

cancer cells should be guaranteed by releasing the anti-cancer drug pre-loaded in EGF-

HMSNs. Therefore, it is important to study the drug loading and release behaviours of

EGF-HMSNs. Figure 6.12 shows the morphology of EGF-HMSNs. After EGF grafting,

EGF-HMSNs maintain the spherical and hollow structure as plain HMSNs in addition to

a good dispersion. The loading and release profiles of EGF-HMSNs and HMSNs are

given in Figure 6.13, where a decreased loading capacity is found in EGF-HMSNs. EGF

appeared on the surface of nanoparticles, which may block some pores of HMSNs and

hinder the drug to get inside the nanocarriers. Therefore, the 5-FU loading capacity of

nanocarriers decreases from 20% to 12.5% after EGF conjugation. On the other hand,

Page 148: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

130

the EGF attachments have limited effect on the final drug release percentage which was

proved by a similar cumulative release percentage found in both HMSNs and EGF-

HMSNs. Nevertheless, the release rate of EGF-HMSNs is slowed down by the EGFs. As

seen in Figure 6.13B, HMSNs can attain a maximum drug concentration in the first 5

hours of the release process while EGF-HMSNs exhibit a much slower release profile

with a highest release rate at 25 hours.

Figure 6. 12 SEM images (a & b) and TEM image (c) of HMSN-EGF

0 5 10 15 20 25

0

5

10

15

20

Load

ing

capa

city

(%)

Loading time (h)

EGF non-EGF

A

0 5 10 15 20 25

0

10

20

30

40

50

Release time (h)

Cul

mat

ive

rele

ase

(%)

non-EGF EGF

B

Figure 6. 13 Loading (a, at pH 8.0) and release (b, at pH 5.5) profiles of HMSNs with

and without EGF attachments

Page 149: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

131

6.4 SiRNA encapsulation based on HMSNs

With targeting ligand EGF on the surface, HMSN-NH2-EGF can precisely target and kill

cancer cells by releasing the anti-cancer drug into target cells. However, the cancer cells

develop drug resistance after the drugs are continuously applied.

Multidrug resistance (MDR) referring to the ability of cancer cells to adapt and survive

from chemotherapy is one of the most complex and challenging problems in cancer

treatments [14, 109, 254, 255]. Although nanopariticle based drug delivery systems are

often used to prevent MDR in cancer treatment by sidestepping drug resistance

mechanisms, the development of drug resistant genes by cancer cells ultimately leads to

unsatisfied outcomes. A novel approach to address MDR is to take advantage of the

ability of nanocarriers to deliver nucleic acids (such as siRNA) to cancer cells, which

knock down genes associated with MDR.

To overcome the 5-FU resistance, HMSNs were modified for siRNA delivery to silence

the gene expression involved in MDR. As siRNA is negatively charged, HMSNs were

pre-functionalized to be positively charged and thus offer the electrostatic force for

siRNA loading.

6.4.1 Increased surface charge of HMSNs for SiRNA loading

HMSNs-NH2 possessing abundant amine groups on the surface thereby render HMSNs

positive surface charge for siRNA loading. An additional advantage of HMSNs-NH2 is

the enhanced loading capacity specifically to 5-FU. In this regard, HMSNs-NH2 were

applied to load siRNA and could serve as carriers for co-delivery of 5-FU and siRNA.

Amino-functionalized HMSNs with an average diameter of ~100 nm were obtained via

grafting 3-aminopropyltriethoxysilane (APTES) on HMSNs (Figure 6.14a). The particles

displayed a high surface area of 680 m2/g, a pore size of 2.5 nm and a pore volume of

1.12 cm3/g. The corresponding values after amino-functionalization were 380 m2/g, 2.3

Page 150: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

132

nm and 0.78 cm3/g respectively (Table 6.4). Zeta potentials of particles before and after

functionalization are -27.1 mV and 0.936 mV respectively (Figure 6.14b). The increase

in zeta potential and decrease in surface area is due to the introduction of –NH2 groups.

-150 -100 -50 0 50 100 150

HMSNs HMSN-NH2

Zata potential (mV)In

tens

ity

Inte

nsity

-27.1 0.936

Figure 6. 14 SEM image (a) and zeta potential (b) of HMSN-NH2

Table 6. 4 Parameters of HMSNs and HMSN-NH2

Samples Pore size

(nm)

Pore volume

(cm3/g)

Surface area

(m2/g)

HMSNs

2.5

1.12

680

HMSN-NH2

2.3

0.78

380

Thereafter, various amount of HMSN-NH2 (0.5, 1, 2, 5, 10 and 20 μg) were incubated

with 0.1 μg of siRNA to obtain particles/siRNA weight ratio from 5 to 200. The

incubation was carried out in phosphate-buffered saline (PBS pH 7.4) for 2 h and

encapsulation efficiency of siRNA by HMSN-NH2 were characterized by agarose gel

electrophoresis. As shown in Figure 6.15, the loading of siRNA to HMSN-NH2 was

Page 151: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

133

negligible (black bands at the bottom of each column indicate unloaded siRNA). With the

highest ratio of 200, a black band with faded colour (decreased intensity) can been seen

in the gel showing very low amount of siRNA was loaded. This result reveals that siRNA

can be adsorbed by HMSN-NH2 but only limited siRNA was encapsulated at a weight

ratio of 200.

Figure 6. 15 siRNA encapsulation efficiency of HMSN-NH2. Various weight ratio of

particles/siRNA from 5 to 200 were applied.

The poor siRNA loading capacity could be due to the low surface charge of HMSN-NH2.

Although HMSN-NH2 presented positive charges, the potential is only 0.936 mV which

might not be high enough to adsorb siRNA. As it is expected a high siRNA loading of

HMSN-NH2 will guarantee an effective delivery of siRNA for the inhibition of drug

resistant genes expression, HMSN-NH2-1, HMSN-NH2-2 and HMSN-NH2-3 with

different surface charge were prepared and used for siRNA loading. Figure 6.16 displayed

the potential value of plain HMSNs, HMSN-NH2-1, HMSN-NH2-2 and HMSN-NH2-3;

the increased charge of HMSN-NH2-2 and HMSN-NH2-3 were achieved by increasing

the amount of APTES used in functionalization stages.

Page 152: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

134

-27.1

0.936

11.1

40.2

-30

-20

-10

0

10

20

30

40

50

HMSN-NH2-3HMSN-NH2-2HMSN-NH2-1

Zeta

-pot

entia

l (m

V)

HMSNs

Figure 6. 16 Zeta potential of HMSNs, HMSN-NH2-1, HMSN-NH2-2

and HMSN-NH2-3

The siRNA encapsulation efficiency was studied using the same method. We

hypothesized that an enhanced loading capacity would come with the increase in surface

charge of nanoparticles. Nevertheless, gel electrophoresis results revealed that loading of

siRNA by particles was still negligible (Figure 6.17 and Table 6.5). These results

suggested that the encapsulation of siRNA by HMSNs might not only associated with the

surface charge of particles. Instead, other parameters such as pore size might also

important for achieving an effective siRNA encapsulation.

Figure 6. 17 siRNA encapsulation efficiency of HMSN-NH2-2 (a) and HMSN-NH2-3

(b). Various weight ratio of particles/siRNA from 5 to 200 were applied.

Page 153: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

135

Table 6. 5 siRNA encapsulation efficiency

samples Pore size (nm) Potential (mv) SiRNA Loading

HMSNs 2.3 -27 None

HMSNs-NH2-1 2.3 0.936 limited

HMSNs-NH2-2 2.3 11.1 limited

HMSNs-NH2-3 2.3 40.2 limited

6.4.2 Pore size expansion of HMSNs for efficient SiRNA encapsulation

On the basis of poor siRNA loading capacity illustrated above, it is suggested that besides

positive charged surface, larger pores might be necessary for a high siRNA loading by

HMSNs. As the length of siRNA is about 3nm, the relative small pore size (2.3 nm) of

current used particles would significantly reduce the adsorption of siRNA. When siRNA

was mixed with nanoparticles, they can only attach to the surface or stay near the surface

of nanoparticles rather than enter the pores. Similar results were found by other

researchers when non-hollow mesoporous silica nanoparticles were employed to load

siRNA [256]. Consequently, pore expansion of HMSNs is required before siRNA

loading.

Three HMSNs (HMSNs-S, HMSNs-M and HMSN-L) with the same particle size, shell

thickness but different pore size (small, middle and large) were prepared and their

structure parameters were presented in Table 6.6. As shown in Figure 6.18, all HMSNs

Page 154: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

136

possess a large cavity inside and mesoporous shell outside, highlighting their distinctive

hollow structure. These HMSNs with very spherical shape exhibited a good dispersion.

Successful pore expansion of HMSNs were confirmed by scanning electron microscope

(SEM) (Figure 6.18 a, c & e) where middle and large pores of HMSNs-M and HMSNs-

L are clearly seen from the images. Similarly, in transmission electron microscopy (TEM)

images (Figure 6.18 b, d & f), small pores of HMSNs-S were hardly seen as white dots

indicating pores were missing in Figure 6.18 b. In contrast, white dots appeared in Figure

6.18 d, and become larger in Figure 6.18 f, revealing an increase in pore size of HMSNs.

Page 155: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

137

Figure 6. 18 SEM and TEM images of HMSNs-S (a & b), HMSNs-M (c & d) and HMSNs-L (e & f)

Page 156: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

138

The corresponding structure parameters of HMSNs-S, HMSNs-M and HMSN-L were

listed in Table 6.6. As expected, three HMSNs achieve the same particle size of ~120 nm

and shell thickness of ~10 nm, which is essential for investigating the effect of pore size

on siRNA loading capacity as keeping other parameters the same are necessary for

effective comparisons. The pore size of HMSNs-S, HMSNs-M and HMSNs-L were 2.5

nm, 4.3 nm and 20.1nm respectively. With an increase in pore size, pore volume

increased from 1.12 cm3/g to 1.45 cm3/g while the surface area decreased from 680 m2/g

to 513 m2/g [113].

Table 6. 6 Structure parameter of HMSNs-S, HMSN-M and HMSNs-L

Samples HMSNs-NH2-S HMSNs-NH2-M HMSNs-NH2-L

Pore size (nm) 2.5 4.3 20.1

Particle size (nm) 120 120 120

Shell thickness (nm)

10 10 10

Pore volume (cm3/g)

1.12 1.23 1.45

Surface area (m2/g)

680 604 513

To evaluate the capacity as siRNA carriers, these three HMSNs were aminated for the

presence of positive charges to adsorb negatively charged siRNA through electrostatic

interaction. The degree of amination of three kinds of HMSNs were evaluated by zeta-

potential and thermogravimetric analysis (TGA). The zeta potential results (Figure 6.19a)

demonstrate that three HMSNs possess a similar positive charge of around ~13 mV

revealing their close amount of amine groups. This amount of grafted amine groups was

further tracked via TGA. As shown in Figure 6.19b, the plain HMSNs have a total weight

loss of ~12.6 wt% under nitrogen condition, which is due to the dehydroxylation of

silanol groups. HMSNs-S, HMSNs-M and HMSNs-L show an increased weight loss

Page 157: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

139

proving their successful modifications. Further, a similar total weight loss percentage (~

18%) of HMSNs-S, HMSNs-M and HMSNs-L again illustrates the fact that they have a

close number of functional groups.

-27.1

13.5 12.7 13.1

-30

-20

-10

0

10

20

HMSNs

HMSNs-LHMSNs-M

Zeta

-pot

entia

l (m

V)

HMSNs-S

100 200 300 400 500 600 70060

80

100

4500C

Temperature (0C)

Wei

ght (

%)

weight loss:HMSNs 14.6%HMSNs-S 19.8%HMSNs-M 19.3%HMSNs-L 19.4%

HMSNs HMSNs-S HMSNs-M HMSNs-L

2500C

Figure 6. 19 Zeta potential (a) and TGA (b) of aminated HMSNs, HMSNs-S,

HMSNs-M and HMSNs-L

Page 158: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

140

Results from siRNA loading (Figure 6.20) demonstrate that HMSNs-M and HMSNs-L

have significantly higher siRNA loadings than HMSNs-S. When the pore size of HMSNs

was expanded form 2.5 nm to 4.3 nm, an obvious increase in siRNA loading was found.

Almost all siRNA could be encapsulated into HMSNs-M at the particle to siRNA ratio

of 100. This can be explained that the 4.3 nm pores of HMSNs-M are large enough for

the siRNA to pass through freely which enable HMSNs-M to encapsulate siRNA into the

pores as well as the hollow cavity (Figure 6.21). On the other hand, when the pore size

was further expanded to 20.1 nm, the siRNA loading of HMSNs-L remain the same as

that of HMSNs-M. This could be due to that both HMSNs-M and HMSNs-L had reached

their loading limit. In other words, the maximum loading capacities of both HMSNs-M

and HMSNs-L were similar and all their active space for siRNA loading had already been

occupied by siRNA at a weight ratio of 100. Therefore, for siRNA loading, pore size is

a significant parameter only when the pore size of the carriers is under 4.3 nm. When

HMSNs have a pore size larger than 4.3 nm, further expansion of the pore size will not

affect the siRNA loading anymore.

Page 159: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

141

Figure 6. 20 siRNA encapsulation efficiency of HMSNs-S, HMSNs-M and HMSNs-L. Various weight ratio of particles/siRNA from 5 to 200

were applied.

Page 160: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

142

Figure 6. 21 Proposed siRNA encapsulation of HMSNs with different pore sizes (on the

surface VS inside the pores and hollow cavity)

6.5 Conclusions

The presence of functional groups with different hydrophobicity and hydrophilicity on

HMSNs led to the variation of 5-FU loading. As 5-FU is more hydrophilic, HMSNs-NH2

showed the highest 5-FU loading capacity of 28.89%, while the presence of the more

hydrophobic group –CH3 on HMSNs have decreased the HMSNs’ loading capacity from

18.34% to 12.73%. In addition, due to the electrostatic attraction between nanoparticles

Page 161: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

143

and 5-FU, the presence of positive charge on HMSN-NH2 contributed to its highest

loading capacity. The loading and release behaviours of HMSNs can also be well

controlled via adjusting the structure parameters of HMSNs: particle size, pore size and

shell thickness. For siRNA encapsulation, pore size, rather than the intensity of positive

charge existing on HMSNs, turned out to be a crucial parameter in determining the

siRNA encapsulation efficiency of HMSNs. Almost all siRNA can be encapsulated into

HMSNs with a pore size of 4.3 nm at the particle to siRNA ratio of 100.

Page 162: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

144

Chapter 7 Conclusions and future works

7.1 Overview of current work

Nanoparticle based drug delivery systems have the potential to revolutionize cancer

therapy, because nanoparticles maintain a prolonged effective drug concentration which

allows cancer cells to be exposed sustainedly to the drug. Owning to their small size and

the characteristic features of tumour biology, like the leaky vasculature and the impair

lymphatic system, nanoparticles can reach certain solid tumours and release biologically

active cargos into the tumour cells. Several therapeutic nanocarriers including carbon

nanotubes, dendrimers, polymeric nanoparticles and inorganic nanoparticles, have been

reviewed in Chapter 2 and it was found that HMSNs possess the ability to maintain

optimum therapeutic efficacy with nontoxicity due to their large surface area, high

volume for drug loading and excellent biocompatibility. However, the non-ionic

surfactant templated HMSNs often have a broad size distribution or a defective

mesoporous structure because of the difficulties involved in controlling the formation

and organization of micelles for the growth of silica framework.

In Chapter 4, a novel “Eudragit assisted” strategy has been developed to fabricate hollow

mesoporous silica nanoparticles (HMSNs) by utilising the Eudrgit nanoparticles as cores

and to assist in the self-assembly of micelle organization. The significant achievements

of this newly developed synthesis route include:

1. Uniform, dispersible hollow mesoporous silica nanoparticles with large internal

cavities were achieved with this newly developed strategy. The method using

Eudragit S100 nanoparticles as both a core template and an assistant in the self-

organization of surfactant micelles proved an excellent method for HMSNs synthesis.

The mesoporous structure, a high surface area and a small pore size of HMSNs were

suggested by BET results, where IV type N2 adsorption- desorption isotherm, a

surface area of 760m2/g and a pore size of 2.5 nm were presented.

Page 163: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

145

2. Eudragit S-100 is a widely used pharmaceutical polymer for preparing enteric

formulations [257, 258] and as carriers for pH dependent drug delivery systems [192-

194] due to its pH dependent solubility [190, 191]. The utilisation of Eudragit S100

nanoparticles as a core template for the fabrication of HMSNs affords the advantage

of non-toxicity [259], potential sites (carboxyl groups) for hydrogen bonding [260],

and most importantly easier core removal. Eudragit core can be easily removed by

soaking the as-synthesized HMSNs (HMSNs with Eudragit cores) in base solution

or in Acetone under room temperature. This solvent extraction process to clean the

Eudragit cores is facile and gentle, which remarkably preserved HMSNs form further

damage or aggregation caused by calcination, where HMSNs subjected to a very high

temperate. To our knowledge, this is the first time to report the use of Eudragit S100

nanoparticle as a core template for HMSNs synthesis.

3. The synthesized HMSNs were found to exhibit great sustained release of anti-cancer

drug 5-FU and nontoxic to colorectal cancer cells. The successful synthesis of

HMSNs in addition to their excellent biocompatibility demonstrate the potential of

HMSNs as effective drug carriers for building versatile cancer-targeted drug delivery

systems.

Many chemotherapeutic drugs for cancer treatments with low specificity often kill

healthy cells as well and thereby causing immense toxicity to the patient. Therefore,

targeting functions are essential for drug delivery systems as they have the potential to

precisely target and kill cancerous cells while leaving normal cells unharmed. To

configure HMSNs for targeted drug delivery, HMSNs were bioconjugated with specific

receptor ligand EGF in Chapter 5. The construction of a targeting molecule (EGF) to the

surface of HMSNs was achieved through a new method, where the carboxyl groups of

EGF conjugated to an amine terminated mesoporous silica surface. This method proved

to be facile and effective to graft EGF on silica surface and can protect HMSNs from

damage during modification. Three principal conclusions have been reached on the

preparation of EGF-HMSNs and their bioactivity in Chapter 5:

Page 164: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

146

1. The current EGF grafting method can tune the quantity of EGF attachments on the

surface of HMSNs by changing the EGF concentration during the grafting stages,

which have not been previously reported in the open literature. A small change in

the properties of targeting ligand in terms of effective immobilization, stability,

concentration and binding efficiency can have a great influence on the subsequent

performance of nanoparticulate targeted drug delivery systems in the biological

environment. Although the utilisation of EGF for targeting delivery is well

established, a precise control of the density of EGF attachments on nanoparticles is

very limited. Therefore, the successful controlled grafting of EGF onto HMSNs by

the strategy illustrated in Chapter 5 proved the superiority of this method to

commonly used ones.

2. Characterization of EGF attachments was first carried out using ToF-SIMS. The

ToF-SIMS data revealed the successful grafting of EGF attachments. PCA was

implemented to further extract and clarify the information from the complex ToF-

SIMs data and illustrated that EGF concentration played a crucial role in the binding

efficiency of the EGF to the nanoparticles. In all cases, the combination of ToF-

SIMS and PCA proved to be very effective in classifying relative surface

concentration of EGF attachments.

3. In vitro experiments present direct evidences that HMSN-EGF can efficiently and

selectively enter the EGFR positive colorectal cancer cells (SW480) and they had no

effect on EGFR negative cancer cell line SW620. More importantly, cellular uptake

results on SW 480 cells and SW 620 cells suggested that the EGF-EGFR endocytosis

pathway accounted for the improved internalisation of HMSN-EGF. Therefore, by

utilizing the EGF-EGFR interactions, the specific delivery of drugs in colorectal

cancer cells can be achieved. This guarantees a more precisely killing of cancer cells

without killing the normal cells.

For drug delivery system, a high drug loading is always being pursued to guarantee an

adequate intracellular concentration of anti-cancer drug in tumour area. The capacity of

nanoparticles to include drugs can be improved by functionalization [248] or by

optimizing the drug encapsulation conditions. In Chapter 6, the focus was on the

Page 165: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

147

modifications of HMSNs for enhanced 5-FU loading. Different functionalization of

HMSNs is introduced in this chapter to make HMSNs more versatile for particular

application. The significant achievements include:

1. The presence of different functional groups on HMSNs led to the variation of 5-FU

loading, which resulted from the different hydrophobicity and hydrophilicity of the

functional groups. As 5-FU is more hydrophilic, the presence of –NH2 groups on the

surface of nanoparticles resulted in the highest 5-FU loading capacity of 28.89%,

while the more hydrophobic group –CH3 decreased the HMSNs’ loading capacity

from 18.34% to 12.73%. In addition, the presence of positive charge on HMSN-NH2

contributed to its highest loading capacity. This is because of the electrostatic

attraction between nanoparticles and 5-FU.

2. The loading and release behaviours of HMSNs can also be controlled via adjusting

the structure parameters of HMSNs: particle size, pore size and shell thickness.

Studies in Chapter 6 revealed that HMSNs with particle size of 100, 200 and 300 nm

achieved similar 5-FU loading capacities of ~20% and cumulative release

percentages of ~40%, indicating particle size is not an important factor of HMSNs

in determining 5-FU loading and release ability. However, the capacity of HMSNs

to encapsulate 5-FU decreased with an increase in shell thickness and HMSNs with

the maximum shell thickness (30 nm) shown an obvious decrease in 5-FU loading

and release. Similarly, pore size is another significant factor for release, with an

increase in pore size, the release rate is accelerated (data was shown in Chapter 4).

3. To overcome the 5-FU resistance, HMSNs were modified for the delivery of siRNA

to silence the genes expression involved in multidrug resistance (MDR). Rather than

the intensity of positive charge existing on HMSNs, pore size turned out to be a

crucial parameter in determining the siRNA encapsulation efficiency of HMSNs.

Poor siRNA encapsulation was observed with HMSNs having a pore size of 2.3 nm,

which was due to the fact that siRNA was only adsorbed on the surface of HMSNs

and cannot pass through the pores to get into the cavity. When the pore size of

HMSNs was expanded to >4 nm, siRNA encapsulation was boosted. Almost all

siRNA can be encapsulated into HMSNs with a pore size of 4.3 nm at the particle to

Page 166: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

148

siRNA ratio of 100. This can be explained that the 4.3 nm pores are large enough for

the siRNA to pass through freely which enables HMSNs to encapsulate siRNA into

the pores as well as the hollow cavity.

7.2 Limitations and future work

Although achievements have been made as outlined in the previous sections, there are

various limitations in the current project mainly due to the shortage of experimental and

characterization equipment. Animal experiments are highly desirable to this work even

though cell experiments were carried out for testing the biological behaviour of

investigated drug delivery system. However, the limitations do not reduce the

significance of the results which have been achieved as the cell experiments are as

important as animal ones in assessing the performances of current targeted drug delivery

system based on HMSNs.

7.2.1 Targeting efficiency of EGF-HMSNs

It is demonstrated in Chapter 5 that HMSNs bioconjugated with EGF can selectively and

efficiently target cancer cells with an overexpression of EGFR. Given that most cancer

cells overexpressed EGFR, these results in addition to the great biocompatibility of

HMSNs suggested the feasibility of future applications of HMSNs in targeted drug

delivery systems for cancer treatment. More importantly, the use of TOF-SIMs provides

opportunities to determine relative EGF concentration on HMSNs. Therefore, the amount

of EGF grafted onto HMSNs can be compared by changing EGF concentration during

modification process. Nevertheless, the subsequent targeting efficiency of EGF-HMSNs

with a varying amount of EGF attachments on the surface in cancer cells have not been

investigated in the current study. It is noteworthy that a precise control of EGF on HMSNs

could lead to an effective regulation of nanoparticles entering the cells and thereby

controlling the drug concentration in targeted cells. Therefore, a further investigation of

targeting efficiency of EGF-HMSNs with different EGF concentration on the surface is

of great importance to control drug concentration in cancer cells. It is known that the

cancer development is a multistep process where different drug concentration and drug

Page 167: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

149

exposure time were required for different stages. Thus, looking at the effect of the number

of EGF ligands on the targeting behaviour of EGF-HMSNs would help engineer a

versatile drug delivery system with controlled targeting efficiency.

7.2.2 siRNA release and knockdown efficiency

In Chapter 6, siRNA delivery has been highlighted as an essential task for HMSNs to

overcome multidrug resistance. Although an effective siRNA loading by functionalized

HMSNs was clearly illustrated in this chapter, the release profile and gene knockdown

efficiency of siRNA loaded HMSNs were not studied, where further study associated with

siRNA delivery in vitro and in vivo is necessary to fully assess the potential of HMSNs

as siRNA carriers.

Drug resistance at the tumour level is complex, often involving multiple and dynamically

acquired multidrug resistance (MDR) mechanisms as a result of the expression of drug

e ux pumps, anti-apoptotic proteins, oncogenes, and regulators of drug metabolism.

Therefore, further study, including how to enhance the cellular uptake efficiency of

siRNA loaded HMSNs and effectively silence the targeted gene in vivo, needs to be

carried out in the near future. Also, how to control the release of siRNA from HMSNs

and protect siRNA from enzymatic degradation in vitro and in vivo should be addressed

to guarantee the most optimal siRNA knocking down.

7.2.3 Preparation multi-anticancer drug loaded HMSNs

As a drug carrier, high drug-loading capacity is being pursued. In Chapter 6, the structure

parameters of HMSNs can be controlled to load as many drug molecules as possible. In

addition, the surface potential and hydrophilicity of HMSNs can also be altered for

attracting drug molecules to enter the pores by precise surface modifications. However,

all the studies mentioned above is designed specifically for only one drug, 5-FU. As

pointed out by preclinical studies, a combination of chemotherapy drugs have led to

improved anti-cancer effects and the use of combination therapy for colon cancer is well

Page 168: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

150

established worldwide. For instance, irinotecan is commonly used with 5-FU, and have

shown superior survival benefits to regimen of 5-FU only. In Australia, most frequently

used first line chemotherapy regimens involve the combination of 5-FU, oxaliplation and

leucovorin, becacizumab or irinotecan. Therefore, there exists a need to further explore

HMSNs as a useful platform for multidrug-combined therapy.

To enable the encapsulation of more than one drug in HMSNs more functionalizations

need to be introduced on the surface or in the pores of HMSNs. However, for a

nanoparticle based drug delivery system, the challenge lies in successfully combining

several functions into one system. For example, an ideal drug delivery system refers to

one that with targeting actions, sufficient therapeutic concentration of drug, nontoxicity

and adequate biological activity. In many cases, it is difficult to couple several functional

groups of sufficient concentration, since the number of attachment sites on the particle

surface is limited. Moreover, each functionalization step might negatively affect the

suspension stability of the particulate system, depending on the physicochemical

properties of the added function. Keep this in mind, further study would focus on the

structure optimisation of HMSNs rather than introducing functional groups. A novel kind

of core–shell dual-mesoporous structure with smaller mesopores in the shells and ordered

larger mesopores in the core has been constructed by Niu [261], which have exhibited the

capacity to simultaneously loaded two kinds of drugs with different molecular sizes and

different hydrophobicity. These core/shell hierarchical structures provide an inspiration

for future investigation into HMSNs. In Chapters 4 and 6, it has been demonstrated that

the pore size, shell thickness as well as particle size of HMSNs can be easily controlled,

which would open up opportunities to build up more versatile HMSNs with tailed

structure for multidrug encapsulation.

Another method proposed for the future co-delivery based on HMSNs is that HMSNs

would be divided into separate groups and each group will undergo respective specific

functionalization and drug loading. Finally, all the groups would mixed up with certain

ratio to achieve a multidrug loaded HMSNs mixture. This method takes the advantage of

high flexibility, specificity and more importantly, avoiding multi-functionalization on one

HMSNs and thereby increasing the efficiency while reducing toxicity.

Page 169: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

151

7.2.4 Animal experiences

In this study, HMSNs have been demonstrated as a vehicle for successful intracellular

delivery of anticancer drug 5-FU. The current HMSNs based targeted drug delivery

system has shown sustained release behaviour, negligible toxicity and specifically

targeting and killing of colorectal cancer cells in vitro, which suggests that this system is

promising in the application for colorectal cancer therapy. However, in vitro study is not

enough for fully evaluate the biological behaviours of a drug delivery system as results

got form cell experiments could be quite different from outcomes obtained from

experiments performed on live mammals. Therefore, further validation of this targeted

drug delivery system in vivo is essential for its clinical translation.

Specifically, even though EGF-HMSNs have been selectively taken up by colorectal

cancer cells, it cannot guarantee a high selective accumulation of these nanoparticles in

solid tumours of live animal models, because the bio-distribution of nanoparticles in vivo

can be further influenced by biological environments [262]. In addition, there are findings

[263] that provide strong evidence that nanoparticles not only actively interact with

cancer cells, but also engage in and mediate in vivo behaviours. Thus, the interplay

between biological e ects and particles will undoubtedly be an important aspect that

needs to be further investigated so as to identify the full potential of the current delivery

system.

Page 170: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

152

Bibliography

1. Walther, A., et al., Genetic prognostic and predictive markers in colorectal cancer.

Nature Reviews Cancer, 2009. 9(7): p. 489-499.

2. Kumar , K., Colon targeting drug delivery system: A review on recent approaches

International Journal of Pharmaceutics 2011. 2(1): p. 11.

3. Rajguru, V.V., et al., An overview on colonic drug delivery system. International

Journal of Pharmaceutical Sciences Review & Research, 2011. 6(2): p. 197-204.

4. Slowing, I.I., et al., Mesoporous silica nanoparticles: Structural design and

applications. Journal of Materials Chemistry, 2010. 20(37): p. 7924-7937.

5. Argyo, C., et al., Multifunctional mesoporous silica nanoparticles as a universal

platform for drug delivery. Chemistry of Materials, 2014. 26(1): p. 435-451.

6. Peer, D., et al., Nanocarriers as an emerging platform for cancer therapy. Nature

Nanotechnology, 2007. 2(12): p. 751-760.

7. Peng, X.H., et al., Targeted delivery of cisplatin to lung cancer using ScFvEGFR-

heparin- cisplatin nanoparticles. ACS Nano, 2011. 5(12): p. 9480-9493.

8. Byrne, J.D., T. Betancourt, and L. Brannon-Peppas, Active targeting schemes for

nanoparticle systems in cancer therapeutics. Advanced Drug Delivery Reviews,

2008. 60(15): p. 1615-1626.

9. Fay, F. and C.J. Scott, Antibody-targeted nanoparticles for cancer therapy.

Immunotherapy, 2011. 3(3): p. 381-394.

10. Zhang, Q., et al., Biocompatible, uniform, and redispersible mesoporous silica

nanoparticles for cancer-targeted drug delivery in vivo. Advanced Functional

Materials, 2014. 24(17): p. 2450-2461.

11. Sandoval, M.A., et al., EGFR-targeted stearoyl gemcitabine nanoparticles show

enhanced anti-tumor activity. Journal of Controlled Release, 2012. 157(2): p. 287-

296.

12. Hompes, R. and C. Cunningham, Colorectal cancer: Management. Medicine,

2011. 39(5): p. 254-258.

13. Eng, C., Toxic effects and their management: Daily clinical challenges in the

treatment of colorectal cancer. Nature Reviews Clinical Oncology, 2009. 6(4): p.

207-218.

Page 171: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

153

14. KAISER, J., Combining Targeted Drugs To Stop Resistant Tumors. Science, 2011.

331: p. 1542.

15. Longley, D.B., D.P. Harkin, and P.G. Johnston, 5-Fluorouracil: mechanisms of

action and clinical strategies. Nature Reviews Cancer, 2003. 3(5): p. 330-338.

16. Guo, Y., et al., Capecitabine plus irinotecan versus 5-FU/leucovorin plus

irinotecan in the treatment of colorectal cancer: A meta-analysis. Clinical

Colorectal Cancer, 2014. 13(2): p. 110-118.

17. Kim, G.P. and A. Grothey, Current Challenges in the Adjuvant Therapy of Colon

Cancer. Challenges in Colorectal Cancer. 2008: Blackwell Publishing Ltd. 133-

152.

18. Enhanced Anti-Cancer Effect of 5-Fluorouracil Loaded into Thermo-Responsive

Conjugated Linoleic Acid-Incorporated Poloxamer Hydrogel on Metastatic

Colon Cancer Models. Journal of Nanoscience and Nanotechnology, 2011. 11: p.

1425-1428.

19. Dev, R.K., V. Bali, and K. Pathak, Novel microbially triggered colon specific

delivery system of 5-Fluorouracil: Statistical optimization, in vitro, in vivo,

cytotoxic and stability assessment. International Journal of Pharmaceutics, 2011.

411(1-2): p.142-151.

20. Feng, S.S., L. Zhao, and J. Tang, Nanomedicine for oral chemotherapy.

Nanomedicine, 2011. 6(3): p. 407-410.

21. Krishnaiah, Y.S.R., et al., In vitro drug release studies on guar gum-based colon

targeted oral drug delivery systems of 5-fluorouracil. European Journal of

Pharmaceutical Sciences, 2002. 16(3): p. 185-192.

22. Wu, J., et al., The mechanisms of 5-FU-PLA-O-CMC-NPS-mediated inhibition of

the proliferation of colorectal cancer cell line SW480. Tumor Biology, 2014.

35(6): p. 6095-9103

23. Neuse, E.W., Synthetic polymers as drug-delivery vehicles in medicine. Metal-

Based Drugs, 2008. 2008.

24. Javadzadeh, Y., L. Musaalrezaei, and A. Nokhodchi, Liquisolid technique as a

new approach to sustain propranolol hydrochloride release from tablet matrices.

International Journal of Pharmaceutics, 2008. 362(1-2): p. 102-108.

25. Asaduzzaman, M., et al., Development of sustain release matrix tablet of

ranolazine based on methocel K4M CR: In vitro drug release and kinetic

approach. Journal of Applied Pharmaceutical Science, 2011. 1(8): p. 131-136.

Page 172: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

154

26. Taylor, H.E. and K.B. Sloan, 1-Alkylcarbonyloxymethyl prodrugs of 5-

fluorouracil (5-FU): Synthesis, physicochemical properties, and topical delivery

of 5-FU. Journal of Pharmaceutical Sciences, 1998. 87(1): p. 15-20.

27. Sloan, K.B., E.F. Sheretz, and R.G. McTiernan, The effect of 5-fluorouracil on

inhibition of epidermal DNA synthesis in vivo: A comparison of the effect of

formulations and a prodrug of 5-FU. Archives of Dermatological Research, 1990.

282(7): p. 484-486.

28. Koehler, K.C., et al., A Diels-Alder modulated approach to control and sustain

the release of dexamethasone and induce osteogenic differentiation of human

mesenchymal stem cells. Biomaterials, 2013. 34(16): p. 4150-4158.

29. Min, Y. and P.T. Hammond, Catechol-modified polyions in layer-by-layer

assembly to enhance stability and sustain release of biomolecules: A bioinspired

approach. Chemistry of Materials, 2011. 23(24): p. 5349-5357.

30. Pavani, E., S. Noman, and I.A. Syed, Liquisolid technique based sustained release

tablet of trimetazidine dihydrochloride. Drug Invention Today, 2013. 5(4): p. 302-

310.

31. Nokhodchi, A., et al., Liquisolid compacts: The effect of cosolvent and HPMC on

theophylline release. Colloids and Surfaces B: Biointerfaces, 2010. 79(1): p. 262-

269.

32. Ganesh, N.S., et al., Formulation and evaluation of sustained release lornoxicam

by liquisolid technique. International Journal of Pharmaceutical Sciences Review

and Research, 2011. 11(1): p. 53-57.

33. Elyagoby, A., N. Layas, and T.W. Wong, Colon-specific delivery of 5-

fluorouracil from zinc pectinate pellets through In Situ intracapsular

ethylcellulose-pectin plug formation. Journal of Pharmaceutical Sciences, 2013.

102(2): p. 604-616.

34. He, W., et al., Study on colon-specific pectin/ethylcellulose film-coated 5-

fluorouracil pellets in rats. International Journal of Pharmaceutics, 2008. 348(1-

2): p. 35-45.

35. Bose, A., A. Elyagoby, and T.W. Wong, Oral 5-fluorouracil colon-specific

delivery through in vivo pellet coating for colon cancer and aberrant crypt foci

treatment. International Journal of Pharmaceutics, 2014. 468(1-2): p. 178-186.

Page 173: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

155

36. Wei, H., et al., In-vitro and in-vivo studies of pectin/ethylcellulose-film-coated

pellets of 5-fluorouracil for colonic targeting. Journal of Pharmacy and

Pharmacology, 2008. 60(1): p. 35-44.

37. Kaiser, N., et al., 5-Fluorouracil in vesicular phospholipid gels for anticancer

treatment: Entrapment and release properties. International Journal of

Pharmaceutics, 2003. 256(1-2): p. 123-131.

38. Cardillo, J.A., et al., An intravitreal biodegradable sustained release naproxen

and 5-fluorouracil system for the treatment of experimental post-traumatic

proliferative vitreoretinopathy. British Journal of Ophthalmology, 2004. 88(9): p.

1201-1205.

39. Kulthe, S.S., et al., Modulated release of 5-fluorouracil from pH-sensitive and

colon targeted pellets: An industrially feasible approach. Drug Development and

Industrial Pharmacy, 2013. 39(1): p. 138-145.

40. Fan, L.F., et al., Biphasic drug release: Permeability and swelling of

pectin/ethylcellulose films, and in vitro and in vivo correlation of film-coated

pellets in dogs. Chemical and Pharmaceutical Bulletin, 2008. 56(8): p. 1118-1125.

41. Won, Y.W., et al., Self-assembled nanoparticles with dual effects of passive tumor

targeting and cancer-selective anticancer effects. Advanced Functional Materials,

2012. 22(6): p. 1199-1208.

42. Galagudza, M.M., et al., Passive targeting of ischemic myocardium with the use

of silica nanoparticles. Nanotechnologies in Russia, 2010. 5(11): p. 844-850.

43. Perše, M., et al., Protective effect of fullerenol nano particles on colon cancer

development in dimethylhydrazine rat model. Digest Journal of Nanomaterials and

Biostructures, 2011. 6(4): p. 1543-1551.

44. Klopman, G. and H. Zhu, Recent methodologies for the estimation of n-

octanol/water partition coefficients and their use in the prediction of membrane

transport properties of drugs. Mini-Reviews in Medicinal Chemistry, 2005. 5(2):

p. 127-133.

45. Smyth, E.C., F. Sclafani, and D. Cunningham, Emerging molecular targets in

oncology: Clinical potential of MEeT/hepatocyte growth-factor inhibitors.

OncoTargets and Therapy, 2014. 7: p. 1001-1014.

46. Etienne-Grimaldi, M.C., et al., Molecular patterns in deficient mismatch repair

colorectal tumours: Results from a French prospective multicentric biological

and genetic study. British Journal of Cancer, 2014. 110(11): p. 2728-2737.

Page 174: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

156

47. Roskoski, R., The ErbB/HER family of protein-tyrosine kinases and cancer.

Pharmacological Research, 2014. 79: p. 34-74.

48. Liu, J., et al., LMO1 is a novel oncogene in colorectal cancer and its

overexpression is a new predictive marker for anti-EGFR therapy. Tumor

Biology, 2014. 35(8): p. 8161-8167.

49. Saif, M.W., Colorectal cancer in review: The role of the EGFR pathway. Expert

Opinion on Investigational Drugs, 2010. 19(3): p. 357-369.

50. Sintov, A., et al., Enzymatic cleavage of disaccharide side groups in insoluble

synthetic polymers: A new method for specific delivery of drugs to the colon.

Biomaterials, 1993. 14(7): p. 483-490.

51. Ji, S.R., et al., Carbon nanotubes in cancer diagnosis and therapy. Biochimica et

Biophysica Acta - Reviews on Cancer, 2010. 1806(1): p. 29-35.

52. Gasser, G., I. Ott, and N. Metzler-Nolte, Organometallic anticancer compounds.

Journal of Medicinal Chemistry, 2011. 54(1): p. 3-25.

53. Baláž, P. and J. Sedlák, Arsenic in Cancer Treatment: Challenges for Application

of Realgar Nanoparticles (A Minireview). Toxins, 2010. 2(6): p. 1568-1581.

54. Das, S., A. Chaudhury, and K.Y. Ng, Preparation and evaluation of zinc-pectin-

chitosan composite particles for drug delivery to the colon: Role of chitosan in

modifying in vitro and in vivo drug release. International Journal of Pharmaceutics,

2011. 406(1-2): p. 11-20.

55. Saboktakin, M.R., et al., Synthesis and characterization of superparamagnetic

chitosan-dextran sulfate hydrogels as nano carriers for colon-specific drug

delivery. Carbohydrate Polymers, 2010. 81(2): p. 372-376.

56. Perera, G., J. Barthelmes, and A. Bernkop-Schnurch, Novel pectin-4-

aminothiophenole conjugate microparticles for colon-specific drug delivery.

Journal of Controlled Release, 2010. 145(3): p. 240-246.

57. Yellela, S., Guar gum-based colon-specific drug delivery systems for the

treatment of Inflammatory Bowel Diseases. Inflammatory Bowel Diseases, 2009.

15(12): p. S56-S56.

58. Shyale, S., et al., Pharmacokinetic evaluation and studies on the clinical efficacy

of guar gum-based oral drug delivery systems of albendazole and albendazole-

beta-cyclodextrin for colon-targeting in human volunteers. Drug Development

Research, 2006. 67(2): p. 154-165.

Page 175: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

157

59. Wang, M.J., et al., Optimizing Preparation of NaCS-Chitosan Complex to Form

a Potential Material for the Colon-Specific Drug Delivery System. Journal of

Applied Polymer Science, 2010. 117(5): p. 3001-3012.

60. Bigucci, F., et al., Pectin-based microspheres for colon-specific delivery of

vancomycin. Journal of Pharmacy and Pharmacology, 2009. 61(1): p. 41-46.

61. Bigucci, F., et al., Chitosan/pectin polyelectrolyte complexes: Selection of suitable

preparative conditions for colon-specific delivery of vancomycin. European

Journal of Pharmaceutical Sciences, 2008. 35(5): p. 435-441.

62. Saboktakin, M.R., et al., Synthesis and Characterization of Chitosan Hydrogels

Containing 5-Aminosalicylic Acid Nanopendents for Colon: Specific Drug

Delivery. Journal of Pharmaceutical Sciences, 2010. 99(12): p. 4955-4961.

63. Elkhodairy, K.A., N.S. Barakat, and F.K. Alanazi, Solubilization and

Amorphization of Non Steroidal Anti-Inflammatory Drug with Low Molecular

Weight Chitosan for a New Guar-Based Colon Delivery Formulation. Letters in

Drug Design & Discovery, 2011. 8(3): p. 292-301.

64. Hiorth, M., T. Skoien, and S.A. Sande, Immersion coating of pellet cores

consisting of chitosan and calcium intended for colon drug delivery. European

Journal of Pharmaceutics and Biopharmaceutics, 2010. 75(2): p. 245-253.

65. Criscione, J.M., et al., Self-assembly of pH-responsive fluorinated dendrimer-

based particulates for drug delivery and noninvasive imaging. Biomaterials, 2009.

30(23-24): p. 3946-3955.

66. Sinha, V.R. and R. Kumria, Polysaccharides in colon-specific drug delivery.

International Journal of Pharmaceutics, 2001. 224(1-2): p. 19-38.

67. Yinghuai, Z., et al., Substituted carborane-appended water-soluble single-wall

carbon nanotubes: New approach to boron neutron capture therapy drug delivery.

Journal of the American Chemical Society, 2005. 127(27): p. 9875-9880.

68. Pastorin, G., et al., Double functionalisation of carbon nanotubes for multimodal

drug delivery. Chemical Communications, 2006. 21(11): p. 1182-1184.

69. Chen, J.Y., et al., Functionalized Single-Walled Carbon Nanotubes as Rationally

Designed Vehicles for Tumor-Targeted Drug Delivery. Journal of the American

Chemical Society, 2008. 130(49): p. 16778-16785.

70. Cox, B.J., N. Thamwattana, and J.M. Hill, Mechanics of spheroidal fullerenes and

carbon nanotubes for drug and gene delivery. Quarterly Journal of Mechanics and

Applied Mathematics, 2007. 60: p. 231-253.

Page 176: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

158

71. Levi-Polyachenko, N,H., Merkel, E.J., Jones, B.T., Carroll, D.L and Setwart IV,

J.H., Nanotubes for rapid photothermal intracellular drug delivery. Molecular

Pharmaceutics, 2009. 6(4): p. 1062-1099.

72. Yang, D., et al., Hydrophilic multi-walled carbon nanotubes decorated with

magnetite nanoparticles as lymphatic targeted drug delivery vehicles. Chemical

Communications, 2009. 7(29): p. 4447-4449.

73. Ciofani, G., et al., Boron Nitride Nanotubes: A Novel Vector for Targeted

Magnetic Drug Delivery. Current Nanoscience, 2009. 5(1): p. 33-38.

74. Cai, K.Y., et al., Temperature-Responsive Controlled Drug Delivery System

Based on Titanium Nanotubes. Advanced Engineering Materials, 2010. 12(9): p.

B565-B570.

75. Bhirde, A.A., et al., Targeted killing of cancer cells in vivo and in vitro with EGF-

directed carbon nanotube-based drug delivery. ACS Nano, 2009. 3(2): p. 307-

316.

76. Liang, Y.Q., et al., Characterization of self-organized TiO2 nanotubes on Ti-4Zr-

22Nb-2Sn alloys and the application in drug delivery system. Journal of Materials

Science-Materials in Medicine, 2011. 22(3): p. 461-467.

77. Yin, J.F., et al., Molecularly imprinted nanotubes for enantioselective drug

delivery and controlled release. Chemical Communications, 2010. 46(41): p.

7688-7690.

78. Cho, S.J., et al., Silica coated titania nanotubes for drug delivery system.

Materials Letters, 2010. 64(15): p. 1664-1667.

79. Zhang, X., et al., Targeted delivery and controlled release of doxorubicin to

cancer cells using modified single wall carbon nanotubes. Biomaterials, 2009.

30(30): p. 6041-6047.

80. Simon-Deckers, A., et al., In vitro investigation of oxide nanoparticle and carbon

nanotube toxicity and intracellular accumulation in A549 human pneumocytes.

Toxicology, 2008. 253(1-3): p. 137-146.

81. Cheng, C., et al., Toxicity and imaging of multi-walled carbon nanotubes in

human macrophage cells. Biomaterials, 2009. 30(25): p. 4152-4160.

82. Kresge, C.T., et al., Ordered mesoporous molecular sieves synthesized by a

liquid-crystal template mechanism. Nature, 1992. 359(6397): p. 710-712.

Page 177: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

159

83. Choi, Y.L., et al., Controlled release using mesoporous silica nanoparticles

functionalized with 18-crown-6 derivative. Journal of Materials Chemistry, 2011.

21(22): p. 7882-7885.

84. Radu, D.R., et al., Gatekeeping Layer Effect: A Poly(lactic acid)-coated

Mesoporous Silica Nanosphere-Based Fluorescence Probe for Detection of

Amino-Containing Neurotransmitters. Journal of the American Chemical Society,

2004. 126(6): p. 1640-1641.

85. Mercier, L. and T.J. Pinnavaia, Access in mesoporous materials: Advantages of a

uniform pore structure in the design of a heavy metal ion adsorbent for

environmental remediation. Advanced Materials, 1997. 9(6): p. 500-503.

86. Corma, A., From microporous to mesoporous molecular sieve materials and their

use in catalysis. Chemical Reviews, 1997. 97(6): p. 2373-2419.

87. Lu, J., et al., Biocompatibility, biodistribution, and drug-delivery efficiency of

mesoporous silica nanoparticles for cancer therapy in animals. Small, 2010.

6(16): p. 1794-1805.

88. Vivero-Escoto, J.L., et al., Mesoporous silica nanoparticles for intracellular

controlled drug delivery. Small, 2010. 6(18): p. 1952-1967.

89. Slowing, I.I., et al., Mesoporous silica nanoparticles as controlled release drug

delivery and gene transfection carriers. Advanced Drug Delivery Reviews, 2008.

60(11): p. 1278-1288.

90. He, Q. and J. Shi, Mesoporous silica nanoparticle based nano drug delivery

systems: Synthesis, controlled drug release and delivery, pharmacokinetics and

biocompatibility. Journal of Materials Chemistry, 2011. 21(16): p. 5845-5855.

91. Liu, X., et al., Adsorption of CO2, CH4 and N2 on ordered mesoporous silica

molecular sieve. Chemical Physics Letters, 2005. 415(4-6): p. 198-201.

92. Polarz, S. and A. Kuschel, Chemistry in confining reaction fields with special

emphasis on nanoporous materials. Chemistry - A European Journal, 2008.

14(32): p. 9816-9829.

93. Zhao, Y.L., et al., PH-operated nanopistons on the surfaces of mesoporous silica

nanoparticles. Journal of the American Chemical Society, 2010. 132(37): p.

13016-13025.

94. Mamaeva, V., et al., Mesoporous Silica Nanoparticles as Drug Delivery Systems

for Targeted Inhibition of Notch Signaling in Cancer. Molecular Therapy, 2011.

19(8): p. 1538-1546

Page 178: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

160

95. Huh, S., et al., Organic Functionalization and Morphology Control of

Mesoporous Silicas via a Co-Condensation Synthesis Method. Chemistry of

Materials, 2003. 15(22): p. 4247-4256.

96. Chen, D., et al., Facile and scalable synthesis of tailored silica "nanorattle"

structures. Advanced Materials, 2009. 21(37): p. 3804-3807+3724.

97. Gao, H., W. Shi, and L.B. Freund, Mechanics of receptor-mediated endocytosis.

Proceedings of the National Academy of Sciences of the United States of America,

2005. 102(27): p. 9469-9474.

98. Li, Y., et al., Hollow spheres of mesoporous aluminosilicate with a three-

dimensional pore network and extraordinarily high hydrothermal stability. Nano

Letters, 2003. 3(5): p. 609-612.

99. Huo, Q., D.I. Margolese, and G.D. Stucky, Surfactant Control of Phases in the

Synthesis of Mesoporous Silica-Based Materials. Chemistry of Materials, 1996.

8(5): p. 1147-1160.

100. Tang, F., L. Li, and D. Chen, Mesoporous silica nanoparticles: Synthesis,

biocompatibility and drug delivery. Advanced Materials, 2012. 24(12): p. 1504-

1534.

101. Feng, Z., et al., A facile route to hollow nanospheres of mesoporous silica with

tunable size. Chemical Communications, 2008. 21(23): p. 2629-2631.

102. Kao, K.C., C.J. Tsou, and C.Y. Mou, Collapsed (kippah) hollow silica

nanoparticles. Chemical Communications, 2012. 48(28): p. 3454-3456.

103. Mandal, M. and M. Kruk, Family of single-micelle-templated organosilica hollow

nanospheres and nanotubes synthesized through adjustment of

organosilica/surfactant ratio. Chemistry of Materials, 2012. 24(1): p. 123-132.

104. Mei, X., et al., Hollow mesoporous silica nanoparticles conjugated with pH-

sensitive amphiphilic diblock polymer for controlled drug release. Microporous

and Mesoporous Materials, 2012. 152: p. 16-24.

105. Zhang, T., et al., Formation of hollow silica colloids through a spontaneous

dissolution-regrowth process. Angewandte Chemie - International Edition, 2008.

47(31): p. 5806-5811.

106. Lou, X.W., L.A. Archer, and Z. Yang, Hollow micro-/nanostructures: Synthesis

and applications. Advanced Materials, 2008. 20(21): p. 3987-4019.

Page 179: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

161

107. Zhao, W., et al., Fabrication of uniform hollow mesoporous silica spheres and

ellipsoids of tunable size through a facile hard-templating route. Journal of

Materials Chemistry, 2009. 19(18): p. 2778-2783.

108. Chen, Y., et al., Hollow/rattle-type mesoporous nanostructures by a structural

difference-based selective etching strategy. ACS Nano, 2010. 4(1): p. 529-539.

109. Gao, Y., et al., Controlled intracellular release of doxorubicin in multidrug-

resistant cancer cells by tuning the shell-pore sizes of mesoporous silica

nanoparticles. ACS Nano, 2011. 5(12): p. 9788-9798.

110. Zhang, Q., et al., Permeable silica shell through surface-protected etching. Nano

Letters, 2008. 8(9): p. 2867-2871.

111. He, Q., et al., An anti-ROS/hepatic fibrosis drug delivery system based on

salvianolic acid B loaded mesoporous silica nanoparticles. Biomaterials, 2010.

31(30): p. 7785-7796.

112. Lu, J., et al., Mesoporous silica nanoparticles as a delivery system for

hydrophobic anticancer drugs. Small, 2007. 3(8): p. 1341-1346.

113. Radu, D.R., et al., Fine-tuning the degree of organic functionalization of

mesoporous silica nanosphere materials via an interfacially designed co-

condensation method. Chemical Communications, 2005. (10): p. 1264-1266.

114. Huh, S., et al., Cooperative Catalysis by General Acid and Base Bifunctionalized

Mesoporous Silica Nanospheres. Angewandte Chemie International Edition,

2005. 44(12): p. 1826-1830.

115. Shi, S., F. Chen, and W. Cai, Biomedical applications of functionalized hollow

mesoporous silica nanoparticles: Focusing on molecular imaging. Nanomedicine,

2013. 8(12): p. 2027-2039.

116. Mei, X., et al., Facile preparation of coating fluorescent hollow mesoporous silica

nanoparticles with pH-sensitive amphiphilic diblock copolymer for controlled

drug release and cell imaging. Soft Matter, 2012. 8(19): p. 5309-5316.

117. Lin, Y.-S., et al., Synthesis of hollow silica nanospheres with a microemulsion as

the template. Chemical Communications, 2009. (24): p. 3542-3544.

118. Yasmin, T. and K. Müller, Synthesis and characterization of surface modified

SBA-15 silica materials and their application in chromatography. Journal of

Chromatography A, 2011. 1218(37): p. 6464-6475.

Page 180: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

162

119. Zhou, X., et al., Synthesis of hollow mesoporous silica nanoparticles with tunable

shell thickness and pore size using amphiphilic block copolymers as core

templates. Dalton Transactions, 2014. 43(31): p. 11834-11842.

120. Kao, K.-C. and C.-Y. Mou, Pore-expanded mesoporous silica nanoparticles with

alkanes/ethanol as pore expanding agent. Microporous and Mesoporous

Materials, 2013. 169(0): p. 7-15.

121. Jia, L., et al., Successfully tailoring the pore size of mesoporous silica

nanoparticles: Exploitation of delivery systems for poorly water-soluble drugs.

International Journal of Pharmaceutics, 2012. 439(1–2): p. 81-91.

122. Selvam, P., S.K. Bhatia, and C.G. Sonwane, Recent advances in processing and

characterization of periodic mesoporous MCM-41 silicate molecular sieves.

Industrial and Engineering Chemistry Research, 2001. 40(15): p. 3237-3261.

123. Zhang, H., et al., Opening and closing of nanocavities under cyclic loading in a

soft nanocomposite probed by real-time small-angle X-ray scattering.

Macromolecules, 2013. 46(3): p. 900-913.

124. Tse, N.M.K., et al., High-throughput preparation of hexagonally ordered

mesoporous silica and gadolinosilicate nanoparticles for use as MRI contrast

agents. ACS Combinatorial Science, 2012. 14(8): p. 443-450.

125. Wang, J., et al., Biphasic synthesis of colloidal mesoporous silica nanoparticles

using primary amine catalysts. Journal of Colloid and Interface Science, 2012.

385(1): p. 41-47.

126. Zhai, W., et al., Degradation of hollow mesoporous silica nanoparticles in human

umbilical vein endothelial cells. Journal of Biomedical Materials Research - Part

B Applied Biomaterials, 2012. 100(5): p. 1397-1403.

127. Morishige, K. and M. Tateishi, Accurate Relations between Pore Size and the

Pressure of Capillary Condensation and the Evaporation of Nitrogen in

Cylindrical Pores. Langmuir, 2006. 22(9): p. 4165-4169.

128. Ravikovitch, P.I. and A.V. Neimark, Experimental Confirmation of Different

Mechanisms of Evaporation from Ink-Bottle Type Pores:  Equilibrium, Pore

Blocking, and Cavitation. Langmuir, 2002. 18(25): p. 9830-9837.

129. Kruk, M. and M. Jaroniec, Gas Adsorption Characterization of Ordered

Organic−Inorganic Nanocomposite Materials. Chemistry of Materials, 2001.

13(10): p. 3169-3183.

Page 181: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

163

130. Shaobin, W., Ordered mesoporous materials for drug delivery. Microporous and

Mesoporous Materials, 2009. 117(1-2): p. 1-9.

131. Sarkisov, L. and P.A. Monson, Modeling of Adsorption and Desorption in Pores

of Simple Geometry Using Molecular Dynamics. Langmuir, 2001. 17(24): p.

7600-7604.

132. Ma, Z., et al., Impact of shape and pore size of mesoporous silica nanoparticles

on serum protein adsorption and RBCS hemolysis. ACS Applied Materials and

Interfaces, 2014. 6(4): p. 2431-2438.

133. Cauda, V., A. Schlossbauer, and T. Bein, Bio-degradation study of colloidal

mesoporous silica nanoparticles: Effect of surface functionalization with organo-

silanes and poly(ethylene glycol). Microporous and Mesoporous Materials, 2010.

132(1-2): p. 60-71.

134. He, Q., et al., The three-stage in vitro degradation behavior of mesoporous silica

in simulated body fluid. Microporous and Mesoporous Materials, 2010. 131(1-3):

p. 314-320.

135. He, Q., et al., In vivo biodistribution and urinary excretion of mesoporous silica

nanoparticles: Effects of particle size and PEGylation. Small, 2011. 7(2): p. 271-

280.

136. Souris, J.S., et al., Surface charge-mediated rapid hepatobiliary excretion of

mesoporous silica nanoparticles. Biomaterials, 2010. 31(21): p. 5564-5574.

137. Kwon, S., et al., Luminescent mesoporous nanoreservoirs for the effective loading

and intracellular delivery of therapeutic drugs. Acta Biomaterialia, 2014. 10(3):

p. 1431-1442.

138. Liu, G., et al., Charged polymer brushes-grafted hollow silica nanoparticles as a

novel promising material for simultaneous joint lubrication and treatment.

Journal of Physical Chemistry B, 2014. 118(18): p. 4920-4931.

139. Sun, Y., et al., Nanoassembles constructed from mesoporous silica nanoparticles

and surface-coated multilayer polyelectrolytes for controlled drug delivery.

Microporous and Mesoporous Materials, 2014. 185: p. 245-253.

140. Lu, F., et al., Iron oxide-loaded hollow mesoporous silica nanocapsules for

controlled drug release and hyperthermia. Chemical Communications, 2013.

49(97): p. 11436-11438.

Page 182: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

164

141. He, Q., et al., Rhodamine B-co-condensed spherical SBA-15 nanoparticles: Facile

co-condensation synthesis and excellent fluorescence features. Journal of

Materials Chemistry, 2009. 19(21): p. 3395-3403.

142. Zhang, K., et al., A continuous tri-phase transition effect for HIFU-mediated

intravenous drug delivery. Biomaterials, 2014. 35(22): p. 5875-5885.

143. Lu, F., et al., Size effect on cell uptake in well-suspended, uniform mesoporous

silica nanoparticles. Small, 2009. 5(12): p. 1408-1413.

144. Slowing, I., B.G. Trewyn, and V.S.Y. Lin, Effect of surface functionalization of

MCM-41-type mesoporous silica nanoparticles on the endocytosis by human

cancer cells. Journal of the American Chemical Society, 2006. 128(46): p. 14792-

14793.

145. Chung, T.H., et al., The effect of surface charge on the uptake and biological

function of mesoporous silica nanoparticles in 3T3-L1 cells and human

mesenchymal stem cells. Biomaterials, 2007. 28(19): p. 2959-2966.

146. He, Q., et al., Intracellular localization and cytotoxicity of spherical mesoporous

silica nano-and microparticles. Small, 2009. 5(23): p. 2722-2729.

147. Munagala, S., et al., Synthesis and evaluation of Strychnos alkaloids as MDR

reversal agents for cancer cell eradication. Bioorganic and Medicinal Chemistry,

2014. 22(3): p. 1148-1155.

148. Radadiya, A., et al., Synthesis and 3D-QSAR study of 1,4-dihydropyridine

derivatives as MDR cancer reverters. European Journal of Medicinal Chemistry,

2014. 74: p. 375-387.

149. Wu, W.D., et al., Cordycepin down-regulates multiple drug resistant (MDR)/HIF-

1α through regulating AMPK/mTORC1 signaling in GBC-SD gallbladder cancer

cells. International Journal of Molecular Sciences, 2014. 15(7): p. 12778-12790.

150. Liu, Z., et al., Sinomenine sensitizes multidrug-resistant colon cancer cells (Caco-

2) to doxorubicin by downregulation of MDR-1 expression. PLoS ONE, 2014.

9(6).

151. Kim, Y.J., et al., Comparison of capecitabine and 5-fluorouracil in

chemoradiotherapy for locally advanced pancreatic cancer. Radiation Oncology,

2013. 8(1): p. 160

152. Mnif, N., et al., Effects of nanoparticle treatment and compatibilizers on the

properties of (PP/EPR)/nano-CaCO3 blends. International Journal of Material

Forming, 2008. 1(SUPPL. 1): p. 639-643.

Page 183: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

165

153. Stylianopoulos, T., EPR-effect: Utilizing size-dependent nanoparticle delivery to

solid tumors. Therapeutic Delivery, 2013. 4(4): p. 421-423.

154. Kim, S.S., et al., A nanoparticle carrying the p53 gene targets tumors including

cancer stem cells, sensitizes glioblastoma to chemotherapy and improves survival.

ACS Nano, 2014. 8(6): p. 5494-5514.

155. Gao, Z., et al., Nanoparticle-based pseudo hapten for target-responsive cargo

release from a magnetic mesoporous silica nanocontainer. Chemical

Communications, 2014. 50(47): p. 6256-6258.

156. Lin, X., et al., Carbon nanoparticle-protected aptamers for highly sensitive and

selective detection of biomolecules based on nuclease-assisted target recycling

signal amplification. Chemical Communications, 2014. 50(57): p. 7646-7648.

157. Yang, C., et al., Target-induced strand release and thionine-decorated gold

nanoparticle amplification labels for sensitive electrochemical aptamer-based

sensing of small molecules. Sensors and Actuators, B: Chemical, 2014. 197: p.

149-154.

158. Chen, Z., et al., Drug resistance reversed by silencing LIM domain-containing

protein 1 expression in colorectal carcinoma. Oncology Letters, 2014. 8(2): p.

795-798.

159. Li, X., et al., Reversal of multidrug resistance of gastric cancer cells by

downregulation of CIAPIN1 with CIAPIN1 siRNA. Molecular Biology, 2008.

42(1): p. 91-97.

160. Shen, D.Y., et al., Inhibition of Wnt/β-catenin signaling downregulates P-

glycoprotein and reverses multi-drug resistance of cholangiocarcinoma. Cancer

Science, 2013. 104(10): p. 1303-1308.

161. Shi, X., et al., AQP5 silencing suppresses p38 MAPK signaling and improves drug

resistance in colon cancer cells. Tumor Biology, 2014. 35(7): p. 7035-7045

162. Xiong, X.B. and A. Lavasanifar, Traceable multifunctional micellar nanocarriers

for cancer-targeted co-delivery of MDR-1 siRNA and doxorubicin. ACS Nano,

2011. 5(6): p. 5202-5213.

163. Hua, J., D.G. Mutch, and T.J. Herzog, Stable suppression of MDR-1 gene using

siRNA expression vector to reverse drug resistance in a human uterine sarcoma

cell line. Gynecologic Oncology, 2005. 98(1): p. 31-38.

164. Logashenko, E.B., et al., Silencing of MDR 1 gene in cancer cells by siRNA.

Nucleosides, Nucleotides and Nucleic Acids, 2004. 23(6-7): p. 861-866.

Page 184: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

166

165. Shanmugam, V., et al., Oligonucleotides-assembled au nanorod-assisted cancer

photothermal ablation and combination chemotherapy with targeted dual-drug

delivery of doxorubicin and cisplatin prodrug. ACS Applied Materials and

Interfaces, 2014. 6(6): p. 4382-4393.

166. Zhou, L., et al., Combination of chemotherapy and photodynamic therapy using

graphene oxide as drug delivery system. Journal of Photochemistry and

Photobiology B: Biology, 2014. 135: p. 7-16.

167. Li, J., et al., Recent advances in delivery of drug-nucleic acid combinations for

cancer treatment. Journal of Controlled Release, 2013. 172(2): p. 589-600.

168. Li, J.M., et al., Multifunctional QD-based co-delivery of siRNA and doxorubicin

to HeLa cells for reversal of multidrug resistance and real-time tracking.

Biomaterials, 2012. 33(9): p. 2780-2790.

169. Kanazawa, T., et al., Prolongation of life in rats with malignant glioma by

intranasal siRNA/drug codelivery to the brain with cell-penetrating peptide-

modified micelles. Molecular Pharmaceutics, 2014. 11(5): p. 1471-1478.

170. Kim, C., et al., Synergistic induction of apoptosis in brain cancer cells by targeted

codelivery of siRNA and anticancer drugs. Molecular Pharmaceutics, 2011. 8(5):

p. 1955-1961.

171. Deng, Z.J., et al., Layer-by-layer nanoparticles for systemic codelivery of an

anticancer drug and siRNA for potential triple-negative breast cancer treatment.

ACS Nano, 2013. 7(11): p. 9571-9584.

172. Vollrath, A., et al., Labeled nanoparticles based on pharmaceutical EUDRAGIT®

S 100 polymers. Macromolecular Rapid Communications, 2010. 31(23): p. 2053-

2058.

173. Guideline for the testing of chemicals - partition coefficient (n-octanol/water):

shake flask method. 1995, Paris: OECD Organisation for Economic Cooperation

and Development

174. Zhang, P., T. Wu, and J.L. Kong, In situ monitoring of intracellular controlled

drug release from mesoporous silica nanoparticles coated with pH-responsive

charge-reversal polymer. ACS Applied Materials and Interfaces, 2014. 6(20): p.

17446-17453.

175. Gui, R., Y. Wang, and J. Sun, Embedding fluorescent mesoporous silica

nanoparticles into biocompatible nanogels for tumor cell imaging and

Page 185: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

167

thermo/pH-sensitive in vitro drug release. Colloids and Surfaces B: Biointerfaces,

2014. 116: p. 518-525.

176. Feng, W., et al., Polyelectrolyte multilayer functionalized mesoporous silica

nanoparticles for pH-responsive drug delivery: Layer thickness-dependent

release profiles and biocompatibility. Journal of Materials Chemistry B, 2013.

1(43): p. 5886-5898.

177. Zhao, C.X., L. Yu, and A.P.J. Middelberg, Magnetic mesoporous silica

nanoparticles end-capped with hydroxyapatite for pH-responsive drug release.

Journal of Materials Chemistry B, 2013. 1(37): p. 4828-4833.

178. Briggs, D., Surface Analysis of Polymers by XPS and Static SIMS. 1998:

Cambridge University Press.

179. Jasieniak, M., et al., Surface Analysis of Biomaterials, in Handbook of Surface

and Interface Analysis. 2009, CRC Press. p. 529-564.

180. Cotts, R.M., et al., Pulsed field gradient stimulated echo methods for improved

NMR diffusion measurements in heterogeneous systems. Journal of Magnetic

Resonance (1969), 1989. 83(2): p. 252-266.

181. Tan, B. and S.E. Rankin, Dual latex/surfactant templating of hollow spherical

silica particles with ordered mesoporous shells. Langmuir, 2005. 21(18): p. 8180-

8187.

182. Meng, H., et al., Development of pharmaceutically adapted mesoporous silica

nanoparticles platform. Journal of Physical Chemistry Letters, 2012. 3(3): p. 358-

359.

183. Patil, A., et al., Encapsulation of water insoluble drugs in mesoporous silica

nanoparticles using supercritical carbon dioxide. Journal of Nanomedicine and

Nanotechnology, 2011. 2(3): 1000111

184. Lin, Y.S., K.R. Hurley, and C.L. Haynes, Critical considerations in the

biomedical use of mesoporous silica nanoparticles. Journal of Physical Chemistry

Letters, 2012. 3(3): p. 364-374.

185. Chang, B., et al., General one-pot strategy to prepare multifunctional

nanocomposites with hydrophilic colloidal nanoparticles core/mesoporous silica

shell structure. Journal of Colloid and Interface Science, 2012. 377(1): p. 64-75.

186. Liu, J., et al., Yolk-shell hybrid materials with a periodic mesoporous organosilica

shell: Ideal nanoreactors for selective alcohol oxidation. Advanced Functional

Materials, 2012. 22(3): p. 591-599.

Page 186: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

168

187. Yokoi, T., et al., Synthesis of mesoporous silica nanospheres promoted by basic

amino acids and their catalytic application. Chemistry of Materials, 2010. 22(13):

p. 3900-3908.

188. He, Q., et al., An anticancer drug delivery system based on surfactant-templated

mesoporous silica nanoparticles. Biomaterials, 2010. 31(12): p. 3335-3346.

189. Xia, X., et al. Resonant micro-cantilever chemical sensor with one-step synthesis

of -COOH functionalized mesoporous-silica nanoparticles for detection of trace-

level organophosphorus pesticide. 2012. Taipei.

190. De Arce Velasquez, A., et al., Novel Pullulan-Eudragit® S100 blend

microparticles for oral delivery of risedronate: Formulation, in vitro evaluation

and tableting of blend microparticles. Materials Science and Engineering C, 2014.

38(1): p. 212-217.

191. Mustafin, R.I., et al., Comparative pharmacokinetic assessment of diclofenac

sodium polycomplex drug delivery systems based on Eudragit copolymers.

Pharmaceutical Chemistry Journal, 2014. 48(1): p. 1-4.

192. Fontana, M.C., et al., Controlled release of raloxifene by nanoencapsulation:

Effect on in vitro antiproliferative activity of human breast cancer cells.

International Journal of Nanomedicine, 2014. 9(1): p. 2979-2991.

193. Hsu, F.Y., D.S. Yu, and C.C. Huang, Development of pH-sensitive

pectinate/alginate microspheres for colon drug delivery. Journal of Materials

Science: Materials in Medicine, 2013. 24(2): p. 317-323.

194. Coco, R., et al., Drug delivery to inflamed colon by nanoparticles: Comparison

of different strategies. International Journal of Pharmaceutics, 2013. 440(1): p. 3-

12.

195. Rosenholm, J.M. and M. Lindén, Towards establishing structure–activity

relationships for mesoporous silica in drug delivery applications. Journal of

Controlled Release, 2008. 128(2): p. 157-164.

196. Lu, Q., D. Chen, and X. Jiao, Fabrication of mesoporous silica microtubules

through the self-assembly behavior β-cyclodextrin and triton X-100 in aqueous

solution. Chemistry of Materials, 2005. 17(16): p. 4168-4173.

197. Singh, R.N., et al., Molecular structure, heteronuclear resonance assisted

hydrogen bond analysis, chemical reactivity and first hyperpolarizability of a

novel ethyl-4-{[(2,4-dinitrophenyl)-hydrazono]-ethyl}-3,5-dimethyl-1H-pyrrole-

Page 187: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

169

2- carboxylate: A combined DFT and AIM approach. Spectrochimica Acta - Part

A: Molecular and Biomolecular Spectroscopy, 2012. 92: p. 295-304.

198. Szafran, M., et al., Molecular structures and hydrogen bonding in the 1 : 1 and

1 : 2 complexes of pyridine betaine with 2,6-dichloro-4-nitrophenol; an example

of strongly coupled hydrogen bonds, O-H···O=C-O-H···O. Journal of Molecular

Structure, 1997. 416(1-3): p. 145-160.

199. Szafran, M., et al., Conformational analysis of bis(trigonelline) hydrochloride,

perchlorate and their monohydrates by the B3LYP calculations, X-ray diffraction

and vibrational spectra. Journal of Molecular Structure, 2006. 784(1-3): p. 98-

108.

200. Barczyński, P., et al., Spectroscopic and structural investigation of 2,5-dicarboxy-

1- methylpyridinium inner salt. Spectrochimica Acta - Part A: Molecular and

Biomolecular Spectroscopy, 2014. 121: p. 586-595.

201. Barczyński, P., et al., Comparison of low-barrier hydrogen bonds in acid salts of

carboxylic acids and basic salts of betaines - FTIR study. Journal of Molecular

Structure, 1999. 484(1-3): p. 117-124.

202. He, Y., B. Zhu, and Y. Inoue, Hydrogen bonds in polymer blends. Progress in

Polymer Science (Oxford), 2004. 29(10): p. 1021-1051.

203. Gorgojo, P., et al., Exfoliated zeolite Nu-6(2) as filler for 6FDA-based

copolyimide mixed matrix membranes. Journal of Membrane Science, 2012. 411-

412: p. 146-152.

204. Wang, Y., et al. Formulation optimization for high drug loading colonic drug

delivery carrier. 2010. Yantai.

205. Moorthy, M.S., et al., Design of a Novel Mesoporous Organosilica Hybrid

Microcarrier: A pH Stimuli-responsive dual-drug-delivery vehicle for

intracellular delivery of anticancer Agents. Particle and Particle Systems

Characterization, 2013. 30(12): p. 1044-1055.

206. Tan, L., et al., Uniform double-shelled silica hollow spheres: Acid/base selective-

etching synthesis and their drug delivery application. RSC Advances, 2013. 3(16):

p. 5649-5655.

207. Chen, B., et al., Hollow mesoporous silicas as a drug solution delivery system for

insoluble drugs. Powder Technology, 2013. 240: p. 48-53.

Page 188: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

170

208. Cao, X., et al., In vitro release and in vitro-in vivo correlation for silybin

meglumine incorporated into Hollow-type mesoporous silica nanoparticles.

International Journal of Nanomedicine, 2012. 7: p. 753-762.

209. Wan, Y., Y. Shi, and D. Zhao, Designed synthesis of mesoporous solids via

nonionic-surfactant-templating approach. Chemical Communications, 2007, 7(9):

p. 897-926.

210. Khimyak, Y.Z. and J. Klinowski, Formation of mesoporous silicates using Triton

XN surfactants in the presence of concentrated mineral acids. Journal of Materials

Chemistry, 2000. 10(8): p. 1847-1855.

211. Hill, M.R., et al., Internal and external surface characterisation of templating

processes for ordered mesoporous silicas and carbons. Journal of Materials

Chemistry, 2009. 19(15): p. 2215-2225.

212. Van Erp, T.S. and J.A. Martens, A standardization for BET fitting of adsorption

isotherms. Microporous and Mesoporous Materials, 2011. 145(1-3): p. 188-193.

213. Ladavos, A.K., et al., The BET equation, the inflection points of N 2 adsorption

isotherms and the estimation of specific surface area of porous solids.

Microporous and Mesoporous Materials, 2012. 151: p. 126-133.

214. Sing, K.S.W., et al., Assessment of Mesoporosity, in Adsorption by Powders and

Porous Solids: Principles, Methodology and Applications: Second Edition. 2013,

Elsevier Inc. p. 269-302.

215. Ritger, P.L. and N.A. Peppas, A simple equation for desciption of solute release I.

Fickian and non-Fickian release from non-swellable devices in the form of slabs,

spheres, cylinders or discs. Journal of Controlled Release, 1987. 5(1): p. 23-36.

216. Ritger, P.L. and N.A. Peppas, A simple equation for description of solute release

II. Fickian and anomalous release from swellable devices. Journal of Controlled

Release, 1987. 5(1): p. 37-42.

217. Salome, A.C., C.O. Godswill, and I.O. Ikechukwu, Kinetics and mechanisms of

drug release from swellable and non swellable matrices: A review. Research

Journal of Pharmaceutical, Biological and Chemical Sciences, 2013. 4(2): p. 97-

103.

218. Kosmidis, K., et al., Analysis of Case II drug transport with radial and axial

release from cylinders. International Journal of Pharmaceutics, 2003. 254(2): p.

183-188.

Page 189: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

171

219. Bhattacharyya, S., et al., Inorganic nanoparticles in cancer therapy.

Pharmaceutical Research, 2011. 28(2): p. 237-259.

220. Seshacharyulu, P., et al., Targeting the EGFR signaling pathway in cancer

therapy. Expert Opinion on Therapeutic Targets, 2012. 16(1): p. 15-31.

221. Kralj, S., et al., Targeting EGFR-overexpressed A431 cells with EGF-labeled

silica-coated magnetic nanoparticles. Journal of Nanoparticle Research, 2013.

15(5): 1666.

222. Waddell, J.T., B. Liesenfeld, and C.D. Batich. Immobilization of EGF onto silica

nanoparticles. 2005. Boston, MA.

223. Xie, M., et al., Hybrid nanoparticles for drug delivery and bioimaging:

Mesoporous silica nanoparticles functionalized with carboxyl groups and a near-

infrared fluorescent dye. Journal of Colloid and Interface Science, 2013. 395: p.

306-314.

224. Sun, L., et al., Study on a carboxyl-activated carrier and its properties for papain

immobilization. Journal of Chemical Technology and Biotechnology. 2012. 87(8):

p. 1083-1088.

225. Kralj, S., M. Drofenik, and D. Makovec, Controlled surface functionalization of

silica-coated magnetic nanoparticles with terminal amino and carboxyl groups.

Journal of Nanoparticle Research, 2011. 13(7): p. 2829-2841.

226. Yiu, H.H.P., P.A. Wright, and N.P. Botting, Enzyme immobilisation using SBA-

15 mesoporous molecular sieves with functionalised surfaces. Journal of

Molecular Catalysis B: Enzymatic, 2001. 15(1–3): p. 81-92.

227. Blin, J.L., et al., Influence of alkyl peptidoamines on the structure of

functionalized mesoporous silica. Chemistry of Materials, 2004. 16(24): p. 5071-

5080.

228. Sevimli, F. and A. Yilmaz, Surface functionalization of SBA-15 particles for

amoxicillin delivery. Microporous and Mesoporous Materials, 2012. 158: p. 281-

291.

229. Shadjou, N. and M. Hasanzadeh, Amino functionalized mesoporous silica

decorated with iron oxide nanoparticles as a magnetically recoverable

nanoreactor for the synthesis of a new series of 2,4-diphenylpyrido[4,3-

d]pyrimidines. RSC Advances, 2014. 4(35): p. 18117-18126.

Page 190: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

172

230. Alesker, M., et al., Hybrid silica nanoparticles traceable by fluorescence and FT-

IR spectroscopy: Preparation, characterization and preliminary biological

studies. Journal of Materials Chemistry, 2011. 21(29): p. 10883-10893.

231. Nedd, S., et al., Using a reactive force field to correlate mobilities obtained from

solid-state 13C NMR on mesoporous silica nanoparticle systems. Journal of

Physical Chemistry C, 2011. 115(33): p. 16333-16339.

232. Vasil'ev, S.G., et al., A Solid-State NMR Investigation of MQ Silicone Copolymers.

Applied Magnetic Resonance, 2013. 44(9): p. 1015-1025.

233. Moreau, J.J.E., et al., Lamellar bridged silsesquioxanes: Self-assembly through a

combination of hydrogen bonding and hydrophobic interactions. Chemistry - A

European Journal, 2005. 11(5): p. 1527-1537.

234. Hu, L., et al., Preparation and characterization of tungsten carbide confined in

the channels of SBA-15 mesoporous silica. Journal of Physical Chemistry B, 2007.

111(14): p. 3599-3608.

235. Huang, J., et al., Multifunctional mesoporous silica supported palladium

nanoparticles as efficient and reusable catalyst for water-medium Ullmann

reaction. New Journal of Chemistry, 2012. 36(6): p. 1378-1384.

236. Shevtsov, M.A., et al., Superparamagnetic iron oxide nanoparticles conjugated

with epidermal growth factor (SPION-EGF) for targeting brain tumors.

International Journal of Nanomedicine, 2014. 9(1): p. 273-287.

237. Matsumoto, R., et al., Targeting of EGF-displayed protein nanoparticles with

anticancer drugs. Journal of Biomedical Materials Research - Part B Applied

Biomaterials, 2014. 102(8): p. 1792-1798.

238. Chen, C., et al., EGF-functionalized single-walled carbon nanotubes for targeting

delivery of etoposide. Nanotechnology, 2012. 23(4). p. 045104.

239. Flavel, B.S., et al., Grafting of poly(ethylene glycol) on click chemistry modified

Si(100) surfaces. Langmuir, 2013. 29(26): p. 8355-8362.

240. Wagner, M.S., T.A. Horbett, and D.G. Castner, Characterization of the structure

of binary and ternary adsorbed protein films using electron spectroscopy for

chemical analysis, time-of-flight secondary ion mass spectrometry, and

radiolabeling. Langmuir, 2003. 19(5): p. 1708-1715.

241. Tidwell, C.D., et al., Static time-of-flight secondary ion mass spectrometry and x-

ray photoelectron spectroscopy characterization of adsorbed albumin and

fibronectin films. Surface and Interface Analysis, 2001. 31(8): p. 724-733.

Page 191: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

173

242. Henry, M. and P. Bertrand, Surface composition of insulin and albumin adsorbed

on polymer substrates as revealed by multivariate analysis of ToF-SIMS data.

Surface and Interface Analysis, 2009. 41(2): p. 105-113.

243. Wagner, M.S. and D.G. Castner, Characterization of Adsorbed Protein Films by

Time-of-Flight Secondary Ion Mass Spectrometry with Principal Component

Analysis. Langmuir, 2001. 17(15): p. 4649-4660.

244. Henry, M., C. Dupont-Gillain, and P. Bertrand, Conformation Change of Albumin

Adsorbed on Polycarbonate Membranes as Revealed by ToF-SIMS. Langmuir,

2003. 19(15): p. 6271-6276.

245. Lhoest, J.B., et al., Fibronectin adsorption, conformation, and orientation on

polystyrene substrates studied by radiolabeling, XPS, and ToF SIMS. Journal of

Biomedical Materials Research, 1998. 41(1): p. 95-103.

246. Ratner, B.D., et al., Biomaterials Science - An Introduction to Materials in

Medicine (2nd Edition). Elsevier.

247. Sardan, M., et al., Noncovalent functionalization of mesoporous silica

nanoparticles with amphiphilic peptides. Journal of Materials Chemistry B, 2014.

2(15): p. 2168-2174.

248. Kamarudin, N.H.N., et al., Role of 3-aminopropyltriethoxysilane in the

preparation of mesoporous silica nanoparticles for ibuprofen delivery: Effect on

physicochemical properties. Microporous and Mesoporous Materials, 2013. 180:

p. 235-241.

249. Noble, A., Partition coefficients (n-octanol-water) for pesticides. Journal of

Chromatography, 1993. 642(1-2): p. 3-14.

250. Sato, S., et al., Enhancing effect of N-dodecyl-2-pyrrolidone on the percutaneous

absorption of 5-fluorouracil derivatives. Chemical and Pharmaceutical Bulletin,

1998. 46(5): p. 831-836.

251. Tomoiaga, A.M., et al., Investigations on nanoconfinement of low-molecular

antineoplastic agents into biocompatible magnetic matrices for drug targeting.

Colloids and Surfaces B: Biointerfaces, 2013. 111: p. 52-59.

252. Wierzchowski, K.L., E. Litonska, and D. Shugar, Infrared and ultraviolet studies

on the tautomeric equilibria in aqueous medium between monoanionic species of

uracil, thymine, 5-fluorouracil, and other 2,4-diketopyrimidines. Journal of the

American Chemical Society, 1965. 87(20): p. 4621-4629.

Page 192: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

174

253. Markova, N., V. Enchev, and G. Ivanova, Tautomeric equilibria of 5-fluorouracil

anionic species in water. Journal of Physical Chemistry A, 2010. 114(50): p.

13154-13162.

254. Meng, H., et al., Codelivery of an optimal drug/siRNA combination using

mesoporous silica nanoparticles to overcome drug resistance in breast cancer in

vitro and in vivo. ACS Nano, 2013. 7(2): p. 994-1005.

255. Zhang, X., et al., Biofunctionalized polymer-lipid supported mesoporous silica

nanoparticles for release of chemotherapeutics in multidrug resistant cancer cells.

Biomaterials, 2014. 35(11): p. 3650-3665.

256. Na, H.K., et al., Efficient functional delivery of siRNA using mesoporous silica

nanoparticles with ultralarge pores. Small, 2012. 8(11): p. 1752-1761.

257. Sun, H., et al., Preparation and in vitro/in vivo characterization of enteric-coated

nanoparticles loaded with the antihypertensive peptide VLPVPR. International

Journal of Nanomedicine, 2014. 9(1): p. 1709-1716.

258. Xu, M., et al., Preparation and evaluation of colon adhesive pellets of 5-

aminosalicylic acid. International Journal of Pharmaceutics, 2014. 468(1-2): p.

165-171.

259. Dandekar, P., et al., Toxicological evaluation of pH-sensitive nanoparticles of

curcumin: Acute, sub-acute and genotoxicity studies. Food and Chemical

Toxicology, 2010. 48(8-9): p. 2073-2089.

260. Maghsoodi, M. and F. Sadeghpoor, Preparation and evaluation of solid

dispersions of piroxicam and Eudragit S100 by spherical crystallization technique.

Drug Development and Industrial Pharmacy, 2010. 36(8): p. 917-925.

261. Niu, D., et al., Synthesis of Core−Shell Structured Dual-Mesoporous Silica

Spheres with Tunable Pore Size and Controllable Shell Thickness. Journal of the

American Chemical Society, 2010. 132(43): p. 15144-15147.

262. Miller, L., et al., Synthesis, characterization, and biodistribution of multiple 89Zr-

labeled pore-expanded mesoporous silica nanoparticles for PET. Nanoscale,

2014. 6(9): p. 4928-4935.

263. Huang, X., et al., The shape effect of mesoporous silica nanoparticles on

biodistribution, clearance, and biocompatibility in vivo. ACS Nano, 2011. 5(7):

p. 5390-5399.

Page 193: Tailored hollow mesoporous silica nanoplatforms with biological …dro.deakin.edu.au/eserv/DU:30073697/she-tailoredhollow-2015A.pdf · Tailored hollow mesoporous silica nanoplatforms

175

Publication List

1. X. D. She, C. Z. He, Z. Peng, L. X. Kong, Molecular-Level Dispersion of Graphene into Epoxidized Natural Rubber: Morphology, Interfacial Interaction and Mechanical Reinforcement, Polymer, 2014, 55 (26), 6803-6810

2. X. D. She, L.J. Chen, L. Velleman, C. P. Li, H. J. Zhu, C. Z. He, T. Wang, S. Shigdar, W. Duan, L. X. Kong, Fabrication of High Specificity Hollow Mesoporous Silica Nanoparticles Assisted by Eudragit for Targeted Drug Delivery, Journal of Colloid and Interface Science, 2015, 445, 151-160

3. C. P. Li, J. Vongsvivut, X. D. She, Y. Z. Li, F. H. She and L. X. Kong, New insight into non-isothermal crystallization of PVA/graphene composites, Physical Chemistry Chemical Physics, 2014, 16 (40), 22145-22158

4. X. D. She, L. Velleman, L. X. Kong, New technologies and tools: Self-Assembly synthesis of hollow mesoporous silica nanoparticles for colon targeting drug delivery systems, Clinical and experimental pharmacology and physiology, 2013, 40(supplement. s1), 52-52

5. C. P. Li, M. She, X. D. She, J. Dai, L. X. Kong, Functionalization of polyvinyl alcohol hydrogels with graphene oxide for potential dye removal, Journal of Applied Polymer Science, 2014, 131(3). DOI: 10.1002/APP.39872