Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

195
Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing Capability for Bone Tissue Engineering Dreidimensionale bioaktive Glasgerüste mit therapeutischer Ionenfreisetzung zur Gewebezüchtung von Knochen Der Technischen Fakultät der Friedrich-Alexander-Universität Erlangen-Nürnberg zur Erlangung des Grades Doktor-Ingenieur vorgelegt von Herrn Dipl.-Ing. Alexander Hoppe aus Kustanaj

Transcript of Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Page 1: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Bioactive Glass Derived Scaffolds with Therapeutic Ion

Releasing Capability for Bone Tissue Engineering

Dreidimensionale bioaktive Glasgerüste mit therapeutischer

Ionenfreisetzung zur Gewebezüchtung von Knochen

Der Technischen Fakultät

der Friedrich-Alexander-Universität Erlangen-Nürnberg

zur Erlangung des Grades

Doktor-Ingenieur

vorgelegt von

Herrn Dipl.-Ing. Alexander Hoppe

aus Kustanaj

Page 2: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Als Dissertation genehmigt

von der Technischen Fakultät

der Friedrich-Alexander-Universität Erlangen-Nürnberg

Tag der mündlichen Prüfung: 14.07.2014

Vorsitzende des Promotionsorgans: Prof. Dr.-Ing. habil. Marion Merklein

Gutachter: Prof. Dr. Aldo R. Boccaccini

Prof. Dr. Uwe Gbureck

Page 3: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Acknowledgment

Acknowledgment

My greatest gratitude and appreciation goes to my supervisor Prof. Boccaccini who

gave me the opportunity to start my PhD research at the newly formed Institute of

Biomaterials. I want to thank him for his support and advices during all stages of

my PhD and also for giving me the freedom to develop my own research ideas and

framing my entire PhD project. Exceptionally, I want to thank him for supporting

and enabling numerous research fellowships and workshops abroad including stays

in Turin, Montreal, São Carlos (Brazil) or Buenos Aires which have had huge

impact not only my scientific career but also on my private life.

I want to thank Dr. Julia Will who has been supporting me from the very beginning

of my scientific career by supervising my diploma thesis whose great outcome

encouraged me to stay in academia and to pursue a PhD. She has always succeeded

to motivate me in all my research efforts and was helping with any issue during my

PhD studies.

My gratitude extends to Dr. Rainer Detsch whom I thank for introducing me to the

“world of cell biology”, for all the fruitful discussions and proof reading of parts of

my thesis. I also thank him for persistently motivating me to finish my PhD and his

unlimited trust in my skills.

I am thankful to Alina Grünewald for teaching me practical skills for the work in

the cell lab and for always being helpful no matter what issue would come up. I also

thank her for creating an extremely nice working atmosphere with many enjoyable

hours in the cell lab.

I want to thank Tobias Zehnder, Bapi Sarker and Raquel Silva and the complete

“Henkestrasse-crew” for the great time sharing the office, the labs and the lunch

breaks while exploring the infinite number of culinary options the nearby Aldi

offered us.

I would like to thank all my fellow PhD students for the grate time in and outside of

the institute.

I thank Dr. Gerhard Frank for his tireless patience and his help with the

management of the institute’s projects and administration issues.

Page 4: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Acknowledgment

I am grateful to Heinz Mahler for his help with technical and computer related

problems.

I would like to express my gratitude to all bachelor and master student I was lucky

to supervise who all contributed to my PhD project: Alexander Kent, Vincent

Bürger, Lukas Weidenbacher, Harald Unterweger, Stefan Grimm, Tobias Reichel,

Florian Ruther and Katharina Rzepka.

All collaborators who contributed to the completion of my PhD work are gratefully

acknowledged, in particular: Eva Springer for countless SEM images, Robert

Meszaros and Prof. Wondraczek for the help with glass melting, Chris Stähli and

Prof. Nazhat for helping with glass structural and degradation studies, Stefan

Romeis and Jochen Schmidt for ICP measurements and fruitful discussion, Jonathan

Lao and Prof. Jallot for micro-PIXE-RBS measurements, Juan Catallini and Prof. V.

Mourino for capillary electrophoreses measurements, Prof. Janackovic and Bojan

Jokic for the collaboration on bioactive glass.

I want to express my deepest gratitude to my parents and my brother for their

support during the last years who never gave up their faith that I would finally

graduate.

Page 5: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Abstract

Abstract

Regarding the need for highly vascularised engineered bone constructs to regenerate

bone defects, loading biomaterials with angiogenic therapeutics, like metallic ions,

has emerged as promising approach to develop novel biomaterials for regenerative

medicine.

In the present work, bioactive silicate glasses based on 45S5 and 1393 compositions

containing the well-known angiogenic ions Cu and Co were fabricated and used for

templating 3D scaffolds by foam replica technique.

It was shown that bioactive glasses can be used as carrier for therapeutic metal ions

with controlled release kinetics tuneable via the glass composition. The degradation

and bioactivity studies revealed that incorporation of metallic ions does not impair

the bioactivity and that traces of metallic ions are incorporated in the calcium

phosphate layer formed on the scaffolds surface in contact with simulated body

fluid. Cu and Co ions in therapeutic range were released and cell culture assays

confirmed high compatibility of 45S5-Cu and 1393-Co derived particulate glasses

and corresponding scaffolds with MG-63 osteoblast-like cells, human bone marrow

derived stem cells (hBMSCs) as well as human dermal micro vascular endothelial

cells (hDMECs). Cellular response, however, was shown to be dependent on the

concentration of the metallic dopant. Furthermore, Cu ions were shown to stimulate

angiogenesis in vitro by enhanced VEGF (vascular-endothelial growth factor)

expression in hBMSCs likely due to the activation of the HIF-1 transcriptional

factor. In a co-culture study of hDMECs and hMSCs 1 wt% Cu containing 45S5

scaffolds also enhanced the expression of endothelial cell specific markers vWF and

VEGFR and stimulated the hDMECs towards formation of prevascular tube-like

structure indicating the overall angiogenic potential of Cu. Altogether, Cu or Co

doped BG scaffolds represent a new family of highly promising materials for

applications in bone regeneration.

Page 6: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Kurzzusammenfassung

Kurzzusammenfassung

Die Vaskularisierung von gezüchtetem Knochengewebe ist von entscheidender

Bedeutung für die klinische Applikation dieser Gewebekonstrukte und stellt hohe

Anforderungen an das als Trägerstruktur (Scaffold) eingesetzte Biomaterial.

Ein vielversprechender Ansatz ist dabei das Dotieren von anorganischen

Biomaterialien mit Spurenelementen mit therapeutischer, angiogener Wirkung um

die Vaskularisierung von Scaffolds zu stimulieren.

In der vorliegenden Arbeit wurden ausgehend von zwei bekannten

Zusammensetzung, 45S5 und 1393, Cu and Co dotierte bioaktive Gläser (BG)

synthetisiert und anschließend zur Herstellung von 3D Poröser Scaffolds verwendet.

Es konnte gezeigt werden, dass bioaktive Glas-Scaffolds als Trägerstrukturen zur

kontrollierten Freisetzung von Metallionen geeignet sind und die

Freisetzungskinetik über die Glaszusammensetzung gesteuert werden kann.

Bioaktivitäts- und Degradationsuntersuchungen in simulierter Körperflüssigkeit

ergaben, dass die dotierten Elemente keinen negativen Einfluss auf die Bioaktivität

haben und Cu bzw. Co in der Hydroxylapatitschicht auf der BG Oberfläche

substituiert sind. Außerdem konnten Cu und Co in einem als therapeutisch wirksam

geltenden Bereich freigesetzt werden, wobei Zelluntersuchungen hohe

Biokompatibilität der Cu und Co dotierten Scaffolds mit osteoblastähnlichen Zellen,

mesenchymalen Stammzellen (MSCs) sowie Endothelzellen (ECs) bestätigten.

Darüber hinaus, wurde deutlich, dass Cu2+

Ionen die Expression des angiogenen

Signalmoleküls VEGF in humanen Stammzellen signifikant erhöhen und in einer

Co-Kultur aus MSCs und ECs die Ausbildung tubulärer prevaskulärer Strukturen

durch ECs stimuliert. Dies bestätigt das angiogene Potential von Cu dotiertem BG.

Insgesamt stellen Cu and Co dotierte bioaktive Gläser aufgrund ihrer hohen

Bioaktivität und stimulierenden Wirkung auf relevante Zelltypen vielversprechende

Materialen für den Einsatz in der regenerativen Medizin dar.

Page 7: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Contents i

Contents

1 Introduction ..................................................................................................... 11

1.1 Motivation: General need for bone regeneration therapies ....................... 11

1.2 Aim of the work ......................................................................................... 13

2 Fundamentals ................................................................................................... 14

2.1 General aspects of bone tissue engineering ............................................... 14

2.1.1 Human bone ....................................................................................... 14

2.1.2 Concept of (bone) tissue engineering (TE) ........................................ 18

2.2 Biocompatibility and Biomaterials ............................................................ 23

2.3 Bioactive silicate glasses ........................................................................... 24

2.3.1 Glass production ................................................................................. 24

Sol-gel process ................................................................................ 24

Melt derived bioactive glasses ........................................................ 25

2.3.2 Structure and general properties of (silicate) bioactive glasses ......... 26

2.3.3 Acellular in vitro bioactivity .............................................................. 28

2.3.4 Cellular response to bioactive glasses (BG) ....................................... 33

2.4 Bioactive glass derived scaffolds (State of the art) .................................... 37

2.5 Metal ions and bioinorganics in silicate glasses ........................................ 41

2.5.1 Role of bioinorganics in bone metabolism ......................................... 41

2.5.2 Biological performance of metal ion containing glasses and

glass ceramics ................................................................................................. 45

3 Materials and Methods ................................................................................... 50

3.1 Glass fabrication ........................................................................................ 50

3.1.1 Cu-containing 45S5 ............................................................................ 50

3.1.2 Co containing 13-93 ........................................................................... 50

3.2 Preparation of the glass powder ................................................................. 51

3.3 Characterization techniques ....................................................................... 51

3.3.1 Scanning electron microscopy (SEM) / Energy dispersive X-ray

spectroscopy (EDS)......................................................................................... 51

3.3.2 Fourier transform infrared spectroscopy (FT-IR) ............................... 52

3.3.3 Raman spectroscopy ........................................................................... 52

3.3.4 Inductively coupled plasma atomic emission spectroscopy

(ICP-OES) ....................................................................................................... 52

3.3.5 Micro-Ion beam .................................................................................. 53

3.3.6 XRD ................................................................................................... 54

3.3.7 X-ray microtomography (µCT) .......................................................... 54

3.4 Scaffold fabrication .................................................................................... 54

3.5 Acellular bioactivity in simulated body fluid (SBF) ................................. 55

Page 8: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Contents ii

3.6 Compressive strength ................................................................................ 55

3.7 Cell tests .................................................................................................... 55

3.7.1 Cells and culture ................................................................................ 56

Osteoblast-like cell line MG-63 ....................................................... 56 Human bone-marrow mesenchymal stem cells (hBMSCs) ............. 57 Human dermal microvascular endothelial cells (hDMECs) ............ 57

3.7.2 Analytical methods ............................................................................ 57

Cell viability and cell proliferation .................................................. 57 Analysis of osteogenic and angiogenic markers / Gene

expression and protein release ......................................................... 58 Cell morphology and cell adhesion ................................................. 59

Evaluation techniques of 2D experiments with EC ......................... 60 Statistical analysis ............................................................................ 62

3.7.3 Powder cytotoxicity ........................................................................... 62

3.7.4 Evaluation of the scaffolds ................................................................ 63

Indirect study (2D) ........................................................................... 63 Direct study (3D) ............................................................................. 63 Co-culture of hBMSCs and ECs ...................................................... 63

3.8 In vivo study .............................................................................................. 64

4 Results and Discussion.................................................................................... 66

4.1 Cu containing 45S5 bioactive glasses ....................................................... 66

4.1.1 Glass properties ................................................................................. 66

Glass structure .................................................................................. 66

Thermal properties ........................................................................... 70

Role of CuO in 45S5 glass structure ................................................ 71

4.1.2 Scaffold properties ............................................................................. 74

Macro-Structure ............................................................................... 74

Mechanical properties ...................................................................... 76 Acellular bioactivity in SBF ............................................................ 77 Degradation of 45S5 derived scaffolds and Cu release ................... 90

4.1.4 In vitro cell response .......................................................................... 94

Powder cytotoxicity ......................................................................... 94 Cell attachment on 2D pellets .......................................................... 97 In vitro cell studies with 3D scaffolds ........................................... 100

4.1.5 In vivo evaluation ............................................................................. 112

4.2 Cobalt containing 13-93 based glasses .................................................... 115

4.2.1 Glass properties ............................................................................... 115

Structure and Thermal Properties .................................................. 115

The effect of Co on 1393 glass structure ....................................... 119

4.2.2 Scaffold properties ........................................................................... 120

Macrostructure ............................................................................... 120 Mechanical properties .................................................................... 122 Acellular Bioactivity in SBF .......................................................... 124

Page 9: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Contents iii

Degradation and ion release in SBF .............................................. 131

4.2.3 In vitro cell response ........................................................................ 137

Powder cytotoxicity ....................................................................... 137 Scaffolds ........................................................................................ 139

4.3 Evaluation of the compressive strength of the 45S5 and 1393 scaffolds

in the context of bone TE .................................................................................. 145

4.4 Degradation behaviour of 45S5 and 1393 glass scaffolds and their

suitability as carrier for therapeutic metal ions ................................................. 148

5 Summary ........................................................................................................ 152

6 Conclusion and Outlook ............................................................................... 154

7 References ...................................................................................................... 160

8 Appendix ........................................................................................................ 188

Page 10: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...
Page 11: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Introduction 11

1 Introduction

1.1 Motivation: General need for bone regeneration

therapies

Bone healing is a biologically optimised process and hence the majority of small

bone defects will heal spontaneously with minimal treatment. However, in certain

clinical situations an additional bone regeneration therapy is required: for example

in order to support compromised bone healing (possibly due to impaired blood

supply or infection) after fracture or to restore bone tissue loss due to osteoporosis,

malignant tumours or osteomyolitis.1 In this context, functional disorder and defects

of bone have become a severe global health care problem with major clinical and

socioeconomic impact.2, 3

The clinical demand for engineered bone tissue has been

growing in recent years in direct relation to the increasing human population and its

aging.4 The use of autologous material for bone regeneration is still considered as

the “gold standard” by the clinicians. However, limited availability and donor

morbidity are major drawbacks of autologous grafts5 while the use of allogenic

transplants bears the risk of cell-mediated immune response and pathogen transfer.

Hence, in last decades there has been large amount of work on developing new

biomaterials and therapy approaches for bone regeneration.

Tissue engineering (TE) is one of the approaches being investigated to tackle this

problem by regenerating bone tissue using a combined cell/material therapy.6, 7

In

common TE strategies a three-dimensional structure, termed “scaffold”, fabricated

from a suitable artificial or natural material and exhibiting high porosity and pore

interconnectivity is used.8-11

Ideally, these scaffolds should not only provide a

passive structural support for bone cells, but they also should favourably affect bone

formation by stimulating osteoblastic cell proliferation and differentiation.12

Additionally, vascularisation of the engineered bone construct is essential for

successful bone healing process and hence induction of vascularisation should be

considered when developing new biomaterials and bone regeneration therapies.

Several approaches have been successfully introduced to construct porous scaffolds

with desired porosity and appropriate mechanical performance from inorganic

Page 12: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Introduction 12

materials such as bioactive ceramics and glasses as well as from biodegradable

polymers and their composites.10, 11

Due to their chemical similarity to the inorganic

phase of bone calcium phosphates (CaP), like hydroxyapatite (HAp) or α- and β-

tricalciumphosphate (TCP) have been widely used as bone engineering

grafts/scaffolds.8, 9, 13, 14

These materials are bioactive, osteoconductive and are able

to bond directly to bone.15

However, the optimal degradation rate of CaP adapted to

the bone forming rate is difficult to achieve while also fabrication of pure CaP phase

may be challenging. Bioactive glasses (BGs), in turn, are considered 3rd

generation

biomaterials16

which have the ability to stimulate specific intrinsic cell responses

resulting in osteoinductive behaviour, ability to bond to soft tissue as well as to hard

tissue and also to form a strong bonding to bone via formation of a surface

carbonated hydroxyapatite layer (HCA) when exposed to biological fluid.9, 15

More

specifically, it was observed that ionic dissolution products from silicate based BGs

(e.g. Si, Ca, P) stimulate expression of osteogenic genes resulting in enhanced bone

formation.17

Overall, these unique features make BGs highly attractive biomaterials

for bone tissue engineering applications. Another advantage of BGs is that they are

biodegradable and may be used as carrier for metallic ions and bioinorganics.

Recently, progress has been made to enhance the biological impact of BGs by

incorporating specific metallic ions in silicate (or phosphate) glasses leading to

novel biomaterials with controlled release of inorganic or metallic species with

specific biological effect depending on the field of biomedical application.18-20

As

many trace elements such as Sr, Cu, Zn or Co present in the human body are known

for their anabolic effects in bone metabolism,21-23

these approaches for enhancing

bioactivity of scaffold materials are being investigated by introducing therapeutic

ions into the scaffold material. The subsequent release of these ions after exposure

to a physiological environment is believed to favourably affect the behaviour of

human cells and to enhance the bioactivity of the scaffolds related to both

osteogenesis and angiogenesis. Since too high concentrations of metallic ions are

known to be toxic24

a controlled release of these ions from a suitable (inorganic)

carrier system is required.19

Page 13: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Introduction 13

1.2 Aim of the work

The aim of this work is, therefore, to incorporate metallic ions (Cu and Co were

selected) in a bioactive silicate glasses and to fabricate 3D glass derived scaffolds

for bone tissue engineering applications. The hypothesis was that upon degradation

of the glass derived scaffolds the metallic ions are released in a controlled manner

into the physiological environment and are suggested to enhance the osteogenic and

angiogenic potential of the material. This approach should lead to a novel group of

inorganic materials with specific functionalities to be used in regenerative medicine.

Fig. 1 gives a scheme illustrating the basic idea of this work. Materials

characterisation of the starting glass and glass derived scaffolds was performed to

determine the effect of metal ion doping on the glass properties as wells as the

macro-/ microstructure and mechanical properties of the scaffolds. The effect of

metal ion doping on the mineralisation behaviour of the scaffold was assessed in an

acellular in vitro study by immersion in simulated body fluid (SBF) and evaluation

of the HAp forming ability. In vitro cell biology studies were performed in order to

assess the biocompatibility of the glasses and corresponding scaffold and their

osteogenic and angiogenic potential. Furthermore, the scaffolds angiogenic

potential was evaluated in an in vivo model.

Fig. 1: Schematic overview of the general idea of this work using bioactive glass derived

scaffolds as carriers for metallic ions. Upon degradation of the scaffolds the released metallic

ions exhibit osteogenic, angiogenic or antibacterial properties [graphical abstract from Hoppe

at al.19

]

Page 14: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 14

2 Fundamentals

2.1 General aspects of bone tissue engineering

2.1.1 Human bone

Bone formation (Ossification)

The bone formation process is known as ossification or osteogenesis. Basically, in

the developmental biology of bone two types of bone formation are discriminated:

intramembranous and endochondral ossification. The former one describes direct

bone formation process from mesenchymal stem cell via cell differentiation and

proliferation and which forms mandible, clavicle, and craniofacial bones.25

Endochondral ossification, in turn, involves cartilage tissue as precursor which is

converted to bone through cartilage degradation and deposition of bone matrix by

osteoblast cells.25

Bone forming cells, osteoblasts, arise from pluripotent mesenchymal stem cells

whereby the differentiation process is regulated by tissue-specific transcription

factors, e.g. Runx2, Cbfa1. After osteoblastic differentiation of the progenitor cells

osteoblasts undergo proliferation, matrix maturation and mineralisation: During

ossification osteoblasts (size ~20-30 μm) secrete an amorphous matrix the so-called

osteoid predominantly consisting of collagen I protein and other proteins like

osteopontin, osteocalcin and sialoprotein.25, 26

Further, osteoblasts facilitate the

mineralisation of the osteoid forming calcified hard tissue which is likely driven by

two mechanisms: i) formation of matrix vesicles that create microenvironment with

Ca and P enrichment in a ratio being optimal for crystallisation followed by

alignment and subsequent mineralisation of hydroxyapatite (the inorganic part of

bone) or ii) vesicle independent initialisation by components of the collagen

molecules.25, 27

In any case, osteoblasts play an critical role in bone mineralisation

as they both secrete collagen and form matrix vesicles. During matrix maturation

and mineralisation osteoblasts are entrapped in the mineralised matrix and become

osteocytes (the most abundant bone cell) which exhibit dendritic-like morphology

and are connected and communicate via gap junctions.28

Page 15: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 15

Fig. 2: Schematic illustration of bone formation /bone remodelling process including osteoblast

and osteoclast differentiation from mesenchymal and hematopoietic stem cell, respectively;

matured osteoblast become osteocytes and are entrapped in the bone matrix.29

Osteoblast proliferation and differentiation is regulated by different factors

including biophysical factors, e.g. mechanical loads or matrix deformation and

biological factors. The latter ones include growth factors, e.g. bone morphogeneic

proteins (BMPs) which are secreted by osteoprogenitor cells and mature osteoblast

and active osteoblastic differentiation in vitro.28

Other growth factors include FGF

(fibroblast growth factor) IGF (insulin growth factor) and PDGF (platelet-derived

growth factor) which may regulate osteoblastic differentiation and proliferation.28

Human bone is a dynamic, highly vascularised tissue with the ability to remodel

throughout the life by regulated activity of bone-forming (osteoblasts) and bone-

resorbing cells (osteoclasts). This process of bone remodelling fulfils two functions

in the human body: adaptation of the skeleton system to metabolic demands and

changes of mechanical stress. Bone remodelling describes the addition of new bone

on the preexisistent bone surface,25

whereby it is regulated by a variety of

systematic and local regulatory agents30

including growth factors, hormones and

stress actions.26, 31

When the remodelling process is in balance it involves

osteoblasts as bone forming cells and also equal participation of bone resorbing

cells, the osteoclast.32

Osteoclasts derive from hemopoetic stem cells and the differentiation process

requires cell-cell interactions with wither osteoblasts or osteoblast progenitor cells.

Page 16: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 16

Mature osteoclasts contain typically 6-8 (or even more) nuclei and are 20-100 µm in

size.32

After migration of the osteoclasts to the resorption site on the bone surface

osteoclast form a sealing zone with a “ruffled boarder” (actin ring). Inside this zone

hydroxyapatite crystals are dissolved are resorbed through creation of acidic

environment by targeted secretion of HCl through the ruffled border into the

resorption lacuna. Then proteolytic enzymes (like metalloproteasen, MMP, e.g.

MMP-9, MMP-13 and cathopsines) degrade the collageneous bone matrix.33

Osteoclasts resorb the bone by forming corrosion pits, the so-called Howship's

Lacunae.26

Bone macro and microstructure

Regarding the microstructure human bone has two forms: woven and lamellar,

whereas the former one is considered primary and immature bone which is resorbed

and turned into lamellar bone during adolescence.25

Lamellar bone is further

structurally organized into can cancellous (spongiosa) and dense cortical bone

(compacta). Cancellous bone is composed of irregular, sinuous, convolutions of

lamellae, whereby, the microstructure of cortical bone is composed of regular

cylindrically shaped lamellae.34

Compacta makes up to 80 % of the human hard tissue in the body. The main

structural unit of the compacta is the osteon; Fig. 3 shows the microscopic structure

of spongy and cortical bone. The osteon, or the haversion system is build of

lamellae (collagen fibres arranged in planes, 3-7 µm thick) forming a cylindrical

structure of about 200- 250 µm in diameter.34

In comparison, the woven form of

bone does not show well-arranged mineralised bone and has somewhat less pattern

which can be distinguished. In the centre of each osteon there are Haversian

channels containing blood vessels and nerves. The singe lamellae consist of

collagen fibrils (1 µm) which have plate-like nanosized hydroxyapatite crystals

within the discrete spaces between the collagen fibres.26, 34, 36

Bone structure is

highly anisotropic which results in broad variations in its mechanical properties. For

compact bone and the mechanical properties depend on porosity, mineralisation

level and organisation of the solid matrix.

Page 17: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 17

Fig. 3: Microscopic structure of human lamellar bone.35

For cancellous bone the differences in mechanical properties are usually much

broader and can vary by a factor of 2-5 depending on the origin of bone. Cancellous

bone is composed of bony trabecular struts which are preferentially oriented

according to the load.26

The anisotropic structure forms according to Wolff´s law

directing the orientation of the trabeculae. The trabecular struts (50-300 µm in

diameter) form a highly porous interconnected network in which they are comprised

in cellular structures.34

Cancellous bone makes up 20 % of the human bone and

creates an highly porous environment with porosity values of 50-90 %.37

The

mechanical properties of human bone are listed in Table 1.

Page 18: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 18

Table 1: Porosity P, compressive strength σC, flexural strength σF, fracture toughness KIC tensile

strength, σT and elastic modulus of human bone.38

P[%] σC [MPa] σF [MPa] KIC[MPam1/2

] E-modulus [GPa]

Cortical bone 5-10 100-150 135-193 2-12 10-20

Cancellous bone 50-90 2-12 10-20 0.1-0.8 0.1-5

2.1.2 Concept of (bone) tissue engineering (TE)

The so called “gold standard” for regeneration of critical size bone defects is the use

of autologous bone materials, usually extracted from healthy bone tissue in the

pelvic ridge of the patient. However, this process involves a complicated two-step

surgical procedure which often results in highly problematic donor morbidity.

Hence, to overcame this: i) synthetic osteostimulating biomaterials are being

investigated for use as bone grafts and ii) bone tissue engineering approaches have

been intensively investigated during the last two decades in order to establish a

robust bone regeneration strategy.

The general aim of tissue engineering (TE) is to find ways to restore and maintain

damaged tissues and organs.39

This approach combines cell biology techniques and

materials engineering whereby cultured cells (primary or isolated cell lines) are

coaxed to grow on bioactive degradable scaffolds which provides physical support

and chemical guidance to cell differentiation and their assembly into 3D tissue. Fig.

4 gives a schematic overview of single steps involved in the tissue engineering

process in vitro. After isolation of the cells from the patient and cultivating and

multiplying cells are seeded on a scaffold which can be loaded with drugs, growth

factors, nano particles, or metallic ions inorganic in order to stimulate the cells.

After proliferation and differentiation of the cells towards the required cell type, the

tissue construct is formed which can be re-transplanted into the patient’s bone

(defect) site.

Common approaches in TE include the use of an inorganic scaffold or organic

matrix with aim to provide an optimal environment for the cells to differentiate,

proliferate and communicate.

Page 19: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 19

Fig. 4: Basic concept of tissue engineering. Inorganic scaffolds based on bioactive glass can

serve as supportive porous 3D structure for cell proliferation and differentiation. Reprinted

from.40

From the materials point of view there is a need for an ideal scaffolds with

appropriate biological and mechanical properties being able to fulfil the tasks

mentioned above. The scaffold should maintain cell attachment, migration,

differentiation and proliferation.41

In order to achieve these goals an ideal scaffold,

specifically used for bone tissue engineering, should exhibit6, 42, 43

High biocompatibility

High bioactivity, the ability to bond to hard tissue

Surface texture and chemistry that enables adsorption of biological moieties

(e.g. proteins) and cell attachment

Interconnected pore system, high total porosity with hierarchical pore

distribution enabling cell migration (micropores >100 µm) but also blood

vessel ingrowth (macro pores >400 µm)

Page 20: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 20

Degradation rates that correspond to the formation of native (bone) tissue

and degradation products which are non-toxic and can be easily excreted by

the body (e.g. by respiratory or urinary systems)

Mechanical properties adapted to the specific tissue being replaced and

sufficient to replace bone tissue in load bearing sites. For biodegradable

materials a sufficient temporal mechanical stability during matching the

growth rate of the native tissue is desired

Vascularisation strategies of engineered bone constructs

Bone formation and regeneration is considered a complex multi-component

biological system which involves two major elements: firstly, recruitment,

differentiation of progenitor cells into osteoblasts and their proliferation and

migration as described in 2.1.1; and secondly, formation of new bold vessels,

(neovascularisation) to provide newly formed tissue with nutrition and

oxygenation.44

Neovascularisation, in turn, includes two components: i)

vasculogenesis, the in situ assembly of capillaries from undifferentiated endothelial

cells and ii) angiogenesis, the sprouting of capillaries from pre-existing blood

vessels.

Vasculature of the hard tissue plays a key role in bone remodelling through acting as

reservoir and conduit for bone cells, growth factors and providing key signals

involved in bone metabolism.45

Therefore, angiogenesis not only precedes

osteogenesis but it is also required for its occurrence. This is accomplished by a

combination of factors, which include adequate oxygen tension, compression

forces, nutrients, and secretion of growth factors.46

Considering this, the successful

clinical application of engineered bone constructs highly depends on a functional

vascularised network which might result in enhanced bone metabolism and bone

formation. Several strategies for induction of vascularisation in engineering bone

tissue have been proposed as shown in Fig. 5. These include for instance:46

- Delivery of angiogenic growth factors (GFs, e.g. VEGF, bFGF):

i) direct approach by local and sustained delivery from natural, organic or

inorganic matrices; In vivo instability of growth factors and controlled

release in critical doses are main challenges for GFs delivery approaches;

Page 21: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 21

ii) indirect approach by applying other growth factors (e.g. HIF-1, BMP-2,

or inorganic molecules acting as angiogenic agents, e.g. Cu, Co which, in

turn, would stimulate secretion of angiogenic factors.20, 46

Indirect approach

enables controlled VEGF release mediated by cells which is adapted to the

physiologically, locally needed conditions avoiding “overdose”.

- Micro fabrication of blood vessels and scaffolding: inclusion of network of

vascular geometry in a 3D scaffold using rapid prototyping, foam replica

and other scaffold fabrication techniques

- Micro surgical techniques: engineered (bone) tissue construct is connected

to host vascular system to induce vascularisation. Usually, the tissue graft is

prevascularised in vivo (e.g. in an AV-loop47

or so-called flap fabrication 48

)

before transplantation to the defect site and connected to local vasculature.

High vascularisation degrees and immediate blood perfusion can be

achieved; however, technical challenges and two surgical steps with donor

morbidity are major drawbacks.

- Pre vascularisation in vitro: i) endothelial cells (ECs) and

ii) co-cultures of ECs and osteoblast cells (OCs) are cultured in natural or

synthetic scaffold to form a prevascular network (micro capillary structures)

in vitro which will connect to the host vascular tissue implantation. Host

blood vessel only need to connect the outer region of the bone construct

which will anastomose to the existing blood vessels. Co-cultures of OC and

ECs can may improve vascularisation by forming a more stable and mature

prevascularised network.

Page 22: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 22

Fig. 5: Vascularisation strategies for engineering bone constructs as described by Santos and

Reis46

including applications of angiogenic growth factors (e.g. VEGF), microfabrication of

blood vessels, micro surgical techniques (e.g. AV-loop model prevascularisation), delivery of

mature or precursor endothelial cells (ECs) and co-cultures of ECs and osteoblast cells.

Modified after Santos and Reis.46

One other possibility for stimulating vascularisation of a scaffold materials would

also include the enhancement of angiogenic properties of the scaffold material.

Hence the research has been expanded towards investigations on the angiogenic

effects of biomaterials.46

One common approach includes the loading and

subsequent release of therapeutic inorganic ions (TII) into inorganic scaffold

materials. Potential TIIs are discussed in 2.5. The general requirements, however,

for a material used as a scaffold for (bone) tissue engineering are summarized in

next chapter.

Page 23: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 23

2.2 Biocompatibility and Biomaterials

A biomaterial is defined as a non-viable material used in a medical device and

intended to interact with host tissue whereas biocompatibility is described as the

“ability of a material to perform with an appropriate host response”.49

However,

more recently considering different implant systems and taking into account that a

biomaterial is always a part of a regenerative therapy D. F. Williams updated and

specified this definition by elaborating that

“Biocompatibility refers to the ability of a biomaterial to perform its desired

function with respect to a medical therapy [...] generating the most appropriate

beneficial cellular or tissue response [...].50

When applied to a biomaterial scaffold

used for bone tissue engineering one can state that such a scaffold should perform

as a “substrate that will support the appropriate (bone cell) activity, including the

facilitation of molecular and mechanical signalling systems, in order to optimise

(bone) tissue regeneration.”50

Regarding the reaction with the surrounding tissue biomaterials are classified as

bioinert, bioresorbable and bioactive. Bioinert materials, e.g. alumina, titania or

most of metallic compounds, do not cause any toxic response from their host, yet

they are known to induce a fibrous tissue, disallowing formation of tissue

bonding.51

Bioresorbable materials, such as tricalcium phosphate or biodegradable

polymers dissolve in physiological environment and can be used as temporary

implants which are resorbed by the body and replaced by the native tissue.

Bioactive materials (which can be bioresorbable as well) can react with the local

tissue and form strong bonds to hard tissue (class B bioactivity). The discovery of

these complex surface reactions that allow for a strong bone bonding have made

glass a suitable implant material lead to the development of large amount of

bioactive materials including 45S5 Bioglass®

, glass-ceramics such as Ceravital®,

A/W glass ceramic, dense hydroxylapatite such as Durapatite® or Calcitite

®, as well

as composites such as polyethylene-Bioglass®, polysulfone-Bioglass

®, and

polyethylene-hydroxylapatite (Hapex®).

52 The physico-chemical reactions occurring

at the interface between a bioactive (glass) surfaces upon implantation in living

body leading to the tissue integration are summarized in 2.3.3. Moreover, bioactive

Page 24: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 24

glasses have been shown to stimulate human cells towards a specific cell response

by activating several genes related to osteogenesis and angiogenesis. This

osteoinductive behavior which promoted the development of so called 3rd

generation biomaterials53

is described in detail in section 2.3.4

2.3 Bioactive silicate glasses

Bioactive glasses (BG) such as 45S5 Bioglass®

(wt.%: 45SiO2-25CaO-25Na2O-

6P2O5) and other silicate based compositions represent an important group of

inorganic, bioactive biomaterials which have been widely used in regenerative

medicine.9 For an overview on the history, the properties and diverse applications of

bioactive glasses the reader is referred to articles by Hench53

and Jones54

. In this

section, however, some fundamental aspects on bioactive glasses fabrication, their

structure, general properties, further processing of BGs to scaffolds and their

biological performance are described.

2.3.1 Glass production

Sol-gel process

Low temperature techniques, like the sol-gel route offer an opportunity to produce

bioactive glasses. Hereby, the synthesis of an inorganic network is processed by

mixing organic precursor (e.g. metal alkoxides) in solution which is followed by

hydrolysis, gelation, and low-temperature firing. Silicate glass alkoxide precursors,

such as tetraethylorthosilicate (TEOS), undergo hydrolysis forming a colloidal

solution (sol). After polycondensation of silanol (Si–OH) groups, a silicate (–Si–O–

Si–) network is formed. While the gel is forming the viscosity of the system

increases as the network connectivity raises. Afterwards the gel is dried and

stabilised during a thermal process at round 600-800 °C.55, 56

Sol-gel derived glasses

exhibit mesoporous characteristics and have larger surface area than melt derived

glasses showing pores from 300 to 800 nm with total porosities of over 60%.57

By

choosing suitable precursor materials novel bioactive glass compositions containing

metal oxides can be produced. For instance, Zn containing sol-gel derived glass,58, 59

Page 25: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 25

and Mg-containing59

as well as Sr - glasses60

for biomedical applications have been

developed. However, long processing times, high costs of the precursor materials,

large shrinkage rates and residual carbon/hydroxyl are some of the disadvantages of

the sol gel process.61

Melt derived bioactive glasses

In the melt derived process conventional glass melting is used where desired

amounts of oxides, carbonates or phosphates are homogenized and melted in a

platinum crucible at 1350-1500 °C. The molten glass is then either poured into

graphite moulds in order to obtain solid glass blocks or quenched in water or oil

which results in a glass frit. Subsequent mechanical grinding (e.g. in a planetary

mill) can be applied to obtain glass powder which can be directly used as bone

defect filler material, as addition to polymer-BG composites or can be further

processed for fabrication of 3D scaffolds by sintering methods, as described in 2.4.

Melting is a flexible technique which allows the production of various different

glass compositions, simply by varying the number and proportion ratio of the raw

materials. Several melt-derived bioactive glass compositions have been used in

order to obtain metal oxide containing BGs including Sr-containing BG,62

Boron

derived BG,63

Co-BG64

and F-containing glasses.65, 66

The melt-derived route for

processing bioactive glasses also shows some disadvantages. For example it might

be difficult to achieve high purity glasses due to the high melting temperatures

involved (impurities from crucible material) and the subsequent use of grinding

steps which could lead to contamination with debris particles of the grinding media.

Moreover, the standard 45S5 Bioglass® obtained by the melting process tends to

crystallise during the sintering process forming a predominantly crystalline phase

(combeite)67

which may reduce the hydroxyapatite (HAp) conversion rat 68

and thus

affect the bioactivity of the material. Other modified bioactive glass compositions,

like “13-93” have been developed which show enhanced viscous flow and can be

densified without crystallisation.69

Despite these drawbacks mentioned above the

melting route was chosen for glass fabrication used in this study which exhibits

Page 26: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 26

advantages like easy controllable chemistry of the glass, low costs (regarding the

raw materials), short processing times and large outcome (amounts).

2.3.2 Structure and general properties of (silicate) bioactive glasses

Since Hench et al.70

has discovered the first bioactive silicate material, now called

45S5 Bioglass®

various different composition of bioactive glasses have been

developed including silicate melt derived glasses phosphate glasses, boro-silicate

glasses and sol-gel derived bioactive glasses.71

45S5 is a silicate glass based on

SiO2 as network former building the 3D structure in which Si is fourfold

coordinated to O. In the presence of Ca and Na cations acting as network modifiers

the 3D structure is disrupted and non-bridging oxygens (NBO) are created in the

glass network.

Common nomenclature describing the coordination of Si atom uses Qn

notation (Q

short for quaternary) where n is the number of bridging O bonds to a Si atom.72

NMR studies and network simulations have shown that 45S5 Bioglass® primarily

consists of chains and rings of Q2 (69%) with lower fractions of Q

3 (31%) providing

some cross linking.73

Even though it is well accepted that the 45S5 Bioglass®

structure is dominated by Q2 and Q

3 units a more precise model is given by

assuming a trinodal distributions with Q1, Q

2 and Q

3 units.

74-76 The exact

proportions of the Q units slightly vary between specific studies and simulation

models: one example is given by Pedone et al. who have shown based on MAS-

NMR studies that 45S5 Bioglass® matrix consists of 67.2 % Q

2 SiO2 tetrahedral

cross-linked by 22.3 % Q3 where the chains and rings are terminated by 10.5 %

Q1.77

Despite suggestions by molecular dynamic simulations77, 78

and spectroscopy

data74, 79

for the existence of Si-O-P bonds Pedone et al. showed in a combined

theoretical and experimental study that P is present in isolated orthophosphate

environment (QP0) charged balanced with cations without forming Si-O-P bonds.

80

Moreover, it has been confirmed that the cations are non-randomly distributed in the

glass compensation the NBOs from the Si species. Also the PO43-

species distract

the cations from the silica network for charge balancing. A fragment of the 3D

structure of 45S5 bioactive glass is given in Fig. 6.

Page 27: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 27

One of the key properties of 45S5 Bioglass® (and related bioactive glass

compositions) is their bioactive and osteostimulating behaviour which is described

in more detail in 2.3.3 and 2.3.4, respectively. One reason why 45S5 Bioglass®

lacks full commercial success compared to other bioactive ceramics like HAp or

TCP is its difficulty to be processed to foams or fibres.54

This is because 45S5 BG

tends to crystallise during high temperature treatment above 1000 °C. This leads to

poor densification of 45S5 solid bodies and to formation of micro cracks resulting

in low mechanical strength of 45S5 glass-ceramic. Overall the workability of 45S5

BG is quite poor.

As one alternative to 45S5 BG the so called 13-93 (wt%: 53SiO2-6Na2O-12K2O-

5MgO-20CaO-4P2O5) has been developed and has been approved for in vivo use in

Europe. Based on 45S5, 1393 has a comparatively higher SiO2 content and addition

network modifiers K2O and MgO. Furthermore, 1393 BG exhibits better processing

characteristics (larger process window with lower Tg and higher Tc) by viscous flow

sintering enabling densification of 1393 derived scaffolds without crystallisation.

Also processing of 1393 tapes or glass fibres is possible.71

Some of the key

parameters regarding the thermal behaviour of 45S5 and 1393 BGs are given in

Table 2.

Table 2: Material characteristics of 45S5 and 13-93 bioactive glass; glass transition point Tg,

crystallisation onset To, crystallisation peak Tc and melting point Tm. Temperatures are given in

°C.81, 82

Tg To Tc Tm

45S5 532 655 708 1180

13-93 606 714

In this work 45S5 and 1393 will be chosen as basic silicate systems for metal ion

doping. The density and some selected specific mechanical characteristics of these

two glasses are given in Table 3.

Page 28: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 28

Fig. 6: 3D structure model of 45S5 BG showing isolated PO4 tetrahedral, Q2 silica fragments

crosslinked with Q3 units and terminated Q

1 species. Na (green) and Ca (yellow) are charge

balancing the non-bridging oxygens.

Table 3: Density ρ, compressive strength σC, flexural strength σF, fracture toughness KIC tensile

strength, σT and elastic modulus of 45S5 and 1393 glass.

ρ [g cm-3

] σC [MPa] σF [MPa] KIC [MPam1/2

] σT [MPa] E-modulus [GPa]

45S5 2.7 500 30-60¥ 0.7-1.1

83 88±32

84*

4285#

3585

13-93 2.65 n.a. n.a. n.a. 440±15186**

n.a.

*for fibres of 165-310 µm;

**for fibres of 93-160 µm;

#annealed bulk;

¥tape cast and sintered

2.3.3 Acellular in vitro bioactivity

As mentioned above the ability of a biomaterial to bond to hard tissue is termed

“bioactivity”. The index of bioactivity (Ib) was proposed in order to quantify a

relative bioreactivity between different biomaterial which marks the time for more

than 50% of the interface to be bonded to bone in vivo.51

𝐼𝐵 =100

𝑡50% Eq. 1

The Ib has been shown to be the largest for 45S5 BG compared to other

bioceramics, e.g. alumina or hydroxyapatite.

Page 29: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 29

Silicate-based bioactive glasses exhibit a highly reactive surface capable of

releasing alkali and alkali earth ions into solution due to low SiO2 content in the

glass. BG show bioactive behavior in a broad range of different silicate glass

composition depending on the SiO2 content, as shown in Fig. 7.

Increasing the network modifier concentration increases the reactivity of the glass

surface up to the limit of a non-glass forming region at about 40 wt% SiO2. In fact,

the glass network is filled with so many cations that the reduced melting

temperature at the BG composition allows for a lower temperature glass production

with standard laboratory, with pure silica having a melting temperature of 1720

°C.87

The ternary diagram for bioactive glasses is depicted by Fig. 7 containing a constant

6 wt% of P2O5. A very high reactive region at low CaO composition does not bond

bone. A low reactivity region at low Na2O composition shows the A/W glass

ceramic. The region A has been reported as gene activating, an important step in the

ossification process.53

As indicated in the figure, this region has also the possibility

to bond to soft tissue,88

which is important for the complex mineralised tissue of

bone.

Fig. 7: Schematic compositional diagram for bioactive glasses at constant 6 wt% P2O5.13, 89

Page 30: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 30

Basically, the mechanism behind bioactive behaviour involves the formation of

carbonated hydroxyapatite (CHA) layer in the surface which is bonded to bone

tissue in vivo.

Table 4 shows the currently reaction events occurring ion BG surface in

physiological environment.9, 11, 90

The first 5 steps describe the physico-chemical

reaction on the surface of bioactive implant, whereas steps 6-11 show further

interaction of the surface with biological moieties and bone tissue formation:

(Step 1): First, a rapid diffusion driven (t0.5

dependent) exchange of sodium (Na+),

potassium (K+), or calcium (Ca

2+) ions from the glass with the proton ions from the

solution (H+ or H3O

+) occurs:

91, 92

𝑆𝑖 − 𝑂 − 𝑁𝑎+ + 𝐻+ + 𝑂𝐻− → 𝑆𝑖 − 𝑂𝐻+ +𝑁𝑎+(𝑎𝑞. ) + 𝑂𝐻− Eq. 2

The cationic alkali content is depleted to a depth > 0.5 pm.51

(Step 2): The second, though not sequential, step involves a loss of SiO2 in the form

of Si(OH)4 from the glass mediated by the hydrolysis of water. The produced

silanols form on the glass surface, exhibiting a t1.0

dependence.51

𝑆𝑖 − 𝑂 − 𝑆𝑖 + 𝐻2𝑂 → 𝑆𝑖 − 𝑂𝐻 + 𝐻𝑂 − 𝑆𝑖 Eq. 3

(Step 3): A low solubility product of the released orthosilicate induces a

condensation and repolymerisation of an amorphous silica-based (SiO2) layer on the

surface without the alkali and alkaline-earth cations present.

(Step 4): Ca2+

and PO43-

ions diffuse through the SiO2 layer forming a CaO-P2O5-

rich amorphous film that is continuing to grow by including ions from the

physiological solutions.

(Step 5): Finally, CHA crystallizes from the amorphous CaO-P2O5 film, the

amorphous silica thought to be acting as a nucleation site.51

Anions form solution

(OH-, CO3

2-, or F

-) can be substituted in the CHA crystals. Other calcium phosphate

Page 31: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 31

phases may co precipitate or other crystalline phases, like calcite, may also be

formed.93, 94

(Step 6): Adsorption of biological moieties in CGA layer. Protein adsorption is an

essential step occurring prior to cell adhesion. CHA is known to support adsorption

of certain proteins and hence enhance cell adhesion. Moreover, it has been

described in literature that an in vitro immersion of a BG surface lead to a

competitive effect of serum adsorption and the growth of the Ca-P rich amorphous

layer95

where a full hydroxyapatite layer did not develop after 3 or more days. It has

been suggest that “adsorbed serum proteins impede the nucleation and growth

reactions by which [calcium phosphate] would transform to carbonated apatite”,

indicating that proteins would quickly cover any BG-derived surface.

(Step 7-11): Action of osteoblast, formation and mineralization of bone matrix and

bone tissue formation.

Considering the network connectivity bioactive behavior of BGs is given for NC

values <2 whereas NC>2 results in non-bioactive behavior.96

Influences of glass

composition on HA forming abilities were investigated by Kim et al. 97

For SiO2

glass ranking from 50-70 mol %, the soaking time in SBF required for the

formation of HA varied from 0.5 days to 28 days. Hence the time of appearance of

the HAp layer can be used as an “bioactivity marker” in vitro to compare the

reactivity of different bioactive glass compositions.

Page 32: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 32

Table 4: Sequence of reactions induced by Bioglass® in physiological conditions.

9, 11, 90.

Time [h] Step Surface reaction stages

100

Bo

ne

form

atio

n

11 Mineralisation of matrix and bone growth

10 Generation of organic matrix

9 Differentiation of stem cells

20 8 Attachment of osteoblast stem cells

7 Action of macrophages

10 6 Adsorption of biological moeities in HCA layer

2

Ch

emo

-ph

ysi

cal

surf

ace

reac

tio

ns

5 Crystallization of hydroxyl carbonate apatite (HCA)

1 4 Migration of Ca

2+ and PO4

3- groups through the SiO2 layer and formaton

of amourphous calcium phoshpate (ACP). Further growth of ACP

through incorporation of Ca amd P-species from the solution

3 Polycondensation of SiOH + SiOH --> Si-O-Si+H2O

2 Network dissolution and formation of silanol (SiOH) bonds

1 Exchange of Na

+ with hydrogen ions from body fluids

Bioactive glass surface

Bioactivity assessment in vitro and degradation

In order to assess the bioactivity of biomaterials in vitro it has been proposes that

the formation of the CHA layer in simulated body fluid (SBF) is an indicator for in

vivo bioactivity.98

SBF attempts to mimic the physiological environment of the body

by simulating the inorganic part of the human blood regarding the ionic

concentration, ionic species, and the pH.99

Different recipes for SBF solution have

been proposed98, 100

in order to find the optimal conditions for assessing the

bioactive of materials in vitro. Despite the criticism of the “SBF-test” concerning

the correct prediction of the bioactivity101, 102

this test remains a widely used tool

used for assessing the mineralization potential of biomaterials in vitro.99

The formation of CHA on the BG surface in SBF can be used as a marker of

bioactivity. Hereby, on easily approachable method to quantify the bioactivity is to

detect the time point of appearance of the HCA layer which can be used to compare

the bioreactivity of newly developed glass compositions.

One other possibility to quantify bioactivity is by using activation energy to for

silicon release from the BG which follows an Arrhenius behaviour. It has been

shown that higher activation energies can be correlated to decreased bioactivity.103

Page 33: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 33

Taking this into account the rate of formation of CHA on 45S5 glass surfaces and

the dissolution behaviour of the glass by means of Si release should be considered

for characterization of the reactivity of BGs in physiological environment.

In first place the reaction scheme of a BG surface reaction (Table 4) in

physiological environment describes the formation of CHA leading to strong

interfacial bonding to bone. However, ionic dissolution products released from BG

have been shown to induce an intrinsic cell response influencing differentiation and

proliferation of cells involved in the bone formation process. These effects are

discussed in the following paragraph.

2.3.4 Cellular response to bioactive glasses (BG)

Beside acellular bioactivity BGs have been shown to stimulate bone cells via

upregulation of osteogenesis and angiogenesis related genes, hence enhancing bone

regeneration. Starting with the Xynos et al.17

investigation in 2001, the research

work has been focussing on the molecular interactions of ionic dissolution products

of BGs and their physiological environment in order to gain greater understanding

of these mechanisms and to be able to fabricate “smart” glasses with tailored

properties for specific tissue engineering applications, the so-called third generation

biomaterials.16, 104

One key finding from Xynos et al. studies in 2001 was that

Bioglass®

dissolution products are able to regulate gene expression in human

osteoblastic cells (HOC).17

Several genes known to play a role in osteoblast

metabolism, proliferation and cell-cell and matrix-cell adhesion were up-regulated

up to 5 fold when HOC were cultured in Bioglass® conditioned medium.

17 These

results were confirmed by Kaufmann et al.105

who observed expression of several

genes (osteocalcin, osteonectin, osteopontin) in osteoblast-like cells and increased

alkaline phosphatase (ALP) activity and collagen I formation. Related studies by

Jell at al.106

showed six and threefold upregulation of osteogenic markers, namely

bone sioloprotein (BSP) and ALP, in osteoblasts treated with BG conditioned

culture medium resulting in enhanced cell differentiation.

Beside 45S5 Bioglass, good cellular compatibility was also observed for 13-93 BG.

It has been shown that 1393 derived scaffolds support osteoblast growth and

differentiation in vitro as well as bone tissue formation in vivo.107-109

Similarly, 1393

Page 34: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 34

derived scaffolds have been shown to support growth and differentiation of

osteoblast-like cells (MC3T3-E1).110

The osteostimulating effect of the 13-93 BG,

where the defects filled with the glass showed high mRNA expression of genes for

both bone formation and bone resorption, thus affecting not only osteoblast but also

osteoclasts function and enhancing bone turn over.111

However, the glass particles

were injected along with bone morphogenic protein 2 (BMP-2), so that synergetic

effects of addition of bioactive glass and BMP-2 on bone formation were evaluated.

The osteogenic potential of the 1393 glass was also observed in vivo by implanting

particles in bone defects of rat tibia and by investigating gene expression at the

defect area.111

Osteogenic potential of bioactive glasses has been also confirmed for sol-gel

derived BGs. For example, sol-gel derived 77S (mol%: 80SiO2-16CaO-4P2O5)

bioactive glass has been shown to induce osteogenic differentiation of bone marrow

stromal cells into osteoblast-like cells and to promote cell mineralisation. 112

Furthermore, ionic dissolution products another sol-gel derived BG composition

(58S, mol%: 60SiO2-36CaO-4P2O5) resulted in cell activation through stimulation

of the proliferation of osteoblasts cells113

and upregulation of the expression of a

number of genes including IGF-I, gpl30 or MAPK3/ERK1.114

Additionally, it has

been observed that sol-gel derived phosphate-free glass 70S30C (mol%: 70SiO2-

30CaO) is able to enhance osteoblast maturation and differentiation as well as

production of bone-like minerals by osteoblasts indicating that P may be not

necessarily required for in vitro mineralization of the extracellular matrix.12

Beyond

that, a large number of additional studies have confirmed the osteogenic effects of

bioactive silicate glasses and their dissolution products which have been

comprehensively reviewed in literature.18, 115

Beside osteogenic potential also angiogenesis-related studies have been carried out

on BGs using endothelial cells and fibroblasts.116-121

Day et al.121

observed

increased neovascularisation into Bioglass®

coated polymer meshes after being

subcutaneously implanted in rats. Furthermore, Day116

showed that Bioglass®

stimulates the secretion of angiogenic growth factors from human fibroblasts cells,

providing further evidence that bioactive glass enhances angiogenesis. In this study,

fibroblast cells (CCD-18Co) cultured in Bioglass® particles containing medium

Page 35: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 35

secreted increased amounts of vascular endothelial growth factor (VEGF) and basic

fibroblast growth factor (bFGF). These and other studies on angiogenic effects of

bioactive glasses have been comprehensively reviewed confirming angiogenic

potential of BGs117

. However, it should be mentioned that the angiogenic effect of

bioactive glasses seems to be dependent on the morphology and the form in which

the BG is applied. The influence of the shape, morphology and size of bioactive

glass particles (which are for instance used as fillers in polymer-BG composites)

should be taken into account and need further investigation. It has been shown that

the angiogenic effect of bioactive glass seems to be more pronounced in bioactive

glass-based scaffolds (i.e. BG loaded collagen sponges,119

discs,122

meshes,121

tubes,123

and porous glass-ceramic scaffolds117, 124, 125

than in composite structures

incorporating and fully embedding bioactive glass particles in polymer matrices

such as microsphere composites126

or foams.127, 128

The large amount of results on the biological performance of bioactive glasses

discussed above confirms the hypotheses that “ionic dissolution products released

from bioactive glasses stimulate the genes of cells towards a path of regeneration

and self-repair” as discussed by Hench 129

. It is now well accepted that bioactive

silicate glasses are able to stimulate osteogenesis as well as angiogenesis and in

some cases exhibit antibacterial properties. Fig. 8 shows an overview of biological

response to bioactive glasses based on literature reports.

Page 36: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 36

Fig. 8: Schematic overview of biological responses to ionic dissolution products from bioactive

glass (BG). Additional release of therapeutic metal ions upon the degradation of the glass

matrix is proposed to enhance the biological performance of BGs. [From Hoppe et al.18

]

However, the exact mechanisms of interaction between the ionic products from

bioactive glasses and cells remain unclear. It has been discussed in literature that

critical concentrations of Si, Ca, and P are required for osteostimulating effects of

BGs. Based on Xynos et al. 17, 130, 131

studies, Hench 129

stated that eluted Ca and Si

concentrations of 60-88 ppm and 17-21 ppm, respectively, are critical for

upregulation of several osteogenic genes. The importance of Si was also

meantioned by Valerio et al. 132

who suggested that higher osteoblastic proliferation

and collagen secretion after treatment with BG60S dissolution products is related to

Si contact, since despite higher Ca concentration no increased osteoblast activity

was observed in presence of biphasic calcium phosphates (BCP) with no Si release.

Indeed, as described in 2.5.1 in more detail Si has been shown to be an essential

element for metabolic processes associated with the formation and calcification of

bone tissue 133, 134

and to be present in early stages of bone matrix calcification 134

.

Moreover, high Ca concentrations (of 88-109 ppm) have been shown to reduce

Saos-2 osteoblast proliferation when exposed to bioactive glass “MBG85” (mol%:

Page 37: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 37

85SiO2-10CaO-5P2O5) conditioned culture, whereas a Si concentration of 60 ppm

did not have any negative effect on the cell proliferation 135

. Despite high Ca

concentrations beeing toxic for human osteoblastic cells, Ca has been shown to

increase the glutamate release of osteoblast cells.136

Since glutamate signalling

pathways are known to play an important role in bone mechanosensitivity,137

the

extracellular Ca concentration must be considered as an important stimulating agent

in bone formation.

Altogether bioactive glasses offer various beneficial features making them

promising material for use in bone tissue replacement and bone regeneration

applications. In this context, BGs have been widely used as filler in polymer

composites,10

as bioactive coatings,138

in dentistry for alveolar augmentation 139

and

in the field of maxillofacial surgery.140

The use of BG to produce 3D porous

scaffold for tissue engineering in described in the following chapter.

2.4 Bioactive glass derived scaffolds (State of the art)

Various fabrication techniques have been described to produce 3D porous bioactive

glass and ceramic foams141

including foam replica68

, diverse rapid prototyping

methods,142, 143

freeze casting144

or freeze extrusion.145, 146

Table 5 provides an

overview of selected techniques currently used for fabrication of bioactive glass

foams and corresponding structural and mechanical properties. A more

comprehensive overview over fabrication methods of bioactive glass scaffolds and

bioglass-polymer composite foams can be found elsewhere.11

Clearly, all these

methods lead to different morphologies and structures of bioactive glass scaffolds.

Related to in vitro bioactivity in simulated body fluid (SBF), early studies involving

dense specimens indicated that crystallinity reduces the bioactivity of bioactive

glass.147

However, later studies focusing on highly porous scaffolds have shown that

the bioactive character remains for crystalline materials and the formation of HAp is

just delayed.68

Page 38: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 38

Table 5: Overview of different methods reported in literature for fabrication of bioactive glass

derived foams and their corresponding properties.

Fabrication

technique

Glass

composition

P [%] Pore size [µm] σc [MPa] Ref.

Foam replica “Fa-GC” 75 ~100 2

148

“13-93” 85 ± 2 ~100-500 11±1

69

“45S5”

∼90 510–720 0.3-0.4

68

Gel casting “ICIE”16

~80 ~380 1.9 149

Freeze extrusion “13-93“ ~50 Pore width:300 µm and

struts diameter 300 µm

~140 145, 146

“45S5” ~53 - -

150

“13-93” 55–60% 90–110 (pore width,

columnar);

20–30 (pore width,

lamellar)

25 ±3

(columnar)

10 ±2 (lamellar)

108

Direct ink „6P53B“ 60 500 μm (pores size),

100 μm (rod diameter)

136 ±22 151

Lithography “45S5” - - 0.33

152

P...relative Porosity; σc...compressive strength; “Fa-GC...(mol%: 50SiO2-18 CaO-9CaF2-Na2O-7K2O-6P2O5-3MgO);

“13-93”...(wt%: 53SiO2-6Na2O-12K2O-5MgO-20CaO-4P2O5); “45S5”...(wt.%: 45SiO2-24.5Na2O-24.5CaO-6-P2O5);

“ICIE”16…(mol%:49.5SiO2, 36.3CaO, 6.7Na2O, 1.1P2O5, 6.6K2O); „6P53B“...(wt%: 52.7SiO2-10.3Na2O-2.8K2O-

10.2MgO-18.0CaO-6P2O5)

Besides the now well established foam replica method to make bioactive glass

scaffolds, other techniques have been considered to fabricate 3D porous glass and

glass-ceramic scaffolds. Organic molecules, such as starch from rice, potato, or corn

grains for instance can be used to introduce porosity by swelling these molecules in

water.153

After the sintering process and burn out of the organic fillers, a highly

interconnected pore system remains which contains pores of size 84 μm and a pore

content of 40% vol., as reported for bioactive glass.153

Fibre-derived scaffolds have been also presented, which are based on the assembly

of bioactive glass fibres to produce porous structures. Melt derived glass fibres are

packed and bonded together in a ceramic mould using a continuous bead of silicone

adhesive110

or sintered together.154

Typically fibre based scaffolds show porosities of

40-60 vol% and compression strength values of 12-18 MPa, notably higher than

values achieved by the foam replica method, albeit at lower porosities.

Other techniques for fabricating glass foams include freeze casting144

and freeze

extrusion.145, 146

Camphene, ice or water and glycerol can be used as freezing

vehicles.150, 155

After mixing the glass powder with the relevant vehicles, the slurries

Page 39: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 39

are cast and frozen at temperatures between -20 and -70°C, followed by a sintering

process.

Freeze casted 13-93 scaffolds with oriented (lamellar and columnar) pores and

equivalent porosity of 55–60% were shown to have a compressive strength of 25±3

MPa, compressive modulus of 1.2 GPa and pore width of 90–110 µm for columnar

scaffolds, compared to values of 10 ± 2 MPa, 0.4 GPa, and 20–30 μm, respectively,

obtained for the lamellar scaffolds.108

Rapid prototyping techniques have also been

described for fabricating porous bioactive glass based scaffolds. Direct ink writing

for instance was used to develop bioactive glass (6P53B composition) scaffolds

exhibiting a compressive strength of 136 ± 22 MPa, which is comparable to the

value for cortical bone (100–150 MPa) with porosity of 60% 151

. In a recent study, a

method based on lithography-based additive manufacturing technologies (AMTs)

was applied to create 45S5 bioactive glass scaffolds152

resulting in scaffolds

showing biaxial strength and compressive strength of ~40 MPa and 0.33 MPa,

respectively.

Using sol-gel derived bioactive glass particles direct foaming methods can be

applied in order to fabricate porous scaffolds.42, 156

Jones et al.157

described sol gel

derived BG foams where the scaffolds are obtained by direct foaming of the sol

using Teepol as foaming agent. After a drying process the gelled foams are aged at

60 °C, dried at 130 °C and stabilized at 600 °C. In a further heat treatment process

the foams are densified at 800 °C. Fig. 2 shows a typical structure of a sol-gel

derived bioactive glass scaffold. By varying the amount of foaming agent the pore

size distribution and overall porosity can be tuned 157

. The mechanical strength of

sol-gel derived bioactive glass scaffolds is usually in the range of 0.3–2.3´MPa (in

compression) limiting their applications to non-load bearing tissue engineering

approaches. Another related technique involving sol-gel derived glasses has been

developed using sugar cane as a template.158

In this study, the foam replica (FR) technique will be used for scaffold fabrication.

The polymer foam replica method was introduced in 2006 to fabricate 3D “45S5”-

based scaffolds for bone tissue engineering and it is widely used since then.68, 69

Briefly, polyurethane (PU) foam used as sacrificial template is infiltrated with a

glass powder containing slurry which adheres to the PU foam surface. Afterwards

Page 40: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 40

the excess slurry is removed and the coated PU foam is dried and then densified in a

sintering step. The polyurethane template determines the macro structure of the

final glass or glass-ceramic foam-like scaffold. Typically glass foams made by the

FR method show total porosities of >85 vol% and pore sizes in the range 100 - 400

µm. The chemical composition and extent of crystallinity depends on the starting

glass powder composition used. While 45S5 Bioglass® derived scaffolds crystallize

during sintering and form silicate and phosphorous rich phases,67, 68

more recently

developed glasses like “13-93”, which contain larger amounts of alkali oxides,

remain amorphous without any crystallization during the densification heat

treatment.69, 159

The structure and chemistry of the scaffolds also determine the

scaffold’s in vitro bioactivity, mechanical properties and has also an effect on the

protein adsorption on scaffold surfaces. For example dense highly crystalline 45S5

Bioglass®

derived scaffolds have compressive strength in the range of 0.25-0.4 MPa

(Table 5), lying even below the lowest compressive strength values reported for

spongy bone. On the other hand for amorphous 13-93 bioactive glass derived

scaffolds compressive strength values up to 11 MPa have been reported.69

The

different values for the strength of these scaffolds might be related to the processing

conditions and the resulting structure of the scaffolds. 13-93 bioactive glass can be

densified in a viscous flow process which should lead to crack-free struts where

sintering is not impaired by the crystallization process, which occurs in 45S5 type

bioactive glass. Fu et al. 69

reported pore size values of 100-500 µm and a

compressive strength of 11 ± 1 MPa for 13-93 derived glass scaffolds made by foam

replica technique. The influence of chemical composition, fabrication method and

scaffold structure on the mechanical properties of bioactive glass derived foams has

been discussed in the literature.38

Page 41: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 41

2.5 Metal ions and bioinorganics in silicate glasses

2.5.1 Role of bioinorganics in bone metabolism

Metallic ions are essential in human metabolism and are also known to play a

critical role in osteogenesis and angiogenesis.22, 160

They have been considered

highly promising for the field of biomedicine.161

Single inorganic ions such as

Calcium (Ca).136, 137, 162, 163

, Phosphorous (P),164

Silicon (Si),133, 134, 165

Strontium

(Sr),30, 166-170

Zinc (Zn),31, 171

as well as Boron (B),172, 173

Vanadium (V),23, 174, 175

,

Cobalt (Co)176-179

and Magnesium (Mg)180-186

are known to be involved in the

bone metabolism and to play a physiological role in angiogenesis, growth and

mineralization of bone tissue. Table 6 gives an overview of biological responses to

single inorganic ions. In particular, metal ions are known to act as enzyme co-

factors and therefore influence signalling pathways and stimulate metabolic

effects occurring during tissue formation.22, 187

Considering these effects and the

fact that metallic ions are cheap, easy to process and are less risky compared to

gene therapy or growth factor based techniques metal ions attractive for use as

therapeutic agents in the fields of hard and soft tissue engineering.188

The

advantages of therapeutic inorganic molecules such as metallic ions for use as

stimulating agents for osteogenesis and angiogenesis have been comprehensively

reviewed in literature.18-20, 160, 188

In the following few key properties of selected

bioinorganics are highlighted demonstrating their importance in bone physiology.

Si is an essential element for metabolic processes and is also associated with the

formation and calcification of hard tissue.133, 134

High Si contents have been

detected in early stages of bone matrix calcification,134

while aqueous Si was

shown to be able to induce hydroxyapatite precipitation, the inorganic phase of

human bone.189

Furthermore, dietary Si intake was shown to increase the bone

mineral density (BMD) in men and premenopausal women.190

In a in vivo study

with Ca-deficient rats showed that Si supplementation caused positive effects on

bone mineral density by reducing bone resorption.191

Regarding its physiological

role Nielsen et al.192

suggested that Si has a biochemical function in bone growth

Page 42: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 42

processes affecting bone collagen turnover and sialic acid-containing extracellular

matrix (ECM) proteins like osteopontin. Moreover, orthosilicate acid (Si(OH)4) at

physiological concentration of 10 µmol has been shown to stimulate collagen I

formation in human osteoblast cells (HOC) and to stimulate osteoblastic

differentiation.165

Table 6: Effect of metallic ions on human bone metabolism and angiogenesis.

Ion Biological response in vivo / in vitro

Si - essential for metabolic processes, formation and calcification of bone tissue

133, 134

- dietary intake of Si increases bone mineral density (BMD) 190

- aqueous Si induces HAp precipitation189

- Si(OH)4 stimulates collagen I formation and osteoblastic differentiation165

Ca - favours osteoblast proliferation, differentiation and extracellular matrix (ECM)

mineralization 162

- activates Ca-sensing receptors in osteoblast cells, increases expression of growth factors,

e.g. IGF-I or IGF-II136, 163

P - stimulates expression of matrix la protein (MGP) a key regulator in bone formation

164

Zn - shows anti-inflammatory effect and stimulates bone formation in vitro by activation

protein synthesis in osteoblasts31

- increases ATPase activity, regulates transcription of osteoblastic differentiation genes,

e.g. collagen I, ALP, osteopontin and osteocalcin193

Mg - stimulates new bone formation

186

- increases bone cell adhesion (probably due to interactions with integrins) 186, 194

Sr - shows beneficial effects on bone cells and bone formation in vivo

30, 170

- promising agent for treating osteoporosis195

Cu - significant amounts of cellular Cu are found in human endothelial cells when undergoing

angiogenesis196

- promotes synergetic stimulating effects on angiogenesis when associated with angiogenic

growth factor FGF-2197

- stimulates proliferation of human endothelial cells198

- induces differentiation of mesenchymal cells towards the osteogenic lineage199

Co - mimics hypoxia hence stabilises HIF-1 factor

200

-In vivo: CoCl2 pre-treated BMSCs induced higher degree of vascularisation and enhanced

osteogenesis within the implants179

B - stimulates RNA synthesis in fibroblast cells

201, 202

- dietary boron stimulates bone formation173

Mg is another essential element for bone metabolism and it has been shown to

have stimulating effects on new bone formation.180-186

Mg is suggested to interact

with integrins of osteoblast cells which are responsible for cell adhesion and

stability.186, 194

Rude et al. 181-183

observed that Mg depletion results in impaired

Page 43: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 43

bone growth, increased bone resorption and loss in trabecular bone underlining

the significant role that Mg plays in bone metabolism.

Because of the chemical analogy to Ca (same main group, similar atomic radius,

Ca2+

- 1.0 A, Sr – 1.16 A), Sr can accumulate in bone by exchanging with Ca in

the hydroxyapatite crystal lattice.203

The therapeutic potential of Sr in bone

metabolism has been investigated by Marie et al., who showed that Sr exhibits

beneficial effects on bone cells and bone formation in vivo.30, 170

Sr has also been

shown to be a promising agent in treating osteoporosis.195

Additionally, Sr based

drug such as strontium renalate enhances bone healing indicated by increased

callus resistance in rat bones which has confirmed Sr as a promising agent in

healing bone fractures.204

Zinc is also known to play an important role in bone metabolism 31

and to have

anti-inflammatory effects.205

Furthermore, Zn stimulates bone formation in vitro

by activating protein synthesis in osteoblast cells and increasing ATPase activity

in bone 31

. Moreover, Zn shows inhibitory effect on bone resorption inhibiting the

formation of osteoclast cells in mouse marrow cultures.31

The regulatory effects of

Zn on bone cells has suggested the importance of Zn in gene expression 206

. More

recently, zinc was identified as regulation agent in transcription of osteoblastic

differentiation genes, such as collagen I, ALP, osteopontin and osteocalcin.193

It

was assumed that Zn can be considered a Runx2 stimulating agent being able to

directly stimulate bone formation through increasing Runx2-targeted osteoblast

differentiation gene transcription.193

Cu has been shown to play a significant role in angiogenesis.196-198

For example,

remarkable distributions of cellular Cu have been found in human endothelial

cells when they were induced to undergo angiogenesis revealing the importance of

the ion as angiogenic agent.196

Another recent study revealed that copper,

associated with angiogenesis growth factor FGF-2, promotes synergetic

stimulating effects on angiogenesis in vitro.197

Moreover, Cu was shown to

stimulate proliferation of human endothelial cells 198

and to induce an increase in

differentiation of mesenchymal stem cells (MSC) towards the osteogenic lineage

199. Cu

2+ at a concentration of 10

-6 mol l

-1 has been also shown to inhibit osteoclast

activity.207

. However, a possible positive effect of Cu on bone metabolism is still

Page 44: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 44

not clearly demonstrated. Cashman et al.208

for instance found that copper

supplements over a period of 4 weeks do not show any effect on biochemical

markers of bone formation or bone resorption.208

Similarly, Lai and Yamaguchi 209

have shown that supplementation with copper induced a significant decrease in

bone tissue of rats showing no anabolic effects on bone formation in vivo and in

vitro and, additionally, it was shown that Cu reduces anabolic effects of Zn. Other

researchers have found that dietary copper depletion causes a reduction of bone

mineral density (BMD), although no biological markers for bone formation and

bone resorption were explicitly affected by Cu.210

Cobalt ions have been considered as important ions in the context of bone

physiology176-179

since Co is a known to induce hypoxia conditions and to

stabilize HIF-1.211, 212

Hypoxia conditions, in turn, are known to activate several

pro-regenerative processes in human body213

via regulation of the hypoxia-

inducible factor 1 (HIF-1). HIF-1 activation has been shown to result in

accelerated bone ingrowth whereby the HIF-1a pathway has been identified being

critical for angiogenesis and skeleton regeneration214

and are important for

development of angiogenesis, stem cell differentiation and fracture repair.64, 215

Furthermore, HIF-1 is stabilized under hypoxic conditions and regulates several

genetic pathways relevant for skeleton repair.216

Hence Co-releasing bioactive glass has been proposed as hypoxia mimicking

material217

to be used for artificial stabilization of HIF-1. These potential benefits

of Co2+

ions were further tested in vitro and in vivo. For instance, Cobalt was

shown promote angiogenesis via activation of hypoxia inducible factor 1 (HIF-1)

in a rat remnant kidney in vivo model when Co was applied by subcutaneous

injections.179

Also in a rat bladder in vivo model Co was shown to enhance

hypoxia response, cell growth and angiogenesis indicated by stimulated

expression of HIF-1α and vascular endothelial growth factor (VEGF).176

Furthermore, bone marrow derived stem cells (BMSCs) pre-treated with CoCl2

induced higher degree of vascularisation and enhanced osteogenesis within

collagen scaffolds implanted in vivo.200

Page 45: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 45

The use of metallic ions incorporated in a silicate glass system has been widely

investigated. In the following chapter relevant studies on the biological

performance of metal ion containing bioactive glasses are highlighted.

2.5.2 Biological performance of metal ion containing glasses and

glass ceramics

In order to enhance the bioactivity of bioactive glasses towards a specific

biological response in relevant physiological environments many approaches have

been investigated incorporating various metal ions in the silicate network. The

main goal is to increase the stimulating effects of bioactive glasses on

osteogenesis, angiogenesis and the promotion of antibacterial properties.

The incorporation of selected metal ions into silicate matrices has resulted in

enhanced bone formation and angiogenesis,18

whereby phosphate based glasses

can also be used as effective carriers for TII.218

These effects have been shown for

various silicate systems including Co-BG,64, 219

Zn-BG,220-224

Cu-BG,224-226

Sr-

BG,62, 227-230

Mg-BG,225, 231-235

, Ag-BG,148, 236-241

Ce-BG242

as well as B-BG222, 243-

245 and F-BG.

66, 246 Comprehensive reviews on the biological impact of such novel

silicate and also phosphate based glass compositions can be found in literature.18,

247 In the following few selected glass systems and their relevance for bone tissue

engineering applications.

Diverse silicate glasses based on CaO–MgO–SiO2 with different additive agents

(B2O3, CaF2, Na2O and P2O5) have been shown to exhibit appropriate level of

acellular bioactivity proved through standard bioactivity testing by soaking the

material in simulated body fluid, according to Kokubo et al.15, 99

, in order to prove

the formation of surface HCA layer.248

In the study of Agathopoulos et al. 248

, for

example, it was observed that increasing amount of phosphates favours the

deposition of HCA while increasing contents of CaO and SiO2 inhibit its

deposition. However no biological data through in vitro cell test were provided for

the glass compositions investigated. Moreover, Vrouwenvelder et al.249

reported

on the osteoblast behaviour on 45S5 bioactive glass doped with iron, titanium,

fluorine and boron giving (to our knowledge) the first overview of the biological

response to ion doped bioactive glasses. It was shown that Ti-BG exhibited higher

Page 46: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 46

proliferation and osteoblast expression probably due to more controlled release of

the ionic glass products, whereas doping with B, Fe and F resulted in lower

osteoblast activity compared to undoped control 45S5 bioactive glass.

Moreover, HAp formation was delayed through the incorporation of CoO, ZnO

and MgO as it was revealed by SBF studies,64

, which is, in fact, desirable for soft

tissue engineering applications, e.g. cartilage regeneration.

In the following paragraphs different studies on particular ions and their effect on

cell behaviour are summarized and the results discussed regarding their relevance

for bone tissue engineering.

Bioactive silicate glasses doped with zinc have been shown to enhance acellular

formation of calcium phosphate layer on the BG surface after soaking BG

substrates in biological fluids (Dulbecco's Modified Eagle's Medium, DMEM)250

and in simulated body fluid (SBF),251, 252

which is a fundamental requirement for

bioactive glasses to bond to bone.9 Independently of the acellular behaviour

investigations, in vitro cell culture studies have shown the anti-inflammatory

effect of Zn-doped bioactive sol-gel derived 58S glass.224

In this study, murine

macrophage cells (RAW 264.7) stimulated with lippopolysaccharide (LPS)

endotoxin were cultivated in DMEM treated with BG suspension (1mg ml-1

) and

glass ionic dissolution products, respectively, resulting in decreased secreted

amount of tumor necrosis factor TNF-α and interleukin-1 which indicated the

inflammatory effect on the cells. Glass compositions with 5mol-% and 11mol-%

Zn did not have any therapeutic effects oncells since the TNF-α and IL-1 contents

were not significantly decreased after treatment with the BG suspension. On the

other hand the pre-treatment of cells with bioactive glasses resulted in a

significantly decreased amount of TNF and IL-1 indicating a prophylactic anti-

inflammatory effect of the Zn-doped glasses compared to the undoped control

glasses. However, cell treatments with ionic dissolution products of the same glass

did not show any positive anti-inflammatory effects of the Zn-doped glasses

compared to undoped control glass, which might be related to the high

concentrations of Zn (up to 16 ppm) released into the cell medium being toxic for

cells. Other studies have also shown that Zn concentrations in the range of 2-8

ppm can cause damage in human osteoblasts via oxidative stress.253

Similar

Page 47: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 47

results were observed for endothelial cells seeded on wells in the presence of Zn-

doped glass slabs revealing that bioactive glasses doped with 20 wt.% Zn increase

adhesion of cells but decrease the proliferation rate probably due to toxic Zn

concentration of 2.7 ppm released into the culture medium, whereas 5Zn-BG

promoted well cell adhesion as well as high cell proliferation of endothelial cells

at Zn concentrations of 1.1 ppm.220

However, results on cell adhesion related to

the undoped glass composition, indicated that all glass samples, undoped 58S,

5%Zn-58S and 20%Zn-58S, exhibited poorer cell adhesion and proliferation than

the silica control samples.

Additionally, Zn-releasing bioactive glass scaffolds have been investigated in

relation to their possible osteogenic effects.58, 222, 252, 254, 255

It has been shown that

sol-gel derived glasses containing 5 mol% ZnO resulted in increased ALP activity

and increased osteoblast proliferation 58

indicating the possible stimulating effect

of Zn-BG on osteoblast cells.

In related research, Oki et al.255

observed that Zn containing bioactive glass (5

mol%) increased AP activity of human fetal osteoblastic cells (hFOB 1.19) when

seeded on glass discs in relation to polystyrene control. However, these results

were not compared to un-doped glass samples, so that no definitive statement on

the stimulatory effect of Zn can be made based on the available experimental

evidence. More recently, Haimi et al.222

investigated the effects of Zn-doped BG

scaffolds on human adipose stem cells proliferation and osteogenic differentiation

revealing no significant effect of Zn-doping on cell activity when cells were

seeded on the Zn-doped BG scaffolds. The authors suggested that the possible

stimulatory effect of Zn is inhibited through the decreased degradation profile of

the bioactive glass caused by the Zn addition. Similar results were observed by

Lusvardi et al.252

, who found that Zn-doping provides no significant effect on

adhesion and proliferation of osteoblast-like cells (MC3T3-E1) compared to un-

doped control glasses, probably due to the slowed degradation of the BG caused

by Zn addition resulting in a negligible amount of Zn ions released. However, the

degradation kinetics was investigated only in SBF (release of 4.6 ppm after 70

days) and no cell cultures treated with ionic dissolution products of the Zn doped

Page 48: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Fundamentals 48

glasses were carried out missing a specific investigation of possible anabolic

effect of Zn. Despite the fact that in vitro studies with systematic supplements of

Sr-doped glasses have been shown to exhibit enhanced acellular bioactivity and to

release critical concentrations of Sr ions in the range of 1-5 ppm into the

dissolution medium.229, 256

Moreover, Gentleman et al. 227

have shown that ion

release from Sr-doped silicate glasses enhances bone cell activity. They

investigated the in vitro behaviour of SiO2-P2O5-Na2O-CaO silicate glass in which

Ca was systematically replaced by Sr up to 100%. In this study, treatment with

ionic dissolution products of Sr-doped glasses (ion release in culture medium in

the range of 5-23 ppm) resulted in enhanced osteoblast (Saos-2 cell line) activity

and inhibited osteoclasts differentiation. Moreover, these Sr-doped glasses

promoted osteoblast proliferation and ALP activity when directly applied in

contact with cells as solid BG-discs. Similarly, it has been shown that Sr doped

BGs produce increased proliferation of rat calvaria osteoblastic cells and enhance

cell differentiation and ALP activity.60

In vivo studies have also confirmed the

high biocompatibility of Sr-containing 45S5 Bioglass® expressed through strong

bonding to bone via HCA layer without any inflammatory affects, though no

differences of the in vivo behaviour compared to undoped Bioglass® control were

observed 228

.

Different approaches have been made in order to incorporate Cu in an inorganic

scaffold material, including Cu addition in calcium phosphates,257, 258

silicate

glasses,226

phosphate based glasses259, 260

and in polymer coatings on bioactive

glass scaffolds.261

Cu release from calcium phosphates was shown to enhance

angiogenesis257

whereas Cu-containing meso-porous bioactive glass scaffolds

were reported to stimulate angiogenesis as well as osteogenesis.226

The direct

incorporation of Cu ions into 45S5 Bioglass® scaffolds has not been investigated

to date and will be part of this work.

Cobalt-containing inorganic biomaterials have been also poropsed for use in

regenerative medicine. For example, cobalt-containing β-tri-calcium phosphate

ceramic (β-TCP) have been shown to stimulate VEGF expression of BMSCs and

to enhance the formation of network structure of human umbilical vein

endothelial cells (HUVECs) compared with pure TCP.262

Co-containing glasses

Page 49: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

49

co-doped with Zn and Mg 64

have been fabricated in order to create “hypoxia-

mimicking” (low pressure environment) biomaterials to be used in bone tissue

engineering. Creating hypoxia conditions is suggested to be a strategy for

activating pro- and anti-angiogenic genes.263

Melt derived Co-containing BGs

have been reported in literature which a have been introduced as suitable carrier

for Co2+

ions.64

. From these glasses Co concentrations in the range of 10-14 ppm

were released in physiological medium, which are believed to be within the

biologically active limits. Sol-gel Co-containing bioactive glasses have been

reported by Wu et al. who used sol-gel derived BG derived scaffolds by foam

replica technique as Co releasing platform.219

However, no Co containing 13-93

bioactive glass (1393-Co) has been presented so far. Also no Co-containing highly

porous scaffolds with sufficient mechanical strength have been reported in

literature. Hence, on of the goals of this study was to fabricate Co-releasing

mechanically robust 3D scaffolds based on 13-93 bioactive glass suitable for

potential applications in regenerative medicine.

Most of the ion-doped bioactive glasses reported in literature have been shown to

exhibit sufficient levels of acellular bioactivity indicated through the formation of

carbonated hydroxyapatite layer after soaking in SBF18

which is an essential

requirement of biomaterials to form a strong bonding to bone.9 Moreover, doping

with different metal ions seems to be a promising approach to enhance the

biocompatibility of the glasses and to stimulate cell proliferation, as it was shown

for a few specific novel glass compositions discussed above.

In this work Cu and Co were chosen as therapeutic ions to be incorporated in BG

glass derived scaffold.

Page 50: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Materials and Methods 50

3 Materials and Methods

3.1 Glass fabrication

3.1.1 Cu-containing 45S51

Melt derived 45S5 bioactive glasses with different CuO contents (45S5-Cu) were

produced by mixing analytical grade silicon oxide (SiO2, Merck, Darmstadt,

Germany), sodium carbonate (Na2CO3, Merck, Darmstadt, Germany), calcium

carbonate (CaCO3, Merck, Darmstadt, Germany), tri-calcium phosphate (Ca3(PO4)2,

Sigma-Aldrich, Germany) and basic Cu carbonate (CuCO3*Cu(OH)2, Sigma

Aldrich, Germany). The raw materials were well homogenized and melted in a

platinum crucible at 1450 °C for 45 min (Furnace LHT 3 KW, C295 Control Unit,

Nabertherm, Germany). The nominal compositions of the glasses investigated in

this study are given in Table 7.

Table 7: Nominal 45S5 Bioglass® derived glass compositions with different Cu contents.

Glass Composition in wt%

SiO2 Na2O P2O5 CaO CuO

45S5 45 24.5 6 24.5 -

45S5-0.1Cu 45 24.5 6 24.4 0.1

45S5-1Cu 45 24.5 6 23.5 1

45S5-2.5Cu 45 24.5 6 22 2.5

3.1.2 Co containing 13-932

Co-containing bioactive glasses based on the 1393 composition (1393-Co) were

fabricated using melt-derived technique. Silicic-acid (pure anhydrated, Riedel de

Haën, Germany), Sodium-carbonate (Riedel de Haën, Germany), Potassium-

carbonate (Riedel de Haën, Germany), Sodium-dihydrogenphosphate (Riedel de

1 Melting of the 45S5-Cu glass series was carried out at the Institute of Glass and Ceramics (Glass group, Prof. Wondraczek) under assistance of R. Meszaros 2 Melting of the 1393-Co glasses was carried out by B, Jokic and Prof. Jankovic at Universtiy of Belgrade

Page 51: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Materials and Methods 51

Haën, Germany), Magnesium-carbonate (Lachner, Czech Republic), Calcium-

carbonate (Lachner, Czech Republic), Cobalt-nitrate (Carlo Erba-Italy) calculated to

the glass composition. All chemicals were p.a. grade. Melting was performed in a

platinum pot at 1300 oC for 2h and rapidly cooled by quenching in water. The glass

compositions are given in Table 8.

Table 8: 13-93 derived glass compositions with varying amounts of CoO (Composition in wt%).

Glass SiO2 Na2O K2O P2O5 MgO CaO CoO

1393 53.0 6.0 12 4 5 20 -

1393-1Co 53.0 6.0 12.0 4.0 5.0 19.0 1.0

1393-5Co 53.0 6.0 12.0 4.0 5.0 15.0 5.0

1393-10Co 53.0 6.0 12.0 4.0 5.0 10.0 10.0

3.2 Preparation of the glass powder

Then, the glass was fritted, pre-milled in a jaw crusher and milled in a planetary

mill to the final particles size of d50 =5 µm. Fritted glass was then pre-milled in a

jaw crasher (BB51, Retsch, Germany) and milled in a planetary mill using ZrO2 jar

and ZrO2 milling balls (PM100, Retsch, Germany) to a final particle size of d50~5

um.

3.3 Characterization techniques

3.3.1 Scanning electron microscopy (SEM) / Energy dispersive X-

ray spectroscopy (EDS)

For SEM analyses the scaffolds were fixed on a sample holder with a dry

conductive adhesive (Leit-C, Fluka Analytical). Before analyzing, the scaffolds

were sputtered with gold. Images were taken with a FEI-Quanta 200 scanning

electron microscope at an operation voltage of 20 kV.

Page 52: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Materials and Methods 52

3.3.2 Fourier transform infrared spectroscopy (FT-IR)

The glass structural properties and the formation of an apatite-like phase on the

scaffold surface were investigated by FT-IR (Impact 420 Nicolet, Thermo

Scientific, Waltham, US). For that the glass or ground scaffolds were pressed into

potassium bromide (KBr) pellets. The pellets were made by mixing 1 mg sample

and 300 mg of KBr (Spectroscopy grade, Merck, Darmstadt, Germany). Spectra

were collected between 2000 and 400 cm-1

with a resolution of 4 cm-1

. In order to

compare the intensities of single bands all spectra were normalized to the silica

band at ~1040 cm-1

.

3.3.3 Raman spectroscopy

Raman spectra (LabRAM HR – Evolution, Horiba Scientific, Germany) of the glass

powders were obtained using a diode pumped solid state laser (YAG. = 532 nm).

Hole, slit and grating were set to 1000 µm, 100 µm, and 600 gr/mm, respectively.

Data acquisition time was 5s and averaging was performed over 20 measurements.

The spectra were normalized to the intensity of the band centered at 630 cm-1

.

3.3.4 Inductively coupled plasma atomic emission spectroscopy

(ICP-OES)

Static conditions3

Ion release from scaffolds soaked in SBF was quantified using an inductively

coupled plasma - optical emission spectrophotometer (ICP-OES) (Thermo

Scientific iCAP 6500). Values were calibrated against certified standards serially

diluted with SBF to 100, 10, 1 and 0.1 and ppm. (4% nitric acid was added to all

samples and standards).

Quasi-dynamic conditions4

Ion release (Cu, Si) under quasi-dynamic conditions from scaffolds soaked in SBF

was measured using the ICP-OES Optima 8300 (Perkin Elmer, USA). Five point

3 ICP measurements of ions released under static conditions were carried out by C. Stähli at McGill University, Montreal, Canada 4 ICP measurements under quasi-dynamic conditions were carried out by S. Romeis and J. Schmidt at the Institute of Particle Technology, University of Erlangen-Nuremberg

Page 53: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Materials and Methods 53

calibrations (25; 10; 5; 1; 0.1 ppm) were performed by diluting certified standards

(Carl Roth, Germany) with SBF. Given errors are estimated by linear regression

analysis. The samples were measured in triplicate and mean values with standard

deviation were derived. For scaffold digestions a of 2 ml hydrofluoric acid (48 %,

supra purity, Roth Chemicals, Germany), 2 ml nitric acid (69 %, supra purity, Roth

Chemicals, Germany), and 4 ml hydrochloric acid (35 %, supra purity, Roth

Chemicals, Germany) was used. After the microwave procedure 2 g of boric acid

(>99.8 %, Roth Chemicals, Germany) was added and the total volume was set to

250 ml.

3.3.5 Micro-Ion beam5

Analyses of the glass derived scaffolds/biological fluids interface were carried out

using nuclear microprobes at CENBG (Centre d’Études Nucléaires de Bordeaux-

Gradignan, France).264

PIXE-RBS analyses were performed on the nanobeam line

with a proton scanning micro-beam of 3 MeV energy and 60 pA in intensity. The

beam diameter was nearly 1 µm. An 80 mm2 Si(Li) detector was used for X-ray

detection, orientated at 135° with respect to the incident beam axis and equipped

with a beryllium window 12 µm thick. An aluminium funny filter of 100 µm in

thickness with a hole of 2 mm in diameter was added on the detector. The software

SUPAVISIO was used to define the different regions of interest with the use of

masks. These masks isolate the spectra corresponding to the region of interest in

order to calculate the elemental composition in that region. Quantification of PIXE

spectra was done using GUPIX software.265

The data was calibrated against NIST

standard reference glass materials. Relating to RBS, a silicon particle detector

placed at 135° from the incident beam axis provided the number of protons that

interacted with the sample. Data were treated with the SIMNRA code.29 The

thickness of reaction layers formed during immersion in SBF was determined using

ImageJ software. Three measurements at different spots were taken per chemical

map and the mean value and standard deviation were derived from 3 chemical

maps.

5 PIXE-RBS analysis was carried out by J. Lao and E. Jallot, fClermont Université, Université

Blaise Pascal

Page 54: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Materials and Methods 54

3.3.6 XRD

X-ray diffraction analysis was performed using a D8 ADVANCE X-ray

diffractometer (Bruker, Madison, US) in a 2Theta range from 20-80°. BG powders

were dispersed in ethanol. Then, the solution was dripped on off-axis cut, low

background silicon wafers (Bruker AXS, Germany). BG derived scaffolds were

powdered and prepared in the same way.

3.3.7 X-ray microtomography (µCT)6

µCT analysis was performed with a SkyScan 1172 (SkyScan, Kontich, Belgium).

Briefly, 45S5 glass-ceramic scaffolds were analyzed through a 360° flat-field

corrected scan at 40 kV and 250 µA, and then reconstructed (NRecon software,

SkyScan) with a beam hardening correction of 10, a ring artifact correction of 20

and an “auto” misalignment correction. The 2D analysis (software CTAn, SkyScan)

of reconstructed microCT transverse cross-sections of 45S5 glass-ceramic scaffolds

was carried out using a greyscale intensity range of 16 to 255 (8 bit images) in order

to remove background noise. 3D reconstruction and visualization of the scaffold

microstructure was achieved using CTVol software, Skyscan.

3.4 Scaffold fabrication

BG derived scaffolds were produced using the foam replica technique as

schematically shown in Fig. 9.68

Briefly, polyurethane foam (45 ppi, Recticel, UK)

was immersed in slurry containing 60 wt% BG-particles and 1.1 wt% PVA as

binder. The green bodies were dried at 60 °C for 24 h and sintered at 1050 °C for 2

h for glass series 45S5, 45S5-0.1Cu and 45S5-1Cu and 1000 °C for 45S5-2.5Cu

glass composition. The 1393-Co glass derived scaffolds were densified at 700 °C,

680 °C, and 670 °C for 1393. 1393-1Co and 1393-5Co, respectively.

Also, multiple coatings were applied by dip-coating the sintered foam and

centrifugation in order to remove the excess slurry.

6 µCT analysis was performed by B. Marelli at McGill Universtiy, Montreal, Canada

Page 55: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Materials and Methods 55

Fig. 9: Scheme of the foam replica technique used for scaffold fabrication.

3.5 Acellular bioactivity in simulated body fluid (SBF)

In order to assess the in vitro bioactivity, the scaffolds (5x5x5 mm3) were immersed

in 50ml simulated body fluid (SBF) for 1, 3 7, 14 and 21 days. SBF was produced

according to the protocol by Kokubo et al.99

After immersion, the samples were

gently washed with deionized water, dehydrated with acetone and dried at 60 °C for

12 h. Analysis of CHA formation was performed with SEM, FT-IR and micro-Ion

beam techniques.

3.6 Compressive strength

Compressive strength of the scaffolds (5x5x5 mm3) was measured using a tensile

testing machine (Z050, Zwick Roell, Germany). The testing speed was 10 mm min-

1. In order to assure homogenous loading of the scaffolds a polymeric rubber

interlayer was placed between the sample and the steel plates. 10 samples were

measured per scaffolds series and the standard deviation was derived.

3.7 Cell tests

In vitro cell investigations were carried out in order to assess the biocompatibility of

the developed Cu and Co containing bioactive glasses. Different cell types were

used in order to evaluate specific cell response analysing the cytotoxicity,

osteogenic differentiation and angiogenic potential. Hereby, 2D and 3D experiments

were conducted using glass powders, dense pellets and 3D scaffolds. Table 9 gives

Page 56: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Materials and Methods 56

an overview of cell types materials modifications used in the cell culture

experiments.

Table 9: Overview of the cell types and material modifications used for biocompatibility

assessment.

Cell type

Osteoblast-like

cells (MG-63)

Human bone marrow

derved stem cells

(hBMSC)

Human dermal

microvascular

endothelial cells

(hDMECs)

Co-culture of

hBMSC and

hDMECs

Aim Cytotoxicity Osteogenic

differentiation Angiogenic potential

Assessment Cell viability

Cell morphology

Cell vitality

Gene expression

Cell vitality, formation of tube-like

structures, VEGF release

Powder , - - -

2D pellets - - -

3D scaffolds

direct , - -

indirect - ,

...45S5-Cu, ...1393-Co

3.7.1 Cells and culture

Osteoblast-like cell line MG-63

MG-63 osteoblast-like cells Human osteosarcoma cell line (Sigma-Aldrich,

Germany) were used cultured at 37 °C in a humidified atmosphere of 95 % air and 5

% CO2, in DMEM (Dulbecco’s modified Eagle’s medium, Gibco, Germany)

containing 10 vol% fetal bovine serum (FBS, Sigma-Aldrich, Germany) and 1 vol%

penicillin/streptomycin (Sigma-Aldrich, Germany). Cells were grown for 48 hours

to confluence in 75 cm2 culture flasks (Nunc, Denmark), washed with phosphate

buffered saline (PBS) and harvested using Trypsin (Sigma, Germany). Cells were

counted by a hemocytometer (Roth, Germany) and diluted to a final concentration

of 100.000 cells ml-1

(if not specified otherwise).

Page 57: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Materials and Methods 57

Human bone-marrow mesenchymal stem cells (hBMSCs)

Human BMSCs were (at passage two) were purchased from PromoCell GmbH,

Heidelberg, Germany. They were cultured in flasks (COSTAR, Cambridge, USA)

by using MSC growth medium with supplemented cytokines as specified by the

manufacturer (PromoCell GmbH, Heidelberg, Germany), in an incubator with a

humidified atmosphere maintained at 37 ºC and 5% CO2. The media were changed

twice weekly. At 80-90% confluency, cells were trypsinized (Trypsin/EDTA, PAA,

Pasching, Austria) and sub-cultured further at 80% confluence until passage five.

For all the experimental protocols cells at passage 5 cells were used. For the 2D and

3D cell experiments the BMSCs were cultured in basal culture media

(DMEM/Ham’s F-12 (1:1) with 10% FCS, 2 mg l-1

of L-glutamine (Biochrom AG,

Berlin, Germany).

Human dermal microvascular endothelial cells (hDMECs)

hDMCs and the specialized endothelial cell growth media (EGM MV2) were

obtained from PromoCell GmbH, Heidelberg, Germany.

3.7.2 Analytical methods

Cell viability and cell proliferation

Mitochondrial activity.

The mitochondrial activity was assessed applying a WST-8 assay (Sigma-Aldrich).

After the time point of interest the culture media was removed from the respective

well culture plate and the cells were washed with PBS. After addition of the WST-8

reagents (1vol% in DMEM) in each well, the plates were incubated for 1.5 h.

Subsequently, the supernatant of all samples was transferred to a 98-well plate (100

µl in each well) to measure the absorbance at 450 nm and 650 nm with an ELISA-

Reader (Perkin Elmer, Multilabel Reader Enspire 2300, Germany).

Page 58: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Materials and Methods 58

LDH-activity

Lactate dehydrogenase (LDH-) activity provides a measurement of the amount of

attached cells on the samples. A commercially available LDH-activity quantification

kit (TOX7, Sigma-Aldrich) was used to quantify cell number by the LDH enzyme

activity in the cell lysate. Cells were washed with PBS and lysed with lysis buffer

for 30 min (1 mL/well). Lysates were centrifuged (5 min, 2500 rpm) and 100 µL

from the supernatant solutions were transferred to a 1 mL cuvettes. 100 µl of

master-mix containing equal amounts of substrate solution, dye solution, and

cofactor solution for LDH assay were added to each cuvette. After 30 min reaction

in the dark and the reaction was stopped with 300 ml 1M HCl per cuvette. The dye

concentration was measured on a spectrophotometer (Specord 40, Jena Analytik,

Germany) at 490 nm.

Metabolic activity

Each week, the cell-seeded scaffolds or the cells in 12-well plates were analyzed by

alamarBlue® (Biosource Int., Camarillo, CA, USA) assay, according to

manufacturer’s protocol. The absorbance of the reduced dye was measured by a

plate reader (SPECTRAmax 190, Molecular Devices, Sunnyvale, CA, USA) and

was calculated according to manufacturer’s instructions.

Analysis of osteogenic and angiogenic markers / Gene expression and protein

release

Specific Alkaline phosphatase (ALP) activity

ALP is a membrane-bound metalloenzyme which catalyzes the hydrolysis of

phosphomonoesters at an alkaline pH. For determining the osteoblastic activity of

the MG-63, ALP was analysed by measuring the specific enzyme activity. The cells

were lysed with a cell lysis buffer containing 20 mM TRIS buffered solution

(Merck) with 0.1 wt% Triton X-100 (Sigma, Germany), containing 1mM MgCl2

(Merck) and 0.1 mM ZnCl2 (Merck). The cell lysate was incubated with a reacting

solution containing 0.1M Tris solution, 2 mM MgCl2 and 9 mM p-Nitrophenyl-

phosphate. After incubation absorption was measured at 405 nm using a

Page 59: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Materials and Methods 59

spectrometer (Specord 40) after 200 min of incubation. The specific activity was

calculated with respect to the protein concentration of the cell lysates. The protein

content of the cell lysates was determined by means of a commercial kit based on

Bradford assay (Sigma).

Real time RT-PCR

Total RNA was isolated from the 2D cultured MSCs (n=4) using TRIzol Reagent

(Invitrogen, Carlsbad, CA, US) followed by RNeasy Mini Kit (Qiagen, Hilden,

Germany) as described elsewhere.266

Total RNA was converted to cDNA

(QuantiTect reverse transcription kit, Qiagen, Hilden, Germany). The amount of

cDNA corresponding to 20 ng of total RNA was then analyzed by semi-quantitative

real-time PCR (iQ SYBR green, Bio-Rad, Munich, Germany) for selected genes

with primers as shown in Appendix Fig. A 2 (CFX 96 real time systems, Bio-Rad,

Munich, Germany). The gene expressions were normalized to internal GAPDH

expression, and the relative fold change was expressed by comparing to that of

45S5 samples.

Cell morphology and cell adhesion

SEM analysis of the cells

After cell culture experiments the scaffolds were washed with PBS, fixed with a

solution containing 3 vol.% glutaraldehyde (Sigma, Germany) and 3 vol.%

paraformaldehyde (Sigma, Germany) in 0.2 M sodium cacodylate buffer (pH 7.4)

and finally rinsed three times with PBS For SEM analysis (ESEM, Quanta 200, FEI,

Netherlands). All samples were dehydrated in a graded ethanol series (30, 50, 75,

90, 95 and 99.8 vol.%). Samples were maintained at 99.8 vol.% ethanol and critical-

point dried (Leica, EM CPD 300). Prior to SEM examination the samples were

sputtered with gold.

Page 60: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Materials and Methods 60

Actin staining

In order to observe and to quantify the cell attachment on the BG samples labelling

of the cell skeleton (Actin labelling) was performed and the spreading behaviour of

the cells was analysed using fluorescence microscopy (Axio Scope.A1, Leica,

Germany). Actin staining was performed using a commercially available dye kit

(Alexa Fluor® 488 Phalloidin, Life Technologies GmbH, Darmstadt, Germany) at 5

units/ml concentration to see the cell spreading on the bio-active glass scaffold. The

nuclei of the cells are counterstained by 300 nM DAPI (SelectFX®, Life

Technologies GmbH, Darmstadt, Germany).

Vybrant™ cell-labeling

To analyse the adherent growth and distribution of cells on the BG scaffold

samples, commercially available Vybrant™ cell-labeling solution (Molecular

Probes, The Netherlands) was used. After 48 hours of incubation, cell culture

medium was removed and staining solution (5 μl dye labelling solution to 1 ml of

growth medium) was added and incubated for 15 min. Afterwards the solution was

removed, the samples were washed with PBS (phosphate buffered saline, Gibco)

and cells on the BG samples were fixed by 3.7 vol.% paraformaldehyde. The

samples were prepared for confocal scanning laser microscopy (CSLM, Leica TCS

SP5 II, Germany) to analyse cell morphology and distribution.

Evaluation techniques of 2D experiments with EC

Flow Cytometry analysis (FACS)

The human endothelial cell surface markers (Cluster of differentiation, CD31; von

Willebrand factor, vWF, Vascular endothelial growth factor 2, VEGFR2) were

stained at 5x104 cells for each antigen. CD31 is a known platelet adhesion protein

which is expressed at cell connecting junctions and is found in early and mature

endothelial cells and is known to be involved in activation of integrins. CD31 was

stained by mouse anti-human CD31-Biotin followed by a second staining step with

Streptavidin PerCP-eFluor® 710 (both from eBioscience, San Diego, CA, USA).

Page 61: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Materials and Methods 61

vWF is an important platelet adhesion factor which is also used as a marker for any

pathological dysfunctions of endothelial cells as increased release of vWF from

endothelial cells is related to endothelia cells damage. vWF was stained by sheep

anti-human vWF-FITC (Abcam, Cambridge, UK). VEGFR2, an important

signalling molecule for endothelial cell mitogenesis and migration, was stained by

mouse anti-human CD309 (VEGFR2)-Alexa Fluor® 647 (Biolegend, San Diego,

CA, USA). All staining steps were performed on ice for 30 minutes in dark. Stained

cells were analyzed by FACS-Calibur (BD Biosciences, San Diego, CA, USA) and

Cell Quest software (Beckton Dickinson, Heidelberg, Germany). Data

quantifications were perfomred by FlowJo software (Tree Star, Inc., Ashland, OR,

USA).

Matrigel™ sprouting assay

For analysis of capillary tube formation, 75 µl of Matrigel™ (Becton Dickinson,

Heidelberg, Germany), an extracellular mouse sarcoma matrix known to produce

pro-angiogenetic stimulus both in vitro and in vivo, was pipetted into each well of a

96-well plate (Falcon, Heidelberg, Germany) and incubated at 37°C for 60 minutes.

HDMECs were harvested at week 1 or week 2 and suspended at 50,000 cells per

150 µl of EGM MV2 media. 150 µl of this media were added to the Matrigel coated

96-well plates and incubated for 24 h at 37°C. Capillary tube formation on Matrigel

was observed with a light microscope (DMIL, Leica, Germany) and images were

processed with Leica application suite software (Leica GmbH, Wetzlar, Germany).

LDL (Low density lipoprotein)-staining

At end time points, the media were removed and the cells were washed once with

PBS to remove non-adherent cells. The wells were incubated with 2.5 µg/ml ac-

LDL-Alexa Fluor®

488 (Life Technologies GmbH, Darmstadt, Germany) for 4 h at

37°C. Afterwards cells were washed twice with PBS and further analysed by a

fluorescence microscope.

Page 62: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Materials and Methods 62

VEGF quantification ELISA

At each time point, the cultured media were collected from 3 different samples,

pooled together, labeled and frozen at -20°C. At the end of all experiments, the

frozen media were thawed overnight at 4°C and the VEGF content of the media was

quantified by using an ELISA kit (Thermo Fisher Scientific GmbH, Schwerte,

Germany).

Statistical analysis

The statistical significant of the results was evaluated by one-way analysis of

variance (ANOVA). The level of the statistical significance was defined at p < 0.05

(Origin 8.1G, OriginLab Corporations, USA).

3.7.3 Powder cytotoxicity

Initial cytotoxicity tests of the glasses fabricated were assessed by culturing BG

powders in direct contact with MG-63 cells. The BG powder was dispersed in

DMEM and ultra-sounded for 5 min to break agglomerates. MG-63 cells were

cultured in direct contact with BG particles at concentrations from 0.1 to 1000 µg

ml-1

) in 600 µl DMEM for 48 hours. As positive control cells cultured without BG

particles, as negative control ZnO was used.

Cell distribution and morphology were evaluated using phase contrast light

microscopy (LM, Nikon Eclipse TE 2000-U, Japan). Cells viability was assessed

through WST-8 test (mitochondrial activity) and LDH assay (cell number) after 48h

of incubation. From these the LC50 value (lethal concentration of the used particles

where the activity of the cells is reduced to 50%) was derived.

Page 63: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Materials and Methods 63

3.7.4 Evaluation of the scaffolds7

Indirect study (2D)

In the indirect setup the cells were seeded on the bottom of the cell culture well

exposed to the ionic dissolution products released from the scaffold which was

placed in a permeable insert as shown in Fig. 10. Indirect studies were performed

with hBMSCs. Cell morphology was detected by light microscopy, ALP activity,

VEGF and Runx2 expression was analysed with PCR technique.

Fig. 10: Indirect setup for cell culture studies with scaffolds.

Direct study (3D)

Additionally, in order to test the interaction of cells and the scaffold surface the cells

were directly seeded on the BG derived scaffolds. MG-63 cells and hBMSCs were

used. Before cell seeding the scaffolds were preconditioned in DMEM for 5d. The

cells attachment and growth was evaluated with SEM, Actin staining, Live staining

(Calcein/Vybrant). Further, ALP activity was detected.

Fig. 11: Direct setup for cell culture studies with scaffolds.

Co-culture of hBMSCs and ECs

A co-culture model of hDMECs and hBMSCs was applied to 45S5-Cu derived

scaffolds as shown in Fig. 11. hBMSCc were seeded directly on the scaffolds and

7 Cell studies with hMBSCs were carried out in collaboration with D. Hiller and S. Rath, U. Kneser and R. Horch, Plastic and Handsurgery Department, University Clinic Erlangen

Page 64: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Materials and Methods 64

the constructs where placed into permeable membranes. hDMECS were seeded in a

concentration of 100.000 cells/ml and on the well bottom being exposed to ionic

dissolution products released from the scaffods as well as effects of the hBMSCs.

Cell morphology of the hDMECS was observed with light microscopy. Further,

HDMSCs specifically the LDL uptake and FACS analysis of endothelial markers

vWF, VEGFR and CD31 was performed.

Fig. 12: Set-up for the co-culture experiment with hBMSc and hDMECSs using permeable

insert containing the scaffold seeded with hBMSCs.8

3.8 In vivo study

In vivo evaluation of the angiogenic potential of Cu-doped glass scaffolds was

carried out at the Department for Hand plastic surgery, University of Erlangen9.

Six male Lewis rats were used and the experiments were carried out in compliance

with the animal care committee of the FAU and the government of Mittelfranken.

45S5-1Cu BG derived scaffolds were chosen as most promising candidates for in

vivo study, to prove their angiogenic potential in relevant model in addition to in

vitro results. Fabricated scaffold were crushed into small pieces with a diameter of ~

2-3 mm in order to create granulate like units. A Teflon chamber with an inner

diameter of 10 mm and a height 10 mm was used. The half of the chamber was first

filled scaffold granules and fibrin gel with a fibrinogen concentration of 10mg ml-1

and thrombin concentration of I.U ml-1

. Afterwards the AV loop was placed on the

BG/fibrin gel matrix and the chamber was completely filled with BG granules/fibrin

gel. The construct was, then, implanted subcutaneously and sutured to the

musculature of the rats. The arrangement of the AV-loop and the BG granules in the

8 Cell studies with hDMECs were carried out in collaboration with A. Brandl and O. Bleizifffer, U. Kneser and R. Horch, Plastic and Handsurgery Department, University Clinic Erlangen 9 In vivo studies were carried out by Ulrike Rothensteiner from the group by Dr. A. Arkudas, Dr. U. Kneser and Prof. Dr. Horch

Page 65: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Materials and Methods 65

teflon chamber are shown in Fig. 13. After 3 weeks of implantation the constructs

were explanted and were investigated by means of microCT, histological and

morphometrical analysis. 3D reconstruction of the vessel sprouting, vascular

density as wells as vessel number and vessel density were derived were derived.

Fig. 13: In vivo setup for assessing the angiogenesis of 45S5 and 45S5-1Cu BG scaffolds. Firstly,

the teflon chamber is filled half with BG granules and the AV-loop is placed (a). Afterwards the

chamber is fully filled with BG granules and implanted subcutaneously (b).

Page 66: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 66

4 Results and Discussion

In this chapter the properties of different bioactive glass compositions and the

respective glass-derived scaffolds are described. The influence of metal ion doping

on the glass structure and glass properties is discussed. Further the structural and

mechanical properties of the glass derived scaffolds are shown. The apatite forming

ability as indicator for acellular bioactivity is assed in simulated body fluid and the

physic-chemical reactions are monitored in detail by SEM, FT-IR and RBS-PIXE

techniques. Moreover, the ion release kinetics from the BG scaffolds is observed

and the suitability of glass derived scaffolds as “carrier” of therapeutic metal ions is

discussed. The cellular biocompatibility is analysed in relevant cell culture studies

and also an in vivo model is applied.

4.1 Cu containing 45S5 bioactive glasses

4.1.1 Glass properties

In this section the glass structure of the synthesised Cu containing glasses is

discussed on the basis of FT-IR, Raman and NMR-analyses and is related to glass

thermal properties derived from DSC measurements.

Glass structure

Fig. 14 shows the FT-IR spectra of as-fabricated glass powder. The bands at ~ 500

cm-1

and ~1080 cm-1

can be assigned to Si-O-Si stretching and Si-O-Si bending

modes, respectively.267

The peak at 1460 cm-1

is related to carbonate species

adsorbed on the BG surface.268

No major differences for Cu-containing glasses can

be observed compared to undoped reference 45S5 glass indicating that the main

structure of the glass network remains unchanged. However, the intensity of the

non-bridging oxygen, Si-ONBO peak (~930 cm-1

), is decreasing with increasing Cu-

content in the 45S5 glass.

Similar observations can be made from Raman analysis, Fig. 15. The bands at ~620

cm-1

and 1060 cm-1

can be assigned to the rocking and stretching mode of Si-O-Si

Page 67: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 67

band, respectively. While no major differences are observed for Cu-containing

glasses the band at 860 cm-1

which is assigned to non-bridging oxygen-silica band

Si-O2NBO is reduced in intensity for the 45S5-2.5Cu. Also the Si-O band at 1060 cm-

1 increases in intensity for the glasses with higher Cu content (>1%).

Fig. 14: FT-IR spectra of as-fabricated Cu-doped 45S5 glasses. A. Hoppe et al. J. Mater. Chem.

B 1 (2013), p. 5659. - Reproduced by permission of The Royal Society of Chemistry.

Fig. 15: Raman spectra of as-fabricated Cu-doped 45S5 glasses. A. Hoppe et al. J. Mater.

Chem. B 1 (2013), p. 5659. - Reproduced by permission of The Royal Society of Chemistry.

Page 68: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 68

Fig. 16 shows the UV-VIS spectra of the glasses. For reference 45S5 glass a

continuous transmission in the range from 300 to 2000 nm is observed. Cu

containing glasses 45S5-0.1Cu, 45S5-1Cu and 45S5-2.5Cu, however, show an

absorption band at 800 nm which can be assigned to the presence of Cu2+

in the

glass network.269, 270

Fig. 16: UV-VIS spectra of 45S5-Cu BGs showing an absorption band at ~800 nm which is

assigned to Cu2+

. A. Hoppe et al. J. Mater. Chem. B 1 (2013), p. 5659. - Reproduced by

permission of The Royal Society of Chemistry.

For 45S5-0.1Cu no NMR measurements were performed since the Cu content is too

low and its influence on the 45S5 glass network structure cannot be resolved by

NMR. Fig. 17 shows the 29

Si MAS NMR spectra of the 45S5and 45S5-2.5Cu

glasses. The effect of Cu on 45S5 silicate glass structure is discussed on the basis of

45S5-2.5Cu data only as the NMR spectra for 45S5-1Cu (data not shown) and

45S5-2.5Cu show almost exactly the same characteristics and are completely

overlapped. An asymmetrical peak is observed for both glasses with the overall

peak position of the Si-related signal slightly shifted to negative values for 45S5-

2.5Cu glasses by 0.75 ppm. The NMR results show a broadening of the 29

Si peak.

Compared to reference 45S5 sample the FWHM (full width at half maximum) of

the Si signal for 45S5-2.5 Cu increased to 13.8 ppm (+ 1ppm).

Page 69: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 69

Table 10: Distribution of Qn units calculated from the peak area of the deconvoluted

29Si NMR

signal and network connectivity NC for the 45S5 and 45S5-2.5Cu samples. A. Hoppe et al. J.

Mater. Chem. B 1 (2013), p. 5659. - Reproduced by permission of The Royal Society of

Chemistry.

45S5 45S5-2.5Cu

Q1 26.9% 12.7%

Q2 39.5% 47.6%

Q3 33.1% 38.9%

NC* 1.99 2.17

Commonly the structure of silicate glasses is described by quantifying the Qn

distribution (a Q species is a Si atom with n bridging oxygen atoms). Hence, for

detailed analysis of the NMR results for the 29

Si NMR signal of 45S5 and 45S5-2.5

Cu the curves were fitted and peak deconvolution applying Gauss profiles was

performed, Fig. 17. The glass structure of both glasses is dominated by Q2 with

smaller fractions of Q1 and Q

3 species as depicted in Fig. 17. For 45S5-2.5Cu a

slightly larger fractions of Q2 and Q

3 species is observed at the expense of Q

1. The

fractions of Qn and resulting network connectivity (NC) can be calculated from the

peak area. The exact values for the Qn distribution obtained for 45S5 and 45S5-

2.5Cu and corresponding values for the network connectivity (NC) are given in

Table 10. NC was calculated according to:

𝑁𝐶 = 1𝑄1 + 2𝑄2 + 3𝑄3 Eq. 4

Page 70: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 70

Fig. 17: MAS NMR 29

Si analysis for 45S5 and 45S5-2.5Cu glass samples. Gauss fit and peak

deconvolution show the distribution of the Qn species in the glass network. A. Hoppe et al. J.

Mater. Chem. B 1 (2013), p. 5659. - Reproduced by permission of The Royal Society of

Chemistry.

Thermal properties

Fig. 18 shows the DSC curves for the 45S5-Cu glasses as fabricated. The

characteristic temperatures of glass transition Tg, crystallization onset and maximum

Tc and Tp and melting temperature Tm are given in Table 11. Both glasses show two

crystallisation events described by Tc1, Tc2 and Tp1, Tp2, respectively. Compared to

the 45S5 reference, for 45S5-2.5Cu a decrease of Tg, Tc1, Tp1 and Tm is observed

while Tc2 increases and Tp2 remains constant. For the lower Cu concentrations,

however, no significant differences in the DSC diagram for the characteristic

Page 71: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 71

temperatures were observed. The origin of the crystalline phases occurring during

the sintering of 45S5 bioactive glass is discussed in 4.1.2.

Table 11: Characteristic features (temperatures in °C) of the Cu-containing glasses observed

with DSC.

Glass Tg Tc1 Tc1-Tg Tp1 Tc2 Tp2 Tm

45S5 522 592 70 682 756 908 1105

45S5-0.1Cu 526 611 85 696 - - 1110

45S5-1Cu 519 610 91 698 783 832 1112

45S5-2.5Cu 496 610 114 645 806 908 1044

Fig. 18: DSC analysis of 45S5-Cu glasses showing the glass transition point and two

crystallisation events at ~650-700 °C and 850 °C, respectively. A. Hoppe et al. J. Mater. Chem.

B 1 (2013), p. 5659. - Reproduced by permission of The Royal Society of Chemistry.

Role of CuO in 45S5 glass structure

The 45S5-1Cu and 45S5-2.5Cu glasses show almost the same characteristics.

Hence, 45S5-2.5Cu glass is used for deeper discussion of the effect of Cu doping on

the structure of 45S5 silicate glass.

FT-IR analysis, Raman spectroscopy, NMR and UV-VIS spectroscopy were applied

in order to investigate the glass structure and the effect of Cu-doping. It was

confirmed through UV-VIS spectroscopy that Cu is present as Cu2+

in the glass

Page 72: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 72

structure which has been also proposed for soda silicate glasses containing CuO

showing Cu2+

valence state of Cu.271

Cu2+

(as well as Cu+) predominantly work as

network modifiers and are incorporated in the glass work matrix in octahedral

coordination.272

According to Abdrakhmanov et al.273

Cu2+

is surrounded by two

non-bridging oxygens (NBO) in order to achieve electro neutrality. Since no major

differences were observed in FT-IR and Raman spectroscopy the Q2 SiO2 network

of the 45S5 glass seems to be dominant and is not significantly affected by the Cu-

doping. Basically, according to FT-IR and Raman analyses a decreased intensity of

the Si-ONBO bond with increasing Cu content and at the same time increased

intensity of asymmetric Si-O vibration mode for glass with high Cu contents in the

glass was observed.

Si-NMR is a well-established experimental technique for characterising the

structure of silicate glasses which allows detailed insight about the distribution of

Qn species. As described in 2.3.2 the best descriptive model describing the structure

of 45S5 BG is achieved by assuming a trinodal distribution of Q1, Q

2 and Q

3 silica

species. According to this for the experimental NMR data the best fit of the 29

Si

peak (R2=0.99) was achieved by applying a ternary peak deconvolution, Fig. 17.

Concerning the 29

Si MAS-NMR analysis a shift of the peak maximum was observed

for 45S5-2.5Cu compared to 45S5 control. According to Elgayar et al.274

the

negative shift of the maximum to lower frequencies of 0.75 ppm between the 45S5

and 45S5-2.5Cu is likely caused by changing the cation in the glass structure which

is accomplished with an increase of the Q2 species in the glass network neutralized

with Cu2+

ions. Furthermore, the increased line width of the 29

Si peak indicate more

structurally disordered Q2 and Q

3 species neutralized by Cu

2+.274

This might be due

to the larger ion radius of Cu2+

replacing Ca2+

.

The higher fractions of Q2 and Q

3 species and hence the resulting higher network

connectivity in 45S5-2.5Cu samples suggests a repolymerisation effect caused by

introducing CuO into the glass network. Cu are also bigger ion, so the SiO2 network

is disrupted, a higher amount of silica chains Q2 are present and also there is

indication that the chains are repolymerised which results in higher Q3 fractions and

less terminating Q1

species. This is likely due to the more covalent character of the

Cu-O bond compared to Ca-O bond allowing repolymerisation of Si-NBOs.

Page 73: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 73

Accordingly, FT-IR and Raman spectra revealed decreased intensity of the NBO

peak which is consistent with higher Q2 and Q

3 fraction and thus lower amounts of

free NBO in the glass network.

Fig. 19: Model of the Cu-containing 45S5 BG structure. Cu2+

(blue) replace Ca2+

ions (yellow)

as network modifiers.

Even though the main glass structure remains similar consisting mainly of Q2 and

Q3 units, the NMR results show higher Q

2 and Q

3 fractions for the 45S5-2.5Cu glass

which indicates higher silicate network distortion. The weakening of the glass

matrix is confirmed by the DSC measurement which revealed a decrease in glass

transition temperature for Cu containing 45S5 glasses. Incorporation of Cu into the

45S5 glass matrix was found to reduce the Tg of 45S5 glass and to stabilize the

amorphous phase during sintering.

The ionicity, iG, of the Me-O bonds within the glass structure can be used as

explanation of the decrease of Tg. With increasing Cu content in the glass the Tg

decreases likely due to the more covalent character of the Cu-O (iG=0.617) bond

compared to Ca-O (iG=0.707) bond resulting in higher degree of structure relaxation

and hence lower glass transition temperature.275

During heat treatment the 45S5 glass tends to crystallise and two crystalline phases,

combeite and silicorhenanite, form which have been described in literature for this

glass.276, 277

Page 74: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 74

4.1.2 Scaffold properties

Macro-Structure

Scaffold macro and microstructure was analysed with SEM and µCT. The porosity

was derived by the Archimedes method.

Fig. 20a-d show the macrostructure of 45S5-Cu derived scaffolds. High pore

interconnectivity as well as pore sizes of ~ 200-300 µm were observed for all glass

compositions. Porosity values of 92 %, 91 %, 92 % and 93% and were observed for

45S5, 45S5-0.1Cu, 45S5-1Cu and 45S5-2.5Cu respectively. Fig. 20e) shows

additionally high magnifications of a 45S5 derived strut showing a hollow strut as it

is typical for polyurethane foam derived scaffolds. Furthermore, Fig. 20f shows the

3D reconstruction of a 45S5 derived scaffold from µCT data confirming the

interconnected pore system. The total porosity was calculated to be 93.5 ±2.0 % and

the strut thickness of ~67 µm was derived. The average pore size was 314 ± 87 m.

99.1±0.4 % of the total porosity was open indicating a completely interconnected

pore system. High porosity values and interconnected pore system of the scaffold

should enable vascularisation and tissue ingrowth when applied as engineered bone

construct. Vascularisation has been shown to be enhanced in scaffolds with pores >

250 um. 278

. However, high interconnectivity is considered even more important for

blood vessel and tissue ingrowth.43, 279

Hence, 45S5 derived scaffolds fulfil these

requirements.

Page 75: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 75

Fig. 20: SEM images of 45S5-Cu derived scaffolds: a) 45S5, b) 45S%-0.1Cu, c) 45S5-1Cu and

d) 45S5-2.5Cu. e) typical hollow strut of 45S5 derived scaffolds and f) microCT reconstruction

of a 45S5 derived scaffold. A. Hoppe et al. J. Mater. Chem. B 1 (2013), p. 5659. - Adapted by

permission of The Royal Society of Chemistry.

Fig. 21 shows the XRD analysis of the 45S5-Cu derived scaffolds after sintering.

Two crystalline phases were observed after the heat treatment of the glass: a sodium

calcium silicate phase (combeite, Na2Ca2Si3O9) and silicorhenanite

(Na2Ca4(PO4)2SiO4) which have been shown to occur during high temperature

treatment of Bioglass®.67, 276, 280

Basically, no qualitative differences among the Cu-

containing and 45S5 reference were observed. All glasses show high crystallinity

after sintering. However, Rietveld analysis was performed in order to quantify the

crystalline phases. Since the sensitivity of the XRD technique is limited only 45S5

with the highest copper content (45S5-2.5Cu) was investigated in order to

qualitatively evaluate the role of copper during the crystallisation of 45S5 glass. The

Page 76: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 76

amount of the amorphous phase was higher for 45S5-2.5 Cu compared to 45S5

reference. The content for the amorphous phase was calculated to 3.5 ±2.2 and 10.5

±2.8 % for 45S5 and 45S5-2.5Cu, respectively.

Fig. 21: XRD analysis of 45S5-Cu derived scaffolds showing two crystalline phases forming

during heat treatment. A. Hoppe et al. J. Mater. Chem. B 1 (2013), p. 5659. - Reproduced by

permission of The Royal Society of Chemistry.

Mechanical properties

For the 45S5-Cu glass series no differences in the compressive strength was

observed between the different glass compositions (data not shown) likely due to

the low amount of CuO presented in the glass. Fig. 22 summarises the mechanical

properties of the 45S5 derived scaffolds for multiple coatings regimes. Fig. 22a and

Fig. 22b shows show the mean values and typical stress-strain curves, respectively,

for 45S5 after multiple coatings with BG slurry were applied to the polyurethane

template. Maximal values of 0.16 ±0.07 MPa were observed for 45S5 scaffolds

after application of 3rd

coating with BG slurry.

Page 77: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 77

Fig. 22: Mechanical properties of the 45S5 BG derived scaffolds: a) mean values calculated

from 10 measurements per scaffold; b) selected typical stress-displacement curves.

The enhancement of the compressive strength through multiple coating of the foam

is not significant and the overall mechanical strength of the scaffolds is rather poor.

The values observed here are in agreement with literature reports where

compressive strength values of 0.3-0.4 MPa (but with lower porosity compared to

scaffolds prepared in this work) were observed for 45S5 derived scaffolds by foam

replica technique. For clinical application, however, the mechanical properties of

the scaffolds should also be considered. The mechanical properties of the 45S5

derived scaffolds are discussed in the context of bone tissue engineering in 0.

Acellular bioactivity in SBF

Despite recent critical discussion in the literature regarding the suitability of SBF

studies for predicting the bioactivity,101, 102

this well-known and established

technique was employed in this work since the objective was to establish a close

comparison with non-doped 45S5 Bioglass® which have been largely investigated

by the SBF test as well as in vitro cell culture studies in the last decades.

The reaction stages of 45S5 bioactive glass derived scaffolds, as documented by

SEM, are summarised in Fig. 23. Fig. 23a shows the typical hollow macrostructure

of the initial 45S5 BG scaffolds strut. After 1d of immersion in SBF, Fig. 23b, the

surface is homogenously covered with calcium phosphate (CaP) precipitates as

indicated through their typical morphology. A higher magnification images of the

cross-section of a scaffold strut, Fig. 23c-d reveal a more detailed view on the

Page 78: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 78

reaction stages. Fig. 23c shows a porous structure of roughly 8-10 µm which is

covered by a thinner dense layer (roughly estimated to be ~1-2 µm). These two

layer s are likely to be identified as a silica rich layer and CaP layer, respectively.

However, at higher magnification according to Fig. 23d no clear discrimination of

the silica and the CaP layer can be made. The CaP enrichment on the BG scaffold

surface is further indicated through small round-shaped particles marked as ACP

precipitates in Fig. 23 c). After 3 d in SBF, the phosphate layer continues to grow

reaching 3-5 µm after 3d (Fig. 23d). After 7 days only minor further growth of the

CaP layer was observed which remained roughly at 5 µm (Fig. 23e). Furthermore,

the inner region of the scaffold (inner part of the struts) shows signs of dissolution

turning into a porous silica structure with the modifier ions Ca and Na leached out.

After 7 days the entire scaffold strut seems to have transformed into pure silica

phase.

However, the analysis of the reaction layers on 45S5 bioactive glass scaffold and

identifying surface reaction layer with SEM is rather speculative. Hence, FT-IR

analysis and micro-PIXE-RBS analysis are presented in the following. Fig. 24

shows the FT-IR spectra of a 45S5 BG derived scaffold after immersion in SBF for

1d, 3d, and 7d. As-fabricated 45S5 scaffold shows bands at 629 cm-1

, 575 cm-1

and

530 cm-1

which can be attributed to apatite-like crystalline phase as the P-O bending

modes are located at 580 and 620 cm-1

.281

This crystalline phase corresponds to the

XRD results showing silicoarhenite (Na2Ca4(PO4)2SiO4) phase which is

isostructural to apatite.277

The bands at 450 and 930 cm-1

can be assigned to

symmetric Si-O-Si vibrations and to non-bridging oxygen Si-ONBO, respectively.267,

282 In contrast to the amorphous glass powder (Fig. 14) the broad silicate band at

~1020 cm-1

is split into two bands which are related to the vibrations of isolated Si

tetrahedral.67

Page 79: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 79

Fig. 23: Reaction stages of HAp formation on crystallized 45S5-0.1Cu derived scaffold: a)

initial strut; b) scaffold after 1 day in SBF; c) and d) scaffold after 1 day in SBF at higher

magnification showing the formation of CaP-SixOy layer and amorphous calcium phosphate

(ACP) precipitates; e) CHA formation and f) further growth of carbonated hydroxyapatite

(CHA) on BG scaffold surface. A. Hoppe et al. J. Mater. Chem. B 1 (2013), p. 5659. -

Reproduced by permission of The Royal Society of Chemistry.

During treatment in SBF, the following changes in the FT-IR spectra occur:

- (after 1d): the P-O bands resulting from the crystalline, rhenanite-like phase

in the BG scaffold transform to a broad peak at 600 cm-1

which is typical for

amorphous calcium phosphate phase (ACP).283

The sharp peak of the Si-O-

Si band at 450 cm-1

is reduced in intensity and becomes broader indicating

the formation of a phosphate phase as O-P-O (bending mode) band absorbs

in this region.284

The shoulder at 960 cm-1

can be attributed to HPO42-

units

incorporated in the calcium phosphate phase.284

Furthermore, a broad peak

appears at 800 cm-1

which corresponds to silica.91

Page 80: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 80

- (after 3d): with increasing immersion time in SBFa a sharp double band was

detected at 560 cm-1

and 600 cm-1

which is assigned to the vibration modes

of the PO43-

groups284

typically observed in crystalline HAp

- (after 7d): the HAp bands increase in intensity and, additionally, CO32-

band

at 870 cm-1

appears which are observed for natural carbonated

hydroxyapatite (CHA)284

but, however, could also be attributed to surface

bonded carbonate compounds.91

Also, double peaks at 1430 and 1500 cm-1

are observed which can be assigned to v3

(CO32-

) of carbonate ions

incorporated in CHA.284

Fig. 24: Evolution of FT-IR spectra of 45S5 derived scaffold during immersion in SBF for 1d,

3d and 7d. A. Hoppe et al. J. Mater. Chem. B 1 (2013), p. 5659. - Reproduced by permission of

The Royal Society of Chemistry.

FTIR-analysis did not reveal any effect of Cu on the CHA formation. Fig. 25 shows

the FTIR-spectra for 45S5-Cu derived scaffolds after 3d. Clear formation of CHA

was detected after 3 days of immersion in SBF independently of Cu content in the

glasses.

Considering the time point of appearance of the CHA layer as marker for bioactivity

one can conclude that the bioactivity of the 45S5 derived scaffolds is not impeded

Page 81: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 81

by the incorporation of Cu. In order to gain more detailed information about the

nature of the CHA layer formed on Cu.

Ion beam method allows a chemical analysis with an excellent sensitivity of several

ppm due to very good signal to background ratio. Compared to other techniques like

SEM/EDS, micro-PIXE-RBS method allows an improvement of the sensitivity up

to 3 orders of magnitude. This is a great advantage to study the distribution and the

role of relevant bone trace elements. In addition to other conventional methods

PIXE-RBS analysis enables to detect the local chemical composition of the reaction

surface layer.

Fig. 25: FT-IR spectra of Cu-containing scaffolds after 3d of immersion in SBF compared to

reference 45S5 glass. A. Hoppe et al. J. Mater. Chem. B 1 (2013), p. 5659. - Reproduced by

permission of The Royal Society of Chemistry.

Chemical maps were acquired for the distribution of Si, Ca, P, (Na is not shown for

simplicity) and Cu in the fabricated Cu-containing scaffolds. 45S5-0.1Cu serves as

representative example for the element distributions in the inner part of the scaffold

and the periphery layer and for the elemental evolution during immersion in SBF.

The glass composition with the lowest Cu amount was also chosen in order to

confirm high sensitivity of the micro ion beam technique enabling monitoring the

evolution of low concentrations of trace elements during the reaction of a bioactive

Page 82: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 82

glass scaffold. Fig. 26 shows typical multichemical elemental maps for Si, Ca, P,

and Cu for an as-fabricated 45S5-0.1Cu cross section (Na is not shown for

simplicity). All elements are homogenously distributed.

Fig. 26: Element distribution of a 45S5-0.1Cu scaffold cross-section in the as-fabricated stage

derived from micro-PIXE-RBS measurements. A. Hoppe et al. J. Mater. Chem. B 1 (2013), p.

5659. - Reproduced by permission of The Royal Society of Chemistry.

Fig. 27-Fig. 29 are representative elemental maps of the scaffolds after immersion

in SBF for different time periods. In correlation to SEM/FT-IR results, during

immersion in SBF the following different regions can be distinguished from the

chemical maps:

i) primary bioactive glass network

ii) silica rich layer on the surface of the scaffolds (clearly distinguishable after 1d)

iii) Ca-P rich layer in the scaffold periphery

For the areas i) and iii) additionally the elemental concentrations were quantified in

these two distinct regions of interests. The elemental evolution for all 45S5-Cu

glass derived scaffolds as function of immersion time in SBF is shown in Fig. 30

and Fig. 31 for the inner region (i) and the periphery region (iii), respectively.

Page 83: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 83

Fig. 27: Element distribution of a 45S5-0.1Cu scaffold after 1 day in SBF monitoring the

formation of a calcium phosphate layer. A. Hoppe et al. J. Mater. Chem. B 1 (2013), p. 5659. -

Reproduced by permission of The Royal Society of Chemistry.

Fig. 28: Element distribution of a 45S5-0.1Cu scaffolds after 3 days in SBF monitoring the

formation of a calcium phosphate layer. A. Hoppe et al. J. Mater. Chem. B 1 (2013), p. 5659. -

Reproduced by permission of The Royal Society of Chemistry.

Page 84: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 84

After 1 day, a silica rich layer is detected at the scaffold surface and Ca and P

enrichment is observed as shown in Fig. 27. The silicon content increased from 21

wt% in the inner part of the scaffold to 33 wt% in the silicon enriched layer (12.0 ±

0.2 µm). The Ca/P ratio for the (iii) periphery layer was determined as 1.3 ± 0.1 µm.

The P layer is rather distinct whereas Ca traces can be also found in the Si-rich

layer. This corresponds to the SEM observation where no clearly distinguishable

CaP but rather a mixed Ca-P-SiO2 layer was detected.

After 3d, Fig. 28, a CaP layer was observed on the pore surface with a Ca/P ratio of

1.88 ± 0.06, Fig. 31. The layer thickness was determined to 10.5 ±0.9 µm. The inner

parts of the BG scaffolds showed further dissolution and depletion in Ca and P. At

the same time the relative amount of silicon present within the glass matrix

increases since Ca, P and Na are leached out. After further immersion in SBF for 7d

no significant changes occured: the Ca-P rich layer thickness remains nearly

constant (10.9 ±1.6 µm), Fig. 29, while the Ca/P ratio slightly increases to 1.9 ±0.1.

Fig. 29: Element distribution of a 45S5-0.1 scaffold after 7 days in SBF showing further growth

the calcium phosphate layer. A. Hoppe et al. J. Mater. Chem. B 1 (2013), p. 5659. - Reproduced

by permission of The Royal Society of Chemistry.

Page 85: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 85

Elemental concentrations for all 45S5-Cu derived scaffolds were calculated in two

areas, the inner area (i) of a strut and the outer layer (surface periphery (iii)).

Hereby, the outer layer was defined as the area of interest once the Ca and P rich

peripheral layer has formed. The evolution of the element concentration during

immersion in SBF is shown in Fig. 30 and Fig. 31 for the inner part and the

peripheral layer, respectively. Si release from the periphery was slowed down for

the 45S5 and 45S5-0.1Cu scaffolds as compared to 45S5-1Cu and 45S5-2.5Cu

indicating slower surface degradation kinetics of these glasses. Accordingly, Ca and

P enrichment is also slower for 45S5 and 45S5-0.1Cu. With higher Cu contents in

the glass (≥ 1wt.-%) Ca and P seem to diffuse faster through the silica layer and are

faster accumulated in the scaffold periphery.

Fig. 30: Evolution if elements in the inner region of the 45S5-Cu derived scaffolds as function

of immersion time in SBF. Lines are for eyes guidance only. A. Hoppe et al. J. Mater. Chem. B

1 (2013), p. 5659. - Reproduced by permission of The Royal Society of Chemistry.

Page 86: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 86

However, the evolution of Ca and P in the inner region of the scaffolds, Fig. 30, is

not significantly different among the glasses indicating that the global dissolution of

the glass is independent of the Cu content in the glass.

The kinetics for Na release from the scaffolds was found to be similar for all 45S5-

Cu derived scaffolds. However, for higher Cu contents (45S5-1Cu and 45S5-2.5Cu)

small amounts of Na can be detected even after 7d of immersion in SBF (in the

inner part of the scaffold), whereby no Na is detectable for 45S5 reference and

45S5-0.1Cu.

Fig. 31: Evolution if elements in the periphery layer of the 45S5-Cu derived scaffolds as

function of immersion time in SBF. Lines are for eyes guidance only. A. Hoppe et al. J. Mater.

Chem. B 1 (2013), p. 5659. - Reproduced by permission of The Royal Society of Chemistry.

Page 87: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 87

Altogether the mechanism for apatite formation on 45S5-Cu derived scaffolds can

be described as follows:

Basically, the present observations confirm the reaction mechanisms of

(amorphous) bioactive 45S5 glass as originally proposed by Hench et al.9 and

Kokubo.285

For Cu containing 45S5 glasses, the relevant physico-chemical reactions

were shown exemplary for the 45S5-0.1Cu composition. However, since no

significant differences in the reaction within the Cu-glass series were detected by

FT-IR (Fig. 25) it is possible to assume that the proposed model is valid also for the

other Cu-containing glass composition.

(1d-3 d): Initial reaction of the BG surface and release of Na+ and Ca

2+ from the

scaffold periphery in an exchange reacting with H+ from the solution according to

Eq. 5.13, 286

. Free Si-NBO are protonated forming silanols groups.

𝑆𝑖 − 𝑂−𝑁𝑎+ + 𝐻2𝑂 → 𝑆𝑖 − 𝑂𝐻 + 𝑂𝐻− + 𝑁𝑎+ Eq. 5

Soluble Si species are released upon breakage of Si-O bonds by hydrolysis in the

surface region of the glass:13, 287

𝑆𝑖 − 𝑂 − 𝑆𝑖 + 𝐻2𝑂 → 𝑆𝑖 − 𝑂𝐻 + 𝐻𝑂 − 𝑆𝑖 Eq. 6

The formation of silanols is triggered by both reactions Eq.5 and Eq.6.288

A SiO2 layer is formed upon condensation of SiOH groups on the BG surface:

2𝑆𝑖𝑂𝐻𝑐𝑜𝑛𝑑.→ 𝑆𝑖 − 𝑂 − 𝑆𝑖 + 𝐻2𝑂 Eq. 7

From SEM images and from micro-PIXE-RBS analysis the thickness of the SiO2

layer was estimated to approx. 8-10 µm. The combination of the two methods

enables a reasonable estimation. The formation of silica-rich region (ii) was indicted

by FT-IR measurements, Fig. 24, and also confirmed in the elemental maps which

revealed increase of Si content. At the same time a thin ACP layer (iii) is formed (2-

3 µm) through migration of Ca and P ions to the BG surface.13

Indeed, the inner

side of BG scaffold (i) is depleted in P, as shown by ion beam measurements.

Page 88: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 88

Interestingly, the inner part of the scaffold is not depleted in Ca. The scaffold

peripheral layer (ii), however, shows depletion in Ca content with a slight increase

in Ca concentration in the region (iii) on top of the Si-enrichment, Fig. 27. From the

ion beam measurements a Ca/P ratio of 1.28 was calculated for this specific region

of interest (iii) which corresponds to typical values of ACP.54

However, as shown in

Fig. 23c-d the silica layer and the ACP layer cannot be clearly distinguished from

the SEM figures. Also ion beam analysis shown in Fig. 27 reveals that Ca is also

found in the (ii) region which is described previously as silica-rich layer. Hence, it

can be concluded, that instead of a distinct ACP layer on top of the silica gel as

proposed by Hench et al.13

more likely a mixed layer of CaP+SixOy is formed on the

scaffolds surface after immersion for 1d in biological fluid as described by Aguiar et

al.289

for bioactive glasses.

(3d-7d) of immersion in SBF, a CHA layer was formed as confirmed by FT-IR

measurements, Fig. 25. Also from micro-PIXE-RBS analysis a Ca/P ratio of 1.85

(after 3d) and 1.77 after 7d was determined which are closer to stoichiometric

apatite (Ca/P~ 1.67) confirming the crystallisation of ACP to CHA. Furthermore,

traces of Cu are incorporated in the CHA layer.

Altogether, the combination of FT-IR, SEM and micro-PIXE-RBS results give an

comprehensive picture of physico-chemical reaction on a 45S5-Cu derived scaffolds

during immersion in SBF. The analysis performed confirmed enhanced apatite

forming ability of the 45S5 derived scaffolds after 3d of immersion in SBF. Despite

some reports in literature on the potential negative impact of crystallinity on the in

vitro bioactivity of BG derived (sintered) scaffolds,147

in this work high bioactivity

was observed which is comparable to those of amorphous bioactive glass products

widely reported in literature.30, 35

Similar rapid CHA formation was also found on

micron sized 45S5 bioactive glass particles after 3 d in SBF has been shown.94

Despite some critical considerations related to the “bioactivity test” by SBF

immersion102

this method can still be considered a relevant technique for assessing

the reactivity of BGs in physiological environment. Particularly, the “SBF test”

serves as an internal control for new developed Cu-containing bioactive glasses in

order to compare the reactivity to the undoped 45S5 bioactive glass as reference.

Page 89: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 89

Moreover, in this work for the first time the reaction stages of a highly crystalline

45S5 derived (glass-ceramic) scaffold in simulated body fluid were observed in

high detail using the micro-PIXE-RBS technique. On top of FT-IR measurements

and SEM observations revealing the reaction on the BG scaffold surface the micro

ion beam method enabled detailed evaluation of elements distributed and evolution

in the 45S5 scaffold. Moreover, the chemical composition of the CHA layer was

derived. Detailed insight into the reaction scheme of 45S5 derived scaffolds was

gained. On top of the conventional methods thickness layer and details of the

composition of the reaction layers formed on the BG scaffold surface was obtained.

In particular for metal ion containing bioactive glass derived scaffolds the

distribution of the dopant element in the BG matrix is of great interest which cannot

be assed with conventional spectroscopic methods.

Additionally, it was confirmed that Cu inclusion in the 45S5 BG does not have any

negative impact on the CHA forming ability of the 45S5 bioactive glasses was

observed. Formation of CHA was observed after 3 days of immersion in SBF for all

glasses investigated,Fig. 25. It was shown that a homogenous CHA layer was

formed covering the BG scaffolds surface completely after 3d in SBF.

These observations are important for better understanding of the physico-chemical

reactions occurring on 45S5 Bioglass® derived scaffolds in physiological fluids.

Since chemistry of the materials surface strongly affects cell adhesion and cell

proliferation (particularly calcium phosphates)290

it is important to consider the

changes of the surface chemistry when testing BG scaffolds in vitro. CaPs for

instance are known to influence osteoblast cell attachment and differentiation.

Hence, the transformation of the silicate bioactive glass to a CaP phase is an

important parameter dictating the biological performance of BG derived scaffolds.

Interestingly, the chemical maps derived from micro-PIXE-RBS analyses revealed

that traces of Cu are also incorporated into the HAp layer formed on the BG

scaffolds surface. Thus, this effect should also be considered as parameter affecting

cell response beside the ionic dissolution products from the BG scaffold. Also, the

kinetics of the transformation of BG to CaP should determine the bone ingrowth

kinetics of BG scaffolds in vivo, which has been shown for ingrowth of 13-93 based

scaffolds.71

Page 90: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 90

Degradation of 45S5 derived scaffolds and Cu release

Two different conditions were used in order to assess the degradation behaviour of

the BG derived scaffolds as it has been shown that degradation behaviour of

bioactive glasses and glass derived scaffolds depend on immersion condition

whether static or dynamic.291

Fig. 32 shows the Cu and Si release from the BG

derived scaffolds when immersed in SBF for a period up to 21d under static

condition. The error bars indicate the standard deviation derived from 3 samples

measured. Depending on the glass composition Cu ions in the range from 0.04 up to

3.4 ppm were released. For all glasses, a rapid increase in Cu release was observed

in the initial stage of degradation reaching a plateau after 7d with ~0.04 ppm, ~0.4

ppm and ~3.4 ppm Cu released from 45S5-0.1Cu, 45S5-1Cu and 45S5-2.5Cu

scaffolds, respectively.

Similar trend was observed for Si released from the 45S5-Cu derived scaffolds

under static conditions. Burst release of Si was detected in the first 7d of reaction in

SBF reaching a plateau at Si levels of ~30-35 ppm.

Depending on the glass composition Cu concentrations in the range from 0.03 up to

3.5 ppm were released in SBF under static conditions. These values were reached

after 7d immersion in SBF and remained constant during further reaction in SBF.

Also, the Si values reach a plateau after 7d immersion in SBF which indicates a

decrease in degradation of the scaffolds due to saturation of the solution. Being a

silicate based glass the release of silicon can be used as marker for degradation of

45S5 derived scaffolds. Since no further increase in Si concentration in SBF was

observed after 7d it can be concluded that the degradation of the scaffolds is

inhibited due to saturation of the solution. This behaviour corresponds to literature

reports which have shown that amorphous BG scaffolds (mol%: 70SiO2-30CaO)

reach a degradation stop after 3d in SBF under static conditions due to saturation of

the solution in Si.291

Variations in Si release from different glasses are detected for

1d and 3d of immersion in SBF, while no significant difference was observed for an

immersion time after longer than 7d. However, the effect of Cu is inconsistent for

this data: 45S5-2.5Cu and reference 45S5 derived scaffolds show highest Si values

for initial stage of immersion (1-3d). Also, micro-PIXE-RBS derived data in Fig. 31

show that 45S5-1Cu and 45S5-2.5Cu exhibit faster Si release from the periphery of

Page 91: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 91

the scaffold indicating enhanced surface reactivity for BG with higher Cu contents

(≥ 1wt%). Altogether it seems that Cu-addition to the 45S5 matrix enhances the

initial surface reactivity of the scaffold while after longer reaction time no effect of

Cu is visible.

Fig. 32: Cu and Si release from 45S5-Cu derived scaffolds in SBF under static conditions.

Fig. 33 shows the Cu and Si release under quasi-dynamic conditions. For all

glasses, the overall released (cumulative) Cu concentrations were found to be higher

compared to the release rates under static conditions.

For 45S5-0.1Cu, 45S5-1Cu and 45S5-2.5Cu glasses, highest Cu levels of ~0.6 ppm,

~2.8 ppm and ~4.6 ppm, respectively, were detected. This was expected as under

quasi-dynamic conditions the concentration gradient and, thus, the ion diffusion is

enhanced. Cu concentrations reached a plateau after 14 days of immersions in SBF.

Accordingly, Si concentrations in the SBF were continuously increasing until 14

days of immersion reaching a plateau suggesting that the overall degradation of the

scaffolds is decelerated. In the same way, Si concentration increased during the first

days of immersion in SBF reaching a saturation level after 7d whereas Si values in

the range of 30-40 ppm depending on the glass composition were obtained.

Considering the standard deviation within the triplicate sample no major differences

in Si release is observed for different glass composition after 7 days of immersion in

Page 92: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 92

SBF as the mean Si values is in the range between 30 ppm and 50 ppm. Solely, in

the initial state of reaction higher values of Si were observed for 45S5 and 45S5-

2.5Cu scaffolds after 1 and 3 days in SBF. Among the Cu-doped glasses the final Si

levels decreased released decreased with higher Cu-content. This might be a result

of higher network connectivity of the silica network in the glass caused by Cu

incorporation which inhibits the release of soluble silica.

Under quasi-dynamic conditions higher absolute Cu as well Si levels were detected

for all glasses indicating overall enhanced degradation of the scaffolds under quasi-

dynamic conditions compared to static setup. These observations should be

considered for in vitro experiment of bioactive glasses.

In literature reports silicate based glass derived scaffolds were shown to follow an

linear degradation profile when cultured in SBF up to 7 days under quasi-dynamic

conditions similar to the parameters used in this work.291

However, after 14 days the

degradation slowed down and Si as well as Cu values reach constant constant

levels. Furthermore, with increasing Cu-content in the glass lower Si values are

released from the scaffolds. Considering the NMR results this is likely due to higher

Q2 and Q

3 fractions in the glass network resulting in higher SiO2 network stability

and less amount of soluble silica being released.

Fig. 33: Cu and Si release from 45S5-Cu derived scaffolds in SBF under quasi-static conditions.

A. Hoppe et al. J. Mater. Chem. B 1 (2013), p. 5659. - Reproduced by permission of The Royal

Society of Chemistry.

Page 93: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 93

After 21 days under both, static and dynamic conditions no significant mass gain

was observed for all 45S5-Cu derived scaffolds. Despite high release rates of

soluble silica the material loss is seemingly compensated with the mass gain

through the formation of hydroxyapatite on the 45S5 scaffolds surface.

Cu ions exhibit dose-dependent effects on human cells. For example, 50 ppm of

CuSO4 (313 µM) were shown to be optimal for stimulation of tube-like structures

formed by ECs when exposed to a Cu concentration range from 0 ppm to 100

ppm.197

Similarly, Wu et al. showed that Cu levels from 60.4 to 152.7 ppm favor

angiogenesis and osteogenesis via expression of osteogenic markers (ALP, OPN,

OCN) and secretion and expression of the angiogenic marker VEGF in human bone

marrow derived stem cells (hBMSCs).226

However, also lower Cu concentrations of

4 ppm released from phosphate glasses (similar to the Cu release ranges observed in

our study) were favorable for HUVEC cells via down regulating apoptosis.260

Furthermore, related values of 1.6 ppm to 8 ppm of CuSO4 were reported to

significantly increase the VEGF expression in keratinocytes.292

In this work Cu

levels from 0.3 ppm to 4.6 ppm were released in SBF from Cu-containing 45S5

bioactive glass derived scaffolds depending on the culturing conditions, hence it can

be concluded that 45S5-Cu bioactive glass might be a potential material for bone

tissue engineering applications where both osteogenic and angiogenic properties are

required.

Page 94: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 94

4.1.4 In vitro cell response

Theoretically the included Cu into the 45S5 Bioglass could affect cell behaviour in

two ways: Firstly, Cu ions released from the 45S5 Cu derived scaffold might

enhance vascularisation as Cu2+

ions are well-known angiogenic agents. As

indicated above, the Cu levels released into physiological environment are within

the therapeutic range according to literature reports for Cu2+

ions.293

Since ionic

dissolution products have been shown to stimulate cells on molecular level towards

osteogenic differentiation18

the Cu doping is expected to impart additional

angiogenic functionalities and enhance the overall biological activity of the 45S5

BG derived scaffolds. Secondly, Cu was incorporated in the CHA layer formed on

the scaffold surface after immersion in SBF, which might influence the attachment

and growth of relevant cell types used in regenerative medicine.290

Metal ion doped

hydroxyapatite, for instance, has been shown to influence osteoblast adhesion and

differentiation.294

The assessment of such biological effects of the incorporation of

Cu in the basic 45S5-Bioglass® material is described in this section.

Powder cytotoxicity

Cu in high dosage might be toxic to human cells and organism and therefore

cytotoxicity of Cu-containing glasses was investigated. Commercial use of

bioactive glass products also includes applications of particulate glass, such as

powder or granula (BonAlive® or NOVABONE

®). Hence in the first place the

cytotoxicity of the particulate 45S5-Cu powder was assessed. Fig. 34 shows light

microscopy evaluation of MG-63 cell morphology after 48 h of incubation with

45S5 bioactive glass particles at different concentrations. The cells show a slightly

elongated triangle morphology which is typically observed for healthy MG-63 cells.

Even at BG particle concentrations of 1 mg ml-1

no negative effects on cell growth

were observed. It seems that the particles sediment during the culturing time and the

cell grow on the top of the particles.

For comparison of the Cu containing glasses the cell morphology of MG-63 cells

cultured with 45S5-Cu containing samples at 100 and 1000 µg ml-1

is shown in Fig.

35. Similarly, no toxic effects are seen for the Cu samples for particle

Page 95: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 95

concentrations up to 100 µg ml-1

. For 1000 µg ml-1

, however, less cells were visible

indicating possible negative effect of the BG particles at 1000 µg ml-1

. Particularly,

for the 45S5-2.5Cu lees cells are visible compared to 45S5-Cu glasses with lower

Cu contents.

.

Fig. 34: Light microscope of MG-63 after 48h of incubation in drect ccontact with 45S5 BG

particles at different concentrations.

Fig. 35: Light microscope images of MG-63 after 48h of incubation in direct contact with 45S5-

Cu BG particles at 100 µg/ml and 1000 µg/ml.

In order to gain more detailed analysis of the cell viability the mitochondrial

activity and the cell number were derived. Fig. 36 shows the mitochondrial activity

and the cell number indicated by LDH activity of osteoblast-like cells after

Page 96: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 96

cultivation time of 48 h as function of BG concentration in the cell culture medium.

All Cu containing 45S5 BGs show high cell mitochondrial activity of > 50% of the

reference control for the entire concentration range investigated. Also the cell

number remains nearly constant for all particle concentrations tested. This indicated

that 45S5-Cu particles do not show any toxic effects on the cells during the first 48

h of incubation. Moreover, statistically significant enhancement of the

mitochondrial activity of cells in contact with 45S5-0.1Cu and 45S5-1Cu for

particles concentration between 0.1 and 100 µg ml-1

were observed compared to

plain 45S5 reference. At 1000 µg ml-1

for the Cu containing glasses a reduction of

the mitochondrial activity to 70 % (for 45S5-0.1Cu and 45S5-1Cu) and ~50 % (for

45S5-2.5Cu) was observed while remaining at ~100 % for the reference 45S5.

Hence, even though the cells remain at high viability higher than 50 % this indicates

possible cytotoxic effects of the 45S5-Cu particles at too high concentrations.

However, this effect could be also assigned to the alkaline pH shift due BG

dissolution when applied at a critical concentration ≥1000 µg ml-1

.

These observations confirm good biocompatibility of Cu-containing 45S5 glass

powders and beyond that stimulating effect on MG-63 cells when applied at

concentration range of 0.1-100 µg ml-1

. In particular, 45S5-0.1Cu and 45S5-1Cu

glass particles enhanced the mitochondrial activity of MG-63 cells compared to the

undoped 45S5 reference.

Fig. 36: Mitochondrial activity and cell number (LDH activity) as function of particle

concentration for 45S5-Cu glass series. *p<0.5 and **p<0.001 compared to 0 µg/ml reference.

Page 97: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 97

It has been shown in literature that Cu can stimulate the cell activity and

proliferation of osteoblastic cells.258

Accordingly, 45S5-0.1Cu and 45S51Cu glass

samples showed enhanced MG-63 cell activity compared to 45S5 reference.

In literature it has reported that hat BG particles of ~100 µm did not have any

significant effect on osteoblastic cell proliferation and metabolic activity.295

On the

other hand nano-sized bioactive glass particles can be cytotoxic as shown by means

of reduced cell activity of mesenchymal stem cells when cultured with BG particles

of ~70 nm at a concentration of 0-200 µg/ml (similar range as tested in this

work).296

The BG glass particles used in this study are ~ 6 µm and were shown to

stimulate the activity of osteoblast like cells indicating good biocompatibility of

micron-sized BG particles.

These findings are important for designing of new in vitro studies incorporating BG

particles. Further, these fundamental studies confirm the biocompatibility of the

45S5-Cu glasses which is a first step in the evaluation of biomaterials for use in

clinical applications. Indeed, commercial use bioactive glasses also involve the

application of particulate bioactive glass (Novabone, BoneAlive).

Cell attachment on 2D pellets

Fig. 37 shows the attachment and growth of MG-63 cells on 45S5-Cu derived

pellets after culturing for 48 hours. A dense cell layer was observed on the 45S5,

45S5-0.1Cu and 45S5-1Cu samples. The cells are widely spread indicating high

compatibility of the samples surface towards MG-63 cells. On 45S5-0.1Cu and

45S5-1Cu even multilayer growth of the cells was observed. However, on the 45S5-

2.5Cu samples no cells were detected indicating possible cytotoxicity of the high

Cu concentration in the 45S5-2.5Cu samples. These results are in agreement with

the cytotoxicity tests carried out on powdered 45S5-Cu glasses which indicated

possible cytotoxic effects of 45S5-2.5Cu glass.

Page 98: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 98

Fig. 37: Fluorescence microscopy of cytoskeleton (red) and cell nucleus (green) staining of

osteoblast-like cells seeded on 45S5-Cu derived pellets for 48h. No cells were observed on 45S5-

2.5Cu.

One well accepted explanation for copper-induced cytotoxicity is related to the

formation of reactive oxygen species (ROS) by Cu ions the via Fenton reaction.297

with the consequence of peroxidative damage of membrane lipids.298

In order to test

this hypothesis a western dot-blot analysis was performed with HOS cells seeded on

45S5-Cu pellets and the formation of 4-hydroxynonenal (HNE), one of the best

known and well-studied products of lipid peroxidation.10

A detailed description of

the experiment is given elsewhere.299

Fig. 38 shows the HNE formation for 2D pellets of Cu-containing 45S5 glasses. For

pure 45S5 (reference material) the cell growth was associated with low HNE

formation which slightly increased from after 7d and 14d of cell culture. With

10

The immuno-blot analysis of the HNE formation was carried out in collaboration with L.

Milkovic and Prof. N. Zarkovic, Laboratory for Oxidative Stress, Rudjer Boskovic Institute, Bijenicka 54, 10000 Zagreb, Croatia. These results are also part of the PhD thesis of L. Milkovic entitled “Beneficial effects of lipid peroxidation in the bone cell growth on bioactive glass - New perspectives in tissue engineering and regenerative medicine”

Page 99: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 99

addition of Cu higher values of HNE were detected whereby the strongest

enhancement was observed for 45S5-1Cu and 45S5-2.5Cu samples.

HNE is an important signaling molecule involved in various cellular processes

including cell proliferation and differentiation.300, 301

However, HNE acts in

concentration-depended manner and hence high HNE levels can also correspond to

cell death.302

This dese-dependent role is in accordance to the results shown in this

study: while for 45S5, 45S5-0.1Cu and 45S5-1Cu the enhanced HNE formation and

(hence peroxidation) are correlated to cell growth, for 45S5-2.5Cu the enhanced

HNE formation is related to cytotoxicity.

From these results it can be concluded lipid peroxidation is involved in the

interaction between 45S5 BG and osteoblast-like cells and that the effect of copper

is dose-depending and is correlated to the formation of lipid peroxidation products.

These correlations should be further explored in future studies in order to get more

information on the mechanism of the interaction of BG and human cells.

Fig. 38: Immuno-blot analyses of the HNE-protein adducts formation in osteoblast-like cells

(HOS) cells after 3, 7 and 14 days. The results are expressed as nmol of HNE-protein

adducts/mg of protein. Enhancement of the lipid peroxidation was in particular pronounced

after 1d (*** p<0.0001) and 14d (** p<0.01) for the cells grown on 45S5-2.5Cu samples.*p<0.05

Page 100: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 100

In vitro cell studies with 3D scaffolds

Response of osteoblast-like cells (MG-63)

From the powder cytotoxicity and the 2D cell attachment studies it was concluded

that 45S5-2.5Cu BG seems to be cytotoxic and, hence, this composition was not

further considered for biological investigations. For the 3D studies with MG-63

cells only 45S5, 45S5-0.1Cu and 45S5-1Cu were tested.

Good cell attachment and growth of osteoblast-like cells was observed with direct

seeding of the cell on the 3D scaffolds. Fig. 39 shows the attachment and growth of

osteoblast-like cells on 45S5-Cu derived scaffolds. Similar to 2D experiments no

signs of cytotoxicity of Cu was observed. The cell can attach and proliferate on the

scaffolds surface. MG-63 are an established cell culture model for assessing

osteoblast-like behaviour of cell in the context of bone tissue engineering.303

In

particular the attachment of human osteoblast cells can be well monitored by using

the MG-63 cell line as they show similar integrin profile as human osteoblasts. Also

it is a widely used cell model to test therapeutic agents and cytocompatibility testing

of materials.303

Hence, it can be concluded that 45S5-Cu derived scaffolds are

suitable for attachment and growth of osteoblast-like cells. Combined with the

results from the studies with powders and dense pellets in general a good

biocompatibility of 45S5-Cu scaffolds was observed. Hence, Cu levels up to 4 ppm

as shown in the degradation studies are not toxic to osteoblast-like cells up to a

period of 21 days. Also independently of the BG morphology applied as particulate,

dense pellets or porous scaffolds 45S5-Cu BG is biocompatible in vitro.

Page 101: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 101

Fig. 39: SEM images of osteoblast-like cells cultured for 21d on 3D scaffolds at different

magnifications: 45S5 (a, b), 45S5-0.1Cu (c, d) and 45S5-1Cu (e, f).

For more detailed evaluation of the cell response to 45S5-Cu derived scaffolds a

more clinically relevant cell type, human bone marrow derived stem cells

(hBMSCs) were used which allow a more accurate evaluation of osteogenic and

angiogenic behaviour of cell.303

Response of hBMSCs to ionic dissolution products of 45S5-Cu BG derived scaffolds

(2D cell culture)

Fig. 40 shows the cell morphology of hBMSCs after exposure to ionic dissolution

products from 45S5-Cu derived scaffolds for 3 weeks. No cytotoxic effects are

indicated since typical elongated morphology of the hBMSCs was observed.

Accordingly, high metabolic activity of the hBMSCs given by AlamarBlue (AB)

Page 102: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 102

reduction was detected as shown in Fig. 41a. For all time points investigated high

AlamarBlue reduction and hence high metabolic activity compared to reference

(hBMSCs only) was observed for hBMSCs when exposed to ionic dissolution

products. This indicates that ions released from the scaffolds do not show any toxic

effects on cells for time period up to 4 weeks. Moreover, after 4 weeks 45S5.1Cu

samples seems to stimulate the metabolism of the cells as indicated by higher AB

reduction compared to plain 45S5 glass. Even though the enhancement is not

statistically significant there is a tendency for Cu to stimulate hBMSCs in 2D

indirect culture model.

Fig. 40: LM images of the morphology of hBMSCs during exposure to ionic dissolution

products from 45S5-Cu derived scaffolds.

Fig. 41b shows the ALP activity of the hBMSCs in the 2D culture experiment. For

all sample groups significant ALP values were detected. However, no differences

for 45S5-Cu were observed compared to plain 45S5 and also the control (hBMSCs

only). In a similar way, the osteogenic marker RUNX2 was expressed after 2 weeks,

Page 103: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 103

Fig. 41d (normalized to their Glyceraldehyde 3-phosphate dehydrogenase

(GAPDH) expression). However, again no significant differences among the 45S5-

Cu derived scaffold groups were observed. Altogether no significant enhancement

of Cu on the expression of osteogenic markers ALP and Runx2 were observed even

though osteogenic potential of Cu has been presented in literature. For example, Cu

has been shown to stimulate MSCs towards the osteogenic cell line199

and has been

shown to upregulate expression of osteogenesis-related genes, e.g. alkaline

phosphatase (ALP), osteopontin (OPN) and osteocalcin (OCN).226

Fig. 41: Evaluation of hBMSCs during exposure to ionic dissolution products from 45S5-Cu

derived scaffolds: (a) metabolic activity by AlamarBlue dye reduction, (b) alkaline phosphatase

activity, (c) relative VEGF and (d) Runx2 expression related to hBMSCs group after 2 weeks. *

statistically significant for p<0.05 compared to all groups.

Interestingly, in this work no significant effect of Cu on osteogenic differentiation

of hBMSCs was observed. This might be due to the fact that in contrast to this work

in the study by Rodríguez et al.199

a osteoinduction medium was used which is

usually essential in order to induce osteogenic differentiation of MSCs in in vitro.

Furthermore, it should be taken into account that plain 45S5 Bioglass is known to

Page 104: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 104

induce osteogenic differentiation of MSCs which has been extensively described in

literature.18, 115

Hence it is likely that the Cu ion levels released from the 45S5-Cu

derived scaffolds did not excess the osteogenic ability of the plain 45S5 BG.

However, the expression of the angiogenic marker VEGF was enhanced for

hBMSCs seeded in contact with 45S5-1Cu scaffolds, Fig. 41c. VEGF expression

was 12 fold, which is 6 fold higher compared to hBMSCs only control and 3fold

stronger than 45S5 and 45S5-0.1Cu. Our observations are in very good accordance

with the well-known role of Cu in the VEGF signalling path ways as described in

literature: Cu ions were shown to stabilize and to upregulate hypoxia-inducible

factor 1 (HIF-1)304, 305

which, in turn, regulates the VEGF expression in MSCs.306

Hence, these results confirm that Cu released from a bioactive 45S5 glass scaffolds

show angiogenic potential in vitro. Since the cell response does not only depend on

the ionic dissolution products released from the scaffolds but also on parameters

like surface chemistry, roughness and pore curvature the biocompatibility 45SS5-Cu

derived scaffolds with hBMSCs was additionally tested in a 3D culture model

whereby the cells were directly seeded on the scaffolds.

Response of hBMSCs in direct contact (3D) with 45S5-Cu BG derived scaffolds

Similar to indirect studies for the hBMSCs seeded directly on the scaffold surface

no signs of cytotoxicity were found as indicated by high values of AB reduction,

Fig. 42. A rapid increase in AlamarBlue reduction was observed after 2 weeks for

all sample groups remaining at this level up to weeks of culture. All samples show

high AlamarBlue reduction values which indicate high metabolic activity and hence

high vitality of the cells. Even though no significant differences among the 45S5-Cu

glass scaffold series were observed the results show that high viability of hBMSCs

is maintained over a period up to 4 weeks and that no signs of toxicity due to Cu

presence were observed. Also the overall metabolic activity was significantly

enhanced after weeks (for all samples) compared to day 1 indicating high

proliferation ability of the cells when seeded on Cu-derived scaffolds.

Similar to the indirect 2D assay no significant differences in ALP activity was

observed among the 45S5-Cu scaffolds. However, all scaffolds induced ALP

Page 105: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 105

activity indicating osteogenic potential of 45S5 BG even though no additional

osteogenic stimulating effect of Cu was detected.

Fig. 42: Evaluation of hBMSCs in direct contact with 45S5-Cu derived scaffolds: (a) metabolic

activity by AlamarBlue dye reduction, (b) alkaline phosphatase (ALP) activity.

Evaluation of co-culture of hDMECs and hBMSCs

Fig. 43 shows the morphology of the ECs observed with a light microscope when

cultured for 2 weeks in presence of BG scaffold/hBMSCs constructs. For all

constructs the ECs show no signs of toxicity, the cells remain vital showing typical

“cobble stone” morphology. However, under the effect of 45S5-1Cu/hBMSC

construct, the cells show endothelial tube formation as indicated in Fig. 43e (dashed

circles).

Further viability and functionality characterisation of the HDMSCs was done by

LDL uptake assay. The LDL uptake was positive for all groups tested as shown in

Fig. 44a-e indicating high functionality of the ECs. Being a viability marker the

high LDL uptake confirms the high vitality of the ECs. In addition to light

microscopic analysis this confirms that the hDMECs retain their phenotype when

cultured in indirect contact. Again, only for the group cultured in presence of 45S5-

1Cu/hBMSC construct a formation of tube-like structures by the HDMSCs was

observed, Fig. 44e-f.

Page 106: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 106

Fig. 43: LM pictures of hDMECs cultured for 2 weeks in the presence of 45S5-Cu/hBMSCs

constructs: a) control ECs only, b) 45S5 c) 45S5-1Cu, d) 45S5/hBMSCs, and e) 45S5-

1Cu/hBMSCs. Only combination of 45S5-1Cu and hBMSCs stimulated hDMSCs towards

formation of tube-like structures.

Fig. 44: FLM analysis of the LDL uptake of hDMECs cultured for 2 weeks in the presence of

45S5-Cu/hBMSCs constructs: A) control ECs only, (B) 45S5 (C) 45S5-1Cu, (D) 45S5/hBMSCs,

and E) 45S5-1Cu/hBMSCs. Only the combination of 45S5-1Cu and hBMSCs leads to

formation of tube-like structures. LDL uptake is directly correlated with light-microscopy

images of the same well (F).

Furthermore, the endothelial phenotype and the ability of the ECs to form tube-like

structure was assessed by seeding the cells on Matrigel (after being trypsinysed).

The results of light-microscopic evaluation are shown in Fig. 45. Basically, for all

groups the hDMECs nicely grow on the Matrigel forming tubular structures. Even

though no major differences among the tested constructs were observed it is visible

Page 107: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 107

that for the 45S5-1Cu/hBMSCs group a more dense tubular structures were formed

indicating higher functionality and viability of the hDMECs.

Fig. 45: LM images of trypsinised ECs seeded for 2 weeks on matrigel for assessment of tube

formation. A) control ECs only, (B) 45S5 (C) 45S5-1Cu, (D) 45S5+hBMSCs, and E) 45S5-

1Cu+hBMSCs. Similar pictures were observed also at week 1.

In order to analyse the detailed EC phenotype flow cytometric analyses (FACS) was

performed identifying vWF (Von Willebrand factor), CD31 and VEGFR2 in

hDMECs. Fig. 46 summarises the results of the ES phenotype analyses by FACS.

Generally, the EC phenotype is better retained in all samples with presence of

hBMSCs compared to non-hBMSCs groups.

Additionally, Cu seems to have a stimulating effect on HDMSCs: for the

45S5/hBMSCs and 45S5-1Cu/hBMSCs 60.3% and 95.1 %, respectively, were in

presence of 45S5-1Cu/hBMSCs (47%) compared to 45S5/hBMSCs (16.5 %).

The CD31 marker was also expresses in the hDMECs even though no significant

difference between the presence of 45S5-1Cu/hBMSCs (98%) and 45S5/hBMSCs

(97.3%). This indicates that the 45S5-1Cu/hBMSCs constructs enhanced the

expression of relevant specific endothelial markers which can be used to

characterise the phenotype of endothelial cells. Usually high expressions of CD31,

vWF and VEGFR2 are correlated with physiological endothelial phenotype while

under pathological conditions the expression of these markers can be inhibited.

Page 108: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 108

According to Fig. 46 addition of Cu to the 45S5 BG enhanced the expression of the

endothelial markers indicating its angiogenic potential.

The VEGF release in culture medium is shown in Fig. 47. After 2 weeks

significantly higher VEGF release was observed for the constructs containing

hBMSCs. After 4 weeks the VEGF values further increase for the constructs

containing BG scaffolds and hBMSCs, while remaining at lower values at ~pg/ml

for the constructs without stem cells. Interestingly, the constructs with plain

45S5/hBMSCs showed higher VEGF release compared to 45S5-1Cu/hBMSCs

construct despite the enhanced expression of VEGF in hBMSCs when cultured in

contact with 45S5-1Cu alone as shown in Fig. 41.

Fig. 46: FACS analysis for vWF antigen, VEGFR2 and CD 31 on surface of hDMECs after 2

weeks of culture. For the green curves the number of positive markers is given in %.

Page 109: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 109

Fig. 47: VEGF release in culture medium from hDMECS cultured indirectly with BG

scaffold+hBMSCs constructs.

Altogether three main observations can be made regarding the effect of Cu on

mesenchymal stem cells and endothelial cells

i) Cu ions stimulate VEGF expression in hBMSCs

ii) Higher VEGF release is observed for hDMECs co-cultured with

hBMSCs seeded on a 45S5 BG scaffolds. But is not enhanced when Cu

is added.

iii) Enhanced endothelial phenotype is observed when hDMECs are exposed

to specifically Cu-containing 45S5-1Cu/hBMSCs construct and are

stimulated to form tube-like structures

Firstly, Cu ions stimulate hBMSCs to express VEGF and thus activate relevant

angiogenesis related pathways. Furthermore, the hBMSCs seeded on a 45S5 derived

scaffold secrete VEGF which stimulates ECs as enhanced EC phenotype was

observed in presence of MSCs indicated by larger numbers of cells being positive

for vWF, CD31, and VEGF2R. However, the enhanced release of VEGF was

observed for both constructs with and without Cu. Nevertheless, the only for 45S5-

1Cu/hBMSCs constructs an enhanced functionality and stimulation of ECs was

observed indicated by increased activity of EC markers and formation of tube-like

structures by the hDMECs confirming their angiogenic activity.

Page 110: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 110

VEGF plays a critical role in angiogenic process as one of the main transcription

mediation blood vessel development.307

Cu, in turn, induces VEGF expression not

only in MSCs but also in ECs like keratocytes213

and cardiomyocites.308

It is

important to mention that Cu has been shown not only to induce VEGF expression

when applied in excessive concentrations (related to physiological values). When

applied at physiological values (5 µM, ~ 0.132 pm), however, it was shown that Cu

is essential for VEGF expression.293

Furthermore, it is known that Cu is not only an

essential co-participant in angiogenesis but can be angiogenic itself.198

McAulan et

al., for example, showed that Cu ions induced neovascularisation in an in vivo rabbit

cornea pocket assay.309

There is also direct evidence of Cu action in increased flap

survival in an in vivo rat model rats due to enhanced VEGF expression. This action

is again due to effects of Cu on nearby cells in the random flap in inducing VEGF.

The mechanisms behind the stimulating effect of Cu ions are usually related to

activation of several transcriptional factors, as for example HIF-1 that is crucial for

VEGF expression.293

Cu ions have been shown to stabilise HIF-1 even under

normoxic conditions and hence activate HIF-1 related pathways.304

However, considering that high VEGF amounts were found also for 45S5/hBMSCs

constructs this indicates that there must be synergetic effects of Cu and the presence

of hMSCs leading to stimulation of ECs. Indeed Cu plays a versatile role in the

physiology of angiogenesis. It is likely that Cu induced expression of some other

angiogenic factors besides VEGF as it is known that expression if VEGF is not the

only possible action of Cu in angiogenesis.293

For example expression off FGF-1

and FGF-2 factors known for regulation of blood vessel functions are also known to

be mediated by Cu as well as fibronectin and angiogenin.293

Indeed, it is known that Cu has direct angiogenic effects in vitro and in vivo.

It has already been reported that VEGF expression in ECs293

can be induced by

copper ions and this property may be exploited to accelerate dermal wound

healing.310

Furthermore, there is evidence in literate that Cu can directly stimulate

ECs. Cu ions are known to regulate endothelial cell proliferation and migration this

having a direct impact on the process of angiogenesis and vascularisation.293

Li et

al., for instance, showed that Cu enhances the proliferation of human umbilical vein

endothelial cells (HUVECs).311

Similarly, Hu showed that application of CuSO4 in

Page 111: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 111

the range of 1-500 µM CuSO4 significantly increased the proliferation of

HUVECs.198

Similar results were observed with 3D printed scaffolds loaded with

CuSO4 and VEGF and FGF-2 growth factor.257

It was shown that Cu combined with

growth factor exhibits a stimulating synergetic effect on angiogenesis in vivo

indicated through formation of tube-like structures and collagen deposition.197

Based on the experimental observation and literature reports the proposed

mechanism for Cu acting in the hBMSCs/hDMECs co-culture model is

schematically shown in Fig. 48.

Fig. 48: Scheme of the mechanism of Cu involved in the angiogenic pathways in a co-culture of

hBMSCs and hDMECs. Cu2+

activate the HIF-1 transcription factor which mediations

expression of VEGF and other angiogenic factors which, in turn, activate signalling pathways

in hDMECs mediating cell migration, proliferation and cell survival. Additionally, Cu ions

might directly stimulate endothelial cells as shown in literature.

The use of co-culture models of osteoblast cells and endothelial cells has been

extensively investigated in the context of bone tissue engineering in the last two

Page 112: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 112

decades.312, 313

The general idea involves the induction of osteogenic hBMSCs

differentiation while in the same time the ECs are stimulated towards formation if

tube-like prevascular structure for providing nutrition and oxygen for osteoblast

cells. In turn, hBMSCs can be used as vesicles for growth factor delivery of

angiogenic factor in order to stimulate the ECs. However, usually additional

amounts of VEGF are supplemented to the cell culture to provide strong initial

stimulation.313

Thus, by stimulating hBMSCs towards higher release of VEGF, which in turn

stimulated ECs, the presented 45S5-Cu scaffold act as an “indirect” angiogenic

growth factor delivery system.46

This indirect approach is advantageous since it

enables controlled VEGF release mediated by cells which is adapted to the

physiologically, local needed conditions avoiding growth factor “overdose”.

It was observed that 45S5 combined with hBMSCs can be used as indirect system

for releasing VEGF as stimulating factor for hDMECs leading to enhanced

angiogenesis. Additionally, incorporation of Cu leads to enhanced expression of

VEGF in hMBSCs hence seemingly activating signalling pathways related to

angiogenesis. Altogether, the combination of 45S5-1Cu scaffolds with hBMSCs has

a stimulating effect on ECs to form tube-like structures and hence can be considered

as a promising biomaterial-cell approach in bone tissue engineering.

45S5-1Cu derived scaffolds showed enhanced expression in MSCs and hence were

chosen as the most promising candidate for the in vivo study.

4.1.5 In vivo evaluation

Angiogenic potential of Cu-containing bioactive glass has been assessed in an AV-

loop model. After 3 weeks the explanted constructs were analysed by means of

vessel density and total vessel cross area, vessel length as well as vessel radius. Fig.

49 shows the microCT reconstruction of the blood vessels in the 45S5 and 45S5-

1Cu constructs. It is clearly visible that both 45S5 and 45S5-1Cu scaffolds support

intrinsic vascularisation as indicated through micro vessel sprouting originating

from the AV-loop. However, no statistically significant stimulation effect of copper

could be shown. The quantification of the microCT data revealed that for 45S5

more vessels with larger radius were formed, whereas for the 45S5-1Cu samples

Page 113: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 113

small radius vessels were present, Fig. 50. It seems that for 45S5-1Cu higher

amount of small radius vessels was formed whereas almost no blood vessel with a

radius below 10 µm were detected. Furthermore, the vessel density was found to be

higher in 45S5-1Cu containing constructs, being 0.16±0.18 mm-2

compared to

0.07±0.05 mm-2

. However, only a trend of Cu stimulating the formation of blood

vessel especially in the initial stage resulting in higher number of small radius

vessels was observed. In this work it was shown that Cu stimulated angiogenesis in

vitro but, however, this effect could not be confirmed in vivo even though there is

also evidence in literature indicating that Cu directly stimulated vascularisation in

vivo.257, 310, 314

This might be due to angiogenic potential of plain 45S5 derived

scaffolds is it was shown in the AV-loop model.315

Hence, the amount released from

the 45S5-1Cu scaffolds under in vivo conditions might not be sufficient enough to

exceed the angiogenic potential of 45S5 BG scaffolds. Also the 3 weeks of

investigation might be to short shown full potential of Cu in the present in vivo

model. Indeed, the formation of small radius vessels was enhanced for the 45S5-

1Cu scaffolds compared to 45S5 reference which might result in a more dense

vascularisation after further maturation.

Fig. 49: MicroCT reconstruction showing the intrinsic vascularisation induced in the AV model

for a) 45S5 and b) 45S5-1Cu scaffolds. Small vessel sprouting originates from the AV loop.

Page 114: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 114

Fig. 50: MicroCT derived quantification of the vessel sprouting in the AV-loop model.

The results also indicate that for in vivo studies even higher Cu concentrations

might be considered biocompatible as 45S5-1Cu did will not impair the angiogenic

potential of 45S5 scaffolds. Higher amounts of Cu doping to 45S5 BG is, hence,

suggested for further investigations. Also constructs seeded with hBMSCs could be

used in in vivo model. This should enhance the angiogenic effect as it has been

shown that combination of 45S5 and cells implanted in vivo result in enhanced

vascularisation.316

Overall, it was shown that 45S5derived scaffolds support intrinsic vascularisation as

confirmed in the AV loop model in rats. Hence, 45S5 derived scaffolds are potential

materials to be used in prevascularisation in vivo models.315

Furthermore, the results

indicate that 45S5-1Cu show a tendency to improve the intrinsic vascularisation.

Even though the results are not statistically significant 45S5-1Cu might be a

candidate to be used in bone regeneration applications where enhanced

angiogenesis is required.

Page 115: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 115

4.2 Cobalt containing 13-93 based glasses

As described in chapter 4.1.2 45S5 BG undergoes crystallisation during sintering

leading to unsatisfied densification ad formation of micro cracks. Hence, 1393 glass

was employed as possible alternative scaffold material and its suitability as carrier

for therapeutic inorganic ions whilst showing appropriate mechanical performance

was assessed.

4.2.1 Glass properties

Structure and Thermal Properties

Fig. 51 shows the FT-IR spectra of Co-containing glasses in the as-fabricated state.

The main bands appearing in the FT-IR spectra are summarised in Table 12.

The intense band at ~1050 cm-1

are attributed to asymmetric stretching vibrations of

the Si-O bond becomes sharper with increasing Co content in the glass.

Furthermore, for 1393-5Co and 1393-10Co additional peak at ~1105 cm-1

and a

weak shoulder at ~1205 cm-1

were detected which are typically observed for silicate

glasses with higher SiO2 content and vitreous silica, respectively.317

With increasing

Co concentration in the glass the Si-NBO (non-bridging oxygen) peak (at ~950 cm-

1)91

is reduced in intensity and is slightly shifted to lower wave numbers. Also, the

intensity of the bending mode of the Si-O band at ~790 cm-1

is increasing with

higher Co content. Also, for 1393-5Co and 1393-10 an additional absorption peak

was observed at ~990 cm-1

which, according to literature, might be attributed to Si-

O-Co bonds as it was shown for metal containing mesoporous alumino-silicates

(zeolithes).318

Hence, the FT-IR analysis gives indication that for concentrations ≥

5wt% Co may be entering the glass network by forming Si-O-Co bonds resulting in

higher polymerized SiO2 network.

Raman spectra (normalised to the Si-O(r) peak) of the 1393-Co glasses are shown

in Fig. 52 with the main bands occurring summarised in Table 12.

Page 116: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 116

According to literature the wide absorption band at ~1100 cm-1

can be assigned to

asymmetric stretching mode of Si-O-Si group, whereas the bands at ~610 cm-1

and

~930 cm-1

can be attributed to Si-O-Si rocking vibrations and Si-NBO bond,

respectively.319, 320

Fig. 51: FT-IR spectra of as fabricated 1393-Co glasses.

Fig. 52: Raman spectra of as fabricated 1393-Co glasses.

Page 117: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 117

The Band at~450 cm-1

is likely assigned to defect lines of vitreous silica typically

observed for silicate ring structures.321

For 393-1Co and 1393-5Co a new broad

peak is evolving at 750 cm-1

which is assigned to the bending mode of Si-O-Si.319

Likely, this band appears due to the longer length of the Co-O bond (192 pm)

compared with the Si-O bond (177 pm) causing stronger bending vibrations of the

Si-O bond.

Table 12: Main absorption bands in the FT-IR spectra observed for as fabricated 1393-Co

glasses. Reprinted with permission from A. Hoppe at al, ACS Appl. Mater. Interfaces 6 (2014),

p. 2865. Copyright (2014) American Chemical Society.

Assignment FT-IR [cm-1

] Raman [cm-1

] Remarks

Symmetric

stretching,

νsym(Si-O-Si)

~490-500 ~450 νsym(Si-O-Si) at ~450 cm-1

is observed

for pure silica. 282

For bioactive

glasses the band is shifted to 500 cm-

1.91

In the Raman spectra this band is

assigned to silica ring structures.321

Asymmetric

stretching,

νasym(Si-O-Si)

1000-1300

~1040/1140*

~1220**

~1060-1080319

In the FT-IR spectra317

the most

intense Si-O band is at ~1040 cm-1

.

With Co doping additional bands at

~1140 cm-1

and ~1220 cm-1

appear

which are typically seen in vitreous

silica. In the Raman spectra319

the

peak maximum is shifted towards

lower frequencies for 1393-5Co and

1393-10Co glasses.

Bending mode

δ(Si-O-Si)

Peak at ~75091

Broad band at

~750319

Intensity is increasing with higher Co

content in the glass.

Rocking

vibrations of Si-

O-Si

n.a. ~620319

--

SiONBO Shoulder at

~930

Peak at ~940 In the FT-IR spectra the SiONBO is

decreasing in intensity and is shifted

to lower frequencies with

incorporation of 5wt% and 10wt%

Co.91

,282

Si-O-Co 992 Appears in the FT-IR spectra of 1393-

5Co and 1393-10Co glass and is

attributed to Si-O-Co bonding.318

*TO1; **TO2; TO=transverse optical groups of Si-O-Si bonds. 317

Page 118: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 118

Thermal properties

Fig. 53 shows the DSC diagrams for the 1393-Co glass powders as fabricated. The

characteristic features glass transition point Tg and the crystallization onset To

derived from the DSC measurement are summarized in Table 13.

Table 13: Glass transition point Tg and crystallisation onset To (in °C) of the Co-containing

glasses derived from DSC measurements. Reprinted with permission from A. Hoppe at al, ACS

Appl. Mater. Interfaces 6 (2014), p. 2865. Copyright (2014) American Chemical Society.

Glass 1393 1393-1Co 1393-5Co 1393-10Co

Tg 626 621 606 587

To 750 848 -* -*

*Outside of the observed temperature range.

Tg decreased continuously with increasing Co2+

content in the glass. For 1393,

1393-1Co, 1393-5Co and 13-93-10Co Tg values of 626 °C, 621 °C, 606 °C and 587

°C, respectively, were determined. In turn the Tc, onset is increasing with higher Co

content: For 1393-1Co the To increased from 750 for 1393 to 848 °C for 1393-1Co,

whereas for Co contents >1 wt% the To was not detected as it was outside of the

measured temperature range. The process window for the glass fabrication is larger

for Co containing glass, and Co is stabilizing the amorphous glass state which

indicated better processing properties of the Co-containing 1393 glasses.

Page 119: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 119

Fig. 53: DSC curves for the 1393-Co glass series showing decrease in Tg with increasing Co

content in the glass. Reprinted with permission from A. Hoppe at al, ACS Appl. Mater.

Interfaces 6 (2014), p. 2865. Copyright (2014) American Chemical Society.

The effect of Co on 1393 glass structure

The effect of Co incorporation on the glass structure and its properties can be

discussed on the basis of FT-IR as well as Raman spectroscopy and correlated to the

glass thermal behaviour as derived from DSC measurement. Co is known to be an

intermediate oxide and, thus, can enter the silicate network or act as modifying

oxide.64

As the Tg decreases with Co incorporation it is likely that Co is entering the

glass network creating Si-O-Co bonds replacing the stronger Si-O-Si bonds.64

Thus,

the glass network becomes weaker and the Tg decreases.64

This observation is in

agreement with FT-IR results which indicated the formation of additional Si-O-Co

bonds in the glass network as a result of Co substitution. However, according to FT-

IR analysis, Fig. 51, the formation of Si-O-Co bonds is evident only for Co contents

higher than 5wt%. Hence, it can be concluded that Co plays a concentration-

dependent role in the glass network acting as network modifier at 1wt% and

network former at CoO concentrations of 5wt% and higher. This corresponds to

literature reports which also showed the concentration dependent behaviour of CoO

in silicate glasses.64

At 1wt% CoO is likely acting as network modifier replacing

Ca2+

ions, while no significant effect on the structure or thermal behaviour was

observed by FT-IR/Raman spectroscopy as well as DSC analysis. A lower Tg is

Page 120: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 120

associated with weaker bonds within the glass network and should therefore result

in enhanced in vitro degradation of the glass scaffolds which will be discussed in

3.3.

Fig. 54: Struture of 1393 glass with 1wt% CoO (left) and >5% wt% CoO (right). Depending on

the concentration CoO oxide acts as network modifier replacing Ca2+

ions (left) or as network

former by entering the network and forming Si-O-Co bonds (right).

4.2.2 Scaffold properties

Macrostructure

Fig. 55 shows the macrostructure of the 1393-Co glass derived scaffolds. Scaffolds

were prepared using the 1393, 1393-1Co and 1393-5Co glasses. For the highest Co

content of 10wt% the glass could not be processed to a powder of suitable particle

size and scaffolds, respectively. No major differences between the glass

compositions were observed. All scaffolds show completely interconnected pore

systems with pore diameters of ~ 200-400 µm. The porosity of the scaffolds after a

2nd

coating was 91%, 90% and 89% for 1393, 1393-1Co and 1393-5Co samples,

respectively, confirming that the main macrostructural features are independent of

the glass composition. Fig. 55d-e additionally show representative higher

magnification images of a strut of a 13-93 sample: smooth and dense scaffold

surface was observed and almost fully densified struts were found.

Furthermore, a µCT reconstruction of a 1393 scaffold is given in Fig. 55f

confirming the interconnected pore system of the scaffolds (98% of the total

porosity was confirmed to be open porosity). From microCT analysis a total

Page 121: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 121

porosity of 92% and average strut thickness of ~74 um were derived. These data are

in good agreement with microscopic analyses and porosity results obtained from

Archimedes measurements.

High porosity values and interconnected pore system of the scaffold should enable

vascularisation and tissue ingrowth when applied as engineered bone construct.

Vascularisation has been shown to be enhanced in scaffolds with pores > 250 µm 278

and also high interconnectivity is considered even a highly important factor for

blood vessel and tissue ingrowth.43, 279

Thus, the 1393-Co derived scaffold meet the

macrostructural requirements for use as bone tissue engineering scaffolds. In Fig.

3d) a 1393 scaffold strut at higher magnification is shown revealing dense struts and

smooth scaffold surface without presence of cracks. This is in agreement with

literature reports which showed that 13-93 bioactive glass can be densified by

viscous flow sintering avoiding crystallization.69

Page 122: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 122

Fig. 55: SEM images of the scaffold macrostructure for (a) 1393, (b) 1393-1Co and (c) 1393-

5Co glass compositions; (d) and (e) show higher magnification of a strut of a 1393 derived with

dense structure and smooth surface; (f) µCT reconstruction of 1393 derived scaffold. Adapted

with permission from A. Hoppe at al, ACS Appl. Mater. Interfaces 6 (2014), p. 2865. Copyright

(2014) American Chemical Society.

XRD analysis of the scaffolds revealed that the scaffolds remain in the amorphous

state; no crystalline phases were detected (see appendix, Fig. A 3). This is in

agreement with DSC results shown above which revealed crystallisation onset

temperature of Tc, onset > 750 °C being below the sintering temperature thus

precluding crystallisation during the densification process.

Mechanical properties

Fig. 56 shows the compressive strength, σc, values measured for the 1393-Co glass

scaffolds series and typical stress-way curves (Fig. 56b). A catastrophic failure of

the scaffolds was observed at a given maximal stress (marked with * in Fig. 56b)

followed by rapid decrease of the stress. The further increase in the stress is

correlated to compression of remaining scaffolds struts.68

However, the

measurement was stopped shortly after the scaffold failure.

For the 1393-Co derived glass scaffolds with high compressive strength values of

2.3±0.4 MPa, 2.3±0.5 MPa and 4.2±0.6 MPa for 1393, 1393-1Co and 1393-5Co,

respectively, were measured. For 1393-5Co the compressive strength is higher (~4

MPa) compared to 1393 and 1393-1Co scaffolds which is likely due to the

influence of Co on the sintering behaviour of the glass. According to DSC

measurement the Tg of the 1393-5Co glass is reduced hence improving the viscous

flow sintering of the scaffolds and leading to higher strength. However, for

comprehensive evaluation of the mechanical strength the porosity of the scaffolds

has to be taken into account. A detailed discussion of the mechanical performance

of the 1393 derived scaffolds in the context of bone tissue engineering is given in 0.

Page 123: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 123

Fig. 56: Mechanical properties of the 1393-Co glass derived scaffolds (double coated): a) mean

values calculated from 10 measurements per scaffold group; b) selected typical stress-strain

curved. Adapted with permission from A. Hoppe at al, ACS Appl. Mater. Interfaces 6 (2014), p.

2865. Copyright (2014) American Chemical Society.

Page 124: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 124

Acellular Bioactivity in SBF

HAp formation

First, the reaction stages occurring on the scaffolds surface during reaction in SBF

are shown exemplary for the plain 13-93 scaffolds and are discussed on the basis of

SEM, FT-IR and micro-PIXE-RBS analysis. Further on, the effect of Co

incorporation in the glass on the hydroxyapatite forming ability and degradation of

the glass scaffolds is presented and discussed based. Fig. 57 shows the SEM images

of 1393 derived scaffolds after immersion in SBF for 1d (a, b), 3d (c, d) and 7d (e,

f), respectively. After 1d first indications of surface reaction were observed on the

scaffold surface visible as a thin reaction layer formed on the scaffolds surface.

However, the inner glass matrix remained intact without any signs of dissolution.

After 3d of immersion the reaction three distinct areas phases can be distinguished

as indicated in Fig. 57d: light grey area showing the inner BG network, a dark grey

layer (SiO2) and thin layer on top of it (CaP). This distinction was made based on

the EDS analysis, as shown in Table 14. The thickness of each layer was estimated

from the SEM figures to 2.4±1.7 µm and 0.61±0.04 um for SiO2 and CaP layers,

respectively.

Table 14: EDS derived elemental concentration (at%) of the reaction phases formed on 1393

scaffolds after 3d in SBF.

Mg K Si P Ca Ca/P

BG 2.79 4.74 20.30 1.34 6.71 5.00

SiO2 0.63 1.23 26.21 1.32 2.02 1.53

CaP 1.21 0.93 14.87 4.87 6.80 1.40

The SiO2 reaction region is quite inhomogeneous and hence the estimation of the

thickness is given with a high standard deviation. In fact, SEM analysis provides

only rough estimations; more reliable data on the dimensions of the reactions layers

is given by PIXE-RBS analysis further below. After 7 days further growth of the

CaP layer occurred: the formation of the typical morphology of hydroxyapatite was

observed. However, again no clearly visible dissolution of the inner region of the

scaffold strut was observed with SEM, Fig 5f.

Page 125: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 125

Fig. 57: SEM analysis of 1393-Co derived scaffolds after immersion in SBF for 1d (a, b), 3d (c,

d) and 7 d (e, f). Reprinted with permission from A. Hoppe at al, ACS Appl. Mater. Interfaces 6

(2014), p. 2865. Copyright (2014) American Chemical Society.

PIXE-RBS derived chemical maps

In order to identify the origin of the reaction phases formed on the BG scaffolds

concentrations maps derived from ion beam measurements which give a more

precise picture of the reactions taking place at the interface scaffold/fluid. The Ion

beam analysis has been shown to be a powerful technique for identifying the

reactions on bioactive glass / fluid interface.322

In the as-fabricated state P, Si, Ca and Mg and Co (Na and K are not shown for

simplicity) are homogenously distributed (see supplementary data).

Fig. 58 - Fig. 60 show PIXE-RBS derived elemental maps for 13-93 scaffolds after

1d, 3d and 7d, respectively. These results can be directly correlated to the SEM

Page 126: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 126

observations shown in Fig. 57. Basically, according to the elemental maps again

three distinct regions during reaction in SBF can be defined:

i) Inner region of the glass scaffolds (BG)

ii) Silica-rich layer (SiO2)

iii) Surface periphery (CaP layer)

Additionally, for the regions of interest i) and iii), the inner regions of the glass

scaffolds and the scaffold periphery, respectively, the quantitative elemental

concentration were derived which can be found in the appendix (Fig. A 4 and Fig. A

5).

Interestingly, the inner part of the scaffolds did not show any signs of degradation:

the elemental concentration in the inner region of the scaffolds remained constant

for immersion times up to 7 days for all glasses investigated. Hence, it is likely that

the degradation of 1393 derived scaffolds occurs preferably at the scaffolds

interface with separate reactions stages as discussed below.

After 1 day of immersion in SBF signs if initial surface reaction were observed: a

thin reaction layer of 2.5 ±0.4 µm on was formed the pore surface (Fig. 58) which is

depleted in Ca, P and Mg indicating fast release of these elements from the scaffold

surface.

After 3 days (Fig. 59) a silica rich layer with a thickness of roughly 5.9 ±0.5 µm is

formed on the scaffolds surface. This value is slightly higher compared to the value

of 2.4±1.7 µm observed with SEM (Fig. 57). Considering the high deviation of the

surface thickness as observed with SEM we conclude that a SiO2 layer with a

thickness ranging between 1-6 µm is formed on 1393 derived scaffold after 3 days

of reaction in SBF. On top of it a Ca and P enriched layer (3.1±0.2 µm) was

detected as already indicated by SEM /EDS analysis. This layer results from

diffusion of Ca and P-species from the glass network. It is clearly visible from the

chemical maps that this Si-rich layer acts as diffusion barrier as Ca is accumulated

in the region right before the silica layer.

Page 127: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 127

Fig. 58: Elemental distribution in the cross-section of a 1393 scaffolds after 1 days. Adapted

with permission from A. Hoppe at al, ACS Appl. Mater. Interfaces 6 (2014), p. 2865. Copyright

(2014) American Chemical Society.

Fig. 59: Elemental distribution in the cross-section of a 1393 scaffolds after 3 days. Adapted

with permission from A. Hoppe at al, ACS Appl. Mater. Interfaces 6 (2014), p. 2865. Copyright

(2014) American Chemical Society.

Page 128: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 128

Fig. 60: Elemental distribution in the cross-section of a 1393 scaffolds after 7 days. Adapted

with permission from A. Hoppe at al, ACS Appl. Mater. Interfaces 6 (2014), p. 2865. Copyright

(2014) American Chemical Society.

After 7 days (Fig. 60) the glass surface is further dissolving; the Si-rich layer is

expanded to 18.0±1.7 µm. Also the CaP layer continues to grow reaching 16.0±0.2

µm. It is worth noticing that the traces of Mg are incorporated in the CaP layer

which is typical for biomimetic hydroxyapatite formed upon reaction in body

fluids.100, 323

Effect of Co

The effect of Co on the bioactivity of the 1393-1Co and 1393-5Co scaffolds is

shown in Fig. 61 and Fig. 63, respectively. In contrast to reference 1393 glass

scaffold it is noticeable from the chemical maps that for 1393-1Co samples no

distinct CaP layer is formed after 7d in SBF. Even though P enrichment in the

surface is clearly detectable, a rather mixed Si-rich CaP layer with traces of

incorporated Co is observed.

For 1393-5Co scaffolds, however, a clearly distinguishable CaP layer was formed

on top of the SiO2 layer. Interestingly, Co is incorporated into the CaP layer. Co ions

appear to diffuse through the SiO2 layer (which itself is depleted in Co) and are

substituted in the CaP layer.

Page 129: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 129

One can speculate that the change of the chemistry of the CaP layer will have an

impact on cellular response. For instance, Co incorporated in calcium phosphates

has been shown to increase osteoclast proliferation and overall mineral

resorption.177

Fig. 61: Elemental distribution in the cross-section of a 1393-1Co scaffolds after 7 days.

Reprinted with permission from A. Hoppe at al, ACS Appl. Mater. Interfaces 6 (2014), p. 2865.

Copyright (2014) American Chemical Society.

In order to prove the formation of carbonate hydroxyapatite FT-IR analysis was

performed for the 1393-Co scaffolds after 7 days in SBF as depicted in Fig. 62

The formation of carbonated hydroxyapatite can be monitored by the appearance of

the triply generated (ν3) bending modes of the O-P-O bands of CHA at 554 cm-1

and

Page 130: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 130

602 cm-1

as detected for 1393 and 1393-1Co scaffolds, respectively.284

However, for

1393-5Co sample a broad band at ~600 cm-1

was observed which is typical for

amorphous CaP.283

Fig. 62: FT-IR spectra of 1393-Co derived scaffolds after 7d of immersion in SBF. At 5wt% Co

inclusion in the glass the crystallisation of CHA seems to be impeded. Adapted with permission

from A. Hoppe at al, ACS Appl. Mater. Interfaces 6 (2014), p. 2865. Copyright (2014) American

Chemical Society.

Based on the SEM/EDS and PIXE-RBS results following scheme can be proposed

for the physico-chemical reactions occurring on the 1393-Co bioactive glass

scaffold surface during immersion in SBF which are comparable to the reactions

described for 45S5-Cu derived scaffolds as shown in 4.1.2. However the kinetics of

the surface reactions also differ.

-1d: initial surface reaction and release of Na, K, Ca and Mg from the scaffold

periphery in an exchange reacting with H+ from the solution, e.g.: according to Eq.

5.13, 286

- 1-3d: Release of soluble Si species upon breakage of Si-O bonds by hydrolysis in

the surface region of the glass and formation of silanol groups triggered by the

reactions Eq. 5 and Eq. 6. In further reaction, surface SiO2 layer is formed upon

recondensation of SiOH groups according to Eq 7.

-7d: Formation of amorphous calcium phosphate (ACP) through migration of Ca2+

and PO43-

groups to the surface through the SiO2-rich layer. Crystallization of the

Page 131: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 131

ACP into carbonated hydroxyapatite (CHA) film by incorporation of OH-, CO3

2-,

Mg2+

. Also Co2+

ions diffuse through the SiO2 layer and are incorporated into the

CHA.324

In this work, however, too high Co-content (and release) lead to inhibition

of CaP layer crystallization remaining in its amorphous state.

Fig. 63: Elemental distribution in the cross-section of a 1393-5Co scaffolds after 7 days.

Reprinted with permission from A. Hoppe at al, ACS Appl. Mater. Interfaces 6 (2014), p. 2865.

Copyright (2014) American Chemical Society.

Degradation and ion release in SBF

Under static conditions

The Co and Si release from the 1393-Co glasses under static conditions is shown in

Fig. 65. From the 1393-1Co scaffolds 0.6±0.1 ppm Co are released after 1 d of

Page 132: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 132

immersion. However, during further immersion in SBF only slight increase in Co

was observed with maximum Co levels of 0.8±0.03 ppm after 21d of immersion.

For 1393-5Co scaffold an initial burst release of Co was detected during first 7d

days reaching 8.3±0.7 ppm. After 14 d the Co concentration drops to 5.4±0.07 ppm

and increases again to 7.3±0.6 pm after 21 d in SBF. The drop of the Co levels

might be due to precipitations of Co in insoluble calcium phosphate phases. Also

the saturated state of the SBF solution might lead to reincorporation of Co in the

glass structure or the surface calcium phosphate layer. It is known that

hydroxyapatite can act is cation exchanger an to incorporate metallic ions like Zn,

Cu, Co in it structure.325

Similar trend was observed for Su release showing burst

release in the first 7 days of immersion reaching levels of ~ 50 ppm independent

from the Co content in the glass. The Si concentration remains nearly constant for

further immersion in SBF. A little decrease in Si concentration is observed after 21

days of immersion. However, this is likely due to the static experiment conditions

leading to precipitation of insoluble salts incorporating Si. Overall, under static

conditions a plateau level of Co and Si release is reached after 7 days of immersion

in SBF.

Fig. 64: Co and Si release from 1393-Co glass scaffolds in SBF under static conditions.

Si release can be considered as marker for the glass degradation, so it can be

concluded that the glass degradation is retarded after 7 days due to the saturation of

Page 133: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 133

the solution in Si. Also, according to FT-IR, SEM and PIXE-RBS data, as discussed

in 4.2.2 , after 7 days a hydroxyapatite layer is formed on the 1393 scaffold surface

which is acting as diffusion barrier. Additionally, the saturation of the solution

decreases the diffusion rate of Si due to lower concentration gradient.

Under quasi-dynamic condition

Si and Co concentrations released from the 1393-Co derived scaffolds under quasi-

dynamic conditions (SBF solutions was changed frequently) are sown in Fig. 65.

Generally, higher Co and Si concentrations were released from 1391-Co derived

scaffolds under quasi-dynamic conditions compared to static conditions. Also the

kinetics of the ion release differ which are described in the following:

The global degradation of the scaffolds can be tracked using the Si release being the

main marker of the glass network dissolution. Two different regions can be

distinguished in the Si release profile of the 1393 scaffolds immersion under quasi-

dynamic conditions: i) initial release of high Si levels reaching a peak during the

first three days of the reaction and ii) drop of Si release to lower release rates. For

1393 glass this initial peak is reached after 3 days, Fig. 65, with maximal Si levels

of 11.3±1.7 ppm and further decrease of absolute Si levels to ~0-1 ppm. The drop of

the Si release rates corresponds to the formation of a CaP layer that acts as diffusion

barrier leading to decreased Si release. Similarly, for the 1393-5Co scaffolds

peaking Si values were observed after 3 days. However, the Si concentration of 18.8

±2.2 ppm was significantly higher compared to 1393 glass which can be explained

with the weaker glass structure caused through Co substitution as discussed in 4.2.1.

For the 1393-1Co samples, in turn, the Si peak is delayed, occurring after 14 days of

immersion in SBF. Maximal absolute Si concentrations of 18.3±1.0 ppm were

released after 14d in SBF.

The exceptional behaviour of the 1393-1Co glass scaffolds might be explained with

with PIXE-RBS results which showed that in contrast to 1393 and 1393-5Co 1393-

1Co scaffolds did not show the formation of a distinct CaP layer until the longest

time period investigated (7 days in SBF). Hence, the passivating CaP layer is likely

to form at a later stage thus retarding the Si release peak from the 1393-1Co

samples.

Page 134: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 134

Co is continuously released from the 1393-1Co and 1393-5Co scaffolds reaching

maximal cumulative values of 1.73±0.03 ppm and 11.4±0.2 ppm, respectively.

Fig. 65: Co and Si release from 1393-Co glass scaffolds in SBF under quasi-dynamic

conditions. Adapted with permission from A. Hoppe at al, ACS Appl. Mater. Interfaces 6

(2014), p. 2865. Copyright (2014) American Chemical Society.

Regarding the absolute values released after each time point of measurement the

highest Co releasing rated occurs in the first 3d of immersion. It is most evident for

the 1393-5Co samples which showed a Co peak of ~4 ppm after 3d of immersion in

SBF. After 7 days of immersion times the Co concentration dropped to ~1 ppm

remaining at this level for further immersion until 21 days. This is likely due to the

formation of a CaP layer on the scaffold surface after 3d as it was observed from

SEM and PIXE-RBS analysis. This layer likely acts as a diffusion barrier leading to

the drop of Co released into SBF.

After 21d in SBF the change in the scaffold mass was +4.3±0.6 %, -18.7±0.6 % and

-39.6±0.6 for 1393, 1393-1Co and 1393-5Co scaffolds, respectively. The slight

mass gain in the 1393 samples is likely due to the formation of the calcium

Page 135: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 135

phosphate layer on the scaffold surface, while at the same time the 1393 scaffold

showed the lowest degradation rates as shown by to ICP OES measurements. After

7d of immersion in SBF the Si release dropped to values close to 0 ppm due the

formation of CaP layer acting as a diffusion barrier. Accordingly, the cumulative

values of Si released in SBF quickly reached an almost constant plateau at ~30

ppm. Hence, it can be assumed that the dissolution of 1393 derived scaffolds was

nearly completely impeded after 7d of reaction in SBF. In contrast, for both cobalt

containing scaffolds1393-1Co and 1393-5Co a continuous release of Si was

detected over a reaction time range of 21d. Even though for these samples also a

drop in absolute Si values released was observed the Si release rates remained at

significant levels of ~1.5-3 ppm/d. Thus, the mass loss of the 1393-1Co and 1393-

5Co is assigned to significantly higher Si release and degradation rates of these

scaffolds. After the quasi-dynamic degradation study the scaffolds were digested

and the fractions of the remaining oxides were analysed. The relative concentrations

of the elements are given in Table 15

Table 15 normalised to the SiO2 content. The relative glass compositions after the

degradations study reflects the dissolution behaviour of the scaffolds.

Table 15: Oxide contents normalized to silica in the glass remaining after the degradation

study.

Glass CaO P2O5 NaO MgO CoO

1393 0.45±0.02 0.51±0.06 0.14±0.00 0.05±0.02 -

1393-1Co 0.37±0.03 0.44±0.04 0.14±0.02 0.04±0.02 0.02±0.00

1393-5Co 0.08±0.02 0.08±0.02 0.06±0.01 0.02±0.01 0.03±0.01

For Co-containing samples higher loss in Ca and P is observed compared to 1393

glass which corresponds to higher Si release rates in SBF. Accordingly, Na and Mg

loss is highest for 1393-5Co scaffolds. For Co-containing samples a residual

amount of CoO of 2 wt% and 3 wt% were observed for 1393-1Co and 1393-5Co

samples, respectively.

The results on the degradation behaviour of the 1393-Co glass indicate that the

1393 glass matrix is suitable for controlled release of Co ions in physiological

Page 136: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 136

environment. For example, Azevedo et al.64

reported Co values of ~13 ppm released

from melt-derived glasses in TRIS buffer after 21 days without refreshing the

solution. Also, Wu et al.219

have shown controlled release of maximum Co levels of

~ 20 ppm from Co containing sol-gel derived scaffolds. Since too high Co

concentrations can be toxic a controlled release mechanism is essential for

applications of such Co-releasing constructs. The results presented here showed that

13-93 glass derived scaffolds can be used for controlled release of Co2+

ions with

release rates (after the initial burst release) of 0.3 ppm/d and 0.1 ppm/d depending

on the glass composition.

The Co concentrations observed in this work are comparable to the values reported

to be within therapeutically active range. For example 50 µM, 100 µM, 200 µM of

CoCl2, which correspond to 12 ppm, 24 ppm and 48 ppm, respectively, stimulated

migration, proliferation, and tubule-like structure formation of umbilical cord

blood-derived CD133(+) cells, hence indicating angiogenic potential of Co inducing

hypoxic conditions.326

Others confirmed treating human microvascular endothelial

cells (HMEC-1) with 12 ppm CoCl2 resulted in binding of HIF-1 hence mediating

transcriptional responses to hypoxia.327

However, Co levels higher than 10 ppm

have been indicated to be cytotoxic as treatment of osteoblast-like cells with 10

ppm of Co2+

resulted in 40% decrease in cell number.328

Similarly, Wu et al.219

observed that Co2+

concentration of ~20 ppm reduced the viability of bone marrow

derived stem cells (BMSCs) even though they did not cause significant cytotoxicity.

Hence, the release of ~2 ppm from 1393-1Co may be considered non-critical

regarding potential cytotoxicity whereas ~12 ppm released from 1393-5Co might be

at the upper edge of the therapeutic, non-toxic range.

Altogether, the addition of Co into the 1393 glass network results in the weakening

of the glass structure leading to faster dissolution of the glass (under quasi-dynamic

conditions). 1393 bioactive glass composition is considered less reactive than 45S5

BG due its high SiO2 content. Hence, through adding CoO in to the glass network it

is possible to enhance the reactivity and the degradation rate of 1393 glass leading

to improved bioactivity. So, degradation of the scaffolds can be adjusted by

tailoring the composition of the glass hence enabling controlled release of ionic

species. CoO acts as network modifier and network former depending on its

Page 137: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 137

concentration in the glass. That way, the 1393 glass matrix might be used as carrier

for controlled release of other therapeutic ions like ZnO or MgO also known to act

as intermediate oxides in silicate glasses.329

4.2.3 In vitro cell response

The potential cytotoxicity of Co ions released from metal alloys used as implant

materials in hip replacement therapies is known problem in the current medicine

which is related to systemic intoxication of the body through long term Co (and Cr)

release as results of wear debris and corrosion.330-332

However, as mentioned in

2.5.1 Co ions are also known to enhance angiogenesis via inducing hypoxic

conditions. Using 13-93 glass derived scaffolds it is possible to control the Co

release which can be adjusted within a therapeutic range. In order to test this

hypothesis n this section the biocompatibility of the 1393-Co glass powders and

glass derived scaffolds is presented and discussed. Firstly, powder cytotoxicity of

the particulate BGs is assessed in direct contact with MG-63 cells. Secondly, the

biocompatibility of the 1393-Co scaffolds is evaluated regarding the response of

MG-63 cells directly seeded on the scaffolds and of hDMECs exposed to ionic

dissolution products of the 1393-Co derived scaffolds.

Powder cytotoxicity

The cell morphology of the osteoblast-like cells after cultivation in direct contact

with 1393-Co glass particles for 48h is shown in Fig. 66. No signs of cytotoxicity

can be observed for 100 mg ml-1

particle concentrations: the cells exhibit slightly

elongated triangle shaped morphology typically found for MG-63 cells. However,

for 1000 µg ml-1

fewer cells seem to be present even though the cell number is

difficult to observe as at 1000 µg ml-1

the BG particles cover a large area of the well

plate.

Page 138: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 138

Fig. 66: Light microscopy images of the MG-63 morphology cultured with 1393-Co glass

particles for 48h.

Fig. 67 shows the mitochondrial activity and cell number of osteoblast-like cells

derived from WST and LDH assays, respectively. For 1393 reference glass the

mitochondrial activity of the cells remains nearly constant for 0.1 and 10 µg ml-1

but is increased for concentrations of 10 and 100 µg ml-1

, while the cell number

stays constant for 0.1-1000 µg ml-1

range.

Compared to the plain 1393 reference, 1393-1Co shows statically significant higher

mitochondrial activity of MG-63 observed in the range from 0.1 µg ml-1

to 100 µg

ml-1

. The increase in the cell activity is statistically even higher (p<0.01) when

compared to 1393 reference. At the same time the cell number remains constant for

0.1 – 10 µg ml-1

whereas for higher particle concentrations a slightly reduced cell

number was observed. This indicated that even though the cell number is slightly

reduced or remains constant the cells show a higher mitochondrial activity detected

by the WST assay stimulated by the 1393-1Co glass particles.

For 1393-5Co glass particles, however, a clear reduction in cell activity and cell

number was observed for particle concentrations > 10 µg ml-1

indicating potential

cytotoxic effect of 1393-5Co glass. Altogether the 1393 glass particle at

concentrations in the range of 0.1-100 µg ml-1

showed good biocompatibility with

osteoblast-like cells with enhanced cell activity with 1393-1Co particulate glass and

indication of cytotoxicity of 1393-5Co glass.

Page 139: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 139

Fig. 67: Mitochondrial activity (WST assay) and cell number (LDH assay) of osteoblast-like

cells seeded for 48 h in direct contact with particulate 1393-Co glasses. *p<0.05; **p<0.01

compared to 0 µg/ml reference.

Scaffolds

Fig. 68 shows the mitochondrial activity of MG-63 cells seeded directly on the

scaffolds. 1393-1Co did not show any cytotoxic effects. In fact after an initial drop

after 7d the cell activity was significantly increased for 1393-1Co samples

compared to 1393 reference. However, 1393-5Co samples show signs of

cytotoxicity from the beginning: after 3 days the cell activity dropped to 20% of the

reference value and further decreased to ~5% after 7d and 14d indicating that

almost no living cells were grown on the scaffolds.

Fig. 69 shows osteoblast-like cells grown on 1393-Co derived scaffolds after 14d

observed with SEM and confocal laser scan microscopy. For 1393 and 1393-1Co

god cell adhesion and cell spreading was observed. The cells fully cover the

scaffold strut and grow following the pore curvature. It has been shown in literature

that cell prefer the growth along the inner side of the pores in case of large pores.333

However, on the 1393-5Co scaffolds no spread cells were observed indicating

cytotoxic effects of the1393-5Co glass scaffold. These observations were confirmed

through WST measurements, as shown in Fig. 68.

Page 140: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 140

Fig. 68: Mitochondrial activity of osteoblast-like cells seeded on 1393-Co scaffolds for 3d, 7d,

and 14d.

The toxic effect of 1393-5Co is likely due to the high concentration of Co ions

releases. Wu et al.219

for instance showed that 22 ppm caused a reduction of cell

vitality of human stem cells after culturing on Co containing BG scaffolds for 1

week. Even though in this study no significant cytotoxicity was observed, high Co

concentration seems to have a negative effect on the vitality of stem cells. Further in

the study of Wu et al.219

only short term effects were considered for a culturing time

of 1 week. Indeed it has been shown in literature that the potential cytotoxicity of

Co is not only dose but also time dependent.328

This might explain the cytotoxicity

of the 1393-5Co scaffolds even though the 12 ppm Co released are within non-toxic

ranges reported for Co ions.328

The cytotoxic effect of Co on MG-63 cells may be

related to oxidative stress caused by Co2+

ions leading to oxidation of proteins.328,

334

Regarding the response of MG-63 cells it can be summarised that 1393 and 1393-

1Co scaffolds show no signs of cytotoxicity when applied as powder and also as 3D

scaffolds. Furthermore, 1393-1Co scaffolds show slight stimulating effects on

mitochondrial activity of MG-63 cells when in contact with particulate glass at

concentrations from 0.1-100 µg ml-1

and seeded on the scaffolds. 1393-5Co glass

powder and corresponding scaffolds, however, show signs of cytotoxicity.

Page 141: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 141

Fig. 69: SEM images (left) and confocal laser scan images (right) of Vybrant live staining (red)

of MG-63 cells cultured for 14d on 1393 (a-b), 1393-1Co (c, d) and 1393-5Co scaffolds.

Response of hDMECs

Fig. 70 shows the light microscopy images of hDMECs culture for 2 weeks exposed

to the ionic dissolution products from 1393-Co series. The 1393 control shows and

1393-1Co scaffolds show good biocompatibility: the cell morphology is retained

while for 1393-1Co indications of tube-like formation of the cells are visible. 1393-

5Co scaffolds, however, seem to be toxic. The morphology of the hDMECs was

impaired when seeded in contact with 1393-5Co scaffolds. These observations were

confirmed through the cell number analysis: the cell number increased continuously

for the hDMECs reference from initial number of 100.000 cells to 160.000 after

1week and 250.00 after 2 weeks of culture. For 1393 samples the cell number

Page 142: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 142

increased to comparable values of 150.000 cells after 1 week and subsequent slight

reduce to 140.000 cells after 2 weeks.

Fig. 70: LM images of hDMECs when seeded for 2 weeks in contact with dissolution products

from 1393, 1393-1Co and 1393-5Co scaffold. hDMECs only served as control.

For 1393-1Co a slight reduction of the cell number from 100.000 to 90.000 and

50.000 after 1 week and 2 weeks, respectively, was observed. For 1393-5Co,

however, already after 1 week no living cells could be detected. These results were

confirmed by LDL uptake analysis which showed high LDL uptake of the hDEMCs

reference, 1393 and 1393-1Co while no LDL uptake was observed for 1393-5Co

samples, Fig. 71. Furthermore, FACS analysis revealed pronounced expression of

vWF, CD31 and VEGFR endothelial cell markers in hDMECS control and 1393

and 1393-1Co samples confirming high functionality and retained endothelial

phenotype of the hDMECs, Fig. 72.

Page 143: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 143

Altogether, 1393 and 1393-1Co scaffolds show good biocompatibitly towards

endothelial cells while 1393-5Co is cytotoxic. As dissolution products from the

scaffolds were investigated in an indirect cell culture set up it can be concluded that

the release of 2 ppm Co from 1393-1Co samples is within compatible range for

endothelial cells. However, the release of ~ 10 ppm Co2+

from 1393-5Co scaffold

caused significant cytotoxicity.

Fig. 71: Florescence images of the LDL uptake of the hDMECs reference (cells only) and in

presence of 1393-1Co and 1393-5Co scaffolds.

In conclusion of 1393-1Co glass and respective scaffolds are not toxic to the

HDMSCs which show high LDL uptake. Also after culturing with 1393-1Co

hDMSCs have been shown to be positive for CD31, vWF and VEGFR-2

confirming the vital phenotype of endothelial cells.

Page 144: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 144

Fig. 72: FACS analysis of endothelial markers expressed in hDMECs cells after 1 week of

culture in presence of 1393-Co scaffolds.

Altogether, good cytocompatibility was confirmed for 1393 and 1393-1Co scaffold

regarding the response of osteoblast-like cells as well as endothelial cells. However,

1393-5Co seems to have a cytotoxic effect on osteoblast-like cells as well as

hDMECs and thus might be not suitable for tissue engineering applications.

Obviously, the release of 12 ppm exceeds the therapeutic range of Co2+

ions

becoming toxic to the cells.

However, 1393-1Co glass derived scaffolds might be potential candidates for use as

hypoxia mimicking biomaterials for bone tissue engineering applications even

though a significant stimulating effect on endothelial cells was not confirmed in this

study. Also, there is convincing evidence in literature reporting stimulating effects

of Co on angiogenesis and bone tissue regeneration processes.179, 200, 326

The

stimulating effect of Co on angiogenesis, however, is suggested to be tested in a co-

culture of ECs and MSCs.

Page 145: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 145

4.3 Evaluation of the compressive strength of the 45S5 and

1393 scaffolds in the context of bone TE

In order to evaluate the mechanical properties of the bioactive glass scaffolds the

influence of the porosity has to be considered. The theoretical strength σtheo of

cellular ceramics with open cells can be estimated by the model of Gibson and

Ashby:335

𝜎𝑡ℎ𝑒𝑜

𝜎𝑓𝑠= 𝐶 (

𝜌𝑓𝑜𝑎𝑚

𝜌𝑠𝑜𝑙𝑖𝑑)

3

2∙1+(

𝑡𝑖𝑡)2

√1−(𝑡𝑖𝑡)2 Eq. 8

= 𝐶(1 − 𝑃)3/2 ∙1+(𝑡𝑖/𝑡)

2

√1−(𝑡𝑖/𝑡)2 Eq. 9

where σfs, is the modulus of rupture of the struts of the foam, C is constant of

proportionality (for brittle ceramics it C=0.2), ρfoam is density of the foam, ρsolid is

the density of the solid ceramic, P is the total porosity of the foam and ti/t is the

ratio of the central void size of the strut to the strut diameter.

From theoretical calculations it can be assumed that for brittle materials σfs =1.1σts

with σts being the tensile strength of the bulk material. With σts of 42 MPa for

annealed 45S5 Bioglass®

and ti/t ratio of 0.5 (derived from Fig. 20) the theoretical

values for compressive strength of 45S5 BG derived scaffolds were calculated as

depicted in Fig. 73. For comparison also literature values for standard HAp bio

ceramic derived scaffolds by foam replica technique are given.

Considering the effect of porosity the values obtained for 45S5 derived scaffolds are

above the values reported for HAp scaffolds also fabricated with foam replica

method. Hence, 45S5 derived scaffolds exhibit comparable compressive strength as

porous scaffolds based on the standard HAp bioactive ceramic. In fact, in order to

match the lower limits of the compressive strength of cancellous bone (2MPa) with

of 45S5 derived scaffolds porosity values < 71% would be required. This, however,

might not be sufficient to meet the requirements for bone tissue engineering

scaffolds regarding porosity needed for cell and blood vessel ingrowth.

Page 146: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 146

Fig. 73: Theoretical values for the compressive strength of open-cell 45S5 BG derived scaffolds

and the experimental values obtained in this work after for multiple coated scaffolds. For

comparison reported values for HAp foam made by replica technique were taken from

literature.336, 337

In other studies it has been reported that higher values for σc for bioactive glass

derived scaffolds can be achieved by modifying the regime of the foam replica

technique resulting in σc values up to 3 MPa.338, 339

However, in those studies the

porosity of the scaffolds were ≤ 70% which is not comparable to porosity levels of

the scaffolds fabricated in this work (~90 %).

Despite the limitations of relatively low mechanical strength of Bioglass®-derived

scaffolds fabricated by the foam replica method, this material remains an important

system in the context of bone tissue engineering due to its high bioactivity and

ability to actively stimulate cells towards osteogenic differentiation via up-

regulation of genes expression which results in enhanced bone regeneration.129

Furthermore, according to literature reports one can assume that in vivo culturing

will increase the mechanical strength and toughness of such scaffolds due to tissue

ingrowth and HAp formation on their surfaces thus leading to formation of a

“biocomposite” as for example shown for hydroxyapatite scaffolds.340

In order to overcome the drawback of limited mechanical stability of 45S5 scaffolds

1393 bioactive glass composition was alternatively investigated in this work. As

shown in the results section (4.2.2) for 13-93 derived scaffolds compressive

strength values of >2 MPa were achieved. The modulus of rapture of a strut is

Page 147: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 147

defined as the maximums stress at failure of the cell wall material. Given the

average strut thickness of ~ 100 µm the tensile strength of 1393 fibers of 440 MPa

(diameter of 93-160 µm) was assumed for σfs.86

According to Fig. 55 the ratio ti/t

was estimated to be between 0 and 0.5. Fig. 74 shows the theoretical strength and

experimental values observed in this work for 1393-Co glass derived scaffolds. Also

compressive strength values reported by Fu et al. for 1393 glass derived scaffolds

made by foam replica technique are given for comparison.69

The experimental

values observed in this work follow the curve for the compressive strength as

predicted by Eq. 9. It is worth noticing that by changing the foam replica regime the

compressive strength of the scaffolds can be enhanced while high porosity of 89-

91% is maintained.

Fig. 74: Theoretical values for the compressive strength of open-cell 1393 derived scaffolds and

the experimental values obtained in this work. Scaffolds with porosities lower than 92% were

made by double-coating. Grey shadowed area marks the region of human spongy bone. For

comparison literature values reported by Fu et al.69

are given. Adapted with permission from

A. Hoppe at al, ACS Appl. Mater. Interfaces 6 (2014), p. 2865. Copyright (2014) American

Chemical Society.

Taking the differences in porosity into account the compressive strength of the

present scaffolds are comparable with values reported by Fu et al.69

for 1393

derived scaffolds even though those values are slightly above theoretical strength.

This might be due to the fact that for theoretical calculation of the strength a

Page 148: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 148

constant modulus of rupture is assumed; in praxis, however, the rapture modulus is

not a constant but varies following the Weibull distribution.335

Also the strength

values reported for dense 1393 material (which is taken for calculations of the

porous foams) are given in literature with a high standard deviation (440±120 MPa)

which allows only rough estimation. Altogether, the σc values obtained for the 1393

derived scaffolds correspond to the lower limits of human spongy bone which could

enable their use for regeneration of moderate load-bearing bone defects.38

4.4 Degradation behaviour of 45S5 and 1393 glass

scaffolds and their suitability as carrier for therapeutic

metal ions

The detailed surface reactions of 45S5 and 1393 BG derived scaffolds in SBF are

described in sections 4.1.2 and 4.2.2, respectively. Despite the different kinetics of

these surface physico-chemical reactions no major differences were observed

between these two glass systems in their overall bioactive character. The surface

reactions of a bioactive glass surface are schematically given in Fig. 75. Basically,

the surface reactions follow the scheme as proposed by Hench13

with the formation

of a SiO2 layer, the precipitation of an amorphous calcium phosphate and

crystallisation of carbonated hydroxyapatite being the main steps, Fig. 75. Hereby,

one important finding concerning the metal ion containing glasses was that metal

ions released from the glass network are incorporated in the carbonated

hydroxyapatite layer on the bioactive glass surface. This model is valid for the

surface reactions of both 45S5 and 1393 BG surfaces. Despite some minor

differences the reaction scheme showed in Fig. 75 is identical for both glass systems

describing the local physical-chemical reactions at the BG scaffold/fluid interface.

Page 149: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 149

Fig. 75: Scheme summarising the surface reactions on a BG derived scaffolds based on the

experimental results given in 4.1.2 and 4.2.2 (from left to right): formation of silanol groups

through protonation of free Si-NBOs and breakage of Si-O-Si by hydrolysis; formation of a

silica layer upon condensation of the silanols and formation of amorphous calcium phosphate,

ACP (or rather mixed layer of SiO-xCaP); crystallisation of ACP to carbonated hydroxyapatite

(CHA) which is enriched in metallic ions Me2+

, e.g. Cu2+

or Co2+

.

However, the global degradation process of these two BG systems seems to differ.

For 45S5 BG derived scaffolds the surface reaction and the degradation is

accompanied by release of soluble silica occurs not only from the surface area but

also from the inner region of the glass network. Clearly, dissolution signs in the

inner region of 45S5 derived scaffolds were observed by SEM. Also micro-PIXE-

RBS measurements clearly indicated the chemical changes in the inner regions of

the scaffolds during immersion in SBF. Furthermore, despite high levels of Si

released, no significant mass loss of the scaffolds was observed for 45S5-Cu series

which indicates that significant amount of silica was released from the inner regions

and the mass loss caused by the Si release was compensated with the mass gain of

the CaP layer. The mechanism for the global degradation of the 45S5 BG scaffold is

shown in Fig. 76.

Page 150: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Results and Discussion 150

Fig. 76: Reaction scheme of a 45S5 BG derived scaffold during immersion in SBF: a) initial

strut of the scaffold; b) first surface reaction and leaching of ions in the scaffold periphery; c)

formation of SiO2 layer; d) formation of CaP layer on the scaffold surface and dissolution of

the glass from the inner region; e) CaP continues to growth and the scaffold further dissolves

from inner region.

Tilloca summarised that the 45S5 bioactive glass network can be easily penetrated

by water molecules in the inner region which is due high fragmentation of the

network and the high hydrophilicity of Ca and Na cations which results in high H2O

affinity.287

This allows easy penetration and migration of H2O molecules without

the need of breakage of silanol groups that bears high amounts of energy. These

effects cause the high reactivity of 45S5 scaffolds.

For 1393 derived scaffolds, in turn, no dissolution signs of the inner region of the

scaffolds were observed. The loss of soluble silica occurs solely in the surface

region of the glass. Also with addition of Co to the 1393 matrix the glass

degradation was enhanced which was monitored by significant mass loss and high

Si release rates. The scheme of the 1393 derived scaffolds is shown in Fig. 77. In

contrast to 45S5 the degradation takes place from the surface of the scaffolds with

corresponding weight loss, while the inner region remains intact. This can be

explained with the network connectivity of the BG:287

45S5 has lower amount of

SiO2 hence, lower network connectivity which enables the migration of water

molecules deep in the glass network. 1393 glass, on the other hand, has a higher

amount of SiO2 (53 wt%). And hence higher network connectivity. Thus, the

migration of the eater molecules reaches only the periphery region. And the

dissolution takes place from the surface region of the scaffold. The reactivity of

BGs with higher network connectivity (which is the case for the 1393 BG with

Page 151: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

151

higher SiO2 content) is reduced since the hydrolysis of silicate units with three or

more bridging oxygens (BOs) bears an excessive amount of energy costs.76

Fig. 77: Reaction scheme of a 1393 BG derived scaffold during immersion in SBF: a) initial

strut of the scaffold; b) first surface reaction and leaching of ions in the scaffold periphery; c)

formation of SiO2 layer; d) formation of CaP layer on the scaffold surface; e) the scaffold

degrades through material loss via dissolution of the surface region.

These are interesting findings as the degradation behaviour dictates the suitability of

the glass network to be used as carrier for therapeutic ions and as a biomaterial in

clinical application. From the finding in this work it can be stated that 45S5 derived

scaffolds remain in its original shape and the degradation occurs in the inner part of

the scaffolds via leaching of cations and soluble Si species. In praxis this means that

45S5 derived material will be integrated in the newly forming bone tissue being

converted to calcium phosphate phase while a core of SiO2 network will remain. In

contrast, 1393 BG derived scaffolds with enhanced degradation achieved through

modification with CoO, for instance, would be resorbed and metabolised by the

body within few weeks.

Page 152: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Summary 152

5 Summary

In this work novel Cu and Co containing bioactive glass compositions based on

45S5 (45S5-Cu) and 1393 BG (1393-Co) have been successfully synthesized and

characterised in terms of glass structure and glass properties. Furthermore, 3D

porous scaffolds have been fabricated using the foam replica technique and their

porosity, acellular bioactivity in simulated body fluid (SBF), degradation and ion

release kinetics as well as the biocompatibility in vitro and in vivo was assessed.

Cu was successfully incorporated as Cu2+

acting as network modifier in 45S5

bioactive glass network which resulted in a decrease of glass transition and lower

melting onset temperature, hence enlarging the process window for this glass.

Highly porous 45S5S-Cu scaffolds with interconnected pore system and total

porosity of ~90 % were fabricated by foam replica technique. Compressive strength

values of ~0.2 MPa were tested which are relatively low but sufficient for the

handling of the scaffold in tissue engineering applications. High acellular in vitro

bioreactivity of the Cu-containing scaffolds was confirmed through SBF test

showing rapid formation of carbonated hydroxyapatite (CHA) layer on the scaffold

surface after 3d without any negative impact of Cu doping on the bioactivity of

45S5 BG scaffolds. Further, detailed micro-PIXE-RBS analysis revealed that traces

of Cu were incorporated in the CHA layer. Degradation studies in SBF showed that

Cu levels from 0.3 to 4.5 ppm are released from 45S5-Cu scaffolds depending on

culturing conditions which are within the therapeutic ranges reported for Cu ions

(see 4.1.4). Cell culture assays confirmed the high compatibility of 45S5-Cu

particulate glasses and corresponding scaffolds with MG-63 osteoblast-like cells,

human bone marrow derived stem cells (hBMSCs) as well as human dermal micro

vascular endothelial cells (hDMECs). Furthermore, a significantly enhanced VEGF

(vascular-endothelial growth factor) expression was found in hBMSCs induced by

Cu ions released from the scaffold likely due to the activation of the HIF-1

transcriptional factor. In a co-culture study of hDMECs and hMSCs 1 wt% Cu

containing 45S5 scaffolds were shown to enhance the expression of endothelial cell

specific markers vFW and VEGFR and to stimulate hDMECs towards formation of

prevascular tube-like structure which is an indication of the overall angiogenic

Page 153: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Summary 153

potential of Cu. The AV-loop in vivo model in the rat showed that 45S5 derived

scaffolds support intrinsic vascularisation through micro vessel sprouting from the

initial AV loop. Qualitatively, 45S5-Cu derived scaffolds showed higher blood

vessel density and higher number of vessels than plain 45S5-derived scaffolds even

though the effect of Cu was not statistically significant.

Complementary to the 45S5 BG, a 1393 BG composition containing CoO was

fabricated. CoO was shown to act in a concentrations depending manner in the glass

network acting as network modifier for concentration of 1 wt% and entering the

glass network by forming Si-O-Co bonds for concentrations ≥5 wt% Co. Inclusion

of Co weakened the glass network due to replacing of the Si-O bonds by weaker Si-

O-Co bonds as indicated by decreased glass transition temperature with increasing

CoO content in the glass. Using the foam replica technique also highly porous

scaffolds with porosities of 89-92 % depending on the foam replica regime were

achieved. 1393-Co glass derived scaffolds revealed relatively high compressive

strength values of > 2MPa, which correspond to the lower boundaries of strength

values of human cancellous bone. Further, acellular in vitro studies revealed rapid

transformation of the 1393 scaffolds surface to CHA after 7d in SBF indicating high

bioactivity even though it is retarded compared to 45S5 derived scaffolds. Similar

to observations with 45S5-Cu derived scaffolds, Co is incorporated in the surface

calcium phosphate layer, as shown from micro-PIXE-RBS analysis. Moreover,

when applied at high concentrations (≥ 5wt% CoO in the glass) Co ions seem to

inhibit the crystallisation of the calcium phosphate surface layer remaining in the

amorphous state. Degradation studies in SBF revealed that Co ions in the range 0.6-

11.4 ppm are released depending on the glass composition and culturing conditions

which is in the therapeutic range (see 4.2.3). In vitro cell assays showed that 1393

and 1393 with 1 wt% CoO glass derived scaffold exhibit high cell compatibly with

a tendency of 1393-1Co glass to stimulate MG-63 cells and endothelial cells.

However, 1393-5Co was shown to be cytotoxic with Co2+

concentration of ~11 ppm

seemingly exceeding the physiologically vital level. Overall, the results of this

research project showed that 45S5 and 13-93 bioactive glass derived scaffolds

represent a new promising family of scaffold for bone regeneration which serve as

inorganic carriers for controlled release of therapeutic metal ions.

Page 154: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Conclusion and Outlook 154

6 Conclusion and Outlook

Applications of metal ion doped glasses in bone tissue engineering and beyond

The aim of this work was to produce robust inorganic scaffolds based on bioactive

glasses with the capability of controlled ion release to enhance bone formation and

angiogenesis. Bioactive glass systems of 45S5 and 1393, two well-known silicate

compositions, were shown to be suitable inorganic materials for incorporation of

biologically active metallic ions, e.g. Cu and Co. These metal ion containing glasses

are suitable carrier for the controlled release of such therapeutic metal ions because

the final ions concentration and release kinetics can be tailored by the glass

composition. Due to high flexibility of the glass melting process a wide variety of

different bioactive metallic ions and other bioinorganics can be incorporated

underlining the high potential of silicate bioactive glasses in the field of

biomaterials and regenerative medicine. Incorporation of osteogenic and angiogenic

agents for example is a promising strategy to enhance the impact of bioactive

glasses for bone tissue engineering applications and to enhance bone regeneration.

Various novel applications of ions for specific applications are possible. Osteogenic

agents like Li or Ga which have not been extensively investigated so far are

promising candidates to be substituted in a BG matrix. Also antibacterial agents

like, Zn, Ga or Ag can be incorporated in glass matrix creating multifunctional

bioactive glasses. Since inorganic therapeutics show many advantages160

this might

have a huge impact on the design and applications of novel osteogenic and

angiogenic materials. Inorganic therapeutics may be considered cheaper and safer

alternatives compared gene therapies or applications of growth factors which have

some drawbacks, such as uncontrolled release, degradation and tendency to diffuse

from the inflammatory site the as well as formation of malign tissue.341

Hence, the

use of therapeutic inorganics might give a cost-effective and safe solution for

enhancing the biological impact of biomaterials in regenerative medicine. Also

related to processing bioinorganics are advantageous since they can be processed at

high temperature using classic technologies of ceramic processing.

Page 155: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Conclusion and Outlook 155

Besides BTE, various other potential fields of medicine can be explored by using

novel metal ion containing bioactive glasses including nerve guidance conducts342,

343 and cancer treatment.

344-349 Ferromagnetic BG particles, for example, can be

used for hyperthermia cancer treatment.350

Hereby, BGs also provide

osteoconductive properties and hence can combine cancer treatment and bone

regeneration in one procedure. Also as guided drug-release systems or in tissue

engineering magnetic F2O3 containing BG particles can be additionally

functionalised with drugs and then specifically targeted to the site of action.

Furthermore, magnetic BG particles can be used as biocompatible targets (as

alternative to potentially toxic iron oxide particles) for placing inside cells for

guiding the cells into particulate shapes and geometries, e.g. assembling endothelial

cells in a 3D vessel-like structure for engineering of vascularised tissue constructs.

Furthermore, ZnO2 and CeO2 which are known to be involved in peripheral nerve

regeneration351, 352

and as neuroprotective agent353

, respectively, can be proposed as

therapeutic agents to be released from BGs used in nerve guidance constructs.

Degradation and reactivity of BG scaffolds in biological fluids

As presented in this work, the kinetics of the ion release from both 45S5 and 1393

show a burst release in the first 7 days of degradation in SBF, followed by further

continuous release over a period of 21d. This profile comes close to a “drug

delivery system” where an initial therapeutic effect is desired through a high “drug”

release in the beginning followed by a continuous release in the so-called

therapeutic zone in the 2-3 weeks following the (bone) defect treatment or

regeneration. Hence, 45S5 and 1393 BGs can be considered clinically relevant

“drug-release” carriers. However, the ion release capability in vivo may be different

to the in vitro situation and, hence, such in vivo studies remain a task for future

investigation. Indeed, there is a clinical need for such carriers of inorganic

therapeutics.

Also the degradation behaviour in SBF of the 45S5 and 1393 glass scaffolds was

assessed in detail and the results were compared revealing major differences in the

degradation mechanisms of these two silicate glass systems: while 45S5 scaffolds

degrade via leaching of cations and release of soluble silica from the inner regions

Page 156: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Conclusion and Outlook 156

of the scaffolds, with the mass loss compensated by calcium phosphate formation

on the surface, the degradation of 1393 scaffolds occurs in the surface region

without signs of dissolution of the inner parts. These observations should help in

designing novel glass compositions based on the 45S5 and 1393 glasses and

respective scaffolds predicting their degradation and ion release kinetics.

Detailed analysis of physico-chemical reactions occurred at the scaffolds /

biological fluid interface showed that Me2+

ions released from the BG matrix are

incorporated in the calcium phosphate layer formed during reaction in biological

fluids. This is an important finding for evaluation of the mineralization behavior and

biological performance of metal ion doped BG derived scaffolds regarding the

material-cell interaction which is known to be strongly influenced by materials

surface chemistry.354

Also this provides new aspects for understanding of

mechanism behind the materials cells interaction which is not only influenced by

ionic dissolution products but also by the surface chemistry. This implies also the

possibility to specifically modify the chemistry of a BGs and respective scaffolds.

Indeed, there is a wide field of biomaterials research which deals with doping of

calcium phosphates with bioinorganics.355

Biological impacts of metal ion containing BGs

Cu ions released from 45S5-Cu BG derived scaffolds were shown to enhance

VEGF expression in mesenchymal stem cells (MSCs) hence activating angiogenesis

related signalling pathways. In a co-culture model with endothelial cells (ECs) and

MSCs Cu simulated formation of tube-like prevascular structures was observed

indicating stimulated vascularisation. Since sufficient vascularisation is essential for

successful clinical application of engineered bone constructs the new developed Cu

containing 45S5 BG derived scaffolds are promising materials for applications in

regenerative medicine including bone tissue engineering and wound healing.

This work revealed some interesting results regarding the role of copper in

angiogenesis. The stimulating effect of Cu on the VEGF expression in stem cells is

described in literature. However, only the combination of 45S5 bioactive glass, Cu

ions and a co-culturing with stem cells seems to lead to stimulation of endothelial

cells. Hence, advanced cell culture studies are further needed to fully unveil the

Page 157: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Conclusion and Outlook 157

detailed intracellular mechanisms and the role of Cu (and other trace elements) in

cell-cell interaction and cell gene expression. Knowing the effect of Cu on a

specific type of cells and combination of those will allow designing bone

engineering constructs with tissue specific functionalities.

Even though a specific stimulating effect on endothelial cells of Co could not be

detected, the 1393-Co glass and its scaffolds remain interesting materials which

should be considered as potential candidate for use as hypoxia mimicking materials

in bone tissue regeneration applications as it is known that HIF-1 stabilised by

hypoxia is one of major transcription factors for regenerative process in bone

defects and wound healing. The use of co-cultures and alternative experimental cell

setups is suggested in order to obtain more information about the hypoxia effect

caused by Co ions indicated in literature.

45S5 vs. 1393 glass compositions

In accordance to literature 45S5 BG derived scaffolds showed high Si release rates

and hence a rapid degradation as well as fast in vitro mineralisation in simulated

body fluid making them promising materials bone tissue engineering applications.

However, 45S5 BG shows some drawbacks as it is difficult to process due to its

tendency to crystallise during high temperature treatment which results in poor

densification and micro-crack formation and hence in poor mechanical properties of

the scaffolds (~0.2-0.3 MPa compressive strength). Hence, depending on the

applications 45S5 based glasses are suggested where high reactivity and rapid ion

release is needed. 1393 BG derived scaffolds, in turn, can be fully densified without

crystallisation. 1393-Co derived scaffolds exhibiting compressive strength > 2 MPa

with high porosities were fabricated which enables the application of such scaffolds

for regeneration of cancellous bone defects.

Considering the lower reactivity of the 1393 glass compared to 45S5 and, thus,

reduced acellular bioactivity in SBF, the inclusion of therapeutic ions in the 1393

glass matrix would give a combination of the good processing ability of this glass

and enhanced biological performance of this system. In this work, introducing CoO

as intermediate oxide in the glass network enhanced the reactivity, degradation and

ion release kinetics of the 1393 glass matrix. In similar way, other therapeutic ions

Page 158: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Conclusion and Outlook 158

like Zn, also known to act as intermediate oxide in glass formation, could be used

for design of novel bioactive glasses with enhanced degradation and cellular

response.356

For improvement of the mechanical performance of the BG derived scaffolds

fabrication of polymer-BG composite10

is one common approach. Polymer coated

scaffolds show enhanced toughness leading to scaffolds with resistance to crack

propagation. Indeed, the brittleness of 3D BG derived scaffolds is a limiting factor

in their application due to the surgeon´s need to be able cut the scaffolds to the right

the right shape and size.54

Drawbacks of a polymer infiltration of BG derived

scaffolds might be the loss of the (at least temporary) bioactive properties and

inhibition of the degradation and ion release. Hence, future studies on polymer/BG

composite materials should also consider the effects of the polymer coating on the

ion release kinetics of bioactive glasses.

Bioactive glasses in medicine

Generally, glasses are a very versatile group of materials which can be easily shaped

and processed enabling a wide range of potential applications in the field of

biomedical engineering. Applications of particulate bioactive glass, glass fibres, 3D

scaffolds, glass coatings and processing of BG / polymer composites offer a wide

range of application opportunities for novel bioactive glass composition with

specific functionalities. BGs are known for their ability to promote scaffold

mineralisation when applied as inorganic filler in an organic matrix mimicking the

extracellular matrix (ECM) as for example shown for collagen derived scaffold.357

Bioactive glass nanoparticles have been shown to have enhanced biological

properties and to improve the mechanical properties of BG/polymer composites

thus making them promising biomaterials for biomedical applications.358

Hence,

applying a top down approach by mechanical comminution359

to metal ion

containing glasses it is possible to explore a new range of bioactive glasses with

advantageous nano-scaled features and advanced glass chemistry with therapeutic

ion release capability.

Beside traditional application fields of bioactive glasses in hard tissue engineering

(bone and teeth), novel areas of biomedical applications are emerging including

Page 159: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Conclusion and Outlook 159

cancer treatment, ophthalmology, wound healing and nerve guidance repair.

Creating novel BG compositions containing specific therapeutic ions adapted to a

certain application with specific functionalities will help to advance these fields, in

particular considering the interaction of BGs and soft tissue. In this work, foam

replica technique was used for scaffold fabrication. However, other techniques, like

rapid prototyping, e.g. additive manufacturing using UV-light lithography or

robocasting will allow smart combinations of novel BG compositions and creation

of defined patient customized biomaterials and devices.

Page 160: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

References 160

7 References

1. Logeart-Avramoglou, D.; Anagnostou, F.; Bizios, R.; Petite, H., Engineering

bone: challenges and obstacles. J. Cell. Mol. Med. 2005, 9, (1), 72-84.

2. Gao, C.; Deng, Y.; Feng, P.; Mao, Z.; Li, P.; Yang, B.; Deng, J.; Cao, Y.; Shuai,

C.; Peng, S., Current progress in bioactive ceramic scaffolds for bone repair and

regeneration. Int. J. Mol. Sci. 2014, 15, (3), 4714-4732.

3. Carrington, J. L., Aging bone and cartilage: Cross-cutting issues. Biochem.

Biophys. Res. Commun. 2005, 328, (3), 700-708.

4. Palangkaraya, A.; Yong, J., Population ageing and its implications on aggregate

health care demand: empirical evidence from 22 OECD countries. Int. J. Health Care

Finance Econ. 2009, 9, (4), 391-402.

5. Younger, E. M.; Chapman, M. W., Morbidity at bone graft donor sites. J.

Orthop. Trauma 1989, 3, (3), 192-195.

6. Hutmacher, D. W., Scaffolds in tissue engineering bone and cartilage.

Biomaterials 2000, 21, (24), 2529-2543.

7. Moroni, L.; de Wijn, J. R.; van Blitterswijk, C. A., Integrating novel

technologies to fabricate smart scaffolds. J. Biomater. Sci. Polym. Ed. 2008, 19, 543-

572.

8. Dorozhkin, S. V., Bioceramics of calcium orthophosphates. Biomaterials 2010,

31, (7), 1465-1485.

9. Hench, L. L., Bioceramics. J. Am. Ceram. Soc. 1998, 81, (7), 1705-1728.

10. Rezwan, K.; Chen, Q. Z.; Blaker, J. J.; Boccaccini, A. R., Biodegradable and

bioactive porous polymer/inorganic composite scaffolds for bone tissue engineering.

Biomaterials 2006, 27, (18), 3413-3431.

11. Gerhardt, L. C.; Boccaccini, A. R., Bioactive glass and glass-ceramic scaffolds

for bone tissue engineering. Materials 2010, 3, (7), 3867-3910.

12. Jones, J. R.; Tsigkou, O.; Coates, E. E.; Stevens, M. M.; Polak, J. M.; Hench, L.

L., Extracellular matrix formation and mineralization on a phosphate-free porous

bioactive glass scaffold using primary human osteoblast (HOB) cells. Biomaterials

2007, 28, (9), 1653-1663.

13. Hench, L. L., Bioceramics: From concept to clinic. J. Am. Ceram. Soc. 1991, 74,

(7), 1487-1510.

Page 161: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

References 161

14. Suchanek, W.; Yoshimura, M., Processing and properties of hydroxyapatite-

based biomaterials for use as hard tissue replacement implants. J. Mater. Res. 1998, 13,

(1), 94-117.

15. Kokubo, T., Apatite formation on surfaces of ceramics, metals and polymers in

body environment. Acta Mater. 1998, 46, (7), 2519-2527.

16. Hench, L. L.; Polak, J. M., Third-generation biomedical materials. Science 2002,

295, (5557), 1014-1017.

17. Xynos, I. D.; Edgar, A. J.; Buttery, L. D. K.; Hench, L. L.; Polak, J. M., Gene-

expression profiling of human osteoblasts following treatment with the ionic products of

Bioglass® 45S5 dissolution. J. Biomed. Mater. Res. 2001, 55, (2), 151-157.

18. Hoppe, A.; Güldal, N. S.; Boccaccini, A. R., A review of the biological response

to ionic dissolution products from bioactive glasses and glass-ceramics. Biomaterials

2011, 32, (11), 2757-2774.

19. Hoppe, A.; Mourino, V.; Boccaccini, A. R., Therapeutic inorganic ions in

bioactive glasses to enhance bone formation and beyond. Biomater. Sci. 2013, 1, (3),

254-256.

20. Mouriño, V.; Cattalini, J. P.; Boccaccini, A. R., Metallic ions as therapeutic

agents in tissue engineering scaffolds: an overview of their biological applications and

strategies for new developments. J. Royal Soc. Interface 2012, 9, (68), 401-419.

21. Saltman, P. D.; Strause, L. G., The role of trace minerals in osteoporosis. J. Am.

Coll. Nutr. 1993, 12, (4), 384-389.

22. Beattie, J. H.; Avenell, A., Trace element nutrition and bone metabolism. Nutr.

Res. Rev. 1992, 5, (01), 167-188.

23. Nielsen, F., New essential trace elements for the life sciences. Biol. Trace Elem.

Res. 1990, 26-27, (1), 599-611.

24. Ash, C.; Stone, R., A Question of Dose. Science 2003, 300, (5621), 925.

25. Bruce, D., Developmental Biology of the Skeletal System. In Bone Tissue

Engineering, CRC Press: 2004; pp 3-26.

26. Jerosch, J.; Bader, A.; Uhr, G., Knochen, curasan Taschenatlas spezial. Thieme

Verlag: Stuttgart, 2002.

27. Marks Jr, S. C.; Odgren, P. R., Chapter 1 - Structure and Development of the

Skeleton. In Principles of Bone Biology (Second Edition), John, P. B.; Lawrence, G. R.;

Gideon A. RodanA2 - John P. Bilezikian, L. G. R.; Gideon, A. R., Eds. Academic Press:

San Diego, 2002; pp 3-15.

28. Basic Bone Biology and Tissue Engineering. In Bone Tissue Engineering, CRC

Press: 2004.

Page 162: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

References 162

29. London, R. C. o. P. o., Osteoporosis: Clinical Guidelines for Prevention and

Treatment. Royal College of Physicians of London: 1999.

30. Marie, P. J.; Ammann, P.; Boivin, G.; Rey, C., Mechanisms of action and

therapeutic potential of strontium in bone. Calcif. Tissue Int. 2001, 69, (3), 121-129.

31. Yamaguchi, M., Role of zinc in bone formation and bone resorption. J. Trace

Elem. Exp. Med. 1998, 11, (2-3), 119-135.

32. Hollinger, J. O.; Einhorn, T. A.; Doll, B.; Sfeir, C., Bone tissue engineering.

CRC Press: 2004.

33. Vaananen, H. K.; Zhao, H.; Mulari, M.; Halleen, J. M., The cell biology of

osteoclast function. J. Cell Sci. 2000, 113, (Pt 3), 377-81.

34. Rho, J.-Y.; Kuhn-Spearing, L.; Zioupos, P., Mechanical properties and the

hierarchical structure of bone. Med. Eng. Phys. 1998, 20, (2), 92-102.

35. Geneser, F., Textbook of Histology. Muuksgcarrel: Kopenhagen, 1986.

36. Junqueira, L. C. U.; Carneiro, J., Histologie. Springer: Berlin, 2002.

37. Kaplan, F.; Hayes, W.; Keaveny, T.; Boskey, A.; Einhorn, T.; Iannotti, J., Form

and function of bone. Orthopaedic Basic Science 1994, 127-185.

38. Fu, Q.; Saiz, E.; Rahaman, M. N.; Tomsia, A. P., Bioactive glass scaffolds for

bone tissue engineering: state of the art and future perspectives. Mater. Sci. Eng., C

2011, 31, (7), 1245-1256.

39. Griffith, L. G.; Naughton, G., Tissue Engineering--Current Challenges and

Expanding Opportunities. Science 2002, 295, (5557), 1009-1014.

40. Dvir, T.; Timko, B. P.; Kohane, D. S.; Langer, R., Nanotechnological strategies

for engineering complex tissues. Nat. Nanotechnol. 2011, 6, (1), 13-22.

41. Berthiaume, F.; Maguire, T. J.; Yarmush, M. L., Tissue engineering and

regenerative medicine: History, progress, and challenges. Annual Review of Chemical

and Biomolecular Engineering 2011, 2, 403-430.

42. Jones, J. R.; Ehrenfried, L. M.; Hench, L. L., Optimising bioactive glass

scaffolds for bone tissue engineering. Biomaterials 2006, 27, (7), 964-973.

43. Karageorgiou, V.; Kaplan, D., Porosity of 3D biomaterial scaffolds and

osteogenesis. Biomaterials 2005, 26, (27), 5474-5491.

44. Jadlowiec, J.; Sfeir, C.; Campbell, P.; Koch, H., Signaling Molecules for Tissue

Engineering. In Bone Tissue Engineering, CRC Press: 2004; pp 125-147.

45. Brandi, M. L.; Collin-Osdoby, P., Vascular Biology and the Skeleton. J. Bone

Miner. Res. 2006, 21, (2), 183-192.

Page 163: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

References 163

46. Santos, M., I.; Reis, R., L., Vascularization in bone tissue engineering:

Physiology, current strategies, major hurdles and future challenges. Macromol. Biosci.

2010, 10, (1), 12-27.

47. Polykandriotis, E.; Arkudas, A.; Euler, S.; Beier, J. P.; Horch, R. E.; Kneser, U.,

Prevascularisation strategies in tissue engineering. Handchirurgie Mikrochirurgie

Plastische Chirurgie 2006, 38, (4), 217-223.

48. Ren, L.-L.; Ma, D.-Y.; Feng, X.; Mao, T.-Q.; Liu, Y.-P.; Ding, Y., A novel

strategy for prefabrication of large and axially vascularized tissue engineered bone by

using an arteriovenous loop. Med. Hypotheses 2008, 71, (5), 737-740.

49. Williams, D. F., Tissue-biomaterial interactions. J. Mater. Sci. 1987, 22, (10),

3421-3445.

50. Williams, D. F., On the mechanisms of biocompatibility. Biomaterials 2008, 29,

(20), 2941-2953.

51. Hench, L. L.; Wilson, J., Introduction. In An introduction to bioceramics, 1 ed.;

Hench, L. L.; Wilson, J., Eds. World Scientific Publishing: Singapore, 1993; Vol. 1, pp

1-24.

52. Hofmann, I.; Müller, L.; Greil, P.; Müller, F. A., Calcium phosphate nucleation

on cellulose fabrics. Surf. Coat. Technol. 2006, 201, (6), 2392-2398.

53. Hench, L. L., The story of Bioglass®. J. Mater. Sci.: Mater. Med. 2006, 17, (11),

967-978.

54. Jones, J. R., Review of bioactive glass: From Hench to hybrids. Acta Biomater.

2013, 9, (1), 4457-4486.

55. Jones, J. R., Bioactive ceramics and glasses. In Tissue engineering using

ceramics and polymers, 1 ed.; Boccaccini, A. R.; Gough, J. E., Eds. Woodhead

Publishing Limited CRC Press: Cambridge, England, 2007; Vol. 1, pp 52-71.

56. Pereira, M. M.; Jones, J. R.; Orefice, R. L.; Hench, L. L., Preparation of

bioactive glass-polyvinyl alcohol hybrid foams by the sol-gel method. J. Mater. Sci.

Mater. Med. 2005, 16, (11), 1045-50.

57. Zhang, K.; Washburn, N. R.; Simon Jr, C. G., Cytotoxicity of three-

dimensionally ordered macroporous sol–gel bioactive glass (3DOM-BG). Biomaterials

2005, 26, (22), 4532-4539.

58. Balamurugan, A.; Balossier, G.; Kannan, S.; Michel, J.; Rebelo, A. H. S.;

Ferreira, J. M. F., Development and in vitro characterization of sol-gel derived CaO-

P2O5-SiO2-ZnO bioglass. Acta Biomater. 2007, 3, (2), 255-262.

59. Du, R. L.; Chang, J., Preparation and characterization of Zn and Mg doped

bioactive glasses. Journal of Inorganic Materials 2004, 19, (6), 1353-1358.

Page 164: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

References 164

60. Hesaraki, S.; Alizadeh, M.; Nazarian, H.; Sharifi, D., Physico-chemical and in

vitro biological evaluation of strontium/calcium silicophosphate glass. J. Mater. Sci.:

Mater. Med. 2010, 21, (2), 695-705.

61. Brinker, C. J.; Scherer, G. W., Sol–Gel Science: The Physics and Chemistry of

Sol–Gel Processing. 1990.

62. Fredholm, Y. C.; Karpukhina, N.; Law, R. V.; Hill, R. G., Strontium containing

bioactive glasses: Glass structure and physical properties. J. Non-Cryst. Solids 2010,

356, (44-49), 2546-2551.

63. Brown, R. F.; Rahaman, M. N.; Dwilewicz, A. B.; Huang, W.; Day, D. E.; Li, Y.;

Bal, B. S., Effect of borate glass composition on its conversion to hydroxyapatite and on

the proliferation of MC3T3-E1 cells. Journal of Biomedical Materials Research. Part A.

2009, 88A, (2), 392-400.

64. Azevedo, M. M.; Jell, G.; O'Donnell, M. D.; Law, R. V.; Hill, R. G.; Stevens, M.

M., Synthesis and characterization of hypoxia-mimicking bioactive glasses for skeletal

regeneration. J. Mater. Chem. 2010, 20, (40), 8854-8864.

65. Lusvardi, G.; Malavasi, G.; Menabue, L.; Aina, V.; Morterra, C., Fluoride-

containing bioactive glasses: surface reactivity in simulated body fluids solutions. Acta

Biomater. 2009, 5, (9), 3548-62.

66. Brauer, D. S.; Karpukhina, N.; O’Donnell, M. D.; Law, R. V.; Hill, R. G.,

Fluoride-containing bioactive glasses: Effect of glass design and structure on

degradation, pH and apatite formation in simulated body fluid. Acta Biomater. 2010, 6,

(8), 3275-3282.

67. Boccaccini, A. R.; Chen, Q.; Lefebvre, L.; Gremillard, L.; Chevalier, J.,

Sintering, crystallisation and biodegradation behaviour of Bioglass-derived glass-

ceramics. Faraday Discuss. 2007, 136, 27-44.

68. Chen, Q. Z.; Thompson, I. D.; Boccaccini, A. R., 45S5 Bioglass®-derived glass-

ceramic scaffolds for bone tissue engineering. Biomaterials 2006, 27, (11), 2414-2425.

69. Fu, Q.; Rahaman, M. N.; Bal, B. S.; Brown, R. F.; Day, D. E., Mechanical and in

vitro performance of 13-93 bioactive glass scaffolds prepared by a polymer foam

replication technique. Acta Biomater. 2008, 4, (6), 1854-1864.

70. Hench, L. L.; Splinter, R. J.; Allen, W. C.; Greenlee, T. K., Bonding mechanism

at the interface of ceramic prosthetic materials. J. Biomed. Mater. Res. 1972, 2, 117-141.

71. Rahaman, M. N.; Day, D. E.; Bal, B. S.; Fu, Q.; Jung, S. B.; Bonewald, L. F.;

Tomsia, A. P., Bioactive glass in tissue engineering. Acta Biomater. 2011, 7, (6), 2355-

2373.

72. Eden, M., NMR studies of oxide-based glasses. Annual Reports Section "C"

(Physical Chemistry) 2012, 108, (1), 177-221.

Page 165: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

References 165

73. FitzGerald, V.; Pickup, D. M.; Greenspan, D.; Sarkar, G.; Fitzgerald, J. J.;

Wetherall, K. M.; Moss, R. M.; Jones, J. R.; Newport, R. J., A neutron and X-ray

diffraction study of bioglass with reverse Monte Carlo modelling. Adv. Funct. Mater.

2007, 17, (18), 3746-3753.

74. Linati, L.; Lusvardi, G.; Malavasi, G.; Menabue, L.; Menziani, M. C.;

Mustarelli, P.; Pedone, A.; Segre, U., Medium-range order in phospho-silicate bioactive

glasses: Insights from MAS-NMR spectra, chemical durability experiments and

molecular dynamics simulations. J. Non-Cryst. Solids 2008, 354, (2–9), 84-89.

75. Pedone, A., Properties Calculations of Silica-Based Glasses by Atomistic

Simulations Techniques: A Review. The Journal of Physical Chemistry C 2009, 113,

(49), 20773-20784.

76. Tilocca, A., Structural models of bioactive glasses from molecular dynamics

simulations. Proc. R. Soc. A 2009, 465, (2104), 1003-1027.

77. Pedone, A.; Malavasi, G.; Menziani, M. C., Computational Insight into the

Effect of CaO/MgO Substitution on the Structural Properties of Phospho-Silicate

Bioactive Glasses. The Journal of Physical Chemistry C 2009, 113, (35), 15723-15730.

78. Tilocca, A.; Cormack, A. N., Structural effects of phosphorus inclusion in

bioactive silicate glasses. J. Phys. Chem. B 2007, 111, (51), 14256-14264.

79. Oliveira, J. M.; Correia, R. N.; Fernandes, M. H., Effects of Si speciation on the

in vitro bioactivity of glasses. Biomaterials 2002, 23, (2), 371-379.

80. Pedone, A.; Charpentier, T.; Malavasi, G.; Menziani, M. C., New Insights into

the Atomic Structure of 45S5 Bioglass by Means of Solid-State NMR Spectroscopy and

Accurate First-Principles Simulations. Chem. Mater. 2010, 22, (19), 5644-5652.

81. Arstila, H.; Vedel, E.; Hupa, L.; Hupa, M., Factors affecting crystallization of

bioactive glasses. J. Eur. Ceram. Soc. 2007, 27, (2–3), 1543-1546.

82. Kolan, K. C. R.; Leu, M. C.; Hilmas, G. E.; Velez, M., Effect of material,

process parameters, and simulated body fluids on mechanical properties of 13-93

bioactive glass porous constructs made by selective laser sintering. J. Mech. Behav.

Biomed. Mater. 2012, 13, (0), 14-24.

83. Clupper, D. C.; Hench, L. L.; Mecholsky, J. J., Strength and toughness of tape

cast bioactive glass 45S5 following heat treatment. J. Eur. Ceram. Soc. 2004, 24, (10-

11), 2929-2934.

84. De Diego, M. A.; Coleman, N. J.; Hench, L. L., Tensile properties of bioactive

fibers for tissue engineering applications. J. Biomed. Mater. Res. 2000, 53, (3), 199-203.

85. Thompson, I. D.; Hench, L. L., Mechanical properties of bioactive glasses,

glass-ceramics and composites. Proceedings of the Institution of Mechanical Engineers,

Part H: Journal of Engineering in Medicine 1998, 212, (2), 127-136.

Page 166: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

References 166

86. Pirhonen, E.; Niiranen, H.; Niemelä, T.; Brink, M.; Törmälä, P., Manufacturing,

mechanical characterization, and in vitro performance of bioactive glass 13–93 fibers. J.

Biomed. Mater. Res., Part B 2006, 77B, (2), 227-233.

87. Scholze, H., Glass-water interactions. J. Non-Cryst. Solids 1988, 102, (1–3), 1-

10.

88. Wilson, J.; Pigott, G. H.; Schoen, F. J.; Hench, L. L., Toxicology and

biocompatibility of bioglasses. J. Biomed. Mater. Res. 1981, 15, (6), 805-817.

89. Tilocca, A., Molecular dynamics simulations of a bioactive glass nanoparticle. J.

Mater. Chem. 2011, 21, (34), 12660-12667.

90. Peitl Filho, O.; LaTorre, G. P.; Hench, L. L., Effect of crystallization on apatite-

layer formation of bioactive glass 45S5. J. Biomed. Mater. Res. 1996, 30, (4), 509-14.

91. Cerruti, M.; Greenspan, D.; Powers, K., Effect of pH and ionic strength on the

reactivity of Bioglass® 45S5. Biomaterials 2005, 26, (14), 1665-1674.

92. Thian, E. S.; Huang, J.; Best, S. M.; Barber, Z. H.; Bonfield, W., Novel silicon-

doped hydroxyapatite (Si-HA) for biomedical coatings: An <I>in vitro</I> study using

acellular simulated body fluid. Journal of biomedical materials research. Part B,

Applied biomaterials. 2006, 76B, (2), 326-333.

93. Jones, J. R.; Sepulveda, P.; Hench, L. L., Dose-dependent behavior of bioactive

glass dissolution. J. Biomed. Mater. Res. 2001, 58, (6), 720-726.

94. Mačković, M.; Hoppe, A.; Detsch, R.; Mohn, D.; Stark, W. J.; Spiecker, E.;

Boccaccini, A. R., Bioactive glass (type 45S5) nanoparticles: in vitro reactivity on

nanoscale and biocompatibility. J. Nanopart. Res. 2012, 14, (7), 1-22.

95. Ducheyne, P.; Qiu, Q., Bioactive ceramics: the effect of surface reactivity on

bone formation and bone cell function. Biomaterials 1999, 20, (23–24), 2287-2303.

96. Hill, R., An alternative view of the degradation of bioglass. J. Mater. Sci. Lett.

1996, 15, (13), 1122-1125.

97. Kim, H.-M.; Miyaji, F.; Kokubo, T.; Ohtsuki, C.; Nakamura, T., Bioactivity of

Na2O-CaO-SiO2 Glasses. J. Am. Ceram. Soc. 1995, 78, (9), 2405-2411.

98. Kokubo, T.; Kushitani, H.; Sakka, S.; Kitsugi, T.; Yamamuro, T., Solutions able

to reproduce in vivo surface-structure changes in bioactive glass-ceramic A-W3. J.

Biomed. Mater. Res. 1990, 24, (6), 721-734.

99. Kokubo, T.; Takadama, H., How useful is SBF in predicting in vivo bone

bioactivity? Biomaterials 2006, 27, (15), 2907-2915.

100. Müller, L.; Müller, F. A., Preparation of SBF with different content and its

influence on the composition of biomimetic apatites. Acta Biomater. 2006, 2, (2), 181-

189.

Page 167: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

References 167

101. Bohner, M.; Lemaitre, J., Can bioactivity be tested in vitro with SBF solution?

Biomaterials 2009, 30, (12), 2175-2179.

102. Pan, H.; Zhao, X.; Darvell, B. W.; Lu, W. W., Apatite-formation ability –

Predictor of “bioactivity”? Acta Biomater. 2010, 6, (11), 4181-4188.

103. Arcos, D.; Greenspan, D. C.; Vallet-Regí, M., A new quantitative method to

evaluate the in vitro bioactivity of melt and sol-gel-derived silicate glasses. J. Biomed.

Mater. Res., Part A 2003, 65, (3), 344-351.

104. Hench, L. L.; Xynos, I. D.; Polak, J. M., Bioactive glasses for in situ tissue

regeneration. J. Biomater. Sci. Polym. Ed. 2004, 15, 543-562.

105. Kaufmann, E.; Ducheyne, P.; Shapiro, I. M., Evaluation of osteoblast response to

porous bioactive glass (45S5) substrates by RT-PCR analysis. Tissue Eng. 2000, 6, (1),

19-28.

106. Jell, G.; Notingher, I.; Tsigkou, O.; Notingher, P.; Polak, J. M.; Hench, L. L.;

Stevens, M. M., Bioactive glass-induced osteoblast differentiation: A noninvasive

spectroscopic study. Journal of Biomedical Materials Research. Part A. 2008, 86A, (1),

31-40.

107. Fu, Q.; Rahaman, M. N.; Bal, B. S.; Brown, R. F., Proliferation and function of

MC3T3-E1 cells on freeze-cast hydroxyapatite scaffolds with oriented pore

architectures. Journal of Materials Science-Materials in Medicine 2009, 20, (5), 1159-

1165.

108. Fu, Q.; Rahaman, M. N.; Bal, B. S.; Brown, R. F., Preparation and in vitro

evaluation of bioactive glass (13-93) scaffolds with oriented microstructures for repair

and regeneration of load-bearing bones. Journal of Biomedical Materials Research Part

A 2010, 93A, (4), 1380-1390.

109. Fu, Q.; Ramahan, M. N.; Bal, B. S.; Kuroki, K.; Brown, R. F., In vivo evaluation

of 13-93 bioactive glass scaffolds with trabecular and

oriented microstructures in a subcutaneous rat implantation model. Biomed. Mat. Res. A

2010, 95A, (1), 235-244.

110. Brown, R. F.; Day, D. E.; Day, T. E.; Jung, S.; Rahaman, M. N.; Fu, Q., Growth

and differentiation of osteoblastic cells on 13-93 bioactive glass fibers and scaffolds.

Acta Biomater. 2008, 4, (2), 387-396.

111. Välimäki, V.-V.; Yrjans, J. J.; Vuorio, E. I.; Aro, H. T., Molecular biological

evaluation of bioactive glass microspheres and adjunct bone morphogenetic protein 2

gene transfer in the enhancement of new bone formation. Tissue Eng. 2005, 11, (3-4),

387-394.

112. Bosetti, M.; Cannas, M., The effect of bioactive glasses on bone marrow stromal

cells differentiation. Biomaterials 2005, 26, (18), 3873-3879.

Page 168: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

References 168

113. Bielby, R. C.; Christodoulou, I. S.; Pryce, R. S.; Radford, W. J. P.; Hench, L. L.;

Polak, J. M., Time- and concentration-dependent effects of dissolution products of 58S

sol–gel bioactive glass on proliferation and differentiation of murine and human

osteoblasts. Tissue Eng. 2004, 10, (7-8), 1018-1026.

114. Christodoulou, I.; Buttery, L. D. K.; Tai, G.; Hench, L. L.; Polak, J. M.,

Characterization of human fetal osteoblasts by microarray analysis following

stimulation with 58S bioactive gel-glass ionic dissolution products. Journal of

biomedical materials research. Part B, Applied biomaterials. 2006, 77B, (2), 431-446.

115. Jell, G.; Stevens, M., Gene activation by bioactive glasses. J. Mater. Sci.: Mater.

Med. 2006, 17, (11), 997-1002.

116. Day, R. M., Bioactive glass stimulates the secretion of angiogenic growth

factors and angiogenesis in vitro. Tissue Eng. 2005, 11, (5-6), 768-777.

117. Gorustovich, A.; Roether, J.; Boccaccini, A. R., Effect of bioactive glasses on

angiogenesis: In-vitro and in-vivo evidence. A review. Tissue Eng Part B 2010, 16, (2),

199-207.

118. Leu, A.; Leach, J., Proangiogenic potential of a collagen/bioactive glass

substrate. Pharm. Res. 2008, 25, (5), 1222-1229.

119. Leu, A.; Stieger, S. M.; Dayton, P.; Ferrara, K. W.; Leach, J. K., Angiogenic

response to bioactive glass promotes bone healing in an irradiated calvarial defect.

Tissue Eng. Part A 2009, 15, (4), 877-885.

120. Deb, S.; Mandegaran, R.; Di Silvio, L., A porous scaffold for bone tissue

engineering/45S5 Bioglass® derived porous scaffolds for co-culturing osteoblasts and

endothelial cells. J. Mater. Sci.: Mater. Med. 2010, 21, (3), 893-905.

121. Day, R. M.; Boccaccini, A. R.; Shurey, S.; Roether, J. A.; Forbes, A.; Hench, L.

L.; Gabe, S. M., Assessment of polyglycolic acid mesh and bioactive glass for soft-

tissue engineering scaffolds. Biomaterials 2004, 25, (27), 5857-5866.

122. Andrade, A. L.; Andrade, S. P.; Domingues, R. Z., In vivo performance of a sol-

gel glass-coated collagen. Journal of Biomedical Materials Research Part B: Applied

Biomaterials 2006, 79B, (1), 122-128.

123. Ross, E. A.; Batich, C. D.; Clapp, W. L.; Sallustio, J. E.; Lee, N. C., Tissue

adhesion to bioactive glass-coated silicone tubing in a rat model of peritoneal dialysis

catheters and catheter tunnels. Kidney Int. 2003, 63, (2), 702-708.

124. Mahmood, J.; Takita, H.; Ojima, Y.; Kobayashi, M.; Kohgo, T.; Kuboki, Y.,

Geometric effect of matrix upon cell differentiation: BMP-induced osteogenesis using a

new bioglass with a feasible structure. J. Biochem. (Tokyo) 2001, 129, (1), 163-171.

125. Nandi, S. K.; Kundu, B.; Datta, S.; De, D. K.; Basu, D., The repair of segmental

bone defects with porous bioglass: An experimental study in goat. Res. Vet. Sci. 2009,

86, (1), 162-173.

Page 169: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

References 169

126. Keshaw, H.; Georgiou, G.; Blaker, J. J.; Forbes, A.; Knowles, J. C.; Day, R. M.,

Assessment of polymer/bioactive glass-composite microporous spheres for tissue

regeneration applications. Tissue Eng. Part A 2009, 15, (7), 1451-1461.

127. Choi, H. Y.; Lee, J. E.; Park, H. J.; Oum, B. S., Effect of synthetic bone glass

particulate on the fibrovascularization of porous polyethylene orbital implants. Ophthal.

Plast. Reconstr. Surg. 2006, 22, (2), 121-125.

128. Day, R. M.; Maquet, V.; Boccaccini, A. R.; Jerome, R.; Forbes, A., In vitro and

in vivo analysis of macroporous biodegradable poly(D,L-lactide-co-glycolide) scaffolds

containing bioactive glass. Journal of Biomedical Materials Research Part A 2005, 75,

(4), 778-787.

129. Hench, L. L., Genetic design of bioactive glass. J. Eur. Ceram. Soc. 2009, 29,

(7), 1257-1265.

130. Xynos, I. D.; Edgar, A. J.; Buttery, L. D. K.; Hench, L. L.; Polak, J. M., Ionic

products of bioactive glass dissolution increase proliferation of human osteoblasts and

induce insulin-like growth factor II mRNA expression and protein synthesis. Biochem.

Biophys. Res. Commun. 2000, 276, (2), 461-465.

131. Xynos, I. D.; Hukkanen, M. V. J.; Batten, J. J.; Buttery, L. D.; Hench, L. L.;

Polak, J. M., Bioglass ®45S5 stimulates osteoblast turnover and enhances bone

formation in vitro: Implications and applications for bone tissue engineering. Calcif.

Tissue Int. 2000, 67, (4), 321-329.

132. Valerio, P.; Pereira, M. M.; Goes, A. M.; Leite, M. F., The effect of ionic

products from bioactive glass dissolution on osteoblast proliferation and collagen

production. Biomaterials 2004, 25, (15), 2941-2948.

133. Carlisle, E., Silicon: A requirement in bone formation independent of vitamin

D1. Calcif. Tissue Int. 1981, 33, (1), 27-34.

134. Carlisle, E. M., Silicon: A Possible Factor in Bone Calcification. Science 1970,

167, (3916), 279-280.

135. Alcaide, M.; Portolés, P.; López-Noriega, A.; Arcos, D.; Vallet-Regí, M.;

Portolés, M. T., Interaction of an ordered mesoporous bioactive glass with osteoblasts,

fibroblasts and lymphocytes, demonstrating its biocompatibility as a potential bone graft

material. Acta Biomater. 2010, 6, (3), 892-899.

136. Valerio, P.; Pereira, M. M.; Goes, A. M.; Leite, M. F., Effects of extracellular

calcium concentration on the glutamate release by bioactive glass (BG60S)

preincubated osteoblasts. Biomed. Mater. 2009, (4), 045011.

137. Hinoi, E.; Takarada, T.; Yoneda, Y., Glutamate signaling system in bone. J.

Pharmacol. Sci. 2004, 94, (3), 215-220.

138. Sola, A.; Bellucci, D.; Cannillo, V.; Cattini, A., Bioactive glass coatings: A

review. Surf. Eng. 2011, 27, (8), 560-572.

Page 170: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

References 170

139. Margonar, R.; Queiroz, T. P.; Luvizuto, E. R.; Marcantonio, É.; Lia, R. C. C.;

Holzhausen, M.; Marcantonio-Júnior, É., Bioactive glass for alveolar ridge

augmentation. J. Craniofac. Surg. 2012, 23, (3), e220-e222.

140. Gosain, A. K., Bioactive glass for bone replacement in craniomaxillofacial

reconstruction. Plast. Reconstr. Surg. 2004, 114, (2), 590-593.

141. Deisinger, U., Generating porous ceramic scaffolds: Processing and properties.

Key Eng. Mater. 2010, 441, 155-179.

142. Comesaña, R.; Lusquiños, F.; del Val, J.; López-Álvarez, M.; Quintero, F.;

Riveiro, A.; Boutinguiza, M.; de Carlos, A.; Jones, J. R.; Hill, R. G.; Pou, J., Three-

dimensional bioactive glass implants fabricated by rapid prototyping based on CO2

laser cladding. Acta Biomater. 2011, 7, (9), 3476-3487.

143. Yang, S.; Leong, K. F.; Du, Z.; Chua, C. K., The design of scaffolds for use in

tissue engineering. Part II. Rapid prototyping techniques. Tissue Eng. 2002, 8, (1), 1-11.

144. Mallik, K. K., Freeze Casting of Porous Bioactive Glass and Bioceramics. J.

Am. Ceram. Soc. 2008, 92, (S1), S85-S94.

145. Doiphode, N. D.; Huang, T. S.; Leu, M. C.; Rahaman, M. N.; Day, D. E., Freeze

extrusion fabrication of 13-93 bioactive glass scaffolds for bone repair. Journal of

Materials Science-Materials in Medicine 2011, 22, (3), 515-523.

146. Huang, T. S.; Rahaman, M. N.; Doiphode, N. D.; Leu, M. C.; Bal, B. S.; Day, D.

E.; Liu, X., Porous and strong bioactive glass (13–93) scaffolds fabricated by freeze

extrusion technique. Mater. Sci. Eng., C 2011, 31, (7), 1482-1489.

147. Filho, O. P.; La Torre, G. P.; Hench, L. L., Effect of crystallization on apatite-

layer formation of bioactive glass 45S5. J. Biomed. Mater. Res. 1996, 30, (4), 509-514.

148. Vitale-Brovarone, C.; Miola, M.; Balagna, C.; Verné, E., 3D-glass-ceramic

scaffolds with antibacterial properties for bone grafting. Chem. Eng. J. 2008, 137, (1),

129-136.

149. Wu, Z. Y.; Hill, R. G.; Yue, S.; Nightingale, D.; Lee, P. D.; Jones, J. R., Melt-

derived bioactive glass scaffolds produced by a gel-cast foaming technique. Acta

Biomater. 2011, 7, (4), 1807-1816.

150. Song, J.-H.; Koh, Y.-H.; Kim, H.-E.; Li, L.-H.; Bahn, H.-J., Fabrication of a

porous bioactive glass–ceramic using room-temperature freeze casting. J. Am. Ceram.

Soc. 2006, 89, (8), 2649-2653.

151. Fu, Q.; Saiz, E.; Tomsia, A. P., Direct ink writing of highly porous and strong

glass scaffolds for load-bearing bone defects repair and regeneration. Acta Biomater.

2011, 7, (10), 3547-54.

Page 171: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

References 171

152. Tesavibul, P.; Felzmann, R.; Gruber, S.; Liska, R.; Thompson, I.; Boccaccini, A.

R.; Stampfl, J., Processing of 45S5 Bioglass® by lithography-based additive

manufacturing. Mater. Lett. 2012, 74, (0), 81-84.

153. Vitale-Brovarone, C.; Verne, E.; Bosetti, M.; Appendino, P.; Cannas, M.,

Microstructural and in vitro characterization of SiO2-Na2O-CaO-MgO glass-ceramic

bioactive scaffolds for bone substitutes. J. Mater. Sci.: Mater. Med. 2005, 16, (10), 909-

917.

154. Moimas, L.; Biasotto, M.; Di Lenarda, R.; Olivo, A.; Schmid, C., Rabbit pilot

study on the resorbability of three-dimensional bioactive glass fibre scaffolds. Acta

Biomater. 2006, 2, (2), 191-199.

155. Liu, X.; Ramahan, M. N.; Fu, Q., Oriented bioactive glass (13-93) scaffolds with

controllable pore size by unidirectional freezing of camphene-based suspensions:

Microstructure and mechanical response. Acta Biomat 2011, 7, (1), 406-416.

156. Jones, J. R.; Lin, S.; Yue, S.; Lee, P. D.; Hanna, J. V.; Smith, M. E.; Newport, R.

J., Bioactive glass scaffolds for bone regeneration and their hierarchical

characterisation. Proceedings of the Institution of Mechanical Engineers, Part H:

Journal of Engineering in Medicine 2010, 224, (12), 1373-1387.

157. Jones, J.; Ehrenfried, L.; Saravanapavan, P.; Hench, L., Controlling ion release

from bioactive glass foam scaffolds with antibacterial properties. J. Mater. Sci.: Mater.

Med. 2006, 17, (11), 989-996.

158. Qian, J.; Kang, Y.; Wei, Z.; Zhang, W., Fabrication and characterization of

biomorphic 45S5 bioglass scaffold from sugarcane. Mater. Sci. Eng., C 2009, 29, (4),

1361-1364.

159. Fu, Q.; Rahaman, M. N.; Bal, B. S.; Huang, W.; Day, D. E., Preparation and

bioactive characteristics of a porous 13-93 glass, and fabrication into the articulating

surface of a proximal tibia. Journal of Biomedical Materials Research Part A 2007, 82,

(1), 222-229.

160. Habibovic, P.; Barralet, J. E., Bioinorganics and biomaterials: Bone repair. Acta

Biomater. 2011, 7, (8), 3013-3026.

161. Thompson, K. H., Orvig C., Boon and Bane of Metal Ions in Medicine. Science

2003, 300, (5621), 936.

162. Maeno, S.; Niki, Y.; Matsumoto, H.; Morioka, H.; Yatabe, T.; Funayama, A.;

Toyama, Y.; Taguchi, T.; Tanaka, J., The effect of calcium ion concentration on

osteoblast viability, proliferation and differentiation in monolayer and 3D culture.

Biomaterials 2005, 26, (23), 4847-4855.

163. Marie, P. J., The calcium-sensing receptor in bone cells: a potential therapeutic

target in osteoporosis. Bone 2010, 46, (3), 571-576.

Page 172: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

References 172

164. Julien, M.; Khoshniat, S.; Lacreusette, A.; Gatius, M.; Bozec, A.; Wagner, E. F.;

Wittrant, Y.; Masson, M.; Weiss, P.; Beck, L.; Magne, D.; Guicheux, J., Phosphate-

Dependent Regulation of MGP in Osteoblasts: Role of ERK1/2 and Fra-1. J. Bone

Miner. Res. 2009, 24, (11), 1856-1868.

165. Reffitt, D. M.; Ogston, N.; Jugdaohsingh, R.; Cheung, H. F. J.; Evans, B. A. J.;

Thompson, R. P. H.; Powell, J. J.; Hampson, G. N., Orthosilicic acid stimulates collagen

type 1 synthesis and osteoblastic differentiation in human osteoblast-like cells in vitro.

Bone 2003, 32, (2), 127-135.

166. Delannoy, P.; Bazot, D.; Marie, P. J., Long-term treatment with strontium

ranelate increases vertebral bone mass without deleterious effect in mice. Metabolism.

2002, 51, (7), 906-911.

167. Grynpas, M. D.; Marie, P. J., Effects of low doses of strontium on bone quality

and quantity in rats. Bone 1990, 11, (5), 313-319.

168. Shahnazari, M.; Sharkey, N. A.; Fosmire, G. J.; Leach, R. M., Effects of

Strontium on Bone Strength, Density, Volume, and Microarchitecture in Laying Hens. J.

Bone Miner. Res. 2006, 21, (11), 1696-1703.

169. Verberckmoes, S. C.; De Broe, M. E.; D'Haese, P. C., Dose-dependent effects of

strontium on osteoblast function and mineralization. Kidney Int. 2003, 64, (2), 534-543.

170. Marie, P. J., Strontium ranelate: A physiological approach for optimizing bone

formation and resorption. Bone 2006, 38, (2, Supplement 1), 10-14.

171. Brandão-Neto, J.; Stefan, V.; Mendonça, B. B.; Bloise, W.; Castro, A. V. B., The

essential role of zinc in growth. Nutr. Res. 1995, 15, (3), 335-358.

172. Choi, M. K.; Kim, M. H.; Kang, M. H., The effect of boron supplementation on

bone strength in ovariectomized rats fed with diets containing different calcium levels.

Food Sci. Biotechnol. 2005, 14, (2), 242-248.

173. Uysal, T.; Ustdal, A.; Sonmez, M. F.; Ozturk, F., Stimulation of bone formation

by dietary boron in an orthopedically expanded suture in rabbits. Angle Orthod. 2009,

79, (5), 984-990.

174. Barrio, D. A.; Etcheverry, S. B., Vanadium and bone development: putative

signaling pathways. Can. J. Physiol. Pharmacol. 2006, 84, (7), 677-686.

175. Cortizo, A. M.; Molinuevo, M. S.; Barrio, D. A.; Bruzzone, L., Osteogenic

activity of vanadyl(IV)-ascorbate complex: Evaluation of its mechanism of action. The

International Journal of Biochemistry & Cell Biology 2006, 38, (7), 1171-1180.

176. Buttyan, R.; Chichester, P.; Stisser, B.; Matsumoto, S.; Ghafar, M. A.; Levin, R.

M., Acute intravesical infusion of a cobalt solution stimulates a hypoxia response,

growth and angiogenesis in the rat bladder. J. Urol. 2003, 169, (6), 2402-2406.

Page 173: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

References 173

177. Patntirapong, S.; Habibovic, P.; Hauschka, P. V., Effects of soluble cobalt and

cobalt incorporated into calcium phosphate layers on osteoclast differentiation and

activation. Biomaterials 2009, 30, (4), 548-555.

178. Peters; Kirsten; Schmidt; Harald; Unger; E., R.; Otto; Mike; Kamp;

KIRKPATRICK; James, C., Software-supported image quantification of angiogenesis in

an in vitro culture system: application to studies of biocompatibility. Elsevier: Oxford,

ROYAUME-UNI, 2002; Vol. 23, p 7.

179. Tanaka, T.; Kojima, I.; Ohse, T.; Ingelfinger, J. R.; Adler, S.; Fujita, T.; Nangaku,

M., Cobalt promotes angiogenesis via hypoxia-inducible factor and protects

tubulointerstitium in the remnant kidney model. Lab. Invest. 2005, 85, (10), 1292-1307.

180. Hartwig, A., Role of magnesium in genomic stability. Mutat. Res., Fundam. Mol.

Mech. Mutagen. 2001, 475, (1-2), 113-121.

181. Rude, R. K.; Gruber, H. E.; Norton, H. J.; Wei, L. Y.; Frausto, A.; Kilburn, J.,

Dietary magnesium reduction to 25% of nutrient requirement disrupts bone and mineral

metabolism in the rat. Bone 2005, 37, (2), 211-219.

182. Rude, R. K.; Gruber, H. E.; Norton, H. J.; Wei, L. Y.; Frausto, A.; Mills, B. G.,

Bone loss induced by dietary magnesium reduction to 10% of the nutrient requirement

in rats is associated with increased release of substance P and tumor necrosis factor-

alpha. J. Nutr. 2004, 134, (1), 79-85.

183. Rude, R. K.; Gruber, H. E.; Wei, L. Y.; Frausto, A.; Mills, B. G., Magnesium

deficiency: Effect on bone and mineral metabolism in the mouse. Calcif. Tissue Int.

2003, 72, (1), 32-41.

184. Staiger, M. P.; Pietak, A. M.; Huadmai, J.; Dias, G., Magnesium and its alloys as

orthopedic biomaterials: A review. Biomaterials 2006, 27, (9), 1728-1734.

185. Tsuboi, S.; Nakagaki, H.; Ishiguro, K.; Kondo, K.; Mukai, M.; Robinson, C.;

Weatherell, J. A., Magnesium distribution in human bone. Calcif. Tissue Int. 1994, 54,

(1), 34-37.

186. Zreiqat, H.; Howlett, C. R.; Zannettino, A.; Evans, P.; Schulze-Tanzil, G.;

Knabe, C.; Shakibaei, M., Mechanisms of magnesium-stimulated adhesion of

osteoblastic cells to commonly used orthopaedic implants. J. Biomed. Mater. Res. 2002,

62, (2), 175-184.

187. Sun, Z. L.; Wataha, J. C.; Hanks, C. T., Effects of metal ions on osteoblast-like

cell metabolism and differentiation. J. Biomed. Mater. Res. 1997, 34, (1), 29-37.

188. Mourino, V.; Boccaccini, A. R., Bone tissue engineering therapeutics: controlled

drug delivery in three-dimensional scaffolds. J. Royal Soc. Interface 2010, 7, (43), 209-

227.

Page 174: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

References 174

189. Damen, J. J. M.; Ten Cate, J. M., Silica-induced precipitation of calcium

phosphate in the presence of inhibitors of hydroxyapatite formation. J. Dent. Res. 1992,

71, (3), 453-457.

190. Jugdaohsingh, R.; Tucker, K. L.; Qiao, N.; Cupples, L. A.; Kiel, D. P.; Powell, J.

J., Dietary Silicon Intake Is Positively Associated With Bone Mineral Density in Men

and Premenopausal Women of the Framingham Offspring Cohort. J. Bone Miner. Res.

2004, 19, (2), 297-307.

191. Kim, M.-H.; Bae, Y.-J.; Choi, M.-K.; Chung, Y.-S., Silicon supplementation

improves the bone mineral density of calcium-deficient ovariectomized rats by reducing

bone resorption. Biol. Trace Elem. Res. 2009, 128, (3), 239-247.

192. Nielsen, F. H.; Poellot, R., Dietary silicon affects bone turnover differently in

ovariectomized and sham-operated growing rats. J. Trace Elem. Exp. Med. 2004, 17,

(3), 137-149.

193. Kwun, I.-S.; Cho, Y.-E.; Lomeda, R.-A. R.; Shin, H.-I.; Choi, J.-Y.; Kang, Y.-H.;

Beattie, J. H., Zinc deficiency suppresses matrix mineralization and retards osteogenesis

transiently with catch-up possibly through Runx 2 modulation. Bone 2010, 46, (3), 732-

741.

194. Yamasaki, Y.; Yoshida, Y.; Okazaki, M.; Shimazu, A.; Uchida, T.; Kubo, T.;

Akagawa, Y.; Hamada, Y.; Takahashi, J.; Matsuura, N., Synthesis of functionally graded

MgCO3 apatite accelerating osteoblast adhesion. J. Biomed. Mater. Res. 2002, 62, (1),

99-105.

195. Meunier, P. J.; Slosman, D. O.; Delmas, P. D.; Sebert, J. L.; Brandi, M. L.;

Albanese, C.; Lorenc, R.; Pors-Nielsen, S.; De Vernejoul, M. C.; Roces, A.; Reginster, J.

Y., Strontium ranelate: Dose-dependent effects in established postmenopausal vertebral

osteoporosis--A 2-year randomized placebo controlled trial. J. Clin. Endocrinol. Metab.

2002, 87, (5), 2060-2066.

196. Finney, L.; Vogt, S.; Fukai, T.; Glesne, D., Copper and angiogenesis:

Unravelling a relationship key to cancer progression. Clin. Exp. Pharmacol. Physiol.

2009, 36, (1), 88-94.

197. Gerard, C.; Bordeleau, L. J.; Barralet, J.; Doillon, C. J., The stimulation of

angiogenesis and collagen deposition by copper. Biomaterials 2010, 31, (5), 824-831.

198. Hu, G.-f., Copper stimulates proliferation of human endothelial cells under

culture. J. Cell. Biochem. 1998, 69, (3), 326-335.

199. Rodríguez, J. P.; Ríos, S.; González, M., Modulation of the proliferation and

differentiation of human mesenchymal stem cells by copper. J. Cell. Biochem. 2002, 85,

(1), 92-100.

200. Fan, W.; Crawford, R.; Xiao, Y., Enhancing in vivo vascularized bone formation

by cobalt chloride-treated bone marrow stromal cells in a tissue engineered periosteum

model. Biomaterials 2010, 31, (13), 3580-3589.

Page 175: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

References 175

201. Dzondo-Gadet, M.; Mayap-Nzietchueng, R.; Hess, K.; Nabet, P.; Belleville, F.;

Dousset, B., Action of boron at the molecular level. Biol. Trace Elem. Res. 2002, 85,

(1), 23-33.

202. Nielsen, F. H., Is boron nutritionally relevant? Nutr. Rev. 2008, 66, (4), 183.

203. Boivin, G.; Deloffre, P.; Perrat, B.; Panczer, G.; Boudeulle, M.; Mauras, Y.;

Allain, P.; Tsouderos, Y.; Meunier, P. J., Strontium distribution and interactions with

bone mineral in monkey iliac bone after strontium salt (S 12911) administration. J. Bone

Miner. Res. 1996, 11, (9), 1302-1311.

204. Habermann, B.; Kafchitsas, K.; Olender, G.; Augat, P.; Kurth, A., Strontium

ranelate enhances callus strength more than PTH 1-34 in an osteoporotic rat model of

fracture healing. Calcif. Tissue Int. 2010, 86, (1), 82-89.

205. Lang, C.; Murgia, C.; Leong, M.; Tan, L.-W.; Perozzi, G.; Knight, D.; Ruffin,

R.; Zalewski, P., Anti-inflammatory effects of zinc and alterations in zinc transporter

mRNA in mouse models of allergic inflammation. Am. J. Physiol. Lung Cell. Mol.

Physiol. 2007, 292, (2), L577-584.

206. Cousins, R. J., A role of zinc in the regulation of gene expression. Proc. Nutr.

Soc. 1998, 57, (02), 307-311.

207. Zhang, J. C.; Huang, J. A.; Xu, S. J.; Wang, K.; Yu, S. F., Effects of Cu2+ and

pH on osteoclastic bone resorption in vitro. Prog. Nat. Sci. 2003, 13, (4), 266-270.

208. Cashman, K. D.; Baker, A.; Ginty, F.; Flynn, A.; Strain, J. J.; Bonham, M. P.;

O'Connor, J. M.; Bugel, S.; Sandstrom, B., No effect of copper supplementation on

biochemical markers of bone metabolism in healthy young adult females despite

apparently improved copper status. Eur. J. Clin. Nutr. 2001, 55, (7), 525-531.

209. Lai, Y. L.; Yamaguchi, M., Effects of copper on bone component in the femoral

tissues of rats: anabolic effect of zinc is weakened by copper. Biol. Pharm. Bull. 2005,

28, (12), 2296-2301.

210. Smith, B. J.; King, J. B.; Lucas, E. A.; Akhter, M. P.; Arjmandi, B. H.; Stoecker,

B. J., Skeletal unloading and dietary copper depletion are detrimental to bone quality of

mature rats. J. Nutr. 2002, 132, (2), 190-196.

211. Peters, K.; Schmidt, H.; Unger, R.; Kamp, G.; Pröls, F.; Berger, B.; Kirkpatrick,

C. J., Paradoxical effects of hypoxia-mimicking divalent cobalt ions in human

endothelial cells in vitro. Mol. Cell. Biochem. 2005, 270, (1-2), 157-166.

212. Chachami, G.; Simos, G.; Hatziefthimiou, A.; Bonanou, S.; Molyvdas, P.-A.;

Paraskeva, E., Cobalt Induces Hypoxia-Inducible Factor-1α Expression in Airway

Smooth Muscle Cells by a Reactive Oxygen Species– and PI3K-Dependent Mechanism.

Am. J. Respir. Cell Mol. Biol. 2004, 31, (5), 544-551.

213. Semenza, G. L., Life with Oxygen. Science 2007, 318, (5847), 62-64.

Page 176: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

References 176

214. Wan, C.; Gilbert, S. R.; Wang, Y.; Cao, X.; Shen, X.; Ramaswamy, G.; Jacobsen,

K. A.; Alaql, Z. S.; Eberhardt, A. W.; Gerstenfeld, L. C.; Einhorn, T. A.; Deng, L.;

Clemens, T. L., Activation of the hypoxia-inducible factor-1α pathway accelerates bone

regeneration. Proc. Natl. Acad. Sci. U. S. A. 2008, 105, (2), 686-691.

215. Emans, P. J.; Spaapen, F.; Surtel, D. A. M.; Reilly, K. M.; Cremers, A.; van

Rhijn, L. W.; Bulstra, S. K.; Voncken, J. W.; Kuijer, R., A novel in vivo model to study

endochondral bone formation; HIF-1α activation and BMP expression. Bone 2007, 40,

(2), 409-418.

216. Dery, M. A. C.; Michaud, M. D.; Richard, D. E., Hypoxia-inducible factor 1:

regulation by hypoxic and non-hypoxic activators. Int. J. Biochem. Cell Biol. 2005, 37,

(3), 535-540.

217. Azevedo, M.; Jell, G.; Hill, R.; Stevens, M. M., Hypoxia-mimicking materials

for bone and cartilage tissue engineering. Eur. Cell. Mater. 2009, 18, (SUPPL. 2), 45.

218. Lakhkar, N. J.; Lee, I.-H.; Kim, H.-W.; Salih, V.; Wall, I. B.; Knowles, J. C.,

Bone formation controlled by biologically relevant inorganic ions: Role and controlled

delivery from phosphate-based glasses. Adv. Drug Delivery Rev., (0).

219. Wu, C.; Zhou, Y.; Fan, W.; Han, P.; Chang, J.; Yuen, J.; Zhang, M.; Xiao, Y.,

Hypoxia-mimicking mesoporous bioactive glass scaffolds with controllable cobalt ion

release for bone tissue engineering. Biomaterials 2012, 33, (7), 2076-2085.

220. Aina, V.; Malavasi, G.; Pla, A. F.; Munaron, L.; Morterra, C., Zinc-containing

bioactive glasses: surface reactivity and behaviour towards endothelial cells. Acta

Biomater. 2009, 5, (4), 1211-1222.

221. Boyd, D.; Carroll, G.; Towler, M. R.; Freeman, C.; Farthing, P.; Brook, I. M.,

Preliminary investigation of novel bone graft substitutes based on strontium-calcium-

zinc-silicate glasses. J. Mater. Sci.: Mater. Med. 2009, 20, (1), 413-420.

222. Haimi, S.; Gorianc, G.; Moimas, L.; Lindroos, B.; Huhtala, H.; Räty, S.;

Kuokkanen, H.; Sándor, G. K.; Schmid, C.; Miettinen, S.; Suuronen, R.,

Characterization of zinc-releasing three-dimensional bioactive glass scaffolds and their

effect on human adipose stem cell proliferation and osteogenic differentiation. Acta

Biomater. 2009, 5, (8), 3122-3131.

223. Murphy, S.; Boyd, D.; Moane, S.; Bennett, M., The effect of composition on ion

release from Ca–Sr–Na–Zn–Si glass bone grafts. J. Mater. Sci.: Mater. Med. 2009, 20,

(11), 2207-2214.

224. Varmette, E. A.; Nowalk, J. R.; Flick, L. M.; Hall, M. M., Abrogation of the

inflammatory response in LPS-stimulated RAW 264.7 murine macrophages by Zn- and

Cu-doped bioactive sol-gel glasses. Journal of Biomedical Materials Research. Part A.

2009, 90A, (2), 317-325.

Page 177: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

References 177

225. Li, X.; Wang, X.; He, D.; Shi, J., Synthesis and characterization of mesoporous

CaO-MO-SiO2-P2O5 (M = Mg, Zn, Cu) bioactive glasses/composites. J. Mater. Chem.

2008, 18, (34), 4103-4109.

226. Wu, C.; Zhou, Y.; Xu, M.; Han, P.; Chen, L.; Chang, J.; Xiao, Y., Copper-

containing mesoporous bioactive glass scaffolds with multifunctional properties of

angiogenesis capacity, osteostimulation and antibacterial activity. Biomaterials 2013,

34, (2), 422-433.

227. Gentleman, E.; Fredholm, Y. C.; Jell, G.; Lotfibakhshaiesh, N.; O'Donnell, M.

D.; Hill, R. G.; Stevens, M. M., The effects of strontium-substituted bioactive glasses on

osteoblasts and osteoclasts in vitro. Biomaterials 2010, 31, (14), 3949-3956.

228. Gorustovich, A. A.; Steimetz, T.; Cabrini, R. L.; López, J. M. P.,

Osteoconductivity of strontium-doped bioactive glass particles: A histomorphometric

study in rats. Journal of Biomedical Materials Research. Part A. 2010, 92A, (1), 232-

237.

229. Lao, J.; Nedelec, J. M.; Jallot, E., New strontium-based bioactive glasses:

physicochemical reactivity and delivering capability of biologically active dissolution

products. J. Mater. Chem. 2009, 19, (19), 2940-2949.

230. O'Donnell, M. D.; Candarlioglu, P. L.; Miller, C. A.; Gentleman, E.; Stevens, M.

M., Materials characterisation and cytotoxic assessment of strontium-substituted

bioactive glasses for bone regeneration. J. Mater. Chem. 2010, 20, (40), 8934-8941.

231. Chen, X.; Liao, X.; Huang, Z.; You, P.; Chen, C.; Kang, Y.; Yin, G., Synthesis

and characterization of novel multiphase bioactive glass-ceramics in the CaO-MgO-

SiO2 system. Journal of biomedical materials research. Part B, Applied biomaterials.

2010, 93B, (1), 194-202.

232. Saboori, A.; Rabiee, M.; Moztarzadeh, F.; Sheikhi, M.; Tahriri, M.; Karimi, M.,

Synthesis, characterization and in vitro bioactivity of sol-gel-derived SiO2-CaO-P2O5-

MgO bioglass. Mater. Sci. Eng., C 2009, 29, (1), 335-340.

233. Watts, S. J.; Hill, R. G.; O'Donnell, M. D.; Law, R. V., Influence of magnesia on

the structure and properties of bioactive glasses. J. Non-Cryst. Solids 2010, 356, (9-10),

517-524.

234. Erol, M.; Ozyuguran, A.; Celebican, O., Synthesis, characterization, and in vitro

bioactivity of sol-gel-derived Zn, Mg, and Zn-Mg co-doped bioactive glasses. Chem.

Eng. Technol. 2010, 33, (7), 1066-1074.

235. Dietrich, E.; Oudadesse, H.; Lucas-Girot, A.; Mami, M., In vitro bioactivity of

melt-derived glass 46S6 doped with magnesium. Journal of Biomedical Materials

Research. Part A. 2009, 88A, (4), 1087-1096.

236. Bellantone, M.; Williams, H. D.; Hench, L. L., Broad-spectrum bactericidal

activity of Ag2O-doped bioactive glass. Antimicrob. Agents Chemother. 2002, 46, (6),

1940-1945.

Page 178: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

References 178

237. Blaker, J. J.; Nazhat, S. N.; Boccaccini, A. R., Development and characterisation

of silver-doped bioactive glasscoated sutures for tissue engineering and wound healing

applications. Biomaterials 2004, 25, (7-8), 1319-1329.

238. Delben, J. R. J.; Pimentel, O. M.; Coelho, M. B.; Candelorio, P. D.; Furini, L.

N.; dos Santos, F. A.; de Vicente, F. S.; Delben, A., Synthesis and thermal properties of

nanoparticles of bioactive glasses containing silver. J. Therm. Anal. Calorim. 2009, 97,

(2), 433-436.

239. Kawashita, M.; Tsuneyama, S.; Miyaji, F.; Kokubo, T.; Kozuka, H.; Yamamoto,

K., Antibacterial silver-containing silica glass prepared by sol-gel method. Biomaterials

2000, 21, (4), 393-398.

240. Lohbauer, U.; Jell, G.; Saravanapavan, P.; Jones, J. R.; Hench, L. L., Indirect

cytotoxicity evaluation of silver doped bioglass Ag-S70C30 on human primary

keratinocytes. In Bioceramics 17, Li, P.; Zhang, K.; Colwell, C. W., Eds. Trans Tech

Publications Ltd: Zurich-Uetikon, 2005; Vol. 284-286, pp 431-434.

241. Verné, E.; Miola, M.; Vitale Brovarone, C.; Cannas, M.; Gatti, S.; Fucale, G.;

Maina, G.; Massé, A.; Di Nunzio, S., Surface silver-doping of biocompatible glass to

induce antibacterial properties. Part I: massive glass. J. Mater. Sci.: Mater. Med. 2009,

20, (3), 733-740.

242. Leonelli, C.; Lusvardi, G.; Malavasi, G.; Menabue, L.; Tonelli, M., Synthesis

and characterization of cerium-doped glasses and in vitro evaluation of bioactivity. J.

Non-Cryst. Solids 2003, 316, (2-3), 198-216.

243. Fu, H.; Fu, Q.; Zhou, N.; Huang, W.; Rahaman, M. N.; Wang, D.; Liu, X., In

vitro evaluation of borate-based bioactive glass scaffolds prepared by a polymer foam

replication method. Mater. Sci. Eng., C 2009, 29, (7), 2275-2281.

244. Gorustovich, A.; Steitnetz, T.; Lopez, J. P.; Guglielmotti, M.; Cabrini, R.,

Osteogenic response to bioactive glass particles modified by boron. Bone 2006, 38, (5),

1.

245. Liang, W.; Rahaman, M. N.; Day, D. E.; Marion, N. W.; Riley, G. C.; Mao, J. J.,

Bioactive borate glass scaffold for bone tissue engineering. J. Non-Cryst. Solids 2008,

354, (15-16), 1690-1696.

246. Bergandi, L.; Aina, V.; Garetto, S.; Malavasi, G.; Aldieri, E.; Laurenti, E.;

Matera, L.; Morterra, C.; Ghigo, D., Fluoride-containing bioactive glasses inhibit

pentose phosphate oxidative pathway and glucose 6-phosphate dehydrogenase activity

in human osteoblasts. Chem. Biol. Interact. 2010, 183, (3), 405-415.

247. Lakhkar, N. J.; Lee, I. H.; Kim, H. W.; Salih, V.; Wall, I. B.; Knowles, J. C.,

Bone formation controlled by biologically relevant inorganic ions: Role and controlled

delivery from phosphate-based glasses. Adv. Drug Delivery Rev. 2013, 65, (4), 405-420.

248. Agathopoulos, S.; Tulyaganov, D. U.; Ventura, J. M. G.; Kannan, S.;

Karakassides, M. A.; Ferreira, J. M. F., Formation of hydroxyapatite onto glasses of the

Page 179: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

References 179

CaO-MgO-SiO2 system with B2O3, Na2O, CaF2 and P2O5 additives. Biomaterials

2006, 27, (9), 1832-1840.

249. Vrouwenvelder, W. C. A.; Groot, C. G.; de Groot, K., Better histology and

biochemistry for osteoblasts cultured on titanium-doped bioactive glass: Bioglass 45S5

compared with iron-, titanium-, fluorine- and boron-containing bioactive glasses.

Biomaterials 1994, 15, (2), 97-106.

250. Courthéoux, L.; Lao, J.; Nedelec, J. M.; Jallot, E., Controlled bioactivity in zinc-

doped sol−gel-derived binary bioactive glasses. J. Phys. Chem. C 2008, 112, (35),

13663-13667.

251. Bini, M.; Grandi, S.; Capsoni, D.; Mustarelli, P.; Saino, E.; Visai, L.,

SiO2−P2O5−CaO glasses and glass-ceramics with and without ZnO: Relationships

among composition, microstructure, and bioactivity. J. Phys. Chem. C 2009, 113, (20),

8821-8828.

252. Lusvardi, G.; Malavasi, G.; Menabue, L.; Menziani, M. C.; Pedone, A.; Segre,

U.; Aina, V.; Perardi, A.; Morterra, C.; Boccafoschi, F.; Gatti, S.; Bosetti, M.; Cannas,

M., Properties of zinc releasing surfaces for clinical applications. J. Biomater. Appl.

2008, 22, (6), 505-526.

253. Aina, V.; Perardi, A.; Bergandi, L.; Malavasi, G.; Menabue, L.; Morterra, C.;

Ghigo, D., Cytotoxicity of zinc-containing bioactive glasses in contact with human

osteoblasts. Chem. Biol. Interact. 2007, 167, (3), 207-218.

254. Du, R. L.; Chang, J.; Ni, S. Y.; Zhai, W. Y.; Wang, J. Y., Characterization and in

vitro bioactivity of Zinc-containing bioactive glass and glass-ceramics. J. Biomater.

Appl. 2006, 20, (4), 341-360.

255. Oki, A.; Parveen, B.; Hossain, S.; Adeniji, S.; Donahue, H., Preparation and in

vitro bioactivity of zinc containing sol-gel-derived bioglass materials. Journal of

Biomedical Materials Research. Part A. 2004, 69A, (2), 216-221.

256. Lao, J.; Jallot, E.; Nedelec, J.-M., Strontium-delivering glasses with enhanced

bioactivity: A new biomaterial for antiosteoporotic applications? Chem. Mater. 2008,

20, (15), 4969-4973.

257. Barralet, J.; Gbureck, U.; Habibovic, P.; Vorndran, E.; Gerard, C.; Doillon, C. J.,

Angiogenesis in calcium phosphate scaffolds by inorganic copper ion release. Tissue

Eng. Part A 2009, 15, (7), 1601-1609.

258. Ewald, A.; Käppel, C.; Vorndran, E.; Moseke, C.; Gelinsky, M.; Gbureck, U.,

The effect of Cu(II)-loaded brushite scaffolds on growth and activity of osteoblastic

cells. Journal of Biomedical Materials Research Part A 2012, 100 A, (9), 2392-2400.

259. Abou Neel, E. A.; Ahmed, I.; Pratten, J.; Nazhat, S. N.; Knowles, J. C.,

Characterisation of antibacterial copper releasing degradable phosphate glass fibres.

Biomaterials 2005, 26, (15), 2247-2254.

Page 180: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

References 180

260. Stähli, C.; Muja, N.; Nazhat, S. N., Controlled copper ion release from

phosphate-based glasses improves human umbilical vein endothelial cell survival in a

reduced nutrient environment. Tissue Eng. Part A 2013, 19, (3-4), 548-557.

261. Erol, M. M.; Mouriňo, V.; Newby, P.; Chatzistavrou, X.; Roether, J. A.; Hupa,

L.; Boccaccini, A. R., Copper-releasing, boron-containing bioactive glass-based

scaffolds coated with alginate for bone tissue engineering. Acta Biomater. 2012, 8, (2),

792-801.

262. Zhang, M. L.; Wu, C. T.; Li, H. Y.; Yuen, J.; Chang, J.; Xiao, Y., Preparation,

characterization and in vitro angiogenic capacity of cobalt substituted beta-tricalcium

phosphate ceramics. J. Mater. Chem. 2012, 22, (40), 21686-21694.

263. Jain, R. K.; Au, P.; Tam, J.; Duda, D. G.; Fukumura, D., Engineering

vascularized tissue. Nat Biotech 2005, 23, (7), 821-823.

264. Incerti, S.; Zhang, Q.; Andersson, F.; Moretto, P.; Grime, G. W.; Merchant, M.

J.; Nguyen, D. T.; Habchi, C.; Pouthier, T.; Seznec, H., Monte Carlo simulation of the

CENBG microbeam and nanobeam lines with the Geant4 toolkit. Nucl. Instrum.

Methods Phys. Res., Sect. B 2007, 260, (1), 20-27.

265. Maxwell, J. A.; Teesdale, W. J.; Campbell, J. L., The Guelph PIXE software

package II. Nucl. Instrum. Methods Phys. Res., Sect. B 1995, 95, (3), 407-421.

266. Rath, S. N.; Strobel, L. A.; Arkudas, A.; Beier, J. P.; Maier, A.-K.; Greil, P.;

Horch, R. E.; Kneser, U., Osteoinduction and survival of osteoblasts and bone-marrow

stromal cells in 3D biphasic calcium phosphate scaffolds under static and dynamic

culture conditions. J. Cell. Mol. Med. 2012, 16, (10), 2350-2361.

267. Cerruti, M.; Bianchi, C. L.; Bonino, F.; Damin, A.; Perardi, A.; Morterra, C.,

Surface Modifications of Bioglass Immersed in TRIS-Buffered Solution. A

Multitechnical Spectroscopic Study. J. Phys. Chem. B 2005, 109, (30), 14496-14505.

268. Cerruti, M.; Morterra, C., Carbonate Formation on Bioactive Glasses. Langmuir

2004, 20, (15), 6382-6388.

269. Ardelean, I.; Cora, S.; Ioncu, V., Structural investigation of CuO-Bi2O3-B2O3

glasses by FT-IR, Raman and UVNIS spectroscopies. Journal of Optoelectronics and

Advanced Materials 2006, 8, (5), 1843-1847.

270. Schultz, P. C., Optical Absorption of the Transition Elements in Vitreous Silica.

J. Am. Ceram. Soc. 1974, 57, (7), 309-313.

271. Andronenko, S. I.; Andronenko, R. R.; Vasil'ev, A. V.; Zagrebel'nyi, O. A., Local

Symmetry of Cu&lt;sup&gt;2+&lt;/sup&gt; Ions in Sodium Silicate Glasses from Data

of EPR Spectroscopy. Glass Phys. Chem. 2004, 30, (3), 230-235.

272. Schreiber, H. D.; Kochanowski, B. K.; Schreiber, C. W.; Morgan, A. B.;

Coolbaugh, M. T.; Dunlap, T. G., Compositional dependence of redox equilibria in

sodium silicate glasses. J. Non-Cryst. Solids 1994, 177, (C), 340-346.

Page 181: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

References 181

273. Abdrakhmanov, R. S.; Ivanova, T. A., The influence of quadrupole effects on

Cu(II) ESR spectra in glasses. J. Mol. Struct. 1978, 46, (0), 229-244.

274. Elgayar, I.; Aliev, A. E.; Boccaccini, A. R.; Hill, R. G., Structural analysis of

bioactive glasses. J. Non-Cryst. Solids 2005, 351, (2), 173-183.

275. Sûowska, J.; Wačawska, I.; Szumera, M., Effect of copper addition on glass

transition of silicate-phosphate glasses. J. Therm. Anal. Calorim. 2012, 109, (2), 705-

710.

276. Bretcanu, O.; Chatzistavrou, X.; Paraskevopoulos, K.; Conradt, R.; Thompson,

I.; Boccaccini, A. R., Sintering and crystallisation of 45S5 Bioglass ® powder. J. Eur.

Ceram. Soc. 2009, 29, (16), 3299-3306.

277. Lefebvre, L.; Chevalier, J.; Gremillard, L.; Zenati, R.; Thollet, G.; Bernache-

Assolant, D.; Govin, A., Structural transformations of bioactive glass 45S5 with thermal

treatments. Acta Mater. 2007, 55, (10), 3305-3313.

278. Druecke, D.; Langer, S.; Lamme, E.; Pieper, J.; Ugarkovic, M.; Steinau, H. U.;

Homann, H. H., Neovascularization of poly(ether ester) block-copolymer scaffolds in

vivo: Long-term investigations using intravital fluorescent microscopy. Journal of

Biomedical Materials Research, Part A 2004, 68, (1), 10-18.

279. Yang, S.; Leong, K. F.; Du, Z.; Chua, C. K., The design of scaffolds for use in

tissue engineering. Part I. Traditional factors. Tissue Eng. 2001, 7, (6), 679-689.

280. Lefebvre, L.; Gremillard, L.; Chevalier, J.; Zenati, R.; Bernache-Assolant, D.,

Sintering behaviour of 45S5 bioactive glass. Acta Biomater. 2008, 4, (6), 1894-1903.

281. Chatzistavrou, X.; Zorba, T.; Kontonasaki, E.; Chrissafis, K.; Koidis, P.;

Paraskevopoulos, K. M., Following bioactive glass behavior beyond melting

temperature by thermal and optical methods. Physica Status Solidi (A) Applied Research

2004, 201, (5), 944-951.

282. Bunker, B. C.; Tallant, D. R.; Headley, T. J.; Turner, G. L.; Kirkpatrick, R. J.,

Structure of leached sodium borosilicate glass. Phys. Chem. Glasses 1988, 29, (3), 106-

120.

283. Layrolle, P.; Ito, A.; Tateishi, T., Sol-Gel Synthesis of Amorphous Calcium

Phosphate and Sintering into Microporous Hydroxyapatite Bioceramics. J. Am. Ceram.

Soc. 1998, 81, (6), 1421-1428.

284. Koutsopoulos, S., Synthesis and characterization of hydroxyapatite crystals: A

review study on the analytical methods. J. Biomed. Mater. Res. 2002, 62, (4), 600-612.

285. Kokubo, T.; Ito, S.; Huang, Z. T.; Hayashi, T.; Sakka, S.; Kitsugi, T.; Yamamuro,

T., Ca, P-rich layer formed on high-strength bioactive glass-ceramic A-W. J. Biomed.

Mater. Res. 1990, 24, (3), 331-343.

Page 182: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

References 182

286. Tilocca, A.; Cormack, A. N., The initial stages of bioglass dissolution: A Car-

Parrinello molecular-dynamics study of the glass-water interface. Proc. R. Soc. A 2011,

467, (2131), 2102-2111.

287. Tilocca, A., Models of structure, dynamics and reactivity of bioglasses: A

review. J. Mater. Chem. 2010, 20, (33), 6848-6858.

288. Tilocca, A.; Cormack, A. N., Modeling the water-bioglass interface by Ab initio

molecular dynamics simulations. ACS Appl. Mater. Interfaces 2009, 1, (6), 1324-1333.

289. Aguiar, H.; Solla, E. L.; Serra, J.; González, P.; León, B.; Almeida, N.;

Cachinho, S.; Davim, E. J. C.; Correia, R.; Oliveira, J. M.; Fernandes, M. H. V.,

Orthophosphate nanostructures in SiO2-P2O5-CaO-Na2O-MgO bioactive glasses. J.

Non-Cryst. Solids 2008, 354, (34), 4075-4080.

290. Dorozhkin, S. V.; Epple, M., Biological and medical significance of calcium

phosphates. Angewandte Chemie - International Edition 2002, 41, (17), 3130-3146.

291. Zhang, D.; Jain, H.; Hupa, M.; Hupa, L., In-vitro Degradation and Bioactivity of

Tailored Amorphous Multi Porous Scaffold Structure. J. Am. Ceram. Soc. 2012, 95, (9),

2687-2694.

292. Sen, C. K.; Khanna, S.; Venojarvi, M.; Trikha, P.; Ellison, E. C.; Hunt, T. K.;

Roy, S., Copper-induced vascular endothelial growth factor expression and wound

healing. Am. J. Physiol. Heart Circ. Physiol. 2002, 282, (5), H1821-H1827.

293. Xie, H.; Kang, Y. J., Role of copper in angiogenesis and its medicinal

implications. Curr. Med. Chem. 2009, 16, (10), 1304-1314.

294. Webster, T. J.; Massa-Schlueter, E. A.; Smith, J. L.; Slamovich, E. B., Osteoblast

response to hydroxyapatite doped with divalent and trivalent cations. Biomaterials

2004, 25, (11), 2111-2121.

295. Silver, I. A.; Deas, J.; Erecinska, M., Interactions of bioactive glasses with

osteoblasts in vitro: effects of 45S5 Bioglass (R), and 58S and 77S bioactive glasses on

metabolism, intracellular ion concentrations and cell viability. Biomaterials 2001, 22,

(2), 175-185.

296. Labbaf, S.; Tsigkou, O.; Müller, K. H.; Stevens, M. M.; Porter, A. E.; Jones, J.

R., Spherical bioactive glass particles and their interaction with human mesenchymal

stem cells in vitro. Biomaterials 2011, 32, (4), 1010-1018.

297. Lynch, S. M.; Frei, B., Mechanisms of copper- and iron-dependent oxidative

modification of human low density lipoprotein. J. Lipid Res. 1993, 34, (10), 1745-53.

298. Gaetke, L. M.; Chow, C. K., Copper toxicity, oxidative stress, and antioxidant

nutrients. Toxicology 2003, 189, (1–2), 147-163.

Page 183: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

References 183

299. Milkovic, L.; Hoppe, A.; Detsch, R.; Boccaccini, A. R.; Zarkovic, N., Effects of

Cu-doped 45S5 bioactive glass on the lipid peroxidation-associated growth of human

osteoblast-like cells in vitro. J. Biomed. Mater. Res., Part A 2013.

300. Valko, M.; Leibfritz, D.; Moncol, J.; Cronin, M. T. D.; Mazur, M.; Telser, J.,

Free radicals and antioxidants in normal physiological functions and human disease.

The International Journal of Biochemistry & Cell Biology 2007, 39, (1), 44-84.

301. Zarkovic, N.; Ilic, Z.; Jurin, M.; Schaur, R. J.; Puhl, H.; Esterbauer, H.,

STIMULATION OF HELA-CELL GROWTH BY PHYSIOLOGICAL

CONCENTRATIONS OF 4-HYDROXYNONENAL. Cell Biochem. Funct. 1993, 11,

(4), 279-286.

302. Awasthi, Y. C.; Yang, Y.; Tiwari, N. K.; Patrick, B.; Sharma, A.; Li, J.; Awasthi,

S., Regulation of 4-hydroxynonenal-mediated signaling by glutathione S-transferases.

Free Radic. Biol. Med. 2004, 37, (5), 607-619.

303. Czekanska, E. M.; Stoddart, M. J.; Richards, R. G.; Hayes, J. S., In search of an

osteoblast cell model for in vitro research. Eur. Cell. Mater. 2012, 24, 1-17.

304. Feng, W.; Ye, F.; Xue, W.; Zhou, Z.; Kang, Y. J., Copper regulation of hypoxia-

inducible factor-1 activity. Mol. Pharmacol. 2009, 75, (1), 174-182.

305. Martin, F.; Linden, T.; Katschinski, D. M.; Oehme, F.; Flamme, I.;

Mukhopadhyay, C. K.; Eckhardt, K.; Tröger, J.; Barth, S.; Camenisch, G.; Wenger, R.

H., Copper-dependent activation of hypoxia-inducible factor (HIF)-1: implications for

ceruloplasmin regulation. Blood 2005, 105, (12), 4613-4619.

306. Okuyama, H.; Krishnamachary, B.; Zhou, Y. F.; Nagasawa, H.; Bosch-Marce,

M.; Semenza, G. L., Expression of Vascular Endothelial Growth Factor Receptor 1 in

Bone Marrow-derived Mesenchymal Cells Is Dependent on Hypoxia-inducible Factor

1. J. Biol. Chem. 2006, 281, (22), 15554-15563.

307. Cébe-Suarez, S.; Zehnder-Fjällman, A.; Ballmer-Hofer, K., The role of VEGF

receptors in angiogenesis; complex partnerships. Cell. Mol. Life Sci. 2006, 63, (5), 601-

615.

308. Jiang, Y.; Reynolds, C.; Xiao, C.; Feng, W.; Zhou, Z.; Rodriguez, W.; Tyagi, S.

C.; Eaton, J. W.; Saari, J. T.; Kang, Y. J., Dietary copper supplementation reverses

hypertrophic cardiomyopathy induced by chronic pressure overload in mice. J. Exp.

Med. 2007, 204, (3), 657-66.

309. McAuslan, B. R.; Reilly, W. G.; Hannan, G. N.; Gole, G. A., Angiogenic factors

and their assay: activity of formyl methionyl leucyl phenylalanine, adenosine

diphosphate, heparin, copper, and bovine endothelium stimulating factor. Microvasc.

Res. 1983, 26, (3), 323-38.

310. Sen, C. K.; Khanna, S.; Venojarvi, M.; Trikha, P.; Christopher Ellison, E.; Hunt,

T. K.; Roy, S., Copper-induced vascular endothelial growth factor expression and

wound healing. Am. J. Physiol. Heart Circ. Physiol. 2002, 282, (5 51-5), H1821-H1827.

Page 184: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

References 184

311. Li, S.; Xie, H.; Kang, Y., Copper stimulates growth of human umbilical vein

endothelial cells in a vascular endothelial growth factor-independent pathway. Exp. Biol.

Med. 2012, 237, (1), 77-82.

312. Fuchs, S.; Hofmann, A.; Kirkpatrick, C. J., Microvessel-like structures from

outgrowth endothelial cells from human peripheral blood in 2-dimensional and 3-

dimensional co-cultures with osteoblastic lineage cells. Tissue Eng. 2007, 13, (10),

2577-2588.

313. Kirkpatrick, C. J.; Fuchs, S.; Unger, R. E., Co-culture systems for

vascularization - Learning from nature. Adv. Drug Delivery Rev. 2011, 63, (4), 291-299.

314. Giavaresi, G.; Torricelli, P.; Fornasari, P. M.; Giardino, R.; Barbucci, R.; Leone,

G., Blood vessel formation after soft-tissue implantation of hyaluronan-based hydrogel

supplemented with copper ions. Biomaterials 2005, 26, (16), 3001-3008.

315. Arkudas, A.; Balzer, A.; Buehrer, G.; Arnold, I.; Hoppe, A.; Detsch, R.; Newby,

P.; Fey, T.; Greil, P.; Horch, R. E.; Boccaccini, A. R.; Kneser, U., Evaluation of

Angiogenesis of Bioactive Glass in the Arteriovenous Loop Model. Tissue Eng Part C

Methods 2013, 16, 16.

316. Rong, M. Z.; Zhang, M. Q.; Zheng, Y. X.; Zeng, H. M.; Friedrich, K.,

Improvement of tensile properties of nano-SiO2/PP composites in relation to

percolation mechanism. Polymer 2001, 42, (7), 3301-3304.

317. Serra, J.; González, P.; Liste, S.; Serra, C.; Chiussi, S.; León, B.; Pérez-Amor,

M.; Ylänen, H. O.; Hupa, M., FTIR and XPS studies of bioactive silica based glasses. J.

Non-Cryst. Solids 2003, 332, (1-3), 20-27.

318. Selvaraj, M.; Kim, B. H.; Lee, T. G., FTIR Studies on Selected Mesoporous

Metallosilicate Molecular Sieves. Chem. Lett. 2005, 34, (9), 1290-1291.

319. González, P.; Serra, J.; Liste, S.; Chiussi, S.; León, B.; Pérez-Amor, M., Raman

spectroscopic study of bioactive silica based glasses. J. Non-Cryst. Solids 2003, 320, (1-

3), 92-99.

320. McMillan, P. F.; Wolf, G. H.; Poe, B. T., Vibrational spectroscopy of silicate

liquids and glasses. Chem. Geol. 1992, 96, (3–4), 351-366.

321. Geissberger, A. E.; Galeener, F. L., Raman studies of vitreous SiO_{2} versus

fictive temperature. Physical Review B 1983, 28, (6), 3266-3271.

322. Jallot, E.; Lao, J.; John, Ł.; Soulié, J.; Moretto, P.; Nedelec, J. M., Imaging

Physicochemical Reactions Occurring at the Pore Surface in Binary Bioactive Glass

Foams by Micro Ion Beam Analysis. ACS Appl. Mater. Interfaces 2010, 2, (6), 1737-

1742.

323. Lao, J.; Nedelec, J. M.; Jallot, E., New insight into the physicochemistry at the

interface between sol-gel-derived bioactive glasses and biological medium: A PIXE-

RBS study. J. Phys. Chem. C 2008, 112, (25), 9418-9427.

Page 185: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

References 185

324. Elkabouss, K.; Kacimi, M.; Ziyad, M.; Ammar, S.; Bozon-Verduraz, F., Cobalt-

exchanged hydroxyapatite catalysts: Magnetic studies, spectroscopic investigations,

performance in 2-butanol and ethane oxidative dehydrogenations. J. Catal. 2004, 226,

(1), 16-24.

325. Matsunaga, K., First-principles study of substitutional magnesium and zinc in

hydroxyapatite and octacalcium phosphate. The Journal of Chemical Physics 2008, 128,

(24), 245101-10.

326. Zan, T.; Du, Z. J.; Li, H.; Li, Q. F.; Gu, B., Cobalt chloride enhances angiogenic

potential of CD133(+) cells. Front. Biosci., Landmark Ed. 2012, 17, 2247-2258.

327. Minchenko, A.; Caro, J., Regulation of endothelin-1 gene expression in human

microvascular endothelial cells by hypoxia and cobalt: Role of hypoxia responsive

element. Mol. Cell. Biochem. 2000, 208, (1-2), 53-62.

328. Fleury, C.; Petit, A.; Mwale, F.; Antoniou, J.; Zukor, D. J.; Tabrizian, M.; Huk,

O. L., Effect of cobalt and chromium ions on human MG-63 osteoblasts in vitro:

Morphology, cytotoxicity, and oxidative stress. Biomaterials 2006, 27, (18), 3351-3360.

329. Watts, S. J.; Hill, R. G.; O’Donnell, M. D.; Law, R. V., Influence of magnesia on

the structure and properties of bioactive glasses. J. Non-Cryst. Solids 2010, 356, (9–10),

517-524.

330. Witzleb, W. C.; Ziegler, J.; Krummenauer, F.; Neumeister, V.; Guenther, K. P.,

Exposure to chromium, cobalt and molybdenum from metal-on-metal total hip

replacement and hip resurfacing arthroplasty. Acta Orthop. 2006, 77, (5), 697-705.

331. Pizon, A. F.; Abesamis, M.; King, A. M.; Menke, N., Prosthetic Hip-Associated

Cobalt Toxicity. J. Med. Toxicol. 2013, 9, (4), 416-417.

332. Schaffer, A. W.; Pilger, A.; Engelhardt, C.; Zweymueller, K.; Ruediger, H. W.,

Increased blood cobalt and chromium after total hip replacement. Journal of Toxicology

- Clinical Toxicology 1999, 37, (7), 839-844.

333. Loh, Q. L.; Choong, C., Three-dimensional scaffolds for tissue engineering

applications: role of porosity and pore size. Tissue Eng Part B Rev 2013, 19, (6), 485-

502.

334. Tsigkou, O.; Labbaf, S.; Stevens, M. M.; Porter, A. E.; Jones, J. R.,

Monodispersed Bioactive Glass Submicron Particles and Their Effect on Bone Marrow

and Adipose Tissue-Derived Stem Cells. Advanced Healthcare Materials 2014, 3, (1),

115-125.

335. Gibson, L. J.; Ashby, M. F., Cellular solids: structure and properties. Cambridge

university press: 1999.

336. Callcut, S.; Knowles, J. C., Correlation between structure and compressive

strength in a reticulated glass-reinforced hydroxyapatite foam. J. Mater. Sci.: Mater.

Med. 2002, 13, (5), 485-489.

Page 186: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

References 186

337. Kim, H. W.; Knowles, J. C.; Kim, H. E., Hydroxyapatite porous scaffold

engineered with biological polymer hybrid coating for antibiotic Vancomycin release. J.

Mater. Sci.: Mater. Med. 2005, 16, (3), 189-195.

338. Baino, F.; Ferraris, M.; Bretcanu, O.; Verné, E.; Vitale-Brovarone, C.,

Optimization of composition, structure and mechanical strength of bioactive 3-D glass-

ceramic scaffolds for bone substitution. J. Biomater. Appl. 2013, 27, (7), 872-890.

339. Vitale-Brovarone, C.; Baino, F.; Verné, E., High strength bioactive glass-ceramic

scaffolds for bone regeneration. J. Mater. Sci.: Mater. Med. 2009, 20, (2), 643-653.

340. Tamai, N.; Myoui, A.; Tomita, T.; Nakase, T.; Tanaka, J.; Ochi, T.; Yoshikawa,

H., Novel hydroxyapatite ceramics with an interconnective porous structure exhibit

superior osteoconduction in vivo. J. Biomed. Mater. Res. 2002, 59, (1), 110-7.

341. Chen, F.-M.; Zhang, M.; Wu, Z.-F., Toward delivery of multiple growth factors

in tissue engineering. Biomaterials 2010, 31, (24), 6279-6308.

342. Zhang, X. F.; Kehoe, S.; Adhi, S. K.; Ajithkumar, T. G.; Moane, S.; O'Shea, H.;

Boyd, D., Composition-structure-property (Zn2+ and Ca2+ ion release) evaluation of

Si-Na-Ca-Zn-Ce glasses: Potential components for nerve guidance conduits. Mater. Sci.

Eng., C 2011, 31, (3), 669-676.

343. Zhang, X. F.; O’Shea, H.; Kehoe, S.; Boyd, D., Time-dependent evaluation of

mechanical properties and in vitro cytocompatibility of experimental composite-based

nerve guidance conduits. J. Mech. Behav. Biomed. Mater. 2011, 4, (7), 1266-1274.

344. Cacaina, D.; Ylanen, H.; Simon, S.; Hupa, M., The behaviour of selected yttrium

containing bioactive glass microspheres in simulated body environments. Journal of

Materials Science-Materials in Medicine 2008, 19, (3), 1225-1233.

345. Shah, S. A.; Hashmi, M. U.; Alam, S., Effect of aligning magnetic field on the

magnetic and calorimetric properties of ferrimagnetic bioactive glass ceramics for the

hyperthermia treatment of cancer. Mater. Sci. Eng., C 2011, 31, (5), 1010-1016.

346. Shah, S. A.; Hashmi, M. U.; Shamim, A.; Alam, S., Study of an anisotropic

ferrimagnetic bioactive glass ceramic for cancer treatment. Applied Physics a-Materials

Science & Processing 2010, 100, (1), 273-280.

347. Jiang, Y. M. J. Y. M.; Ou, J.; Zhang, Z. H.; Qin, Q. H., Preparation of magnetic

and bioactive calcium zinc iron silicon oxide composite for hyperthermia treatment of

bone cancer and repair of bone defects. Journal of Materials Science-Materials in

Medicine 2011, 22, (3), 721-729.

348. Li, G.; Feng, S.; Zhou, D., Magnetic bioactive glass ceramic in the system CaO-

P(2)O (5)-SiO (2)-MgO-CaF (2)-MnO (2)-Fe (2)O (3) for hyperthermia treatment of

bone tumor. J. Mater. Sci. Mater. Med. 2011, 22, (10), 2197-206.

Page 187: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

187

349. Wu, C.; Fan, W.; Zhu, Y.; Gelinsky, M.; Chang, J.; Cuniberti, G.; Albrecht, V.;

Friis, T.; Xiao, Y., Multifunctional magnetic mesoporous bioactive glass scaffolds with a

hierarchical pore structure. Acta Biomater. 2011, 7, (10), 3563-3572.

350. Wang, T.-W.; Wu, H.-C.; Wang, W.-R.; Lin, F.-H.; Lou, P.-J.; Shieh, M.-J.;

Young, T.-H., The development of magnetic degradable DP-Bioglass for hyperthermia

cancer therapy. Journal of Biomedical Materials Research Part A 2007, 83A, (3), 828-

837.

351. Williams, D., Benefit and risk in tissue engineering. Materials Today 2004, 7,

(5), 24-29.

352. Guarino, V.; Causa, F.; Ambrosio, L., Bioactive scaffolds for bone and ligament

tissue. Expert Rev. Med. Devices 2007, 4, (3), 405-418.

353. Schubert, D.; Dargusch, R.; Raitano, J.; Chan, S.-W., Cerium and yttrium oxide

nanoparticles are neuroprotective. Biochem. Biophys. Res. Commun. 2006, 342, (1), 86-

91.

354. Hayes, J. S.; Czekanska, E. M.; Richards, R. G., The cell-surface interaction.

Adv. Biochem. Eng. Biotechnol. 2012, 126, 1-31.

355. Yang, L.; Perez-Amodio, S.; Barrère-de Groot, F. Y. F.; Everts, V.; van

Blitterswijk, C. A.; Habibovic, P., The effects of inorganic additives to calcium

phosphate on in vitro behavior of osteoblasts and osteoclasts. Biomaterials 2010, 31,

(11), 2976-2989.

356. Zhang, J.; Zhao, S.; Zhu, Y.; Huang, Y.; Zhu, M.; Tao, C.; Zhang, C., Three-

Dimensional Printing of Strontium-Containing Mesoporous Bioactive Glass Scaffolds

for Bone Regeneration. Acta Biomater., (0).

357. O'Donnell, M. D.; Hill, R. G., Influence of strontium and the importance of glass

chemistry and structure when designing bioactive glasses for bone regeneration. Acta

Biomater. 2010, 6, (7), 2382-2385.

358. Brunner, T. J.; Stark, W. J.; Boccaccini, A. R., Nanoscale Bioactive Silicate

Glasses in Biomedical Applications. In Nanotechnologies for the Life Sciences, Wiley-

VCH Verlag GmbH & Co. KGaA: 2007.

359. Romeis, S.; Hoppe, A.; Eisermann, C.; Schneider, N.; Boccaccini, A. R.;

Schmidt, J.; Peukert, W., Enhancing In Vitro Bioactivity of Melt-Derived 45S5

Bioglass® by Comminution in a Stirred Media Mill. J. Am. Ceram. Soc. 2013, n/a-n/a.

Page 188: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Appendix 188

8 Appendix

Fig. A 1: pH evolution of 45S5 bioactive glass derived scaffolds during immersion in different

solutions.

Fig. A 2: The primers used for real time RT-PCR analysis.

Page 189: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Appendix 189

Fig. A 3: XRD graphs of 1393-Co derived scaffolds confirming the amorphous state.

Fig. A 4: Elemental evolution in the inner region of the 1393-Co derived scaffolds during immersion

in SBF.

Page 190: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Appendix 190

Fig. A 5: Elemental evolution in the periphery layer of the 1393-Co derived scaffolds during

immersion in SBF.

Page 191: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Curriculum vitae

Alexander Hoppe

Henkestrasse91 91052 Erlangen +49 91318525525 [email protected]

Citizenship: German Date/place of birth: 1st July 1984 in Kustanaj/Kasachstan

Education

November 2009

(on-going)

PhD research (Institute of Biomaterials, University of Erlangen-Nuremberg) Topic: Bioactive glass based scaffolds with therapeutic ion release for bone tissue engineering Supervisor: Prof. Aldo R. Boccaccini

2004-2009

Study of Materials Science and Engineering (University of Erlangen-Nuremberg) completed with diploma degree (grade 1.3=”very good”) Thesis: Metal Ion doped biomimetic hydroxyapatite coatings Supervisor: Prof. P. Greil

2001-2004 Secondary education leading to Abitur (A levels); grade: 1.2 (=very good)

Main Research Activities / Expertise

Fabrication and characterisation of biomaterials

Synthesis and structural characterisation of melt-derived bioactive glasses and fabrication of 3D glass derived scaffolds

Compositional design of (bioactive) glasses

In vitro bioactivity and biomineralisation studies in biological fluids Materials surface characterisation (roughness, topography, wettability)

Biomimetic hydroxyapatite coatings; surface functionalisation Development of bioactive glass/polymer composites Characterisation of nano-scaled bioactive glass particles Degradation and ion release assessment of degradable (bio)materials

Cell biology / Cell-material interactions

Cell culture studies with osteoblast-like cells (MG-63, HOS), mesenchymal stem cells, osteoclasts (RAWs 264.7), endothelial cells (HDMECs)

Biocompatibility assessment of biomaterials (cell viability, cell attachment and cell morphology characterisation); Live/dead cell fluorescence staining

Osteogenic differentiation of stem cells (ALP activity, osteogenic gene-expression, ECM mineralisation)

In vitro angiogenesis assay with endothelial cells (cell viability,

Page 192: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

vessel-like tube formation, expression of angiogenic marker

Characterisation techniques

Materials: X-ray diffraction (XRD), Fourier-Transform Infrared and Raman Spectroscopy (FT-IR), Inductively Coupled Plasma Optical Emission Spectrometry (ICP-OES), Scanning electron microscopy (SEM)+Energy Dispersive Spectroscopy (EDS), X-ray Photoelectron spectroscopy (XPS), Nuclear magnetic resonance spectroscopy (NMR) Cell biology: Gene expression (RT-PCR), Laser Confocal Fluorescence Microscopy (LCFM), SEM, Protein assay (Bradford), Viability (WST-8, AlamarBlue), ELISA assay (VEGF), DNA quantification (PicoGreen, BrdU assay)

International Experience / Fellowships / Grants

Mar 2014 Short Term Scientific Mission (STSM) within a COST (European Cooperation in Science and Technology) action (Project NAMABIO) at 3B's Research Group, University of Minho, Braga, Portugal Supervisor: Prof. Rui Reis

Nov-Dec 2012 and Nov-Dec 2013

Research stay at the University of Buenos Aires, Argentina in the framework of the project: Design and development of novel matrices for tissue engineering and/or drug delivery for pathologies of high social and economic impact Supervisor: Prof. Viviana Murino

2012-2014 (on-going)

Project Assistant for the bilateral research exchange programme with Rudjer Boskovic Institute, Zagreb Coratia, Prof. Neven Zarkovic Research topic: Effect of novel Cu-doped 45S5 bioactive glass and lipid peroxidation products on bone regeneration

Mar 2012 Travel grant to the First São Carlos School of Advanced Studies in Materials Science and Engineering (SanCAS–MSE), São Carlos, Brazil

Aug-Nov 2011

Research stay at McGill University, Montreal, Canada in the framework of the Quebec-Bavarian project Mesenchymal stem cell seeded nano-composite constructs for bone tissue engineering Supervisor: Prof. Showan N. Nazhat

Oct-Nov 2010 KMM-VIN Research Fellowship at Polytechnic University of Turin Topic: Glass-ceramic scaffolds with antibacterial capability for bone tissue engineering Supervisor: Dr. Enrica Verne

Teaching and Supervising

Page 193: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Teaching 2011-2014 (summer terms) 2013 (summer term)

Tutorial lecture on Materials for Biomedical Engineering within the Masters programme Life Science Engineering

Teaching assistant in undergraduate laboratory courses on physical optics

Supervising activities (selected)

Florian Ruther, Bachelor thesis, 2013: Metal ion containing bioactive glass derived 3D scaffolds with enhanced mechanical strength

Stefan Grimm, Master thesis 2011: Fabrication and characterization of 3D highly porous scaffolds with enhanced biocompatibility based on ion doped glasses

Vincent Bürger, Master thesis 2010: Effects of residual surface stresses on the bioactivity of bioactive silicate glasses (Winner of the 2011 Oldfield Award)

Alexander Kent, Bachelor thesis 2010: Fabrication and characterisation of highly porous three-dimensional bioactive glass-based scaffolds for bone tissue engineering

Language and other skills

Language German (native speaker), Russian (native speaker), English (fluent), Spanish (intermediate)

Memberships Member of the American Ceramic Society (ACerS), Member of the German Biomaterials Society (DGBM)

Reviewer activity Reviewer in international peer to peer journals: Materials Letters, Acta Biomaterialia, Surface and Coating Technology, Journal of American Ceramic Society, Colloids and Surfaces B: Biointerfaces, NANO

Computer Scientific software (Origin, ChemBioDraw, EndNote), Graphic software (Photoshop, CorelDraw), MS Office (Word, Excel, Powerpoint)

Other interests I am interested in playing guitar. I also enjoy various kinds of sports including squash, basketball, running, racing bicycle and trekking.

Publications

Publications in peer-reviewed journals Hoppe A, Boccaccini A. R. On the degradation of bioactive glass scaffolds. 2014 (in preparation) Hoppe A, Brandl A, Bleiziffer O, Jokic B, Janackovic D, Boccaccini A R. In vitro cell response to Co-containing 1393 bioactive glass. 2014 (submitted). Hoppe A, Jokic B, Janackovic D, Fey T, Greil P, Romeis S, et al. Cobalt-Releasing 1393 Bioactive Glass-Derived Scaffolds for Bone Tissue Engineering Applications. ACS Appl Mater Interfaces. 2014;6(4):2865-77. Hoppe A, Sarker B, Detsch R, Hild N, Mohn D, Stark WJ, et al. In vitro reactivity of Sr-containing bioactive glass (type 1393) nanoparticles. J Non-Cryst Solids. 2014;387: 41-6. Hoppe A, Will J, Detsch R, Boccaccini AR, Greil P. Formation and in vitro biocompatibility of biomimetic hydroxyapatite coatings on chemically treated carbon substrates. J Biomed Mater Res, Part A. 2014;102: 193-203. Hoppe A, Meszaros R, Stähli C, Romeis S, Schmidt J, Peukert W, et al. In vitro reactivity of Cu doped 45S5 Bioglass® derived scaffolds for bone tissue engineering. J Mater Chem B. 2013;1: 5659-74. Hoppe A, Mourino V, Boccaccini AR. Therapeutic inorganic ions in bioactive glasses to enhance bone formation and beyond. Biomater. Sci. 2013;1:254.

Page 194: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

Hoppe A, Mačković M, Detsch R, Mohn D, Stark WJ, Spiecker E, et al. Bioactive glass (type 45S5) nanoparticles: in vitro reactivity on nanoscale and biocompatibility. J Nanopart Res. 2012;14: 1-22. Hoppe A, Güldal NS, Boccaccini AR. A review of the biological response to ionic dissolution products from bioactive glasses and glass-ceramics. Biomaterials. 2011;32: 2757-74. Rath S N, Brandl A, Hiller D, Hoppe A, Gbureck U, Horch R E, Boccaccini A R, Kneser U; Copper doped 45S5 bioactive glass scaffolds stimulate endothelial cells in a co-culture with mesenchymal stem cells. Biomaterials. Acta Biomater 2014 (submitted). Papageorgiou GZ, Papageorgiou DG, Chrissafis K, Bikiaris D, Will J, Hoppe A, et al. Crystallization and melting behavior of poly(butylene succinate) nanocomposites containing silica-nanotubes and strontium hydroxyapatite nanorods. Ind Eng Chem Res. 2014;53(2):678-92. Grigoriadou I, Nianias N, Hoppe A, Terzopoulou Z, Bikiaris D, Will J, et al. Evaluation of silica-nanotubes and strontium hydroxyapatite nanorods as appropriate nanoadditives for poly(butylene succinate) biodegradable polyester for biomedical applications. Composites Part B: Engineering. 2014;60: 49-59. Milkovic L, Hoppe A, Detsch R, Boccaccini AR, Zarkovic N. Effects of Cu-doped 45S5 bioactive glass on the lipid peroxidation-associated growth of human osteoblast-like cells in vitro. J Biomed Mater Res, Part A. 2013. Romeis S, Hoppe A, Eisermann C, Schneider N, Boccaccini AR, Schmidt J, et al. Enhancing In Vitro Bioactivity of Melt-Derived 45S5 Bioglass® by Comminution in a Stirred Media Mill. J Am Ceram Soc. 2014;97: 150-6. Arkudas A, Balzer A, Buehrer G, Arnold I, Hoppe A, Detsch R, et al. Evaluation of angiogenesis of bioactive glass in the arteriovenous loop model. Tissue Eng, Part C. 2013;19: 479-86. Strobel LA, Hild N, Mohn D, Stark WJ, Hoppe A, Gbureck U, et al. Novel strontium-doped bioactive glass nanoparticles enhance proliferation and osteogenic differentiation of human bone marrow stromal cells. J Nanopart Res. 2013;15: 1-9. Strobel LA, Rath SN, Hoppe A, Beier JP, Arkudas A, Boccaccini AR, et al. Influence of leptin on osteogenic differentiation of human marrow stromal cells (hMSC) and modulation of BMP-2-mediated osteoinduction. J Tissue Eng Regen Med. 2012;6: 271-. Cabal B, Malpartida F, Torrecillas R, Hoppe A, Boccaccini AR, Moya JS. The Development of Bioactive Glass-Ceramic Substrates with Biocide Activity. Adv Eng Mater. 2011;13(12):B462-B6. Will J, Hoppe A, Müller FA, Raya CT, Fernández JM, Greil P. Bioactivation of biomorphous silicon carbide bone implants. Acta Biomater. 2010;6: 4488-94 Book chapters A. Hoppe A, AR. Boccaccini. 7 - Bioactive glass foams for tissue engineering applications. In: Netti PA, editor. Biomedical Foams for Tissue Engineering Applications: Woodhead Publishing; 2014. p. 191-212. A. Hoppe, A. R. Boccaccini. Biological response to ionic dissolution products from bioactive glasses and glass ceramics. In S. Deb, editor. Karger Frontiers in Biology: Biomaterials in Regenerative Medicine and Dentistry. Basel: Karger. (submitted) Presentations in international conferences (selected) A. Hoppe, A. Brandl, O. Bleiziffer, D. Janackovic and A. R. Boccaccini. Cobalt releasing bioactive glass (type 13-93) derived scaffolds for bone tissue engineering applications. Oral presentation. Annual meeting of the German Biomaterials Society Sep 2013, Erlangen, Germany. A. Hoppe. 3D scaffolds for bone tissue engineering based on Cu-releasing 45S5 Bioglass®. Invited oral presentation at the Symposium on “Bioceramics as Carrier” Jun 2013, Medical Center, University of Freiburg, Germany. A. Hoppe, J. Will, R. Detsch, A.R. Boccaccini und P. Greil. Formation and biocompatibility of biomimetic hydroxyapatite on chemically treated carbon substrates. Poster presentation. Annual meeting of the German Biomaterials Society Nov 2012, Hamburg Germany.

Page 195: Bioactive Glass Derived Scaffolds with Therapeutic Ion Releasing ...

A. Hoppe, D. Hiller, S. Narayan Rath, A. Arkudas, U. Kneser, A. R. Boccaccini. In vitro and in vivo studies of Cu-doped 45S5 bioactive glass derived scaffolds. Oral presentation. 3rd International Conference "Strategies in Tissue Engineering" May 2012, Würzburg Germany. A. Hoppe, T. Reichel, R. Detsch, A. Lenhart, A. R. Boccaccini. Ion doped bioactive glass derived scaffolds for bone tissue engineering. Poster presentation. 3rd International Conference "Strategies in Tissue Engineering" May 2012, Würzburg Germany A. R. Boccaccini, A. Hoppe. Cellular response to ionic dissolution products from bioactive glasses and glass ceramics. Invited oral presentation. Meeting of the American Ceramic Society Jan 2012, Daytona Beach (FL), USA. A. Hoppe, D. Hiller, U. Kneser, A. R. Boccaccini Novel scaffolds made from metallic ion doped bioactive glasses: fabrication and characterization. Oral presentation in the young researcher forum. Annual meeting of the American Ceramic Society Jan 2012, Daytona Beach (FL) USA A. Hoppe, D. Hiller, S. Narayan Rath, U. Kneser, A. R. Boccaccini Novel Cu-doped bioactive glass (45S5) derived scaffolds for bone tissue engineering. Oral presentation, Annual meeting of the American Ceramic Society Jan 2012, Daytona Beach (FL) USA A. Hoppe, M. Miola, E. Verné, A.R. Boccaccini. Novel Zn-doped bioactive glasses developed by ion-exchange. Poster presentation, EUROMAT Sep 2011, Montpellier, France A. Boccaccini, A. Hoppe. Oral presentation, The biological effect of ionic dissolution products from bioactive glasses, EUROMAT Sep 2011, Montpellier, France