184
University of Connecticut OpenCommons@UConn Doctoral Dissertations University of Connecticut Graduate School 1-12-2018 Fabrication and Characterization of Collagen- Apatite Composites for Bone Tissue Engineering Changmin Hu University of Connecticut - Storrs, [email protected] Follow this and additional works at: hps://opencommons.uconn.edu/dissertations Recommended Citation Hu, Changmin, "Fabrication and Characterization of Collagen-Apatite Composites for Bone Tissue Engineering" (2018). Doctoral Dissertations. 1695. hps://opencommons.uconn.edu/dissertations/1695

Fabrication and Characterization of Collagen-Apatite

  • Upload
    others

  • View
    7

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Fabrication and Characterization of Collagen-Apatite

University of ConnecticutOpenCommons@UConn

Doctoral Dissertations University of Connecticut Graduate School

1-12-2018

Fabrication and Characterization of Collagen-Apatite Composites for Bone Tissue EngineeringChangmin HuUniversity of Connecticut - Storrs, [email protected]

Follow this and additional works at: https://opencommons.uconn.edu/dissertations

Recommended CitationHu, Changmin, "Fabrication and Characterization of Collagen-Apatite Composites for Bone Tissue Engineering" (2018). DoctoralDissertations. 1695.https://opencommons.uconn.edu/dissertations/1695

Page 2: Fabrication and Characterization of Collagen-Apatite

Fabrication and Characterization of Collagen-Apatite Composites

for Bone Tissue Engineering

Changmin Hu, PhD

University of Connecticut, 2018

Abstract

Due to the high demand of scaffolds for treating bone fractures every year,

biomimetic coatings and scaffolds resembling the composition and structure of natural

bone are of keen interest of bone tissue engineers. In this research, a series of

biomimetic coatings and scaffolds have been developed, including biomimetic apatite

coatings, biomimetic collagen-apatite composite coatings, intrafibrillar calcified

collagen fibrils, intrafibrillar silicified collagen fibrils, and intrafibrillar calcified

collagen scaffolds.

First, a crack-free apatite coating was deposited on the surface-treated Ti6Al4V

substrate using a biomimetic process, in which the surface-treated titanium substrate

was immersed in a modified simulated body fluid (m-SBF). Dual-beam focused ion

beam/scanning electron microscopy (FIB/SEM) was used to characterize the

cross-sectional microstructures of the ceramic coating. Cross-sectional SEM images

and TEM images revealed that crack-free apatite coatings have been achieved by

adjusting the pH of the m-SBF. The coating formed demonstrates a gradient porous

structure, which is dense close to the substrate and then it becomes more and more

porous towards the surface of the coating.

Second, a biomimetic collagen-apatite composite coating was also successfully

Page 3: Fabrication and Characterization of Collagen-Apatite

Changmin Hu-University of Connecticut, 2018

formed on the surface-treated Ti6Al4V alloy using collagen-containing m-SBF.

During the composite coating formation, collagen and apatite co-precipitate to form

the composite coatings. Dual-beam FIB/SEM was used to characterize the

cross-sectional microstructures of the composite coatings. The result indicates that the

cross-section of the collagen-apatite composite coating also exhibits a gradient porous

structure, and that collagen-apatite composite coating is thinner and less porous than

the pure apatite coating.

Third, inspired by the structure of natural bone, collagen fibrils with intrafibrillar

calcification were prepared for bone tissue engineering. Poly(acrylic acid) (PAA) was

used as a sequestration analog of non-collagenous proteins (NCPs) for stabilizing

amorphous calcium phosphate (PAA-ACP) nanoprecursors to infiltrate into collagen

fibrils. Additionally, sodium tripolyphosphate (TPP) was applied as a templating

analog of NCPs to regulate the orderly deposition of minerals within the gap zone of

collagen fibrils. The effect of PAA concentration on the intrafibrillar mineralization of

reconstituted collagen fibrils was also investigated.

Fourth, silicon is an essential element contributing significantly to the health of

bone, and great efforts have been made to produce silica-containing biomaterials.

Poly (allylamine) hydrochloride (PAH) was used as an analog of zwitterionic proteins

to stabilize silicic acid (SA) to produce fluidic silica (PAH-SA) nanoprecursors, and

TPP was used as a templating analog of zwitterionic proteins. Silicified collagen

fibrils with core-shell, twisted and banded structures were produced by controlling the

zwitterionic proteins analogs. Intrafibrillar silicified collagen fibrils were produced

Page 4: Fabrication and Characterization of Collagen-Apatite

Changmin Hu-University of Connecticut, 2018

using PAH-SA as fluidic silica nanoprecursors and TPP as the templating analog to

modulate the deposition of silica within the gap zone of collagen fibrils.

Fifth, with the success in reproducing intrafibrillar mineralized collagen fibrils,

biomimetic collagen/apatite scaffolds consisting of intrafibrillar mineralized collagen

fibrils were prepared via a bottom-up approach, resembling the composition and

structure of natural bone from the nanoscale to the macroscale. At the macro-scale,

unidirectional macro-pores are aligned across the entire scaffold, and at the

micro-scale, each layer is comprised of aligned and well stacked lamella. Finally, at

the nano-scale, apatite minerals deposit within the gap zone of collagen fibrils and

orient along the long axis of these fibrils. The biomimetic collagen-apatite scaffolds

demonstrate a good biocompatibility in vitro, indicating a great potential to be used

for bone tissue engineering applications.

In summary, a novel methodology has been developed for the preparation and

characterization of biomimetic coatings and scaffolds with a hierarchical structure,

demonstrating a great potential for bone tissue engineering.

Page 5: Fabrication and Characterization of Collagen-Apatite

i

Fabrication and Characterization of Collagen-Apatite Composites

for Bone Tissue Engineering

Changmin Hu

B.S., Jinan University, 2010

M.S., Shanghai Jiao Tong University, 2013

A Dissertation

Submitted in Partial Fulfillment of the

Requirements for the Degree of

Doctor of Philosophy

at the

University of Connecticut

2018

Page 6: Fabrication and Characterization of Collagen-Apatite

ii

Copyright by

Changmin Hu

2018

Page 7: Fabrication and Characterization of Collagen-Apatite

iii

APPROVAL PAGE

Doctor of Philosophy Dissertation

Fabrication and Characterization of Collagen-Apatite Composites

for Bone Tissue Engineering

Presented by

Changmin Hu, B.S., M.S.

Major Advisor ______________________________________________________

Mei Wei

Associate Advisor ___________________________________________________

Mark Aindow

Associate Advisor ___________________________________________________

Yu Lei

University of Connecticut

2018

Page 8: Fabrication and Characterization of Collagen-Apatite

iv

Acknowledgement

First and foremost, my special and greatest thanks go to my advisor Prof. Mei

Wei for her invaluable guidance and support though my Ph.D. study and research.

Prof. Wei is a great advisor, she not only teaches me how to do research, but also

helps me with my future development. She is very supportive when I have a chance to

develop my career in industry. She cares about my research, study, life, and future

development. I would want to express my greatest appreciation to her.

I would also like to thank my associate advisors: Prof. Mark Aindow and Prof. Yu

Lei for their valuable comments, suggestions and help for my research projects. I

would want to thank my committee members: Prof. Seok-woo Lee and Prof.

Montgomery Shaw for their support of my work. I am also grateful for my previous

committee members: Prof. Harold D. Brody and Prof. Luyi Sun for their suggestions

and comments for my research. I would also want to thank our collaborators Prof.

David Rowe and Dr. Liping Wang from the University of Connecticut Health Center

(UCHC) for all their help.

I would like to thank Dr. Lichun Zhang and Dr. Roger Ristau for their technical

assistance on electron microscopy. I would also like to thank my previous and current

group members: Dr. Zengmin Xia, Dr. Erica Kramer, Dr. Max Villa, Dr. Michael Zilm,

Mr. Jonathan Russo, Dr. Stephanie Bendtsen, Mr. Drew Clearfield, Mr. Bryant

Heimbach, and Mr. Le Yu. Especially, I would want to thank Zengmin Xia. She is so

nice and helpful. She not only helped me with my research, but also helped me to

adapt to the new environment.

Page 9: Fabrication and Characterization of Collagen-Apatite

v

Last but not least, I would like to thank my parents and my sister for their

unconditioned support of all my choices.

Page 10: Fabrication and Characterization of Collagen-Apatite

vi

Table of Contents

1. Introduction ........................................................................................................... 1

1.1 Background ...................................................................................................... 1

1.2 Structures and composition of natural bone ...................................................... 2

1.3 Intrafibrillar mineralization .............................................................................. 3

1.4 Implant coating ................................................................................................ 5

1.5 Dual beam focused ion beam/scanning electron microscopy (FIB/SEM) .......... 5

2. Cross-sectional microstructure study of apatite coating on titanium alloy............. 19

2.1 Introduction.................................................................................................... 20

2.2 Materials and methods.................................................................................... 23

2.2.1 Biomimetic HA coating preparation ......................................................... 23

2.2.2 Microstructural characterization ............................................................... 24

2.3 Results and discussion .................................................................................... 26

2.4 Conclusions.................................................................................................... 39

3. Cross-sectional microstructure study of collagen-apatite composite coating on

titanium alloy .......................................................................................................... 47

3.1 Introduction.................................................................................................... 48

3.2 Materials and methods.................................................................................... 50

3.2.1 Biomimetic Col-HA coating formation..................................................... 50

3.2.2 Materials characterization ........................................................................ 52

3.3 Results and discussion .................................................................................... 54

3.4 Conclusions.................................................................................................... 66

4. Fabrication of intrafibrillar calcified collagen fibrils ............................................ 74

4.1 Introduction.................................................................................................... 75

Page 11: Fabrication and Characterization of Collagen-Apatite

vii

4.2 Materials and methods.................................................................................... 78

4.2.1 Materials .................................................................................................. 78

4.2.2 Preparation of intrafibrillar and extrafibrillar mineralized collagen fibrils 78

4.2.3 Characterization ....................................................................................... 79

4.3 Results ........................................................................................................... 80

4.4 Discussion ...................................................................................................... 84

4.5 Conclusions.................................................................................................... 89

5. Intrafibrillar silicification of collagen fibrils ........................................................ 97

5.1 Introduction.................................................................................................... 98

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

5.2.1 Materials .................................................................................................100

5.2.2 Preparation of hydrolyzed silicic acid ......................................................100

5.2.3 Preparation of PAH-stabilized silica (PAH-SA) precursors ......................100

5.2.4 Controlled biomimetic silicification of collagen fibrils via an OCSS

approach ..........................................................................................................101

5.2.5 Preparation of silicified collagen fibrils via a two-step approach .............102

5.2.6 Materials characterization .......................................................................103

5.2.7 In vitro cell culture and proliferation .......................................................104

5.2.8 Statistical analysis ...................................................................................105

5.3 Results and discussion ...................................................................................105

5.3.1 Col-SA fibrils ..........................................................................................105

5.3.2 Col-PAH-SA fibrils .................................................................................107

5.3.3 Col-inSA fibrils .......................................................................................109

5.3.4 Silicified collagen fibrils via a two-step approach ................................... 112

5.3.5 In vitro cell culture .................................................................................. 115

Page 12: Fabrication and Characterization of Collagen-Apatite

viii

5.4 Conclusions................................................................................................... 116

6. Intrafibrillar calcified collagen scaffolds .............................................................125

6.1 Introduction...................................................................................................127

6.2 Materials and methods...................................................................................131

6.2.1 Materials .................................................................................................131

6.2.2 Fabrication of extrafibrillar mineralized collagen-apatite (E-Col-Ap)

scaffolds ..........................................................................................................131

6.2.3 Fabrication of intrafibrillar and extrafibrillar mineralized collagen apatite

(IE-Col-Ap) scaffolds ......................................................................................132

6.2.4 Materials characterization .......................................................................133

6.2.5 In vitro cell culture ..................................................................................134

6.3 Results and discussions .................................................................................135

6.4 Conclusions...................................................................................................146

7. Conclusions and future work ..............................................................................159

7.1 Conclusions...................................................................................................159

7.2 Future work ...................................................................................................161

Appendice A: List of publications...........................................................................164

Page 13: Fabrication and Characterization of Collagen-Apatite

ix

List of Figures

Chapter 1

Figure 1.1. Hierarchical structures of natural bone: starting from sub-nanostructure of

collagen molecules and tropocollagen triple helix, to nanostructure of ordered HA

crystals, and to sub-microstructure of collagen fibrils and macrostructure (right) (a).

The arrangement of HA crystals with collagen fibrils.

Figure 1.2. The specimen is tilted towards the ion beam (left), a cross-section is

milled that can be imaged with the electron beam (right).

Chapter 2

Figure 2.1. SE SEM images of HA coatings formed at different initial pH values:

HA-6.4 (A), HA-6.5 (B), HA-6.6 (C), and HA-6.7 (D). The black arrows in (B)

indicate cracks formed on the surface of the coating.

Figure 2.2. XRD spectra of HA coatings on Ti-6Al-4V substrates after 24 h of

immersion in m-SBF.

Figure 2.3. Cross-sectional SE SEM images of HA coatings on Ti-6Al-4V substrates:

HA-6.5 (A) and HA-6.6 (B). White arrow in image (A) is a crack induced by the ion

beam of PFIB and black arrow in image (A) is a crack between the HA-6.5 coating

and the Ti-6Al-4V substrate.

Figure 2.4. Gas-assisted deposition of Pt layer at 8 kV and 30 kV ion beam voltage in

PFIB.

Figure 2.5. SE SEM images of HA-6.6 coating on Ti-6Al-4V substrate from

cross-sections prepared using GFIB (A, B) and PFIB (C, D). The cross-sections were

Page 14: Fabrication and Characterization of Collagen-Apatite

x

polished using CCS pattern (A, C) or Rec pattern (B, D). Cracks indicated by white

arrows in HA coating in image (C) were induced by the CCS process in PFIB.

Figure 2.6. An extraction sequence of a thin lamella from the bulk in GFIB (A-H).

The thickness of the TEM specimen was confirmed by STEM in GFIB (I).

Figure 2.7. The thin lamella from the site of interest before extraction (A) and after

mounted onto copper omni grids (B) in PFIB.

Figure 2.8. Data obtained from the TEM specimens prepared by GFIB: BF TEM

image (A) and HAADF STEM image (B). The inset to (A) is the SAED pattern from

the coating. C-F: Compositional maps obtained from the region in (B), showing the

distribution of Ca (C), P (D), Ti (E), and Ga (F).

Figure 2.9. Data obtained from the TEM specimens prepared by PFIB: BF TEM

image (A) and HAADF STEM image (B). C-F: Compositional maps obtained from

the region in (B), showing the distribution of Ca (C), P (D), Ti (E), and Xe (F).

Chapter 3

Figure 3.1. Illustration of the formation process of biomimetic Col-HA composite

coatings on the surface-treated Ti-6Al-4V substrates: Surface morphology of the

Ti-6Al-4V substrate after polishing using silicon carbide abrasive paper (A) and SLA

treatment (B), and coating formation by immersing surface-treated Ti-6Al-4V

substrate in the collagen-containing m-SBF solution (C).

Figure 3.2. SE SEM images of Col-HA composite coatings on Ti-6Al-4V substrates.

Col-HA coated Ti-6Al-4V alloy was titled 52° (C).

Figure 3.3. SE SEM images of the surface morphologies of HA (A) and Col-HA (B)

Page 15: Fabrication and Characterization of Collagen-Apatite

xi

coatings on Ti-6Al-4V substrates.

Figure 3.4. XRD (A) and ATR-FTIR (B) spectra of HA and Col-HA coated Ti-6Al-4V

substrates.

Figure 3.5. SE SEM images of the cross-sections of Col-HA composite coatings on

Ti-6Al-4V substrates prepared using GFIB (A1, A2, A3) and PFIB (B1, B2, B3). The

cross-sections were polished using Rec pattern (A1, B1), CCS pattern (A2, B2), and

RCS pattern (A3, B3). White arrows in B1, B2, and B3 indicate the cracks between Pt

layer and Col-HA coatings.

Figure 3.6. SE SEM images of the cross-sections of Col-HA (A) and HA (B) coatings

on the Ti-6Al-4V substrates.

Figure 3.7. TEM sample preparation sequence from Col-HA coated Ti-6Al-4V alloy

using GFIB (A-K). The TEM sample was confirmed to be electron transparent by

STEM in GFIB (L).

Figure 3.8. TEM sample preparation sequence of Col-HA coated Ti-6Al-4V alloy

using PFIB (A-K). Col-HA coated Ti-6Al-4V sample was mounted on TEM grid

holders for preparing a TEM sample using PFIB (L).

Figure 3.9. Characterization of the TEM specimens prepared by GFIB: BFTEM (A)

and HAADF STEM image (B). C-F: Compositional maps obtained from the region in

(B), showing the distribution of Ca (C), C (D), Ti (E), and Ga (F).

Figure 3.10. Characterization of the TEM specimens prepared by PFIB: BFTEM (A)

and HAADF STEM image (B). C-F: Compositional maps obtained from the region in

(B), showing the distribution of Ca (C), C (D), Ti (E), and Xe (F).

Page 16: Fabrication and Characterization of Collagen-Apatite

xii

Chapter 4

Figure 4.1. TEM images of unstained mineralized collagen fibrils with different PAA

concentrations: (A) 5.6 mg/mL, (B) 2 mg/mL (insert: SAED result), (C) 0.8 mg/mL

and (D) 0.4 mg/mL.

Figure 4.2. TEM image (A) and SEM image (B) of mineralized collagen fibrils

without PAA. SAED result for mineralized collagen fibrils without PAA (Insert, A).

White arrow in (B): HA minerals on the surface. The cracks on collagen fibrils were

induced by high voltage employed during SEM operation.

Figure 4.3. FESEM images of mineralized collagen fibril bundles with different PAA

concentrations (white arrows indicate D-banding): (A) 5.6 mg/mL, (B) 2 mg/mL, (C)

0.8 mg/mL (the inset is the enlarged image of the spot indicated by the white arrow),

and D) 0.4 mg/mL (white dotted arrow and white arrow heads show minerals on the

collagen surface), minerals aligned on the surface of collagen fibrils with an almost 67

nm spacing distance (white dotted arrow and thin arrows in inset image). The inset is

the enlarged image of the spot indicated by the white dotted arrow.

Figure 4.4. TEM images of collagen solution containing PAA-ACP nanoprecursor

droplets on copper grid. (A) Low magnification TEM image showing PAA-ACP

nanoprecursor droplets that were formed in the close vicinity of the collagen fibrils.

White arrow shows the collagen fibrils that have been removed. (B) & (C) High

magnification images of PAA-ACP nanoprecursor droplets. (C) One electron dense

Page 17: Fabrication and Characterization of Collagen-Apatite

xiii

PAA-ACP nanoprecursor droplet highlighted in the white circle consisting of many

amorphous calcium phosphate nanoparticles.

Figure 4.5. Schematic illustration of the intrafibrillar mineralization process induced

by PAA as a sequestration analog and TPP as a templating analog.

Chapter 5

Figure 5.1. TEM images (A1 and A2), FESEM images (B1 and B2) of Col-SA fibrils

with different SA/Col ratios: 3:1 (A1, B1), and 4:1 (A2, B2). ATR-FTIR spectra (C)

of Col-SA fibrils (Stars indicate the peaks for Si-O-Si, and triangle shows the peak for

Si-OH). Schematic illustration (D) of the process to form Col-SA fibrils, and the TEM

image shows Col-SA fibrils with a SA/Col ratio of 3:1.

Figure 5.2. TEM images of Col-PAH-SA fibrils with different SA/Col ratios: 1:1 (A),

4: 3 (B), and 2:1 (C). ATR-FTIR spectra (D) of Col-PAH-SA fibrils (Stars indicate the

peaks for Si-O-Si, triangle shows the peak for Si-OH, and arrow indicates amine

hydrogen bonding). Schematic illustration (E) of the process to form Col-PAH-SA

fibrils, and the TEM image shows Col-PAH-SA fibrils with a SA/Col ratio of 4:3.

Figure 5.3. TEM images (A1, B1, and C1) and FESEM images (A2, B2, and C2) of

Col-inSA fibrils with different SA/Col ratios: 1:1 (A1, A2), 4:3 (B1, B2), and 2:1 (C1,

C2).

Figure 5.4. ATR-FTIR spectra (A) and TGA spectra (B) of Col-inSA fibrils with

different SA/Col ratios. (Stars indicate the peaks for Si-O-Si, triangle shows the peak

for Si-OH, and arrow indicates amine hydrogen bonding.)

Figure 5.5. The initial stage of PAH-SA precursors interacted with collagen fibrils in

Page 18: Fabrication and Characterization of Collagen-Apatite

xiv

the presence of TPP as a regulator. Low (A) and high (B) magnification of the fluidic

silica precursors infiltrated into the collagen fibrils during the collagen self-assembly

process. White arrows in B: Vague D-banding can be observed.

Figure 5.6. Schematic illustration of the intrafibrillar mineralization process of

Col-inSA fibrils. TEM image shows Col-inSA fibrils with a SA/Col ratio of 4:3.

Figure 5.7. Col-SA fibrils produced using either a two-step process (T-Col-SA (A1),

T-Col-PAH-SA (B1), and T-Col-inSA (C1)) or a one-step approach (Col-SA (A2),

Col-PAH-SA (B2), and Col-inSA (C2)).

Figure 5.8. Proliferation of MC3T3-E1 cells on mineralized collagen membranes:

Collagen-apatite (Col-Ap, control), Col-SA (SA/Col=4:3), Col-PAH-SA

(SA/Col=4:3), and Col-inSA (SA/Col=4:3) over the course of 14 days. Data is

presented as mean ± standard error. Statistically significant differences (p) between

various groups were calculated using two-way RM ANOVA analysis of variance, and

*: p<0.05, **: p<0.01, ***: p<0.001.

Chapter 6

Figure 6.1. TEM images of a small piece of IE-Col-Ap (A) and E-Col-Ap (B)

hydrogel fabricated using the two-temperature process. Inset images in A and B:

SAED results. The mineralized collagen fibrils were unstained.

Figure 6.2. SEM images of IE-Col-Ap (A, A1, A2) and E-Col-Ap (B, B1, B2)

scaffolds. Double headed arrows in A2 and B2 indicate the freezing direction of

self-compressed hydrogels. Black circle in B2 shows the randomly dispersed minerals

on the surface.

Page 19: Fabrication and Characterization of Collagen-Apatite

xv

Figure 6.3. Freeze-dried IE-Col-Ap (A) and E-Col-Ap (B) scaffolds were ground into

small pieces in liquid nitrogen, and one of these small pieces were observed using

TEM. Inset images in A and B: SAED results. White arrows in A and B indicate the

extrafibrillar minerals.

Figure 6.4. XRD spectra of IE-Col-Ap (A) and E-Col-Ap (B) scaffolds.

Figure 6.5. FT-IR spectra of IE-Col-Ap (A) and E-Col-Ap (B) scaffolds.

Figure 6.6. TGA of IE-Col-Ap (A) and E-Col-Ap (B) scaffolds.

Figure 6.7. Proliferation of MC3T3-E1 cells on collagen, E-Col-Ap, and IE-Col-Ap

scaffolds over the course of 7 days. Data is represented as the mean ± standard error.

# denotes statistical significance (p < 0.05) between collagen scaffolds and

experimental scaffolds.

Chapter 7

Figure 7.1. TEM images of a small piece of hydrogel composed of intrafibrillar apatite (A),

silica (B), and apatite/silica (C) mineralized collagen fibrils. The mineralized collagen fibrils

were unstained.

Page 20: Fabrication and Characterization of Collagen-Apatite

1

1. Introduction

1.1 Background

In the United States, approximately one million bone fractures are treated every

year. [1, 2, 3] Various materials have been used, such as autografts, allografts, and

artificial grafts, including metals, polymers and ceramics. Autografts have long been

considered as the gold standard for bone grafting procedures. However, the amount of

bone that can be harvested is limited, and there is a high potential for donor site

morbidity. Allografts are alternatives to autografts, but there are concerns of disease

transmission and immuno-rejection. [4, 5, 6, 7] Artificial grafts can overcome the

resource problem of autografts and disease transmission issues of allografts; however,

many bone grafts are not bioactive. A promising strategy to address these problems is

bone tissue engineering, where a combination of supporting scaffolds, cells and/or

biological signals is used to regenerate defected bone. [4, 8, 9]

Supporting scaffolds play an important role in bone tissue engineering, and the

ideal supporting scaffolds for bone regeneration should have the following properties:

(1) biocompatibility, that does not induce immunogenic responses; (2)

biodegradability, that degrades at a suitable rate to match the regeneration rate of new

tissue; (3) osteoconductive, that facilitates the cell attachment and proliferation; (4)

osteoinductive, that stimulates the osteoprogenitor cells to differentiate into osteoblast

cells and induces new bone formation. [1]

Based on these criteria, bioinspired or biomimetic scaffolds have been prepared,

which have demonstrated great promises and attracted great attention from the

Page 21: Fabrication and Characterization of Collagen-Apatite

2

biomaterial community. [10, 11] Polymer scaffolds or metal implants were coated

with hydroxyapatite, the main inorganic composition of bone, to render these

materials with excellent bioactivity. [12, 13] Pek et al. produced the collagen-apatite

nanocomposite foams by mixing collagen fibers with apatite nanocrystals, mimicking

the composition of natural bone. [14] Xia et al. developed a biomimetic

collagen-apatite scaffold with multi-lamellar structure in order to mimic the

composition and structure of natural bone. [15]

1.2 Structures and composition of natural bone

Natural bone is a hierarchical composite mainly composed of apatite and collagen.

Type I collagen accounts for 95% of the organic material in bone, and has good

resorbability and high affinity to cells. [16, 17] Apatite constitutes about 65 wt% of

natural bone, exhibiting biocompatibility, osteoconductivity, and bone-bonding ability.

[3, 18] Additionally, silicon has been found in active bone growth area, such as the

osteoid of the bone of young rats, [19] indicating osteoinductive property of silicon

for new bone formation and stimulative for neovascularization without supplements

of growth factors. [20, 21]

Page 22: Fabrication and Characterization of Collagen-Apatite

3

Figure 1.1. Hierarchical structures of natural bone: starting from sub-nanostructure of

collagen molecules and tropocollagen triple helix, to nanostructure of ordered HA

crystals, and to sub-microstructure of collagen fibrils and macrostructure (a). The

arrangement of HA crystals with collagen fibrils (b). [22]

As shown in Figure 1.1, natural bone is a complex hierarchical composite having

multi-scale microstructure. Apatite nanocrystals deposit within the gap zone of

collagen fibrils (intrafibrillar mineralization) at an early stage of mineralization and

then on the surface of the collagen fibrils (extrafibrillar mineralization) at a later stage.

[23, 24, 25, 26, 27, 28] Bundles and arrays of mineralized collagen fibrils further

arrange into a multilayered structure rotating across concentric lamellae, which pack

densely into compact bone or loosely into spongy bone. [29, 30, 31]

1.3 Intrafibrillar mineralization

In order to mimic the composition and structure of natural bone, extensive

researches have been done to replicate the intrafibrillar mineralization process in

nature. [32, 33, 34] Non-collagenous proteins (NCPs) are found to play an important

Page 23: Fabrication and Characterization of Collagen-Apatite

4

role in regulating the intrafibrillar mineralization of collagen fibrils. [35, 36] Their

functions are to sequester calcium and phosphate ions to form liquid amorphous

calcium phosphate nanoprecursors, and then template the nucleation and growth of

the nanocrystals at the gap zone of the collagen fibrils. [35, 37, 38] According to

Kerschnitzki et al., abundant vesicles containing small mineral particles were found in

the lumen of blood vessels, in the bone-forming tissue between the blood vessels and

the forming bone surfaces, and in the cells associated with bone formation, proving

the important role of nanoprecursors in bone formation. [39, 40] As NCPs play an

important role in regulating intrafibrillar mineralization, polyanionic polymers such as

PAMAM-COOH dendrimer, poly (aspartic acid) and poly (acrylic acid) have been

used as the sequestration analog of the NCPs. [32, 41, 42] Liu et al. employed poly

(acrylic acid) as the sequestration analog and sodium tripolyphosphate as the

templating analog to induce collagen intrafibrillar mineralization, and proved that the

intrafibrillar mineralized collagen fibers have improved nanomechanics and

cytocompatibility. [43] Based on the success of apatite intrafibrillar mineralization,

silica was also deposited within the gap zone of collagen fibrils using poly (allylamine

hydrochloride) or choline chloride as the analogs of zwitterionic proteins. [44] Niu et

al. produced intrafibrillar silicified collagen scaffolds by incubating collagen scaffolds

in the choline-stabilized silica precursor, and the intrafibrillar silicified collagen

scaffolds promoted the osteocalcin expression and calcium deposition after

transplantation in vivo. [45]

Page 24: Fabrication and Characterization of Collagen-Apatite

5

1.4 Implant coating

Titanium and its alloy (Ti-6Al-4V) have been widely used as orthopedic and

dental implants because of their good biocompatibility, good corrosion resistance and

mechanical strength. [46, 47] However, titanium implants do not have a good bone

bonding and integration ability, resulting in implant failures. [48] Thus, coating

hydroxyapatite, the main inorganic composition of natural bone, on titanium has been

widely applied to impart the bioactivity and osteoconductivity to titanium implants

and improve the bone-bonding ability of the implants. [49, 50] After implanting, a

bone-like apatite layer forms on the bioactive coating, acting as an intermediate layer

connecting new bone and implants. [51] A lot of approaches have been used to form a

hydroxyapatite coating on the titanium implant, such as electrodeposition, [52] cold

spray, [53] electrostatic spraying, [54] plasma spray, [55] sol-gel, [56] pulse laser

deposition [57] and biomimetic approach. [58] The biomimetic process closely

mimics apatite formation in vivo, which significantly improves the biological

performance of the biomaterial. This coating technique has attracted extensive

attention because it is conducted at a condition close to body temperature (37 °C) and

pH (7.4), and it is able to co-precipitate with biologically active molecules without

compromising their bioactivities. [59]

1.5 Dual beam focused ion beam/scanning electron microscopy (FIB/SEM)

Focused ion beam (FIB) was developed in the middle 1970’s mainly for

semiconductor applications. The basic principal is to sputter atoms from the materials

using accelerated heavy ions. [60] FIB is able to mill into the bulk of a specimen,

Page 25: Fabrication and Characterization of Collagen-Apatite

6

exposing the hidden internal microstructure without causing significant damages. As a

result, it has been used to prepare TEM samples, cross-sections, and metal

depositions.

Dual-beam focused ion beam/scanning electron microscopy (FIB/SEM) was

introduced to include an FIB column for material etching and an SEM column for

imaging to facilitate simultaneous milling, imaging, serial sectioning, and many other

functionalities (Figure 1.2). [61] Both ion beam and electron beam are focused at the

same point, enabling the electron column to monitor the process while the ion beam is

millings the specimen. The dual-beam FIB/SEMs not only have all the capabilities as

single beam instruments, but also allow a more precise cross-sectioning of

heterogeneous, fragile, and/or porous materials and structures, 3-D imaging and

preparation of TEM specimens.

The majority of FIB/SEMs are equipped with a Ga liquid metal ion source on the

FIB column (GFIB); however, they are limited by the low material removal rate. [62]

This problem has been addressed by the introduction of a new generation of

dual-beam FIB/SEM, which are equipped with plasma ion source on the FIB column

(PFIB). [63, 64] The PFIB offers milling rates up to two orders of magnitude higher

than those of GFIB. [65, 66]

Page 26: Fabrication and Characterization of Collagen-Apatite

7

Figure 1.2. The specimen is tilted towards the ion beam (left), and a cross-section is

milled to be imaged by the electron beam (right). [67]

Page 27: Fabrication and Characterization of Collagen-Apatite

8

References

[1] Z. Xia, M. Wei. Biomimetic fabrication of collagen-apatite scaffolds for bone

tissue regeneration. Journal of Biomaterials and Tissue Engineering. 2013;

369-384.

[2] L. Meinel, V. Karageorgiou, R. Fajardo, B. Snyder, V. Shinde-Patil, L. Zichner,

D. Kaplan, R. Langer, G. Vunjak-Novakovic. Bone tissue engineering using

human mesenchymal stem cells: effects of scaffold material and medium flow.

Annals of Biomedical Engineering. 2004; 32:112-122.

[3] A. A. Al‐Munajjed, N. A. Plunkett, J. P. Gleeson, T. Weber, C. Jungreuthmayer,

T. Levingstone, J. Hammer, F. J. O'Brien. Development of a biomimetic

collagen‐hydroxyapatite scaffold for bone tissue engineering using a SBF

immersion technique. Journal of Biomedical Materials Research Part B: Applied

Biomaterials. 2009; 90:584-591.

[4] J. R. Porter, T. T. Ruckh, K. C. Popat. Bone tissue engineering: a review in bone

biomimetics and drug delivery strategies. Biotechnology Progress. 2009;

25:1539-1560.

[5] C. Szpalski, M. Barbaro, F. Sagebin, S. M. Warren. Bone tissue engineering:

current strategies and techniques—part II: cell types. Tissue Engineering Part B:

Reviews. 2012; 18:258-269.

[6] J. O. Smith, A. Aarvold, E. R. Tayton, D. G. Dunlop, R. O. Oreffo. Skeletal tissue

regeneration: current approaches, challenges, and novel reconstructive strategies

Page 28: Fabrication and Characterization of Collagen-Apatite

9

for an aging population. Tissue Engineering Part B: Reviews. 2011; 17:307-320.

[7] S. Liao, F. Cui, W. Zhang, Q. Feng. Hierarchically biomimetic bone scaffold

materials: nano‐HA/collagen/PLA composite. Journal of Biomedical Materials

Research Part B: Applied Biomaterials. 2004; 69:158-165.

[8] L. H. Nguyen, N. Annabi, M. Nikkhah, H. Bae, L. L. Binan, S. Park, Y. Kang, Y.

Yang, A. Khademhosseini. Vascularized bone tissue engineering: approaches for

potential improvement. Tissue Engineering Part B: Reviews. 2012; 18:363-382.

[9] I. Marcos-Campos, D. Marolt, P. Petridis, S. Bhumiratana, D. Schmidt, G.

Vunjak-Novakovic. Bone scaffold architecture modulates the development of

mineralized bone matrix by human embryonic stem cells. Biomaterials. 2012;

33:8329-8342.

[10] X. Su, K. Sun, F. Cui, W. Landis. Organization of apatite crystals in human

woven bone. Bone. 2003; 32:150-162.

[11] T. Dvir, B. P. Timko, D. S. Kohane, R. Langer. Nanotechnological strategies for

engineering complex tissues. Nature Nanotechnology. 2011; 6:13-22.

[12] S. S. Kim, M. S. Park, S. J. Gwak, C. Y. Choi, B. S. Kim. Accelerated bonelike

apatite growth on porous polymer/ceramic composite scaffolds in vitro. Tissue

Engineering. 2006; 12:2997-3006.

[13] H. Qu, M. Wei. Improvement of bonding strength between biomimetic apatite

coating and substrate. Journal of Biomedical Materials Research Part B: Applied

Biomaterials. 2008; 84:436-443.

Page 29: Fabrication and Characterization of Collagen-Apatite

10

[14] Y. Pek, S. Gao, M. M. Arshad, K. J. Leck, J. Y. Ying. Porous collagen-apatite

nanocomposite foams as bone regeneration scaffolds. Biomaterials. 2008;

29:4300-4305.

[15] Z. Xia, M. Villa, M. Wei. A biomimetic collagen–apatite scaffold with a

multi-level lamellar structure for bone tissue engineering. Journal of Materials

Chemistry B. 2014; 2:1998-2007.

[16] Z. Xia, X. Yu, X. Jiang, H. D. Brody, D. W. Rowe, M. Wei. Fabrication and

characterization of biomimetic collagen–apatite scaffolds with tunable structures

for bone tissue engineering. Acta Biomaterialia. 2013; 9:7308-7319.

[17] J. M. Holzwarth, P. X. Ma. Biomimetic nanofibrous scaffolds for bone tissue

engineering. Biomaterials. 2011; 32:9622-9629.

[18] S. Yunoki, H. Sugiura, T. Ikoma, E. Kondo, K. Yasuda, J. Tanaka. Effects of

increased collagen-matrix density on the mechanical properties and in vivo

absorbability of hydroxyapatite–collagen composites as artificial bone materials.

Biomedical Materials. 2011; 6:015012.

[19] E. Zhang, C. Zou, S. Zeng. Preparation and characterization of silicon-substituted

hydroxyapatite coating by a biomimetic process on titanium substrate. Surface

and Coatings Technology. 2009; 203:1075-1080.

[20] A. J. Mieszawska, N. Fourligas, I. Georgakoudi, N. M. Ouhib, D. J. Belton, C. C.

Perry, D. L. Kaplan. Osteoinductive silk–silica composite biomaterials for bone

regeneration. Biomaterials. 2010; 2010:8902-8910.

Page 30: Fabrication and Characterization of Collagen-Apatite

11

[21] W. Zhai, H. Lu, L. Chen, X. Lin, Y. Huang, K. Dai, K. Naoki, G. Chen, J. Chang.

Silicate bioceramics induce angiogenesis during bone regeneration. Acta

Biomaterialia. 2012; 8:341-349.

[22] B. Basu. Natural Bone and Tooth: Structure and Properties. Biomaterials for

Musculoskeletal Regeneration. 2016; 45-85.

[23] X. Su, K. Sun, F. Cui, W. Landis. Organization of apatite crystals in human

woven bone. Bone. 2003; 32:150-162.

[24] L. Chen, R. Jacquet, E. Lowder, W. J. Landis. Refinement of collagen–mineral

interaction: A possible role for osteocalcin in apatite crystal nucleation, growth

and development. Bone. 2015; 71:7-16.

[25] Y. Wang, T. Azaïs, M. Robin, A. Vallée, C. Catania, P. Legriel, G.

Pehau-Arnaudet, F. Babonneau, M.-M. Giraud-Guille, N. Nassif. The

predominant role of collagen in the nucleation, growth, structure and orientation

of bone apatite. Nature materials. 2012; 11:724-733.

[26] X. Wang, F. Cui, J. Ge, Y. Wang. Hierarchical structural comparisons of bones

from wild-type and liliput dtc232 gene-mutated Zebrafish. Journal of structural

biology. 2004; 145:236-245.

[27] F. Nudelman, P. H. Bomans, A. George, N. A. Sommerdijk. The role of the

amorphous phase on the biomimetic mineralization of collagen. Faraday

Discussions. 2012; 159:357-370.

[28] F. Nudelman, K. Pieterse, A. George, P. H. Bomans, H. Friedrich, L. J. Brylka, P.

Page 31: Fabrication and Characterization of Collagen-Apatite

12

A. Hilbers, G. de With, N. A. Sommerdijk. The role of collagen in bone apatite

formation in the presence of hydroxyapatite nucleation inhibitors. Nature

Materials. 2010; 9:1004-1009.

[29] W. Landis, M. M. Song, A. Leith, L. McEwen, B. McEwen. Mineral and organic

matrix interaction in normally calcifying tendon visualized in three dimensions

by high-voltage electron microscopic tomography and graphic image

reconstruction. Journal of Structural Biology. 1993; 110:39-54.

[30] T. T. Thula, D. E. Rodriguez, M. H. Lee, L. Pendi, J. Podschun, L. B. Gower. In

vitro mineralization of dense collagen substrates: a biomimetic approach toward

the development of bone-graft materials. Acta Biomaterialia. 2011; 7:3158-3169.

[31] W. Wagermaier, H. Gupta, A. Gourrier, M. Burghammer, P. Roschger, P. Fratzl.

Spiral twisting of fiber orientation inside bone lamellae. Biointerphases. 2006;

1:1-5.

[32] J. Li, J. Yang, J. Li, L. Chen, K. Liang, W. Wu, X. Chen, J. Li. Bioinspired

intrafibrillar mineralization of human dentine by PAMAM dendrimer.

Biomaterials. 2013; 34:6738-6747.

[33] S. S. Jee, L. Culver, Y. Li, E. P. Douglas, L. B. Gower. Biomimetic

mineralization of collagen via an enzyme-aided PILP process. Journal of Crystal

Growth. 2010; 312:1249-1256.

[34] S. S. Jee, R. K. Kasinath, E. DiMasi, Y. Y. Kim, L. Gower. Oriented

hydroxyapatite in turkey tendon mineralized via the polymer-induced

Page 32: Fabrication and Characterization of Collagen-Apatite

13

liquid-precursor (PILP) process. CrystEngComm. 2011; 13:2077-2083.

[35] S. Gajjeraman, K. Narayanan, J. Hao, C. Qin, A. George. Matrix macromolecules

in hard tissues control the nucleation and hierarchical assembly of

hydroxyapatite. Journal of Biological Chemistry. 2007; 282:1193-1204.

[36] Y. K. Kim, L. Gu, T. E. Bryan, J. R. Kim, L. Chen, Y. Liu, J. C. Yoon, L. Breschi,

D. H. Pashley, F. R. Tay. Mineralisation of reconstituted collagen using

polyvinylphosphonic acid/polyacrylic acid templating matrix protein analogues

in the presence of calcium, phosphate and hydroxyl ions. Biomaterials. 2010;

31:6618-6627.

[37] G. He, S. Gajjeraman, D. Schultz, D. Cookson, C. Qin, W. T. Butler, J. Hao, A.

George. Spatially and temporally controlled biomineralization is facilitated by

interaction between self-assembled dentin matrix protein 1 and calcium

phosphate nuclei in solution. Biochemistry. 2005;44:16140-16148.

[38] G. He, T. Dahl, A. Vei, A. George. Nucleation of apatite crystals in vitro by

self-assembled dentin matrix protein 1. Nature Materials. 2003; 2:552-558.

[39] M. Kerschnitzki, A. Akiva, A. B. Shohamb, N. Koifman, E. Shimoni, K. Rechav,

A. A. Arraf, T. M. Schultheiss, Y. Talmon, E. Zelzer, S. Weiner, L. Addadi.

Transport of membrane-bound mineral particles in blood vessels during chicken

embryonic bone development. Bone. 2016; 83:65-72.

[40] M. Kerschnitzki, A. Akiva, A. B. Shohamb, Y. Asscher, W. Wagermaier, P. Fratzl,

L. Addad, S. Weiner. Bone mineralization pathways during the rapid growth of

Page 33: Fabrication and Characterization of Collagen-Apatite

14

embryonic chicken long bones. Journal of Structural Biology. 2016; 195:82-92.

[41] S. S. Jee, T. T. Thula, L. B. Gower. Development of bone-like composites via the

polymer-induced liquid-precursor (PILP) process. Part 1: Influence of polymer

molecular weight. Acta Biomaterialia. 2010; 6:3676-3686.

[42] Y. Liu, Y. K. Kim, L. Dai, N. Li, S. O. Khan, D. H. Pashley, F. R. Tay.

Hierarchical and non-hierarchical mineralisation of collagen. Biomaterials. 2011;

2011:1291-1300.

[43] Y. Liu, S. Liu, D. Luo, Z. Xue, X. Yang, L. Gu, Y. Zhou, T. Wang. Hierarchically

staggered nanostructure of mineralized collagen as a bone-grafting scaffold.

Advanced Materials. 2016; 28:8740-8748.

[44] L. Niu, K. Jiao, Y. Qi, C. K. Yiu, H. Ryou, D. D. Arola, J. Chen, L. Breschi, D. H.

Pashley, F. R. Tay. Infiltration of silica inside fibrillar collagen. Angewandte

Chemie. 2011; 123:11892-11895.

[45] L. Niu, J. Sun, Q. Li, K. Jiao, L. Shen, D. Wu, F. Tay, J. Chen.

Intrafibrillar-silicified collagen scaffolds enhance the osteogenic capacity of

human dental pulp stem cells. Journal of Dentistry. 2014; 42:839-849.

[46] S. B. Goodman, Z. Yao, M. Keeney, F. Yang. The future of biologic coatings for

orthopaedic implants. Biomaterials. 2013; 34:3174-3183.

[47] Z. Xia, X. Yu. Biomimetic collagen/apatite coating formation on Ti6Al4V

substrates. Journal of Biomedical Materials Research Part B: Applied

Biomaterials. 2012; 100:871-881.

Page 34: Fabrication and Characterization of Collagen-Apatite

15

[48] L. M. Bjursten, L. Rasmusson, S. Oh, G. C. Smith, K. S. Brammer, S. Jin.

Titanium dioxide nanotubes enhance bone bonding in vivo. Journal of

Biomedical Materials Research Part A. 2010; 92:1218-1224.

[49] E. Sandrini, C. Giordano, V. Busini, E. Signorelli, A. Cigada. Apatite formation

and cellular response of a novel bioactive titanium. Journal of Materials Science:

Materials in Medicine. 2007; 18:1225-1237.

[50] T. Kokubo, H. Kim, M. Kawashita, T. Nakamura. Bioactive metals: preparation

and properties. Journal of Materials Science: Materials in Medicine. 2004;

15:99-107.

[51] J. Forsgren, F. Svahn, T. Jarmar, H. Engqvist. Formation and adhesion of

biomimetic hydroxyapatite deposited on titanium substrates. Acta Biomaterialia.

2007; 3:980-984.

[52] S. H. Chen, W. L. Tsai, P. C. Chen, A. Fang, W. C. Say. Influence of applied

voltages on mechanical properties and in-vitro performance of electroplated

hydroxyapatite coatings on pure titanium. Journal of the Electrochemical Society.

2016; 163:D305-D308.

[53] E. Kergourlay, D. Grossin, N. Cinca, C. Josse, S. Dosta, G. Bertrand, I. Garcia, J.

M. Guilemany, C. Rey. First cold spraying of carbonated biomimetic

nanocrystalline apatite on Ti6Al4V: physical-chemical, microstructural, and

preliminary mechanical characterizations. Advanced Engineering Materials.

2016; 18:496-500.

Page 35: Fabrication and Characterization of Collagen-Apatite

16

[54] P. J. D. Manders, J. G. C. Wolke, J. A. Jansen. Bone response adjacent to calcium

phosphate electrostatic spray deposition coated implants: an experimental study

in goats. Clinical Oral Implants Research. 2006; 17:548-553.

[55] D. Chen, E. H. Jordan, M. Gell, M. Wei. Apatite formation on alkaline treated

dense tio2 coatings deposited using the solution precursor plasma spray process.

Acta Biomaterialia. 2008; 4:553-559.

[56] W. Zhang, W. Liu, Y. Liu, C. Wang. Tribological behaviors of single and dual

sol-gel ceramic films on Ti6Al4V. Ceramics International. 2009; 35:1513-1520.

[57] O. Blind, L. H. Klein, B. Dailey, L. Jordan. Characterization of hydroxyapatite

films obtained by pulsed laser deposition on ti and ti6al4v substrates. Dental

Materials. 2005; 21:1017-1024.

[58] H. Teng, C. Yang, J. Lin, Y. Huang, F. Lu. A simple method to functionalize the

surface of plasma electrolytic oxidation produced tio2 coatings for growing

hydroxyapatite. Electrochimica Acta. 2016; 193:216-224.

[59] X. Yu, H. Qu, D. A. Knecht, M. Wei. Incorporation of bovine serum albumin into

biomimetic coatings on titanium with high loading efficacy and its release

behavior. Journal of Materials Science: Materials in Medicine. 2009; 20:287-294.

[60] R. Wirth. Focused Ion Beam (FIB) combined with SEM and TEM: Advanced

analytical tools for studies of chemical composition, microstructure and crystal

structure in geomaterials on a nanometre scale. Chemical Geology. 2009;

261:217-229.

Page 36: Fabrication and Characterization of Collagen-Apatite

17

[61] A. J. Bushby, K. M. P'ng, R. D. Young, C. Pinali, C. Knupp, A. J. Quantock.

Imaging three-dimensional tissue architectures by focused ion beam scanning

electron microscopy. Nature Protocols. 2011; 6:845-858.

[62] R. Young, C. Rue, S. Randolph, C. Chandler, G. Franz, R. Schampers, A.

Klumpp, L. Kwakman. A comparison of xenon plasma FIB technology with

conventional gallium LMIS FIB: imaging, milling, and gas-assisted applications.

Microscopy and Microanalysis. 2011; 17:652-653.

[63] N. Smith, W. Skoczylas, S. Kellogg, D. Kinion, P. Tesch, O. Sutherland, A.

Aanesland, R. Boswell. High brightness inductively coupled plasma source for

high current focused ion beam applications. Journal of Vacuum Science &

Technology B. 2006; 24:2902-2906.

[64] N. S. Smith, J. A. Notte, A. V. Steele. Advances in source technology for focused

ion beam instruments. MRS Bulletin. 2014; 39:329-335.

[65] T. Burnett, R. Kelley, B. Winiarski, L. Contreras, M. Daly, A. Gholinia, M.

Burke, P. Withers. Large volume serial section tomography by Xe Plasma FIB

dual beam microscopy. Ultramicroscopy. 2016; 161:119-129.

[66] M. N. Kelly, K. Glowinski, N. T. Nuhfer, G. S. Rohrer. The five parameter grain

boundary character distribution of α-Ti determined from three-dimensional

orientation data. Acta Materialia. 2016; 111:22-30.

[67] D. Stokes, F. Morrissey, B. Lich. A new approach to studying biological and soft

materials using focused ion beam scanning electron microscopy (FIB SEM).

Page 37: Fabrication and Characterization of Collagen-Apatite

18

Journal of Physics: Conference Series. 2006; 26:50.

Page 38: Fabrication and Characterization of Collagen-Apatite

19

2. Cross-sectional microstructure study of apatite coating on titanium

alloy

Abstract

Biomimetic hydroxyapatite (HA) coating has been used to render titanium

implants with excellent osteoconductivity and osseointegration in orthopedic and

dental applications. However, to date it has not been possible to reveal the fine details

of the coating structure or the coating/substrate interface. In this study, a crack-free

biomimetic HA coating has been successfully formed on sand-blasted, acid-etched

Ti-6Al-4V discs by simply immersing the discs in a modified simulated body fluid

(m-SBF). Dual-beam FIB/SEMs with either gallium ion source (GFIB) or xenon

plasma ion source (PFIB) were used to prepare the cross-sections and thin

transmission electron microscopy (TEM) specimens of the coating, and the

capabilities of GFIB and PFIB were compared. Both the cross-sectional SEM and

TEM images confirmed that a well bonded biomimetic HA coating has been formed

on the Ti-6Al-4V substrate. The coating exhibits a gradient structure with a dense

microstructure adjacent to the titanium substrate to facilitate strong bonding, while

there is a porous structure at the surface, which is beneficial to bone cell attachment

and subsequent new bone integration.

Keywords: Biomimetic HA coating; Cross-section; TEM specimen; Dual-beam

FIB/SEM; Gradient structure.

Page 39: Fabrication and Characterization of Collagen-Apatite

20

2.1 Introduction

The titanium alloy Ti-6Al-4V (Ti – 6% Al – 4% V (wt.%)) has been used

extensively in orthopedic and dental implants because of its good biocompatibility,

superior mechanical properties, and good corrosion resistance. [1, 2] However,

Ti-6Al-4V implants do not have a good bone-integration ability. [3] Thus, it has

become common practice to coat the Ti-6Al-4V implant with bioactive

hydroxyapatite (HA) to impart good bioactivity, to induce osteoconductivity, and to

promote osseointegration. [4] A variety of different deposition techniques has been

used to produce HA coatings on titanium alloy substrates. These include: plasma

spraying, sol-gel coating, electrophoretic deposition, and biomimetic immersion. [5-9]

The biomimetic coating approach has attracted extensive attention due to its mild

conditions (low temperature and neutral pH), ability to produce coatings with good

bonding strength, and capability to incorporate bioactive molecules during the coating

process. [2, 10-15] The surface morphology of the biomimetic coating has been

studied extensively using scanning electron microscopy (SEM). [16-18] However,

there are very few publications that describe the fine details of the internal

microstructure for these biomimetic coatings, or of the interface between the coating

and the alloy substrate.

Cross-sectional studies of the coating microstructure and the coating/substrate

interface are essential for investigations of crack formation, coating dissolution, ion

diffusion and distribution, and interfacial bonding. [19] For SEM studies the samples

are typically produced by fracturing or by cutting, grinding and polishing the coated

Page 40: Fabrication and Characterization of Collagen-Apatite

21

substrate. Both approaches have severe limitations. Fracture surfaces often have

complex topographies that complicate the interpretation of structural detail, and there

can be problems associated with plastic deformation in the substrate leading to

interfacial delamination. Cutting, grinding and polishing to produce

metallographic-type cross sections often induces significant mechanical damage and

limits the useful detail that can be extracted from such samples. [20-22] More serious

complications arise if one attempts to study such coatings by transmission electron

microscopy (TEM). These studies require extremely thin (typically <100nm) samples

free of mechanical damage or other artefacts. Traditional approaches for TEM

specimen preparation such as ultramicrotomy, or dimple grinding followed by ion

beam milling, are simply not feasible for HA-coated Ti alloy substrates. The

mechanical stresses introduced by these approaches would cause significant structural

damage both to the coating and to the coating/substrate interface. [23-25]

Focused ion beam (FIB) microscopy is a well-established technique for both

imaging and modifying samples on the micro- or nano-scale. The ability of FIB to

remove material from specific locations without inducing significant ion beam

damage makes it an ideal approach to prepare cross-sections of various materials and,

in particular, to produce site-specific cross-sectional TEM specimens. [26] Most

modern instruments include both an FIB column and an SEM column to facilitate

simultaneous milling and imaging, serial section tomography and many other

functionalities. [23] These dual-beam FIB/SEMs are capable of precise

cross-sectioning of heterogeneous, fragile, and/or porous materials and structures, and

Page 41: Fabrication and Characterization of Collagen-Apatite

22

preparation of TEM specimens. The vast majority of FIB/SEMs are equipped with a

Ga liquid metal ion source on the FIB column. While the latest generation of Ga FIB

(GFIB) columns is capable of operating over a wide range of probe sizes and beam

currents, there are restrictions on the material removal rate that can be achieved; this

limits the volume that can be interrogated in a reasonable milling time. [27] This

problem has been addressed by the introduction of a new generation of dual-beam

FIB/SEMs which utilize an inductively coupled plasma ion source with a heavy inert

gas feed (typically Xe) on the ion column. [28, 29] These plasma FIB (PFIB) columns

offer milling rates up to two orders of magnitude higher than those that can be

achieved using GFIB. [30, 31] They also offer advantages for materials that undergo

undesirable chemical interactions with the Ga ions during GFIB milling. However, to

date there have been very few studies that consider the relative merits of the GFIB

and PFIB approaches for specific complex materials systems.

Here we present a study of the microstructural details in a crack-free

biomimetic HA coating deposited onto a Ti-6Al-4V substrate by immersing it in a

modified simulated body fluid (m-SBF) at body temperature and pH. Conventional

microstructural studies on such coatings have been reported previously. [22] In the

current study we have concentrated on the use of dual-beam FIB/SEM techniques to

obtain high quality cross-sectional specimens for SEM and TEM experiments. These

specimens were prepared using both GFIB and PFIB instruments to reveal the relative

advantages of these approaches for such coated samples.

Page 42: Fabrication and Characterization of Collagen-Apatite

23

2.2 Materials and methods

2.2.1 Biomimetic HA coating preparation

Ti-6Al-4V discs with a thickness of 2 mm and a diameter of 15 mm were surface

treated as described previously. [2, 22] The Ti-6Al-4V discs were sandblasted with 46

grit SiC, followed by acid etching with a mixed solution of HCl and HNO3 and then

with a hot acid mixture of HCl and H2SO4. The acid-treated discs were then exposed

to UV irradiation for 8 h. Biomimetic HA coating was produced by immersing the

surface-treated Ti-6Al-4V in m-SBF. The ion concentrations of m-SBF and human

blood plasma are listed in Table 2.1. The initial pH of the m-SBF solution was

adjusted to 6.4, 6.5, 6.6, and 6.7 using HCl and HEPES

(4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid), and the coatings formed at

different initial pH values were labeled as HA-6.4, HA-6.5, HA-6.6, and HA-6.7,

respectively. The coating formation was carried out by soaking the Ti-6Al-4V discs in

the m-SBF at a temperature of 37 °C for 24 h. The HA coated Ti-6Al-4V discs were

then washed gently with deionized water and air-dried overnight.

Page 43: Fabrication and Characterization of Collagen-Apatite

24

Table 2.1 Ion concentrations (mM) of human blood plasma and m-SBF.

Ion Human blood plasma (mM) m-SBF (mM)

Na+ 142.0 109.9

K+ 5.0 6.0

Mg2+ 1.5 1.5

Ca2+ 2.5 7.9

Cl- 103.0 111.2

HCO3- 27.0 17.5

HPO42- 1.0 3.0

SO42- 0.5 0

2.2.2 Microstructural characterization

The phase(s) present in the coatings were evaluated by X-ray diffraction (XRD)

using a Bruker AXS D2 Phaser diffractometer with CuKα radiation. Data were

acquired at 2θ values of 5° to 60° using a step size of 0.02° and a scan rate of

0.5 °/min.

Observations of surface morphology and FIB sectioning were performed in two

FEI Helios dual-beam FIB/SEMs: a Nanolab 460F1 GFIB and a Xenon PFIB in the

UConn/Thermo Fisher Scientific Center for Advanced Microscopy and Materials

Analysis (CAMMA). Both instruments are equipped with EasyLift nano-manipulators,

and the GFIB also has a flip-stage and a scanning transmission electron microscopy

(STEM) detector for improved final thinning of TEM specimens. To reduce charging

effects during SEM imaging and FIB milling, a thin conducting layer of Au was

sputtered onto the HA coating surfaces using a Polaron E5100 coating unit.

Secondary electron (SE) SEM images of the HA coating surface morphology

Page 44: Fabrication and Characterization of Collagen-Apatite

25

were obtained using the Everhart-Thornley and in-lens detectors on the electron

columns. All sectioning was performed with the coating surface perpendicular to the

optic axis of the ion column (at 52˚ to the electron column). Prior to cross-sectional

milling, a 3-µm-thick Pt layer was deposited onto the chosen region of the HA coating

surface to protect the coating from ion beam damage. This Pt deposition was

performed in two steps. Firstly, a 1-µm-thick Pt layer was deposited using a 5 kV

electron beam (E-beam) to crack the organometallic Pt precursor. A further

2-µm-thick Pt layer was then deposited using the ion beam (I-beam). [32] The ion

column accelerating voltages used in this process were 30 kV for the GFIB and 8 kV

for the PFIB.

For cross-sectional SEM studies a stepped, wedge-shape was first milled out

using a regular cross-section (RCS) pattern, so that the microstructure could be

observed on the vertical wedge wall using the electron column. The RCS pattern is a

multi-pass scan method used to mill the initial wedge. In this approach, the beam is

scanned through the entire pattern and then repeats for a set number of passes. For the

GFIB sectioning study, the wedge was 20 µm long, 10 µm wide, and 5 µm deep,

giving a wedge angle of ≈ 45˚. In the PFIB study, larger wedges with the same angle

were produced. In these cases, the dimensions were 60 µm long, 12 µm wide and 6

µm deep. After the wedge was milled, the vertical wedge wall was polished using an

ion beam at a reduced beam current to eliminate curtaining and other milling artefacts.

Two different pre-set milling patterns were used for the process, namely cleaning

cross-section (CCS) pattern, and rectangle (Rec) pattern. The CCS pattern is

Page 45: Fabrication and Characterization of Collagen-Apatite

26

processed line by line with each line containing a set number of passes, and the Rec

pattern repeatedly scans the beam over a rectangular array of equally spaced

positions.

For TEM studies, specimens were produced in a similar manner, but a second

wedge was milled on the other side of the protective Pt strap so that the vertical

wedge walls defined a pre-thinned FIB lamella. The lamella was then attached to the

EasyLift W probe, cut free from the surrounding material and transferred to a Cu

omni-grid. Final thinning of the lamellae to electron transparency was performed on

the grid using the ion beam with iteratively decreased ion currents to minimize any

residual ion damage.

The TEM specimens were examined using an FEI Talos F200X S/TEM operating

at an accelerating voltage of 200 kV. This instrument is equipped with a Super-X

silicon drift detector (SDD) energy dispersive X-ray spectrometry (EDXS) system,

which allows for a rapid acquisition of spectrum images for elemental mapping.

2.3 Results and discussion

Figure 2.1 is a series of SE SEM images showing the surface morphologies of the

HA coatings produced by immersing surface-treated Ti-6Al-4V substrates in an

m-SBF solution at different initial pH values for 24 h. When the initial pH was 6.4, a

thin, crack-free dense mineral coating was formed (HA-6.4, Figure 2.1A). As the

coating is very thin, the morphology of the underlying Ti-6Al-4V substrate can be

observed clearly. When the initial pH was 6.5, a smooth coating with a few minor

cracks due to dehydration shrinkage was formed, as indicated by the black arrows in

Page 46: Fabrication and Characterization of Collagen-Apatite

27

Figure 2.1B (HA-6.5). [33] A crack-free coating was formed when the initial pH was

6.6 (HA-6.6, Figure 2.1C). In contrast, the coating produced at an initial pH of 6.7

exhibits mudflat cracking with globules of HA scattered on the surface (HA-6.7,

Figure 2.1D). It has been observed previously that the HA-6.5, HA-6.6, and HA-6.7

coatings exhibit a porous and rough coating surface, which has been discovered to be

beneficial to cell attachment and the initiation of bone formation and integration. [34]

Figure 2.1. SE SEM images of HA coatings formed at different initial pH values:

HA-6.4 (A), HA-6.5 (B), HA-6.6 (C), and HA-6.7 (D). The black arrows in (B)

indicate cracks formed on the surface.

Examples of the corresponding XRD patterns are shown in Figure 2.2. These data

indicate that an HA coating can be formed at all initial pH values selected. However,

very weak HA peaks together with strong peaks from the hexagonal-close-packed

(HCP) α phase of the Ti-6Al-4V substrate were observed for the HA-6.4 coating. This

Page 47: Fabrication and Characterization of Collagen-Apatite

28

indicates that a very thin layer of HA coating was formed on the Ti-6Al-4V substrate,

which is in agreement with the SEM observation in Figure 2.1A. Strong peaks for HA

at 2θ=26° ((002) plane) and 31-34° (overlap of (211), (112), and (300) planes) were

observed for HA-6.5, HA-6.6, and HA-6.7 coatings, indicating that a thick HA

coating has been formed on the Ti-6Al-4V substrate. However, the broad diffraction

peak around 31-34° suggests that the HA coatings were not fully crystalline. A

preferred c-axis orientation of the HA crystals can be inferred from the high intensity

of the (002) peak compared to the (112) peak. [35] Diffraction peaks from the HCP α

phase of the Ti-6Al-4V substrate were present in all of the XRD patterns.

Figure 2.2. XRD spectra of HA coatings on Ti-6Al-4V substrates after 24 h of

Page 48: Fabrication and Characterization of Collagen-Apatite

29

immersion in m-SBF.

Based on the SEM and XRD data, it was found that the best-quality coatings

were HA-6.5 and HA-6.6 coatings. As such, these two coatings were selected for

more detailed cross-sectional investigation using FIB techniques. As shown in Figure

2.3, HA-6.5 and HA-6.6 coatings exhibit gradient structures: the HA coating adjacent

to the substrate was dense, and it became more and more porous towards the coating

surface. It is the first time that we clearly observed a gradient, porous cross-sectional

microstructure of the coating. When cross-sections produced are observed by the

conventional method (fracturing/grinding), only a dense structure was observed, but

the porous gradient structure was completely ruined by the sample preparation

process. [22, 36] Moreover, it has been shown that the pore size of HA-6.5 coating at

the outermost surface was larger than that of the HA-6.6 coating. Also, the HA-6.5

coating was thicker than HA-6.6 coating during the same immersion time (24 h),

which is in agreement with our previous observations.[12]

Figure 2.3. Cross-sectional SE SEM images of HA coatings on Ti-6Al-4V substrates:

HA-6.5 (A) and HA-6.6 (B). White arrow in image (A) is a crack induced by the ion

beam of PFIB and black arrow in image (A) is a crack between the HA-6.5 coating

Page 49: Fabrication and Characterization of Collagen-Apatite

30

and the Ti-6Al-4V substrate.

The formation of the HA coating can be expressed by Equations (1) and (2). As

discussed in our previous work, the pH of the m-SBF solution is determined by the

initial pH of the solution and the amount of HCO3- in the solution. The pH profile of

the m-SBF during the coating process has been investigated in details as reported

previously. [12, 22] Briefly, at the initial stage of coating formation, the pH of the

solution increases due to the decomposition of HCO3-, as shown in Equation (2).

Simultaneously, Ca2+ and HPO42- ions are adsorbed onto the substrate surface to form

HA nuclei, as expressed in Equation (1). Previous research reported that a high

concentration of HCO3- ions in m-SBF resulted in a dense apatite coating. [2, 22, 37]

Thus, in the initial stage of coating formation, a lower pH leads to slower coating

formation, resulting in a dense coating adjacent to the substrate. With the

decomposition of HCO3-, the increase in pH and the decrease in the HCO3

-

concentration leads to the formation of a porous HA coating toward the coating

surface. When the initial pH of m-SBF was 6.5, more HCO3- ions were needed to

bring up the pH than for an initial pH=6.6, thus less HCO3- ions remained,

contributing to the larger pore size observed toward the surface of the HA coating.

5Ca2++3HPO42-+4OH-→Ca5(PO4)3OH+3H2O (1)

HCO3-→CO2+OH- (2)

As shown in Figure 2.3A, there is a large crack between the HA-6.5 coating and

the Ti-6Al-4V substrate (black arrow in Figure 2.3A); this is consistent with the

cracks observed on the coating surface, as indicated in Figure 2.1B. Cracks at the

Page 50: Fabrication and Characterization of Collagen-Apatite

31

surface are due to the dehydration shrinkage, which significantly reduces the

interfacial bonding strength. [21] Thus, the cracks on the surface indicate weak

bonding between the HA-6.5 coating and the Ti-6Al-4V substrate. However, no

cracks were observed between the HA-6.6 coating and Ti-6Al-4V substrate, indicating

a strongly bonded HA coating was formed (Figure 2.3B). This might be because

HA-6.6 coating is thinner and denser compared to HA-6.5 coating. Because a

well-bonded HA coating formed at an initial pH=6.6, HA-6.6 coating system was

selected for further cross-sectional microscopy investigations.

In principle, cross-sections of the coating can provide critical information

regarding the interface as well as about the coating itself. However, due to the

significant differences between the stiffness and density of the coating and substrate,

it becomes extremely difficult to create the cross-sectional samples without damaging

the microstructure both at the interface and within the coating itself. FIB instruments

are powerful tools for studies of interfaces between materials with such different

properties, but there are certain drawbacks and instrumental artefacts. A common

artefact produced during FIB milling is the “curtaining” effect in which fluctuations in

the ion beam current, position and/or milling efficiency (due to sample

inhomogeneities) lead to undulations on the surface of an ostensibly planar FIB cut.

This effect can significantly degrade the resolution of images obtained from FIB-cut

cross-sections. Ideally, cross-sectional FIB cuts should be free of such curtaining

effects. After the initial wedge milling using the RCS pattern, the microstructure of

the cross-section was obscured by the curtaining (Figure 2.4). Then CCS and Rec

Page 51: Fabrication and Characterization of Collagen-Apatite

32

patterns were applied to obtain cross-sections with reduced curtaining in both GFIB

and PFIB, and the efficiencies of these two approaches were compared.

Figure 2.4. Gas-assisted deposition of Pt layer at 8 kV and 30 kV ion beam voltage in

PFIB.

Figure 2.5. SE SEM images of HA-6.6 coating on Ti-6Al-4V substrate from

cross-sections prepared using GFIB (A, B) and PFIB (C, D). The cross-sections were

polished using CCS pattern (A, C) or Rec pattern (B, D). Cracks indicated by white

Page 52: Fabrication and Characterization of Collagen-Apatite

33

arrows in HA coating in image (C) were induced by the CCS process in PFIB.

Figure 2.5 shows the cross-sectional images of HA-6.6 coated Ti-6Al-4V

substrates cleaned by Rec and CCS patterns in GFIB and PFIB. First, a 3-µm-thick Pt

layer was deposited onto the area of interest by gas-assisted Pt deposition to protect

the HA coating surface, allowing the features close to or at the surface to be analyzed

without damage. [38] In GFIB, an ion beam with accelerating voltage of 30 kV was

used to deposit the Pt-protective layer, however, in PFIB a dense Pt layer was

obtained at 8 kV. Increasing the ion beam accelerating voltage to 30 kV in PFIB led to

the formation of a porous Pt layer (Figure 2.4). In GFIB, wedges 20 µm in length, 10

µm in width and 5 µm in depth were milled using RCS, and then cleaned up using the

CCS and Rec patterns as shown in Figures 2.5A and 2.5B, respectively. In PFIB,

wedges 60 µm in length, 12 µm in width and 6 µm in depth were milled using RCS.

The images obtained from after cleaning up using CCS and Rec patterns are shown in

Figures 2.5C and 2.5D, respectively. Even though much larger wedges were milled

using PFIB, the total time for milling and the subsequent clean-up was less than those

used in GFIB. As shown in Figure 2.5, in GFIB, the Rec pattern proved to be more

suitable for the production of curtaining-free cross-sections than the CCS pattern at

the same ion beam current. However, the Rec pattern is as efficient as the CCS pattern

in creating a reasonable cross-section using PFIB. When CCS was applied to clean up

PFIB cross-sections, cracks were formed in the HA coating (white arrows in Figure

2.5C). Overall, the detailed microstructure of the HA coating and the interface

between the coating and the substrate have been well preserved by both GFIB and

Page 53: Fabrication and Characterization of Collagen-Apatite

34

PFIB, but GFIB produces a cross-section with higher quality than PFIB, which may

be due to the smaller ion beam spot size achievable in GFIB. [27]

Figure 2.6. An extraction sequence of a thin lamella from the bulk in GFIB (A-H).

The thickness of the TEM specimen was confirmed by STEM in GFIB (I).

Figure 2.7. The thin lamella from the site of interest before extraction (A) and after

mounted onto copper omni grids (B) in PFIB.

TEM experiments were performed to study the interface between the HA-6.6

coating and the Ti-6Al-4V substrate on the nano-scale. The TEM specimens were

Page 54: Fabrication and Characterization of Collagen-Apatite

35

prepared as thin lamellae from the sites of interest following the extraction sequence

presented in Figure 2.6. Firstly, a 3-µm-thick Pt strap was deposited onto the area of

interest (Figure 2.6A); this served as a protective layer for the HA coating during the

milling process. Wedge-shaped trenches were then milled out on either side of the Pt

strap (Figure 2.6B), defining a lamella that was then thinned further in preparation for

lift-out (Figures 2.6C and 2.7A). Due to the larger beam size and higher beam current

in the PFIB, some of the lamellae sustained significant damage during thinning (data

not shown), as indicated by the cracks in the HA coating in Figure 2.5C. As a result,

for PFIB a thicker lamella was prepared (~4 µm; Figure 2.7) than for GFIB (~900 nm,

Figures 2.6C and 2.6H). In both cases, the lamella was cut free at the bottom and on

one side, the W EasyLift probe was attached to the Pt strap by depositing additional Pt

at the point of contact, the lamella was cut free on the other side and was then lifted

out (Figure 2.6D). The lamella was then transferred to a copper omni grid (Figure

2.6E), attached to the grid at two corners by depositing Pt in these locations (Figure

2.6F), and then cut free from the W EasyLift probe (Figure 2.6G). The TEM lamella

was finally thinned to electron transparency using the ion beam with iteratively

decreasing ion currents. In GFIB, STEM imaging at an electron column accelerating

voltage of 30 kV was used to confirm that the specimen was thin enough for TEM

analysis (Figure 2.6I).

Page 55: Fabrication and Characterization of Collagen-Apatite

36

Figure 2.8. Data obtained from the TEM specimens prepared by GFIB: BF TEM

image (A) and HAADF STEM image (B). The inset to (A) is the SAED pattern from

the coating. C-F: Compositional maps obtained from the region in (B), showing the

distribution of Ca (C), P (D), Ti (E), and Ga (F).

Page 56: Fabrication and Characterization of Collagen-Apatite

37

Figure 2.9. Data obtained from the TEM specimens prepared by PFIB: BF TEM

image (A) and HAADF STEM image (B). C-F: Compositional maps obtained from

the region in (B), showing the distribution of Ca (C), P (D), Ti (E), and Xe (F).

Page 57: Fabrication and Characterization of Collagen-Apatite

38

The nano-scale microstructure of the HA coatings on Ti-6Al-4V was investigated

by S/TEM analysis of the TEM specimens prepared by both GFIB and PFIB, and

examples of the data are shown in Figures 2.8 and 2.9, respectively. In each case,

these include a bright field (BF) TEM image, a high angle annular dark field

(HAADF) STEM image, and compositional maps for Ca, P, Ti and the appropriate

milling ion (Ga and Xe, respectively). The maps were obtained from spectrum

imaging experiments in which full EDXS spectra were acquired at each pixel, and

then standard-less quantification was performed at each point. Thus, Figures 2.8C-F

and 2.9C-F are compositional maps (rather than X-ray maps) and the intensities in

these maps are proportional to the measured concentrations of the various species.

The BF TEM and HAADF STEM images showed that the thickness of the

coating was about 1.5-2.5 µm, and no structural defects were observed between the

HA coating and the Ti-6Al-4V substrate, indicating that a strong bond has been

formed at the interface. These images also reveal the details of the gradient structure

in the HA coating, which is consistent with the SEM observations in Figures 2.3 and

2.5. A dense coating is formed adjacent to the substrate to facilitate the strong bonding

between the coating and the substrate, and there is a highly porous surface that could

contribute to coating dissolution and cell attachment. [33] The preferential crystallite

alignment was also observed by STEM (Figure 2.8). Previous research reported that

the HA crystal dimensions decreased with increasing HCO3- ion concentration, and

that the crystals were more equi-axed and less needle-like at a higher concentrations.

[39] In the initial stage of coating formation, a smaller crystal size and denser coating

Page 58: Fabrication and Characterization of Collagen-Apatite

39

was formed adjacent to the substrate due to a higher HCO3- ion concentration and a

lower pH. [2, 22, 35, 39] As the immersion time increases, the HCO3- ions decompose

to bring up the pH, thus the dimensions of the HA crystal increased, and crystallites

elongated preferentially perpendicular to the surface, which is in accordance with the

XRD results (Figure 2.2). [35, 40] The corresponding mineral phase of the coating

was evaluated by selected area electron diffraction (SAED) coupled to TEM, and the

inset pattern in Figure 2.8A indicates that the HA coating is poly-crystalline, which

resembles the apatite structure in natural bone. [33] The compositional maps show

that the distributions of Ca and P in the coating are uniform, and thus there is no

change in composition associated with the structural gradient. These maps also reveal

that there is some implantation of Ga and Xe ion in the TEM specimens prepared by

GFIB and PFIB, respectively (Figures 2.8F and 2.9F). Such implantation is typical in

FIB analyses,[41] and does not appear to have caused any significant changes in the

structure of the coatings. We note, however, that the data from the TEM specimen

produced by PFIB (Figure 2.9) appears to be at slightly lower resolution and shows

somewhat more residual curtaining than the specimen produced by GFIB (Figure 2.8).

This is consistent with there being low-level damage and artefacts due to the coarser,

more powerful Xe ion beam in the PFIB.

2.4 Conclusions

A crack-free biomimetic HA coating was formed by immersing a surface-treated

Ti-6Al-4V substrate in an m-SBF solution at an initial pH of 6.6. Both GFIB and

PFIB were used to create curtaining-free cross-sections and TEM specimens of the

Page 59: Fabrication and Characterization of Collagen-Apatite

40

coating, and the capabilities of these two tools were compared. Although GFIB and

PFIB share a lot of similarities, there are also significant differences, including the

optimal voltage of ion-beam-induced Pt deposition, the TEM sample preparation

process, and the efficiency and the quality of the cross-sectional cuts produced.

Particularly, the GFIB is more beneficial for preparing the high-quality final finish of

the sample, while the PFIB efficient for creating a large cut of the sample. Also, both

SEM and TEM observations confirmed that a well-bonded HA coating has been

formed on the Ti-6Al-4V substrate. The coating demonstrates a unique gradient

structure, with a dense layer adjacent to the substrate to facilitate a strong bonding,

and a porous structure at the surface to allow for bone cell attachment.

Page 60: Fabrication and Characterization of Collagen-Apatite

41

References

[1] S. B. Goodman, Z. Yao, M. Keeney, F. Yang. The future of biologic coatings for

orthopaedic implants. Biomaterials. 2013; 34:3174-3183.

[2] Z. Xia, X. Yu. Biomimetic collagen/apatite coating formation on Ti6Al4V

substrates. Journal of Biomedical Materials Research Part B: Applied Biomaterials.

2012; 100:871-881.

[3] L. M. Bjursten, L. Rasmusson, S. Oh, G. C. Smith, K. S. Brammer, S. Jin.

Titanium dioxide nanotubes enhance bone bonding in vivo. Journal of Biomedical

Materials Research Part A. 2010; 92:1218-1224.

[4] E. Sandrini, C. Giordano, V. Busini, E. Signorelli, A. Cigada. Apatite formation

and cellular response of a novel bioactive titanium. Journal of Materials Science:

Materials in Medicine. 2007; 18:1225-1237.

[5] P. Yu, F. Lu, W. Zhu, D. Wang, X. Zhu, G. Tan, et al. Bio-inspired citrate

functionalized apatite coating on rapid prototyped titanium scaffold. Applied Surface

Science. 2014; 313:947-953.

[6] Y. Huang, Q. Ding, S. Han, Y. Yan, X. Pang. Characterisation, corrosion resistance

and in vitro bioactivity of manganese-doped hydroxyapatite films electrodeposited on

titanium. Journal of Materials Science: Materials in Medicine. 2013; 24:1853-1864.

[7] C. Aparicio, D. Rodriguez, F. Gil. Variation of roughness and adhesion strength of

deposited apatite layers on titanium dental implants. Materials Science and

Engineering: C. 2011; 31:320-324.

[8] D. Chen, E. H. Jordan, M. Gell, M. Wei. Apatite formation on alkaline-treated

Page 61: Fabrication and Characterization of Collagen-Apatite

42

dense TiO 2 coatings deposited using the solution precursor plasma spray process.

Acta Biomaterialia. 2008; 4:553-559.

[9] H. W. Kim, Y. H. Koh, L. H. Li, S. Lee, H. E. Kim. Hydroxyapatite coating on

titanium substrate with titania buffer layer processed by sol–gel method. Biomaterials.

2004; 25:2533-2538.

[10] L. N. Luong, K. M. McFalls, D. H. Kohn. Gene delivery via DNA incorporation

within a biomimetic apatite coating. Biomaterials. 2009; 30:6996-7004.

[11] G. Ciobanu, O. Ciobanu. Investigation on the effect of collagen and vitamins on

biomimetic hydroxyapatite coating formation on titanium surfaces. Materials Science

and Engineering: C. 2013; 33:1683-1688.

[12] H. Qu, M. Wei. The effect of temperature and initial pH on biomimetic apatite

coating. Journal of Biomedical Materials Research Part B: Applied Biomaterials. 2008;

87:204-212.

[13] X. Yu, H. Qu, D. A. Knecht, M. Wei. Incorporation of bovine serum albumin into

biomimetic coatings on titanium with high loading efficacy and its release behavior.

Journal of Materials Science: Materials in Medicine. 2009; 20:287-294.

[14] X. Yu, M. Wei. Preparation and evaluation of parathyroid hormone incorporated

CaP coating via a biomimetic method. Journal of Biomedical Materials Research Part

B: Applied Biomaterials. 2011; 97:345-354.

[15] X. Yu, L. Wang, X. Jiang, D. Rowe, M. Wei. Biomimetic CaP coating

incorporated with parathyroid hormone improves the osseointegration of titanium

implant. Journal of Materials Science: Materials in Medicine. 2012; 23:2177-2186.

Page 62: Fabrication and Characterization of Collagen-Apatite

43

[16] S. E. Bae, J. Choi, Y. K. Joung, K. Park, D. K. Han. Controlled release of bone

morphogenetic protein (BMP)-2 from nanocomplex incorporated on

hydroxyapatite-formed titanium surface. Journal of Controlled Release. 2012;

160:676-684.

[17] J. Faure, A. Balamurugan, H. Benhayoune, P. Torres, G. Balossier, J. Ferreira.

Morphological and chemical characterisation of biomimetic bone like apatite

formation on alkali treated Ti6Al4V titanium alloy. Materials Science and

Engineering: C. 2009; 29:1252-1257.

[18] A. W. Nijhuis, M. R. Nejadnik, F. Nudelman, X. F. Walboomers, J. te Riet, P.

Habibovic, et al. Enzymatic pH control for biomimetic deposition of calcium

phosphate coatings. Acta Biomaterialia. 2014; 10:931-939.

[19] S. Jalota, S. Bhaduri, S. B. Bhaduri, A. C. Tas. A protocol to develop crack-free

biomimetic coatings on Ti6Al4V substrates. Journal of Materials Research. 2007;

22:1593-1600.

[20] C. Auclair‐Daigle, M. Bureau, J. G. Legoux, L. H. Yahia. Bioactive

hydroxyapatite coatings on polymer composites for orthopedic implants. Journal of

Biomedical Materials Research Part A. 2005; 73:398-408.

[21] E. Zhang, C. Zou, S. Zeng. Preparation and characterization of silicon-substituted

hydroxyapatite coating by a biomimetic process on titanium substrate. Surface and

Coatings Technology. 2009; 203:1075-1080.

[22] H. Qu, M. Wei. Improvement of bonding strength between biomimetic apatite

coating and substrate. Journal of Biomedical Materials Research Part B: Applied

Page 63: Fabrication and Characterization of Collagen-Apatite

44

Biomaterials. 2008; 84:436-443.

[23] A. J. Bushby, K. M. P'ng, R. D. Young, C. Pinali, C. Knupp, A. J. Quantock.

Imaging three-dimensional tissue architectures by focused ion beam scanning electron

microscopy. Nature Protocol. 2011; 6:845-858.

[24] E. Coutinho, T. Jarmar, F. Svahn, A. Neves, B. Verlinden, B. Van Meerbeek, et al.

Ultrastructural characterization of tooth–biomaterial interfaces prepared with broad

and focused ion beams. Dental Materials. 2009; 25:1325-1337.

[25] M. Corazza, S. B. Simonsen, H. Gnaegi, K. T. Thydén, F. C. Krebs, S. A.

Gevorgyan. Comparison of ultramicrotomy and focused-ion-beam for the preparation

of TEM and STEM cross section of organic solar cells. Applied Surface Science. 2016;

389:462-468.

[26] M. Schaffer, B. Schaffer, Q. Ramasse. Sample preparation for atomic-resolution

STEM at low voltages by FIB. Ultramicroscopy. 2012; 114:62-71.

[27] R. Young, C. Rue, S. Randolph, C. Chandler, G. Franz, R. Schampers, et al. A

comparison of xenon plasma FIB technology with conventional gallium LMIS FIB:

imaging, milling, and gas-assisted applications. Microscopy and Microanalysis. 2011;

17:652-653.

[28] N. Smith, W. Skoczylas, S. Kellogg, D. Kinion, P. Tesch, O. Sutherland, et al.

High brightness inductively coupled plasma source for high current focused ion beam

applications. Journal of Vacuum Science & Technology B. 2006; 24:2902-2906.

[29] N. S. Smith, J. A. Notte, A. V. Steele. Advances in source technology for focused

ion beam instruments. MRS Bulletin. 2014; 39:329-335.

Page 64: Fabrication and Characterization of Collagen-Apatite

45

[30] T. Burnett, R. Kelley, B. Winiarski, L. Contreras, M. Daly, A. Gholinia, et al.

Large volume serial section tomography by Xe Plasma FIB dual beam microscopy.

Ultramicroscopy. 2016; 161:119-129.

[31] M. N. Kelly, K. Glowinski, N. T. Nuhfer, G. S. Rohrer. The five parameter grain

boundary character distribution of α-Ti determined from three-dimensional orientation

data. Acta Materialia. 2016; 111:22-30.

[32] H. Yu, Y. Sun, S. P. Alpay, M. Aindow. Solidification microstructures in Ag3Sn–

Cu3Sn pseudo-binary alloys. Journal of Materials Science. 2016; 51:6474-6487.

[33] E. Zhang, C. Zou, G. Yu. Surface microstructure and cell biocompatibility of

silicon-substituted hydroxyapatite coating on titanium substrate prepared by a

biomimetic process. Materials Science and Engineering: C. 2009; 29:298-305.

[34] Y. Huang, S. Han, X. Pang, Q. Ding, Y. Yan. Electrodeposition of porous

hydroxyapatite/calcium silicate composite coating on titanium for biomedical

applications. Applied Surface Science. 2013; 271:299-302.

[35] L. Müller, E. Conforto, D. Caillard, F. A. Müller. Biomimetic apatite

coatings—Carbonate substitution and preferred growth orientation. Biomolecular

Engineering. 2007; 24:462-466.

[36] D. O. Costa, B. A. Allo, R. Klassen, J. L. Hutter, S. J. Dixon, A. S. Rizkalla.

Control of surface topography in biomimetic calcium phosphate coatings. Langmuir.

2012; 28:3871-3880.

[37] F. Barrere, P. Layrolle, C. Van Blitterswijk, K. De Groot. Biomimetic calcium

phosphate coatings on Ti6Al4V: a crystal growth study of octacalcium phosphate and

Page 65: Fabrication and Characterization of Collagen-Apatite

46

inhibition by Mg 2+ and HCO 3−. Bone. 1999; 25:107S-111S.

[38] A. Thorfve, A. Palmquist, K. Grandfield. Three-dimensional analytical

techniques for evaluation of osseointegrated titanium implants. Materials Science and

Technology. 2014.

[39] N. Blumenthal, F. Betts, A. Posner. Effect of carbonate and biological

macromolecules on formation and properties of hydroxyapatite. Calcified Tissue

Research. 1975; 18:81-90.

[40] B. Feng, X. Chu, J. Chen, J. Wang, X. Lu, J. Weng. Hydroxyapatite coating on

titanium surface with titania nanotube layer and its bond strength to substrate. Journal

of Porous Materials. 2010; 17:453-458.

[41] R. Wirth. Focused Ion Beam (FIB) combined with SEM and TEM: Advanced

analytical tools for studies of chemical composition, microstructure and crystal

structure in geomaterials on a nanometre scale. Chemical Geology. 2009;

261:217-229.

Page 66: Fabrication and Characterization of Collagen-Apatite

47

3. Cross-sectional microstructure study of collagen-apatite composite

coating on titanium alloy

Abstract

A biomimetic bone-like collagen-hydroxyapatite (Col-HA) composite coating

was formed on a surface-treated Ti-6Al-4V alloy substrate via simultaneous collagen

self-assembly and hydroxyapatite nucleation. The coating process has been carried

out by immersing sand-blasted, acid-etched and UV irradiated Ti-6Al-4V alloy in type

I collagen-containing modified simulated body fluid (m-SBF). The surface

morphology and phase composition of the coating were characterized using scanning

electron microscopy (SEM), X-ray diffraction (XRD), and Fourier transformation

infrared spectroscopy (FTIR). More importantly, dual-beam FIB/SEMs with either

gallium ion source (GFIB) or xenon plasma ion source (PFIB) were used to

investigate the cross-sectional features of the biomimetic Col-HA composite coating

in great details. As a result, the cross-sectional images and thin transmission electron

microscopy (TEM) specimens were successfully obtained from the composite coating.

Both the cross-sectional SEM and TEM results confirmed that the Col-HA coating

had a similar microstructure to pure hydroxyapatite coating that consists of a gradient

coating at the cross-section, with a dense layer adjacent to the interface between the

titanium substrate and the coating, while a porous structure at the coating surface.

Keywords: Collagen-hydroxyapatite coating, Dual Beam FIB, TEM sample,

cross-section, microstructure

Page 67: Fabrication and Characterization of Collagen-Apatite

48

3.1 Introduction

Ti-6Al-4V (Ti-6%Al-4%V (wt.%)) alloy has been widely used as load-bearing

implants due to its excellent mechanical properties, corrosion resistance and

biocompatibility. [1, 2] However, there is still concern of the poor bone-implant

integration due to the formation of a fibrous tissue layer surrounding the Ti-6Al-4V

implants. [3] Thus, a layer of hydroxyapatite (HA) coating is usually deposited on the

surface of the Ti-6Al-4V implants to improve their bioactivity, osteoconductivity and

osteointegration. [4-6] HA, the main component of natural bone, exhibits excellent

biocompatibility, osteoconductivity and bone-bonding ability. [7] A variety of

deposition techniques, such as plasma spray process, hot isostatic pressing,

electro-deposition, sol-gel coating, and biomimetic method have been applied to

successfully deposit HA coatings on the Ti-6Al-4V implants. [5, 8-13] Collagen, the

most abundant protein in bone, is effective in accelerating cellular adhesion and

promoting cell proliferation. [14, 15] Thus, a composite coating comprised of both

collagen and HA will better mimic the composition of natural bone and thereby

improve the bone-implant integration.

Collagen/HA (Col-HA) composite coating has been deposited onto the surface of

Ti-6Al-4V substrates via electrolytic deposition, spin coating, and biomimetic

immersion over the past few years. [16-18] Due to the sensitivity of collagen, there

are limited methods available to deposit Col-HA composite coatings. For example,

plasma spraying and thermal spray, which are two widely used techniques to deposit

HA coatings, are impractical to deposit collagen composite coating due to the high

Page 68: Fabrication and Characterization of Collagen-Apatite

49

operation temperatures. [3] Biomimetic coating method, on the other hand, is a

promising method due to its low cost, easy operation, low temperature, and the ability

to incorporate biomolecules such as proteins and collagens. [7, 17, 19, 20] Thus,

Col-HA composite coatings have been successfully deposited on the Ti-6Al-4V

substrates. [7, 17, 18, 21]

The reliability of the coating and the bonding strength between coating and the

substrate are very important for the success of the implantation. [3] Because most of

the implant failures result from the debonding between the coating and the substrate,

as well as the dissolution of the coating. [22, 23] Cross-sectional studies of the

microstructure of the coating and the interface between the coating and the substrate

are essential to investigate the coating stability and bonding strength. [5, 24] However,

traditional techniques for producing cross-sections, such as coating fracturing or

coating cutting, grinding and polishing, suffer from losing microstructure details or

resulting in the debonding of the coating due to the plastic deformation. [5, 7, 25]

Meanwhile, focused ion beam (FIB) microscopy has been widely used as a

modern tool for structural microanalysis and micro-machining of materials. The most

important application of FIB is to prepare cross-sectional milling and transmission

electron microscopy (TEM) sample, since the microstructure of the sample is

maintained without damage. [26] The most advanced FIB instrument incorporates

both an FIB column and a scanning electron microscopy (SEM) column in a single

system, allowing simultaneous milling and imaging, serial sectioning and many other

functionalities. [27] The dual beam FIB/SEM is capable of preparing cross-sections

Page 69: Fabrication and Characterization of Collagen-Apatite

50

and TEM specimens made of heterogeneous, fragile, and porous materials and

structures. [5] The most widely used FIB/SEM is equipped with gallium liquid ion

source (GFIB), however, is limited with slow removal rate. [28] Thus, a dual beam

FIB/SEM system with plasma ion source (PFIB), which offers two orders of

magnitude higher milling rate than that of GFIB, was used in the current study. [29,

30] In our previous work, a comparison study has been conducted to investigate the

microstructure of the biomimetic HA coating on Ti-6Al-4V substrates using dual

beam FIB/SEM. The capabilities of GFIB and PFIB were also compared. [5] However,

due to ion bombardment and beam heating effects of FIB, the cross-sectioning the

sample using FIB is challenging when the sample involves polymer, which has a low

heating conductivity. [31]

In the present work, a biomimetic Col-HA coating was deposited on the

surface-treated Ti-6Al-4V alloy through simultaneous collagen self-assembly and HA

crystal nucleation. Dual-beam FIB/SEM techniques were used to produce high quality

cross-sections and TEM samples to study the cross-sectional microstructures of the

composite coating and the interface between the coating and the substrate. Finally, the

microstructure of the composite coating was compared with pure HA coating.

Additionally, a comparison between GFIB and PFIB in characterization of Col-HA

composite coatings was conducted.

3.2 Materials and methods

3.2.1 Biomimetic Col-HA coating formation

Ti-6Al-4V discs with a thickness of 2 mm and a diameter of 15 mm were

Page 70: Fabrication and Characterization of Collagen-Apatite

51

polished using 240 grit silicon carbide abrasive paper (LECO Corporation, USA), and

then treated by sandblasting and acid etching (SLA) techniques, followed by UV

irradiation for 8 h. SLA is a widely used surface treatment technique for Ti-based

implants. [5] Briefly, Ti-6Al-4V discs were sandblasted with 46 grit SiC, and then

acid etched by a mixed acid solution of HCl and HNO3, followed with a hot acid

mixture of HCl and H2SO4 for 5 min. After the surface treatment, the Ti-6Al-4V discs

were immersed in the m-SBF containing collagen (0.05 g/L) to produce biomimetic

Col-HA coatings. Type I collagen was extracted from rat tails and purified based on

the protocol reported by Rajan et al.. [32] The m-SBF was prepared as reported

previously (Table 3.1), [5] and the initial pH of the solution was adjusted to 6.75 by

using NaOH and HEPES (4-(2-hydroxyethyl)-1-piperazinneethanesulfonic acid). [7]

During the coating process, urea with a concentration of 0.5 mol/L was added to the

m-SBF containing collagen solution to slow down the collagen polymerization

process. [7] The coating formation was carried out by soaking the surface-treated

Ti-6Al-4V discs in the collagen-containing m-SBF solution at a temperature of 37 °C

for 24 h, as shown in Figure 3.1. As a control, the HA coating was formed by

immersing the surface treated Ti-6Al-4V substrates in the m-SBF solution with an

initial pH of 6.6 at 37 °C for 24 h. Such formed Col-HA and HA coated Ti-6Al-4V

discs were then washed gently with deionized water and dried in air for 24 h.

Page 71: Fabrication and Characterization of Collagen-Apatite

52

Table 3.1. Ion concentration (mM) of human blood plasma and m-SBF.

Ion Human blood plasma (mM) m-SBF (mM)

Na+ 142.0 109.9

K+ 5.0 6.2

Mg2+ 1.5 1.5

Ca2+ 2.5 7.9

Cl- 103.0 111.2

HCO3- 27.0 17.5

HPO42- 1.0 3.1

SO42- 0.5 0

3.2.2 Materials characterization

The phases of the Col-HA and HA coated Ti-6Al-4V alloys were evaluated by

X-ray diffraction (XRD) using a Bruker AXS D2 Phaser Diffractometer with CuKα

radiation. The diffraction data was collected using a step size of 0.02° and a scan

speed of 0.5 ° /min, and the 2θ range is 5° to 60°. The Col-HA and HA coatings were

scratched from the Ti-6Al-4V alloys, and the coatings were then characterized using

attenuated total reflection-Fourier transformation infrared spectroscopy (ATR-FTIR).

A Nicolet Magna 560 FTIR spectrophotometer (Artisan Technology Group®, IL) with

an ATR setup was used to collect the infrared spectra in the range of 4,000-400 cm-1.

The surface morphology and surface roughness of the Ti-6Al-4V substrates,

Col-HA coatings and HA coatings on the substrate were observed using FEI Helios

dual-beam FIB/SEMs: a Nanolab 460F1 GFIB and a Xenon PFIB in the

UConn/Thermo Fisher Scientific Center for Advanced Microscopy and Materials

Page 72: Fabrication and Characterization of Collagen-Apatite

53

Analysis (CAMMA). The milling and observation of the cross-sections of the Col-HA

coating on Ti-6Al-4V alloy were also performed in the Helios dual-beam FIB/SEMs.

Both GFIB and PFIB are equipped with EasyLift nano-manipulators to facilitate TEM

sample preparation, and the GFIB also equipped with a flip-stage and a scanning

transmission electron microscopy (STEM) detector for improved final TEM sample

thinning process. Before SEM imaging and FIB milling, both the Col-HA and HA

coatings were sputter coated with a conducting layer of Au using a Polaron E5100

coating instrument.

The applications of GFIB and PFIB to prepare the cross-sections and TEM

samples have been described previously, [5] and the detailed process is as following:

SEM images were obtained using through-lens detector (TLD) or Everhart-Thornley

detector (ETD) in the electron columns. All the sectioning process was performed

with the coating surface normal to the ion column, and 38° to the electron column.

Firstly, a 2-µm-thick Pt strap was deposited on the coating surface at the site of

interest, with the aim to protect the coating from ion beam damage. The Pt layer was

deposited using either electron beam or ion beam to crack organometallic Pt precursor.

[33] The ion beam voltages used in the Pt deposition were 30 kV in GFIB and 8 kV in

the PFIB.

Secondly, a stepped, wedge-shape was milled out using regular cross-section

(RCS) pattern, so that the microstructure of the coating and the interface between the

coating and the substrate could be exposed on the vertical wedge wall, and be

observed by SEM column. In the GFIB, the size of the wedge was 20 µm in length,

Page 73: Fabrication and Characterization of Collagen-Apatite

54

10 µm in width and 5 µm in depth, and in the PFIB, a large wedge was milled, and the

dimensions were 100 µm in length, 10 µm in width and 5 µm in depth.

Thirdly, after the initial wedge was milled, the vertical wedge wall was polished

by GFIB and PFIB using an ion beam with reduced accelerating current to reduce the

curtaining effects and other milling artefacts. Three different pre-set milling patterns

were used for the polishing process: RCS pattern, cleaning cross-sectional (CCS)

pattern, and rectangle (Rec) pattern.

For the TEM sample preparation, 3-µm-thick Pt strap was deposited at the site of

interest, and then two wedges were milled on both sides of the Pt strap, producing a

pre-thinned FIB lamella. A U-shape was cut on the vertical wedge wall of the

pre-thinned lamella, leaving 2 small pieces on the surface still attached to the bulky

substrate. Then this lamella was welded to the EasyLift W probe, cut totally free from

surrounding material and transferred to a Cu omni-grid. Finally, the lamella was

further thinned to electron transparency on the grid using reduced ion current to

minimize ion damage.

The TEM samples were examined using an FEI Talos F200X S/TEM operating at

an accelerating voltage of 200 kV in the UConn/Thermo Fisher Scientific CAMMA.

This instrument is equipped with a Super-X silicon drift detector (SDD) energy

dispersive X-ray spectrometry (EDS) system, allowing for a rapid acquisition of

spectrum images for elemental mapping.

3.3 Results and discussion

Collagen-HA (Col-HA) composite coatings have been obtained by a biomimetic

Page 74: Fabrication and Characterization of Collagen-Apatite

55

coating process, in which Col-HA coatings were deposited on the Ti-6Al-4V

substrates by immersing them in the collagen molecules containing m-SBF at 37 °C

and initial pH of 6.75 for 24 h (Figure 3.1). Figure 3.1 shows the process of the

formation of biomimetic Col-HA coatings on the surface-treated Ti-6Al-4V substrates.

Ti-6Al-4V substrates were polished using a 240 grit silicon carbide abrasive paper,

and the scratches were observed due to the silicon carbide abrasive paper (Figure

3.1A). Then the polished substrates were sand-blasted with 46 grit silicon carbide and

acid treated (SLA treatment, Figure 3.1B). The acid treatment produced a uniform

rough surface, which consists of sharp peaks at the microscale level, resulting in

higher surface areas. [35, 36] The acid treated Ti-6Al-4V substrate was then irradiated

using UV light at a wavelength of 254.8 nm for 8 h, resulting in a photocatalytic

chemical reaction. The surface chemistry alteration by photochemical reaction

decontaminates hydrocarbon and increase hydrophilicity by converting Ti4+ sites to

Ti3+ sites. [7, 35, 37] Thus abundant Ti-OH or Ti-O¯ groups are formed on the surface

of the substrate when in contact with SBF solution, imparting the surface with

negative charge and high surface energy. The negative charge and high surface area of

the pre-treated surface will attract and absorb Ca2+ to the surface, contributing to the

precipitation and deposition of coating on the Ti-6Al-4V substrate. [25] Then the

Ti-6Al-4V substrates with a bioactive surface were immersed in the

collagen-containing m-SBF (Figure 3.1C).

Page 75: Fabrication and Characterization of Collagen-Apatite

56

Figure 3.1. Illustration of the formation process of biomimetic Col-HA composite

coatings on the surface-treated Ti-6Al-4V substrates: Surface morphology of the

Ti-6Al-4V substrate after polishing using silicon carbide abrasive paper (A) and SLA

treatment (B), and coating formation by immersing surface-treated Ti-6Al-4V

substrate in the collagen-containing m-SBF solution (C).

Figure 3.2 shows the surface morphologies of the Col-HA coatings on the

Ti-6Al-4V substrates which is produced as described above. After 24 h of immersion

in the collagen-containing m-SBF, a crack-free homogeneous coating was formed on

the surface of the treated Ti-6Al-4V substrates at 37 °C with an initial pH of 6.75

(Figures 3.2A and 3.2B). The coating demonstrates a porous and rough surface with

numerous globules, and the roughness of the surface was clearly observed when the

Col-HA coated Ti-6Al-4V substrate was tilted 52° in the Dual Beam FIB/SEM

(Figure 3.2C). Compared with the HA coatings on the Ti-6Al-4V substrates, the

globule size of Col-HA coating was significantly smaller than that of HA coating

(Figure 3.3). In our previous research, a crack-free HA coating was formed at an

initial pH of 6.6 (Figure 3.3A). [5] In contrast, higher initial pH was required to

induce Col-HA composite coating, which is in accordance with the report by Xia et al..

[7]

Page 76: Fabrication and Characterization of Collagen-Apatite

57

Figure 3.2. SE SEM images of Col-HA composite coatings on Ti-6Al-4V substrates.

Col-HA coated Ti-6Al-4V alloy was titled 52° (C).

Figure 3.3. SE SEM images of the surface morphologies of HA (A) and Col-HA (B)

coatings on Ti-6Al-4V substrates.

Figure 3.4 shows the XRD spectra and ATR-FTIR patterns of HA and Col-HA

coatings formed after 24 h of immersion. Characteristic peaks of HA phase at 2θ=26.7°

((002) plane), 31.7-34.3° (overlap of (211), (112), and (300) planes), and 53.8° ((004)

plane) were observed for Col-HA and HA coatings. The XRD spectra indicate that the

crystalline HA presents in both the Col-HA and HA coatings. Further, two additional

peaks at 28.9° ((102) plane) and 50.2° ((213) plane) were only observed in the HA

coating. [38, 39] However, the broad diffraction peak around 31.7-34.3° indicates that

the HA phase in both the Col-HA and HA coatings are poorly crystallized. Weaker HA

peaks together with strong peaks from the hexagonal phase of the Ti-6Al-4V substrate

were observed for the Col-HA coatings. In comparison, the intensity of the Ti peaks

Page 77: Fabrication and Characterization of Collagen-Apatite

58

detected in the HA coating are much lower (Figure 3.4A). This indicates that a thinner

Col-HA coating was formed on the Ti-6Al-4V substrate compares with the HA

coating. Diffraction peaks for the hexagonal phase of the Ti-6Al-4V substrate were

observed in all of the XRD spectra.

Figure 3.4. XRD (A) and ATR-FTIR (B) spectra of HA and Col-HA coated

Ti-6Al-4V substrates.

HA and Col-HA coatings on the Ti-6Al-4V substrates were scratched from the

substrates and ATR-FTIR was also applied to characterize the coatings. Typical

characteristic peaks of the HA phase were shown in Figure 3.4B. The P=O stretching

mode (~1019.5 cm-1) and P-O bending mode (~597.8 cm-1 and ~560.6 cm-1) indicate

the phosphate ions, and the peak at 873.8 cm-1 is due to the joint contribution from

carbonate and HPO42- ions. [40] The peaks at 1448.0 cm-1 and 1414.5 cm-1 were

assigned to the CO32- groups. The absorption peak at 1638.4 cm-1 was attributed to the

stretching vibration of amide I, which is the characteristics of collagen, indicating the

presence of collagen in the coating. [20, 41]

Page 78: Fabrication and Characterization of Collagen-Apatite

59

Figure 3.5. SE SEM images of the cross-sections of Col-HA composite coatings on

Ti-6Al-4V substrates prepared using GFIB (A1, A2, A3) and PFIB (B1, B2, B3). The

cross-sections were polished using Rec pattern (A1, B1), CCS pattern (A2, B2), and

RCS pattern (A3, B3). White arrows in B1, B2, and B3 indicate the cracks between Pt

layer and Col-HA coatings.

FIB was then used to mill the Col-HA coated Ti-6Al-4V substrate to study the

details of cross-sectional microstructure of the coating and the interface between the

coating and the substrate. As described in our previous work, [5] a 2-µm-thick dense

Pt protective layer was deposited onto the surface of Col-HA coating at the area of

interest by gas-assisted Pt deposition, protecting the surface of the coating from ion

damage. [42] The ion beam accelerating voltages were 30 kV for GFIB and 8 kV for

PFIB. Then RCS pattern was used to mill a stepped, wedge-shape trench, exposing

the cross-sections of the Col-HA coated Ti-6Al-4V substrates and the interface

between the coatings and substrates. By applying RCS pattern, which is a multi-pass

scan method, the beam scanned through the entire pattern and then repeats for four

passes. Then Rec, CCS, and RCS patterns with reduced current (40 pA for GFIB and

Page 79: Fabrication and Characterization of Collagen-Apatite

60

24 pA for PFIB) were applied to reduce the “curtaining” effect, which is one of the

main drawbacks and artefacts of FIB milling. The curtaining effect is a result from the

fluctuations of the ion beam current and position, and the inhomogeneity of sample,

[5] which significantly reduces the resolution of the cross-sectional images after the

initial wedge cutting. Thus, eliminating curtaining effect is usually followed after the

initial wedge cutting using RCS pattern. Rec pattern is a pre-set milling pattern, in

which beam repeatedly scans over a rectangular array of equally spaced positions. The

CCS pattern is processed line by line with each line containing a set number of passes.

Figure 3.5 shows the cross-sectional images of Col-HA coated Ti-6Al-4V

substrates polished by Rec, CCS and RCS patterns using a reduced ion beam current

of 40 pA in GFIB (Figure 3.5 (A1-A3)) and an even lower ion beam current of 24 pA

in PFIB (Figure 3.5 (B1-B3)). With stepwise polish of Rec, CCS and RCS patterns in

reduced ion beam current, the curtaining effect on the cross-sectional images of

Col-HA coatings can be attenuated in both GFIB and PFIB, disclosing more details of

the coating cross-sections. Whereas it is also obvious that even low ion beam current

can cause curtaining effect on the cross-sectional images of Col-HA coating. This

issue is predominant in PFIB mainly owing to its larger ion beam spot size compared

to that of GFIB. The cracks between Pt protective layer and Col-coating (marked by

white arrows in Figure 3.5 (B1-B3)) were generated by ion beam of PFIB. [5]

Nevertheless, the cross-sectional details of Col-HA coating have been preserved after

a serial of milling in both GFIB and PFIB, indicating their capability of revealing

inner coating microstructures.

Page 80: Fabrication and Characterization of Collagen-Apatite

61

Figure 3.6. SE SEM images of the cross-sections of Col-HA (A) and HA (B) coatings

on the Ti-6Al-4V substrates.

In a previous study, our research group for the first time managed to clearly

observe a gradient, porous HA coating using dual beam GFIB and PFIB. [5] In this

work, a similar characteristic was observed in Col-HA coatings. Figure 3.6 shows a

cross-sectional comparison of Col-HA and HA coatings. Both coatings demonstrate a

relatively dense structure adjacent to the interface of the coating and the substrate,

whereas they become more and more porous towards the coating surfaces. The

thicknesses of Col-HA and HA coatings were 1-1.5 μm and 1.5-2.5 μm, respectively.

The cross-sectional SEM images also show that Col-HA coating exhibits a less porous

structure compared to the HA coating possibly owing to the fact that collagen

molecule polymerization occurred simultaneously with the HA nucleation. During this

process, the collagen molecules penetrated into the coating, and therefore filled some

of the pores in the coating, resulting in a less porous structure.

Page 81: Fabrication and Characterization of Collagen-Apatite

62

Figure 3.7. TEM sample preparation sequence from Col-HA coated Ti-6Al-4V alloy

using GFIB (A-K). The TEM sample was confirmed to be electron transparency by

STEM in GFIB (L).

Figure 3.8. TEM sample preparation sequence of Col-HA coated Ti-6Al-4V alloy

using PFIB (A-K). Col-HA coated Ti-6Al-4V sample was mounted on TEM grid

holder for preparing a TEM sample using PFIB (L).

To study the coating inner structure and the interface between coating and

Page 82: Fabrication and Characterization of Collagen-Apatite

63

substrate at a nanoscale, TEM samples were prepared from Col-HA coating using

both GFIB and PFIB. A representative extraction sequence of TEM sample

preparation in GFIB is shown in Figure 3.7. Briefly, a 3-μm-thick Pt protective layer

was firstly deposited onto the area of interest (Figure 3.7B). Then two wedge-shaped

trenches were milled out along with either edge of the Pt stripe and further thinned

and cleaned (Figures 3.7C and 3.7D). The sample stage was then tilted to allow a 52°

angle between ion beam and coating cross-section in order to completely cut through

the TEM lamella (Figure 3.7E). The W Easylift probe was then moved adjacent to the

top of the TEM lamella, and an extra Pt layer was deposited to connect the probe and

TEM lamella (Figures 3.7F and 3.7G). After being completely free from the pristine

sample, the TEM lamella was lifted out and welded onto a copper omni-grid (Figures

3.7H and 3.7I). The W Easylift probe was then removed after cutting free from the

TEM lamella (Figure 3.7J), which was finally thinned using ion beam to be electron

transparent (Figures 3.7K and 3.7L). A similar process was conducted on the same

Col-HA sample in PFIB to achieve a TEM lamella (Figure 3.8).

The nanoscale microstructure of the TEM specimens fabricated using GFIB and

PFIB were investigated using S/TEM and EDS. In each case, the representative

images including a bright field (BF) TEM image, a high angle annular dark field

(HAADF) STEM image, and EDS mappings for Ca, C, Ti and the milling ion (Ga and

Xe, respectively) are shown in Figures 3.9 and 3.10, respectively. The BF and

HAADF images confirm that the thickness of the Col-HA coating is 1-1.5 μm, and it

consists of a gradient structure with a more porous layer on the surface, and a denser

Page 83: Fabrication and Characterization of Collagen-Apatite

64

structure towards the interface of the coating and the substrate. By comparing the Ca,

C, and Ti maps, it is clear that a uniform, defect-free interface between the substrate

and the coating was formed. There was very limited amount of milling ions (Ga or Xe)

detected within the coating that underneath the Pt protective layer. The porous Pt layer

on the Col-HA coating was due to the deposition voltage (30 kV) chosen in PFIB

(Figure 3.10B), which is in accordance with our previous work. [5] There are many

studies reported that a porous surface topography is beneficial to cell adhesion,

proliferation as well as drug loading and delivery, [43-45] while a dense interface

could facilitate a strong bonding between the coating and the substrate, which is an

important parameter for in vivo surgical handling. [7, 46] Moreover, our previous

result suggested that with the incorporation of collagen protein molecules in

hydroxyapatite coating could significantly enhance the biocompatibility of the coating.

[7] Therefore, the Col-HA coated Ti-6Al-4V could be used as a hard tissue

replacement material, and the disclosure of the cross-sectional details in this work

provides an insight on the great potential of utilizing Col-HA composite coating in

clinical applications.

Page 84: Fabrication and Characterization of Collagen-Apatite

65

Figure 3.9. Characterization of the TEM specimens prepared by GFIB: BFTEM (A)

and HAADF STEM image (B). C-F: Compositional maps obtained from the region in

(B), showing the distribution of Ca (C), C (D), Ti (E), and Ga (F).

Figure 3.10. Characterization of the TEM specimens prepared by PFIB: BFTEM (A)

Page 85: Fabrication and Characterization of Collagen-Apatite

66

and HAADF STEM image (B). C-F: Compositional maps obtained from the region in

(B), showing the distribution of Ca (C), C (D), Ti (E), and Xe (F).

3.4 Conclusions

A biomimetic collagen-containing hydroxyapatite composite coating was

successfully fabricated on the surface of Ti-6Al-4V by simply immersing the

surface-treated Ti-6Al-4V substrates into a collagen-containing m-SBF. The

cross-sectional microstructure of the Col-HA coating has been mainly investigated by

a dual-beam FIB/SEMs with either gallium (GFIB) or xenon plasma (PFIB) ion

source and the capabilities of these two systems were compared. Both techniques

were able to acquire cross-sectional images and obtain TEM samples from the

Col-HA coating with good quality. It was revealed that the Col-HA coating consists of

a dense layer adjacent to the interface between the coating and the substrate, while a

porous structure towards the surface. Compared to GFIB, PFIB is more time efficient

even at a lower ion beam voltage and current. However, GFIB process induces less

damage to coating and the curtaining effect could be better attenuated.

Page 86: Fabrication and Characterization of Collagen-Apatite

67

References

[1] S. B. Goodman, Z. Yao, M. Keeney, F. Yang. The future of biologic coatings for

orthopaedic implants. Biomaterials. 2013; 34:3174-3183.

[2] L. T. de Jonge, S. C. Leeuwenburgh, J. J. van den Beucken, J. te Riet, W. F.

Daamen, J. G. Wolke, D. Scharnweber, J. A. Jansen. The osteogenic effect of

electrosprayed nanoscale collagen/calcium phosphate coatings on titanium.

Biomaterials. 2010; 31:2461-2469.

[3] E. Mohseni, E. Zalnezhad, A. R. Bushroa. Comparative investigation on the

adhesion of hydroxyapatite coating on Ti–6Al–4V implant: A review paper.

International Journal of Adhesion and Adhesives. 2014; 48:238-257.

[4] E. Sandrini, C. Giordano, V. Busini, E. Signorelli, A. Cigada. Apatite formation

and cellular response of a novel bioactive titanium. Journal of Materials Science:

Materials in Medicine. 2007; 18:1225-1237.

[5] C. Hu, M. Aindow, M. Wei. Focused ion beam sectioning studies of biomimetic

hydroxyapatite coatings on Ti-6Al-4V substrates. Surface and Coatings Technology.

2017; 313:255-262.

[6] H. Daugaard, B. Elmengaard, J.E. Bechtold, T. Jensen, K. Soballe. The effect on

bone growth enhancement of implant coatings with hydroxyapatite and collagen

deposited electrochemically and by plasma spray. Journal of Biomedical Materials

Research Part A. 2010; 92:913-921.

[7] Z. Xia, X. Yu. Biomimetic collagen/apatite coating formation on Ti6Al4V

substrates. Journal of Biomedical Materials Research Part B: Applied Biomaterials.

Page 87: Fabrication and Characterization of Collagen-Apatite

68

2012; 100:871-881.

[8] P. Yu, F. Lu, W. Zhu, D. Wang, X. Zhu, G. Tan, X. Wang, Y. Zhang, L. Li, C. Ning.

Bio-inspired citrate functionalized apatite coating on rapid prototyped titanium

scaffold. Applied Surface Science. 2014; 313:947-953.

[9] Y. Huang, Q. Ding, S. Han, Y. Yan, X. Pang. Characterisation, corrosion resistance

and in vitro bioactivity of manganese-doped hydroxyapatite films electrodeposited on

titanium. Journal of Materials Science: Materials in Medicine. 2013; 24:1853-1864.

[10] C. Aparicio, D. Rodriguez, F. Gil. Variation of roughness and adhesion strength

of deposited apatite layers on titanium dental implants. Materials Science and

Engineering: C. 2011; 31:320-324.

[11] D. Chen, E. H. Jordan, M. Gell, M. Wei. Apatite formation on alkaline-treated

dense TiO 2 coatings deposited using the solution precursor plasma spray process.

Acta Biomaterialia. 2008; 4:553-559.

[12] H. W. Kim, Y. H. Koh, L. H. Li, S. Lee, H. E. Kim. Hydroxyapatite coating on

titanium substrate with titania buffer layer processed by sol–gel method. Biomaterials.

2004; 25:2533-2538.

[13] M. Auger, B. Savoini, A. Muñoz, T. Leguey, M. Monge, R. Pareja, J. Victoria.

Mechanical characteristics of porous hydroxyapatite/oxide composites produced by

post-sintering hot isostatic pressing. Ceramics International. 2009; 35:2373-2380.

[14] Z. Xia, X. Yu, X. Jiang, H.D. Brody, D.W. Rowe, M. Wei. Fabrication and

characterization of biomimetic collagen–apatite scaffolds with tunable structures for

bone tissue engineering. Acta Biomaterialia. 2013; 9:7308-7319.

Page 88: Fabrication and Characterization of Collagen-Apatite

69

[15] J. M. Holzwarth, P. X. Ma. Biomimetic nanofibrous scaffolds for bone tissue

engineering. Biomaterials. 2011; 32:9622-9629.

[16] S. Manara, F. Paolucci, B. Palazzo, M. Marcaccio, E. Foresti, G. Tosi, S.

Sabbatini, P. Sabatino, G. Altankov, N. Roveri. Electrochemically-assisted deposition

of biomimetic hydroxyapatite–collagen coatings on titanium plate. Inorganica

Chimica Acta. 2008; 361:1634-1645.

[17] S. H. Teng, E. J. Lee, C. S. Park, W. Y. Choi, D. S. Shin, H. E. Kim. Bioactive

nanocomposite coatings of collagen/hydroxyapatite on titanium substrates. Journal of

Materials Science: Materials in Medicine. 2008; 19:2453-2461.

[18] K. Hu, X. J. Yang, Y. L. Cai, Z. D. Cui, Q. Wei. Preparation of bone-like

composite coating using a modified simulated body fluid with high Ca and P

concentrations. Surface and Coatings Technology. 2006; 201:1902-1906.

[19] H. Qu, M. Wei. The effect of temperature and initial pH on biomimetic apatite

coating. Journal of Biomedical Materials Research Part B: Applied Biomaterials. 2008;

87:204-212.

[20] H. Qu, Z. Xia, D. A. Knecht, M. Wei. Synthesis of dense collagen/apatite

composites using a biomimetic method. Journal of American Ceramic Society. 2008;

91:3211-3215.

[21] Y. Fan, K. Duan, R. Wang. A composite coating by electrolysis-induced collagen

self-assembly and calcium phosphate mineralization. Biomaterials. 2005;

26:1623-1632.

[22] W. Tong, Z. Yang, X. Zhang, A. Yang, J. Feng, Y. Cao, J. Chen. Studies on

Page 89: Fabrication and Characterization of Collagen-Apatite

70

diffusion maximum in x‐ray diffraction patterns of plasma‐sprayed hydroxyapatite

coatings. Journal of Biomedical Materials Research Part A. 1998; 40:407-413.

[23] F. Garcia, J. Arias, B. Mayor, J. Pou, I. Rehman, J. Knowles, S. Best, B. Leon, M.

Pérez‐Amor, W. Bonfield. Effect of heat treatment on pulsed laser deposited

amorphous calcium phosphate coatings. Journal of Biomedical Materials Research

Part A. 1998; 43:69-76.

[24] S. Jalota, S. Bhaduri, S.B. Bhaduri, A. C. Tas. A protocol to develop crack-free

biomimetic coatings on Ti6Al4V substrates. Journal of Materials Research. 2007;

22:1593-1600.

[25] E. Zhang, C. Zou, S. Zeng. Preparation and characterization of silicon-substituted

hydroxyapatite coating by a biomimetic process on titanium substrate. Surface and

Coatings Technology. 2009; 203:1075-1080.

[26] M. Iliescu, V. Nelea, J. Werckmann, I. Mihailescu. Transmission electron

microscopy investigation of pulsed-laser deposited hydroxylapatite thin films

prepared by tripod and focused ion beam techniques. Surface and Coatings

Technology. 2004; 187:131-140.

[27] M. Groeber, B. Haley, M. Uchic, D. Dimiduk, S. Ghosh. 3D reconstruction and

characterization of polycrystalline microstructures using a FIB–SEM system.

Materials Characterization. 2006; 57:259-273.

[28] R. Young, C. Rue, S. Randolph, C. Chandler, G. Franz, R. Schampers, A.

Klumpp, L. Kwakman. A comparison of xenon plasma FIB technology with

conventional gallium LMIS FIB: imaging, milling, and gas-assisted applications.

Page 90: Fabrication and Characterization of Collagen-Apatite

71

Microscopy and Microanalysis. 2011; 17:652-653.

[29] N. Smith, W. Skoczylas, S. Kellogg, D. Kinion, P. Tesch, O. Sutherland, A.

Aanesland, R. Boswell. High brightness inductively coupled plasma source for high

current focused ion beam applications. Journal of Vacuum Science & Technology B.

2006; 24:2902-2906.

[30] T. Burnett, R. Kelley, B. Winiarski, L. Contreras, M. Daly, A. Gholinia, M. Burke,

P. Withers. Large volume serial section tomography by Xe Plasma FIB dual beam

microscopy. Ultramicroscopy. 2016; 161:119-129.

[31] S. Kim, M.J. Park, N.P. Balsara, G. Liu, A.M. Minor. Minimization of focused

ion beam damage in nanostructured polymer thin films. Ultramicroscopy. 2011;

111:191-199.

[32] N. Rajan, J. Habermehl, M. F. Coté, C. J. Doillon, D. Mantovani. Preparation of

ready-to-use, storable and reconstituted type I collagen from rat tail tendon for tissue

engineering applications. Nature Protocol. 2006; 1:2753-2758.

[33] P. Munroe. The application of focused ion beam microscopy in the material

sciences. Materials characterization. 2009; 60:2-13.

[34] H. Qu, M. Wei. Improvement of bonding strength between biomimetic apatite

coating and substrate. Journal of Biomedical Materials Research Part B: Applied

Biomaterials. 2008; 84:436-443.

[35] H. Aita, W. Att, T. Ueno, M. Yamada, N. Hori, F. Iwasa, N. Tsukimura, T. Ogawa.

Ultraviolet light-mediated photofunctionalization of titanium to promote human

mesenchymal stem cell migration, attachment, proliferation and differentiation. Acta

Page 91: Fabrication and Characterization of Collagen-Apatite

72

Biomaterialia. 2009; 5:3247-3257.

[36] A. Nagaoka, K. I. Yokoyama, J. I. Sakai. Evaluation of hydrogen absorption

behaviour during acid etching for surface modification of commercial pure Ti, Ti–

6Al–4V and Ni–Ti superelastic alloys. Corrosion Science. 2010; 52:1130-1138.

[37] W. Att, N. Hori, F. Iwasa, M. Yamada, T. Ueno, T. Ogawa. The effect of

UV-photofunctionalization on the time-related bioactivity of titanium and chromium–

cobalt alloys. Biomaterials. 2009; 30:4268-4276.

[38] R.K. Brundavanam, G.E.J. Poinern, D. Fawcett. Modelling the crystal structure

of a 30 nm sized particle based hydroxyapatite powder synthesised under the

influence of ultrasound irradiation from X-ray powder diffraction data. American

Journal of Materials Science. 2013; 3:84-90.

[39] C. Hu, M. Zilm, M. Wei. Fabrication of intrafibrillar and extrafibrillar

mineralized collagen/apatite scaffolds with a hierarchical structure. Journal of

Biomedical Materials Research Part A. 2016; 104:1153-1161.

[40] X. Hu, H. Shen, K. Shuai, E. Zhang, Y. Bai, Y. Cheng, X. Xiong, S. Wang, J.

Fang, S. Wei. Surface bioactivity modification of titanium by CO 2 plasma treatment

and induction of hydroxyapatite: in vitro and in vivo studies. Applied Surface Science.

2011; 257:1813-1823.

[41] Z. Xia, M. Wei. Biomimetic fabrication of collagen-apatite scaffolds for bone

tissue regeneration. Journal of Biomaterials and Tissue Engineering. 2013; 3:369-384.

[42] E. Solla, B. Rodríguez-González, H. Aguiar, C. Rodríguez-Valencia, J. Serra, P.

González. Revealing the nanostructure of calcium phosphate coatings using

Page 92: Fabrication and Characterization of Collagen-Apatite

73

HRTEM/FIB techniques. Materials Characterization. 2016; 122:148-153.

[43] L. Yu, S. Qian, Y. Qiao, X. Liu. Multifunctional Mn-containing titania coatings

with enhanced corrosion resistance, osteogenesis and antibacterial activity. Journal of

Materials Chemistry B. 2014; 2:5397-5408.

[44] H. Hu, W. Zhang, Y. Qiao, X. Jiang, X. Liu, C. Ding. Antibacterial activity and

increased bone marrow stem cell functions of Zn-incorporated TiO2 coatings on

titanium. Acta Biomaterialia. 2012; 8:904-915.

[45] H. W. Kim, J.C. Knowles, H. E. Kim. Hydroxyapatite/poly (ε-caprolactone)

composite coatings on hydroxyapatite porous bone scaffold for drug delivery.

Biomaterials. 2004; 25:1279-1287.

[46] G. Hu, L. Zeng, H. Du, X. Fu, X. Jin, M. Deng, Y. Zhao, X. Liu. The formation

mechanism and bio-corrosion properties of Ag/HA composite conversion coating on

the extruded Mg-2Zn-1Mn-0.5 Ca alloy for bone implant application. Surface and

Coatings Technology. 2017; 325:127-135.

[47] M. Stigter, J. Bezemer, K. De Groot, P. Layrolle. Incorporation of different

antibiotics into carbonated hydroxyapatite coatings on titanium implants. release and

antibiotic efficacy. Journal of Controlled Release, 2004; 99:127-137.

Page 93: Fabrication and Characterization of Collagen-Apatite

74

4. Fabrication of intrafibrillar calcified collagen fibrils

Abstract

Bone is an organic-inorganic hierarchical biocomposite. Its basic building block

is mineralized collagen fibrils with both intrafibrillar and extrafibrillar mineralization,

which is believed to be regulated by non-collagenous proteins (NCPs) with

polyanionic domains. In this study, collagen fibrils with both intrafibrillar and

extrafibrillar mineralization were successfully prepared and the mechanism of

biomineralization was proposed. To achieve intrafibrillar mineralization, polyacrylic

acid (PAA) was used to sequester calcium and phosphate ions to form fluidic

PAA-amorphous calcium phosphate (PAA-ACP) nanoprecursors. At the presence of

sodium tripolyphosphate (TPP), PAA-ACP nanoprecursors were modulated to orderly

deposit within the gap zone of collagen fibrils. The effect of PAA concentration on

the intrafibrillar and extrafibrillar mineralization of reconstituted collagen fibrils was

investigated. It was found that with the decrease of PAA concentration, the inhibitory

effect of PAA on mineralization and the stability of ACP nanoprecursors decreased.

As a result, more minerals were deposited both within and on the surface of the

collagen fibrils.

Keywords: Intrafibrillar mineralization; extrafibrillar mineralization;

non-collagenous proteins; amorphous calcium phosphate nanoprecursors; hierarchical

structure

Page 94: Fabrication and Characterization of Collagen-Apatite

75

4.1 Introduction

Bone is an outstanding organic-inorganic nanocomposite, which is made of

mineralized collagen fibrils. It consists of seven hierarchical levels of organization,

[1] ranging from the nanometer scale to the millimeter scale. Nanoscale

hydroxyapatite (HA) crystals are embedded within the interstices of the collagen

fibrils and aligned along the long axis of collagen fibrils, which are in turn present in

bundles or arrays. [2, 3] These aligned mineralized collagen fibrils are arranged into a

multilayered cylindrical structure rotating across the concentric lamellae and forming

osteons. [4] Finally, the osteons are packed densely into compact bone or loosely into

a spongy bone. [5, 6] Since mineralized collagen fibrils are the primary unit of the

complex bone structure, it is important to understand the mineralization mechanism of

collagen fibrils in order to replicate biomineralization of collagen fibrils in vitro.

Biomineralization of collagen fibrils includes oriented HA crystals within and on

the surface of self-assembled collagen fibrils. Triple helical type I collagen molecules

self-assemble into collagen fibrils in a quarter-staggered manner, which leads to 67

nm periodicity with alternating gap and overlap zones. [7, 8] During the

mineralization of collagen fibrils, HA crystals first form within the gap zone, and then

spread through the fibrils, leading to arrays of HA nanocrystals embedded within and

oriented along the collagen fibrils, namely intrafibrillar mineralization. [9-11]

Following intrafibrillar mineralization, HA crystals grow on the surface of the

collagen fibrils, leading to extrafibrillar mineralization. [9] It has been reported that

intrafibrillar mineralization of collagen fibrils attributes to the high mineral content in

Page 95: Fabrication and Characterization of Collagen-Apatite

76

bone and the improvement of the mechanical and biological properties of the collagen

matrix. [5, 6, 12-14] Meanwhile, extrafibrillar mineralization of collagen fibrils plays

an important role in improvement of mineral content and mechanical properties. [13]

The HA minerals in the bone contribute to better biocompatibility, osteoconductivity

and osteointegration. [15, 16]

Substantial research have been done to deposit HA crystals onto the surface of

collagen fibrils (extrafibrillar mineralization) and within collagen fibrils (intrafibrillar

mineralization), [17-19] and it has been widely accepted that collagen matrix cannot

induce intrafibrillar mineralization by itself. Anionic non-collagenous proteins

(NCPs) are found to play an important role in modulating HA crystal nucleation and

the alignment of apatite deposited within collagen fibrils. [20-23] Owing to the

limited availability and high cost of NCPs, polyanionic electrolytes, such as

polyacrylic acid (PAA) and polyaspartic acid (PAsP), are often used to mimic the

functional groups of NCPs. Their functions are to sequester calcium and phosphate

ions to form fluid amorphous calcium phosphate nanoprecursors (polymer induced

liquid precursors, PILP). [21, 24, 25] It has been identified recently that bone

mineralization initially nucleates from an amorphous phase and then grows into

crystals. [15, 26-29] The advantages of using fluid amorphous precursors are that they

are moldable and infiltrate easily into collagen fibrils. [10, 27] However, polyanionic

macromolecules induced amorphous precursors alone could not reproduce highly

ordered intrafibrillar mineralization. Inorganic polyphosphate (sodium

tripolyphosphate, TPP) is normally used to replicate the template function of

Page 96: Fabrication and Characterization of Collagen-Apatite

77

phosphoproteins, which modulate HA deposition due to its excellent binding capacity

for calcium ions and its strong affinity to collagen fibrils. [14, 24, 30]

With considerable progresses in unveiling mechanisms of biomineralization of

collagen fibrils in vivo and in vitro replicating the process, most of the recent research

has been focused on the biomineralization of 3-D constructs through a PILP process.

For example, collagen scaffolds, turkey tendon, eggshell membrane and

demineralized human dentine have been biomineralized with intrafibrillar and

extrafibrillar minerals via polyanionic macromolecules-induced liquid precursors. [19,

22, 29, 31-34] Dai et al. have proved that the presence of polyaspartic acid mimicking

NCPs is essential for inducing intrafibrillar mineralization. [35] Meanwhile, Jee et al.

have revealed that oriented HA crystals in turkey tendon can be formed via the PILP

process, in which polyanionic polyaspartic acid is used to mimic the role of NCPs to

facilitate the formation of fluid amorphous nanoprecursors. [19] Li et al. have also

used PAA to mimic NCPs to stabilize amorphous calcium phosphate precursors to

mineralize eggshell membrane with intrafibrillar mineralization. [34]

Although many studies have been conducted to investigate intrafibrillar

mineralization via the PILP process to mimic the structural organization in natural

bone or teeth, little work has been done to investigate the concentration of polyanionic

polymers on the formation of intrafibrillar or extrafibrillar minerals, and there is no

report on the bottom-up process to produce intrafibrillar mineralized collagen fibrils.

In this study, we actively investigated the effect of polyanionic polymer

Page 97: Fabrication and Characterization of Collagen-Apatite

78

concentrations on the intrafibrillar mineralization of collagen fibrils using a bottom-up

approach in order to develop better understanding of the mineralization process in

natural bone.

4.2 Materials and methods

4.2.1 Materials

Type I collagen was extracted from rat tails, purified based on the protocol

reported by Rajan et al.. [36] Polyacrylic acid (PAA, Mw 2000 Da) and sodium

tripolyphosphate (TPP) were purchased from Sigma-Aldrich. All chemicals were of

analytical chemical grade.

4.2.2 Preparation of intrafibrillar and extrafibrillar mineralized collagen fibrils

Based on our previously established protocol for biomimetic collagen apatite

scaffolds, [37] intrafibrillar and extrafibrilar mineralized collagen fibrils were

fabricated at different PAA concentrations. Briefly, PAA, the sequestration analog,

was added to a modified simulated body fluid (m-SBF) to form stabilized amorphous

calcium phosphate (ACP) nanoprecursors (PAA-ACP). The m-SBF was prepared as

reported previously: NaOH, K2HPO4·3H2O, MgCl2·6H2O, CaCl2 were added in

deionized water (DIW) in the order as they are listed, and the ion concentrations were

adjusted according to Table 4.1. [38-40] Concentrations of PAA in PAA-ACP

nanoprecursors were 0, 0.4, 0.8, 2, and 5.6 mg/mL, and the pH of each PAA-ACP

nanoprecursor was adjusted to 7.2 by addition of HEPES

(4-(2-hydroxyapatiteethyl)-1-piperazineethanesulfonic acid) and NaOH. Type I

Page 98: Fabrication and Characterization of Collagen-Apatite

79

collagen stock solution (4.5 mg/mL) was diluted to a given concentration (0.2 mg/mL)

using PAA-ACP nanoprecursors. TPP (1.2 wt%), the templating analog, was added to

the collagen solution to template the hierarchical arrangement of hydroxyapatite

within collagen fibrils. The ion concentrations of all chemicals in the suspensions are

listed in Table 4.1. All of the reactions were taken under 4 ℃, and then transferred to

a water bath at 37 ℃ for 24 hr.

Table 4.1. Ion concentrations of suspension and body plasma

Ion Human blood plasma (mM) Suspension (mM)

Na+ 142.0 439.0

K+ 5.0 25.0

Mg2+ 1.5 6.0

Ca2+ 2.5 31.7

Cl- 103.0 444.6

HCO3- 27.0 70

HPO42- 1.0 12.5

SO42- 0.5 0

4.2.3 Characterization

PAA-ACP nanoprecursors with a PAA concentration of 0.4 mg/mL were added

to the collagen solution, and TPP (1.2 wt%) was then added to template minerals

deposition within collagen fibrils. The pH of the solution was adjusted to 7.2 using

HEPES and NaOH. After pH adjustment, the solution was dropped onto Formvar

carbon coated copper grid, rinsed with DIW, dried under room temperature, and then

observed using transmission electron microscopy (TEM, JEOL JEM-2010) at an

Page 99: Fabrication and Characterization of Collagen-Apatite

80

accelerating voltage of 80 kV. Five TEM images were analyzed using Image J

software to achieve the average size of ACP nanoprecursors and amorphous calcium

phosphate nanoparticles.

In order to examine the mineralization of collagen fibrils, intrafibrillar and

extrafibrillar mineralized collagen fibrils using PAA-ACP nanoprecursors at different

PAA concentrations were deposited on copper grid and observed using TEM. The

corresponding mineral phase was evaluated by selected area electron diffraction

(SAED) coupled to TEM. The surface morphology of the mineralized collagen fibrils

using PAA-ACP nanoprecursors at different PAA concentrations was also imaged

using field emission scanning electron microscopy (FESEM, JEOL JSM-6335F,

Japan).

4.3 Results

To investigate the effect of PAA concentration in PAA-ACP nanoprecursors on

the biomineralization of collagen fibrils, different PAA concentrations (0, 0.4, 0.8, 2,

5.6 mg/mL) were used to induce PAA-ACP nanoprecursor formation. PAA-ACP

nanoprecursors were added to the type I collagen stock solution and TPP was used to

modulate mineral arrangement in the collagen fibrils. Figure 4.1 shows the TEM

images of unstained collagen fibrils with different PAA concentrations in PAA-ACP

nanoprecursors. An increasingly distinct banding pattern (D-banding) in

electron-dense nucleation sites can be observed in collagen fibrils with decreasing

PAA concentration, implying that the mineral content formed within the gap zone of

Page 100: Fabrication and Characterization of Collagen-Apatite

81

collagen fibrils increased with the reduction of PAA concentration (Figure 4.1). At a

PAA concentration of 5.6 mg/mL, vague D-periods are observed (Figure 4.1A),

indicating electron-dense minerals deposited, poorly reproducing the D-banding

pattern of naturally mineralized collagen fibrils normally observed in bone. With the

decrease of PAA concentration, the intrafibrillar mineralization becomes more

obvious and homogeneous (Figure 4.1A-D), which is attribute to the higher mineral

content formed within the collagen fibrils. The inserted SAED result of the

mineralized collagen fibrils formed at the PAA concentration of 2 mg/mL (Figure

4.1B) does not exhibit a ring pattern, which indicates that the minerals formed within

the collagen fibrils are either amorphous or nanocrystallined. The control is the

sample with no PAA in the collagen-containing calcium phosphate solution. As such,

no D-banding is observed, suggesting that minerals are not deposited in specific

nucleating sites within collagen fibrils. Both SEM and TEM results reveal that HA

minerals (white arrow, Figure 4.2B) only formed on the surface of collagen fibrils

(Figure 4.2). SAED result (Figure 4.2A, insert) exhibits rings of typical low

crystalline apatite. These results collectively suggest that ordered mineralization

within collagen fibrils cannot be achieved without the presence of NCPs analogs.

Page 101: Fabrication and Characterization of Collagen-Apatite

82

Figure 4.1. TEM images of unstained mineralized collagen fibrils with different PAA

concentrations: (A) 5.6 mg/mL, (B) 2 mg/mL (insert: SAED result), (C) 0.8 mg/mL

and (D) 0.4 mg/ mL.

Page 102: Fabrication and Characterization of Collagen-Apatite

83

Figure 4.2. TEM image (A) and SEM image (B) of mineralized collagen fibrils

without PAA. SAED result for mineralized collagen fibrils without PAA (Insert, A).

White arrow in (B): HA minerals on the surface. The cracks on collagen fibrils were

induced by high voltage employed during SEM operation.

Figure 4.3 shows FESEM images of mineralized collagen fibrils with different

PAA concentrations. Vague D-banding is observed at a PAA concentration of 5.6

mg/mL (Figure 4.3A, white arrow). This suggests that few minerals have deposited

within collage fibrils. The D-banding becomes more distinguished with the decrease

of the PAA concentration, indicating more minerals are incorporated into the collagen

fibrils (Figures 4.3A-D), which is in accordance with the TEM result (Figure 4.1).

White dotted arrow and white arrow heads in Figure 4.3D show minerals formed on

the surface of collagen fibrils (extrafibrillar mineralization) with low PAA

concentration in PAA-ACP nanoprecursors. The inset image in Figure 4.3D is the

enlarged image indicated by white dotted arrow, and white thin arrows show minerals

aligned on the surface of collagen fibrils with an almost 67 nm spacing distance. This

suggests that this specific site is the entering sites for calcium phosphate nanoclusters

(Figure 4.3D). Moreover, intrafibrillar and extrafibrillar mineralized collagen fibrils

are arranged into bundles or fused to form collagen fibers (Figure 4.3).

Page 103: Fabrication and Characterization of Collagen-Apatite

84

Figure 4.3. FESEM images of mineralized collagen fibril bundles with different PAA

concentrations (white arrows indicate D-banding): (A) 5.6 mg/mL, (B) 2 mg/mL, (C)

0.8 mg/mL (the insert is the enlarged image of the spot indicated by the white arrow),

and D) 0.4 mg/mL (white dotted arrow and white arrow heads show minerals on the

collagen surface), minerals aligned on the surface of collagen fibrils with an almost 67

nm spacing distance (white dotted arrow and thin arrows in insert image). The insert

is the enlarged image of the spot indicated by the white dotted arrow.

4.4 Discussion

Since mineralized collagen fibrils are the primary unit of the complex bone

structure, investigation of the intrafibrillar and extrafibrillar mineralization of collagen

fibrils has become essential. It is widely accepted that biomineralization is well

regulated by NCPs, such as DMP 1, [41] osteonectin, osteocalcin and bone

Page 104: Fabrication and Characterization of Collagen-Apatite

85

sialoprotein. [8] Two functional motifs in NCPs have been found to play an important

role in regulation of nucleation, dimension and order of HA deposition within natural

bone. First, acidic sequestrating motif of NCPs binds to calcium to form

calcium-bound organic materials, and thereby inhibits the precipitation of calcium

phosphate. This in turn forms stabilized amorphous calcium phosphate

nanopresursors. [42] PAA was used in our study to replicate the sequestration

functional groups in NCPs. The –COO- groups of PAA attract calcium ions from

solution, which further recruit phosphate ions into the fluid droplets to form

supersaturated amorphous calcium phosphate nanoprecursors (Figures 4.4 and 4.5).

[35] Electron dense nanoprecursors were observed in Figure 4.4, and amorphous

calcium phosphate nanoparticles were also shown, indicating nanophase separation

has formed in the PAA-ACP nanoprecursors. The mean size of nanoprecursors is

37.61±6.46 nm. Amorphous calcium phosphate nanoparticles were observed in the

PAA-ACP nanoprecursor droplets with a mean size of 6.70±1.04 nm measured by

Image J. The small size of ACP nanoprecursors and ACP nanoparticles makes it

possible for PAA-ACP to infiltrate into collagen fibrils. It has been reported that

matrix vesicles (MVs) play a crucial role in bone mineralization. During bone

mineralization, MVs with a diameter of 40-200 nm are secreted during osteoid

formation, and observed at sites where mineralized tissue is synthesized de novo. [42]

As a result, MVs are considered the initial mineralization site. MV membranes

enriched with acidic phospholipids have high affinities for calcium cations, which can

induce the formation of stable calcium phosphate-phospholipid complex. Obviously,

Page 105: Fabrication and Characterization of Collagen-Apatite

86

each polymer stabilized amorphous calcium phosphate nanoprecursor has a similar

function as an MV, so PAA-ACP is regarded as an analog to MV. [42, 43]

Besides, templating motif of NCPs regulates the alignment of HA within collagen

fibrils through its excellent binding capacity for calcium ions and collagen fibrils. [22]

Due to the low cost and availability of the chemical, sodium tripolyphosphate (TPP)

has been used as an analog to NCPs in templating the HA deposition within collagen

fibrils in the current study. [24] Highly negatively charged TPP accumulates in the

gap zone of collagen fibrils dominated by highly positive net charge through

electrostatic attraction. [8, 11, 24] The high binding capacity of TPP to calcium and

collagen makes it possible to act as an ionic bridge to attract PAA-ACP

nanoprecursors to enter the gap zone of collagen fibrils (Figure 4.5). [24] PAA-ACP

nanoprecursors were found to accumulate in the close vicinity of collagen fibrils and

in contact with collagen fibrils, demonstrating the attraction of PAA-ACP to the

collagen fibrils (Figure 4.4A). Once the PAA-ACP nanoprecursors penetrate the gap

zone of collagen fibrils via capillary action and electrostatic attraction, the desorption

of calcium from PAA is expected because TPP is a much stronger pentanionic

chelating agent to ACP comparing to the carboxylate groups of PAA. [44] When a

high concentration of TPP (2%) is added to PAA-ACP nanoprecursors, white

precipitates are formed in the solution (data not shown). This also suggests that TPP

has attracted calcium ions from PAA-ACP and induced the aggregation of ACP in the

gap zone of collagen fibrils (see Figure 4.4A). The free polyanionic macromolecules

would exit the gap zone and re-enter the collagen and PAA containing m-SBF

Page 106: Fabrication and Characterization of Collagen-Apatite

87

solution to continuously recruit free calcium and phosphate ions to form PAA-ACP

nanoprecursors and re-infiltrate the fibrils to form ACP minerals. [35] These ACP

minerals are then crystalized under the influence of the amino acid side chains of

collagen fibrils. [33, 45, 46] A large fraction of charged amino acid side chains in the

gap zone helps to recruit and bind calcium and phosphate ions, and the resulting

calcium-phosphate prenucleation cluster become crystallized HA with time. [8, 34,

46] Thus the gap zone of collagen fibrils serves as the entering (as inferred from the

inset image in Figure 4.3D) and mineralization sites, which forms the D-banding of

collagen fibrils (Figure 4.5). This is in agreement with the finding by Nudelman et al.,

who reported that the gap region is the infiltrating and nucleating site for the

intrafibrillar mineraliation of collagen fibrils. [11]

Figure 4.4. TEM images of collagen solution containing PAA-ACP nanoprecursor

droplets on copper grid. (A) Low magnification TEM image showing PAA-ACP

nanoprecursor droplets that were formed in the close vicinity of the collagen fibrils.

White arrow shows the collagen fibrils that have been removed. (B) & (C) High

magnification images of PAA-ACP nanoprecursor droplets. (C) One electron dense

Page 107: Fabrication and Characterization of Collagen-Apatite

88

PAA-ACP nanoprecursor droplet highlighted in the white circle consisting of many

amorphous calcium phosphate nanoparticles.

Fibrils with intrafibrillar and extrafibrillar mineralization are fabricated using

PAA as a sequestration analog to induce amorphous calcium phosphate

nanoprecursors and TPP as a templating analog, and the mechanisms involved have

been illustrated in Figure 4.5. [47, 48] The rate of collagen mineralization is

dependent on the concentration of PAA. The lower the concentration of PAA, the

weaker the inhibition of PAA for ACP nanoprecursor precipitation, the faster the

mineralization and crystallization process, and the more minerals deposited. [10, 49,

50] When the PAA concentration is high, vague D-banding has been observed,

implying that only a small number of minerals are formed in the gap zone of collagen

fibrils (Figures 4.1A and 4.2A). This is due to that the ACP nanoprecursors are highly

stabilized by PAA (Figures 4.1 and 4.3). With the decreasing concentration of PAA,

PAA-ACP precursors became less stable and crystallization process could be quick

enough to form both intra- and extra-minerals, which can be proved by the increase of

the mineral amount and formation of extrafibrillar mineralization (Figures 4.1, 4.2 and

4.3).

Page 108: Fabrication and Characterization of Collagen-Apatite

89

Figure 4.5. Schematic illustration of the intrafibrillar mineralization process induced

by PAA as a sequestration analog and TPP as a templating analog.

4.5 Conclusions

Self-assembly of collagen fibrils and intrafibrillar and extrafibrillar

mineralization were achieved using PAA as sequestration analog and TPP as a

templating analog. The effect of the PAA concentration on intrafibrillar and

extrafibrillar mineralization was realized by adjusting the concentration of NCPs

analog during the biomineralization process. Thus, this study demonstrates a

promising approach to produce intrafibrillar mineralized collagen fibrils using a

bottom-up approach.

Page 109: Fabrication and Characterization of Collagen-Apatite

90

References

[1] S. Weiner, H. D. Wagner. The material bone: structure-mechanical function

relations. Annual Review of Materials Science. 1998; 28:271-298.

[2] W. Landis, M. Song, A. Leith, L. McEwen, B. McEwen. Mineral and organic

matrix interaction in normally calcifying tendon visualized in three dimensions by

high-voltage electron microscopic tomography and graphic image reconstruction.

Journal of Structural Biology. 1993; 110:39-54.

[3] T. T. Thula, D. E. Rodriguez, M. H. Lee, L. Pendi, J. Podschun, L. B. Gower. In

vitro mineralization of dense collagen substrates: a biomimetic approach toward the

development of bone-graft materials. Acta Biomaterialia. 2011; 7:3158-3169.

[4] W. Wagermaier, H. Gupta, A. Gourrier, M. Burghammer, P. Roschger, P. Fratzl.

Spiral twisting of fiber orientation inside bone lamellae. Biointerphases. 2006; 1:1-5.

[5] I. Jäger, P. Fratzl. Mineralized collagen fibrils: a mechanical model with a

staggered arrangement of mineral particles. Biophysical Journal. 2000; 79:1737-1746.

[6] M. J. Olszta, X. Cheng, S. S. Jee, R. Kumar, Y. Y. Kim, M. J. Kaufman, et al. Bone

structure and formation: a new perspective. Materials Science and Engineering: R:

Reports. 2007; 58:77-116.

[7] F. Z. Cui, Y. Li, J. Ge. Self-assembly of mineralized collagen composites.

Materials Science and Engineering: R: Reports. 2007; 57:1-27.

[8] Q. Wang, X. Wang, L. Tian, Z. Cheng, F. Cui. In situ remineralizaiton of partially

demineralized human dentine mediated by a biomimetic non-collagen peptide. Soft

Matter. 2011; 7:9673-9680.

Page 110: Fabrication and Characterization of Collagen-Apatite

91

[9] X. Wang, F. Cui, J. Ge, Y. Wang. Hierarchical structural comparisons of bones

from wild-type and liliput dtc232 gene-mutated Zebrafish. Journal of Structural

Biology. 2004; 145:236-245.

[10] F. Nudelman, P. H. Bomans, A. George, N. A. Sommerdijk. The role of the

amorphous phase on the biomimetic mineralization of collagen. Faraday Discuss.

2012; 159:357-370.

[11] F. Nudelman, K. Pieterse, A. George, P. H. Bomans, H. Friedrich, L. J. Brylka, et

al. The role of collagen in bone apatite formation in the presence of hydroxyapatite

nucleation inhibitors. Nature materials. 2010; 9:1004-1009.

[12] R. B. Martin, D. B. Burr, N. A. Sharkey. Skeletal tissue mechanics. 1998.

[13] A. K. Nair, A. Gautieri, M. J. Buehler. Role of intrafibrillar collagen

mineralization in defining the compressive properties of nascent bone.

Biomacromolecules. 2014; 15:2494-2500.

[14] Y. Liu, D. Luo, X. X. Kou, X. D. Wang, F. R. Tay, Y. L. Sha, et al. Hierarchical

intrafibrillar nanocarbonated apatite assembly improves the nanomechanics and

cytocompatibility of mineralized collagen. Advanced Functional Materials. 2013;

23:1404-1411.

[15] Z. Xia, M. Wei. Biomimetic fabrication of collagen-apatite scaffolds for bone

tissue regeneration. Journal of Biomaterials and Tissue Engineering. 2013; 3:369-384.

[16] A. J. W. Johnson, B. A. Herschler. A review of the mechanical behavior of CaP

and CaP/polymer composites for applications in bone replacement and repair. Acta

Biomaterialia. 2011; 7:16-30.

Page 111: Fabrication and Characterization of Collagen-Apatite

92

[17] Z. Xia, M. Villa, M. Wei. A biomimetic collagen–apatite scaffold with a

multi-level lamellar structure for bone tissue engineering. Journal of Materials

Chemistry B. 2014; 2:1998-2007.

[18] S. S. Jee, L. Culver, Y. Li, E. P. Douglas, L. B. Gower. Biomimetic

mineralization of collagen via an enzyme-aided PILP process. Journal of Crystal

Growth. 2010; 312:1249-1256.

[19] S. S. Jee, R. K. Kasinath, E. DiMasi, Y. Y. Kim, L. Gower. Oriented

hydroxyapatite in turkey tendon mineralized via the polymer-induced liquid-precursor

(PILP) process. CrystEngComm. 2011; 13:2077-2083.

[20] S. Gajjeraman, K. Narayanan, J. Hao, C. Qin, A. George. Matrix macromolecules

in hard tissues control the nucleation and hierarchical assembly of hydroxyapatite.

The Journal of Biological Chemistry. 2007; 282:1193-1204.

[21] S. S. Jee, T. T. Thula, L. B. Gower. Development of bone-like composites via the

polymer-induced liquid-precursor (PILP) process. Part 1: Influence of polymer

molecular weight. Acta Biomaterialia. 2010; 6:3676-3686.

[22] Y. K. Kim, L. Gu, T. E. Bryan, J. R. Kim, L. Chen, Y. Liu, et al. Mineralisation of

reconstituted collagen using polyvinylphosphonic acid/polyacrylic acid templating

matrix protein analogues in the presence of calcium, phosphate and hydroxyl ions.

Biomaterials. 2010; 31:6618-6627.

[23] H. A. Goldberg, K. J. Warner, M. C. Li, G. K. Hunter. Binding of bone

sialoprotein, osteopontin and synthetic polypeptides to hydroxyapatite. Connective

Tissue Research. 2001; 42:25-37.

Page 112: Fabrication and Characterization of Collagen-Apatite

93

[24] Y. Liu, Y. K. Kim, L. Dai, N. Li, S. O. Khan, D. H. Pashley, et al. Hierarchical

and non-hierarchical mineralisation of collagen. Biomaterials. 2011; 32:1291-1300.

[25] R. Q. Song, H. Colfen, A. W. Xu, J. Hartmann, M. Antonietti.

Polyelectrolyte-directed nanoparticle aggregation: systematic morphogenesis of

calcium carbonate by nonclassical crystallization. Acs Nano. 2009; 3:1966-1978.

[26] L. Addadi, D. Joester, F. Nudelman, S. Weiner. Mollusk shell formation: a source

of new concepts for understanding biomineralization processes. Chemistry-A

European Journal. 2006; 12:980-987.

[27] J. Mahamid, A. Sharir, L. Addadi, S. Weiner. Amorphous calcium phosphate is a

major component of the forming fin bones of zebrafish: Indications for an amorphous

precursor phase. Proceedings of the National Academy of Sciences. 2008;

105:12748-12753.

[28] W. J. Landis, F. H. Silver. Mineral deposition in the extracellular matrices of

vertebrate tissues: identification of possible apatite nucleation sites on type I collagen.

Cells, Tissues, Organs. 2008; 189:20.

[29] J. Mahamid, B. Aichmayer, E. Shimoni, R. Ziblat, C. Li, S. Siegel, et al.

Mapping amorphous calcium phosphate transformation into crystalline mineral from

the cell to the bone in zebrafish fin rays. Proceedings of the National Academy of

Sciences. 2010; 107:6316-6321.

[30] Y. Liu, N. Li, Y. p. Qi, L. Dai, T. E. Bryan, J. Mao, et al. Intrafibrillar collagen

mineralization produced by biomimetic hierarchical nanoapatite assembly. Advanced

Materials. 2011; 23:975-980.

Page 113: Fabrication and Characterization of Collagen-Apatite

94

[31] B. Antebi, X. Cheng, J. N. Harris, L. B. Gower, X. D. Chen, J. Ling. Biomimetic

collagen–hydroxyapatite composite fabricated via a novel perfusion-flow

mineralization technique. Tissue Engineering Part C: Methods. 2013; 19:487-496.

[32] J. Li, J. Yang, J. Li, L. Chen, K. Liang, W. Wu, et al. Bioinspired intrafibrillar

mineralization of human dentine by PAMAM dendrimer. Biomaterials. 2013;

34:6738-6747.

[33] B. Marelli, C. E. Ghezzi, A. Alessandrino, J. E. Barralet, G. Freddi, S. N. Nazhat.

Silk fibroin derived polypeptide-induced biomineralization of collagen. Biomaterials.

2012; 33:102-108.

[34] N. Li, L. Niu, Y. Qi, C. K. Yiu, H. Ryou, D. D. Arola, et al. Subtleties of

biomineralisation revealed by manipulation of the eggshell membrane. Biomaterials.

2011; 32:8743-8752.

[35] L. Dai, E. P. Douglas, L. B. Gower. Compositional analysis of a polymer-induced

liquid-precursor (PILP) amorphous CaCO 3 phase. Journal of Non-Crystalline Solids.

2008; 354:1845-1854.

[36] N. Rajan, J. Habermehl, M. F. Coté, C. J. Doillon, D. Mantovani. Preparation of

ready-to-use, storable and reconstituted type I collagen from rat tail tendon for tissue

engineering applications. Nature Protocol. 2006; 1:2753-2758.

[37] Z. Xia, X. Yu, X. Jiang, H. D. Brody, D. W. Rowe, M. Wei. Fabrication and

characterization of biomimetic collagen–apatite scaffolds with tunable structures for

bone tissue engineering. Acta Biomaterialia. 2013; 9:7308-7319.

[38] Z. Xia, X. Yu. Biomimetic collagen/apatite coating formation on Ti6Al4V

Page 114: Fabrication and Characterization of Collagen-Apatite

95

substrates. Journal of Biomedical Materials Research Part B: Applied Biomaterials.

2012; 100:871-881.

[39] H. Qu, Z. Xia, D. A. Knecht, M. Wei. Synthesis of dense collagen/apatite

composites using a biomimetic method. Journal of American Ceramic Society. 2008;

91:3211-3215.

[40] X. Yu, L. Wang, F. Peng, X. Jiang, Z. Xia, J. Huang, et al. The effect of fresh

bone marrow cells on reconstruction of mouse calvarial defect combined with

calvarial osteoprogenitor cells and collagen–apatite scaffold. Journal of Tissue

Engineering and Regeneration Medicine. 2013; 7:974-983.

[41] G. He, S. Gajjeraman, D. Schultz, D. Cookson, C. Qin, W. T. Butler, et al.

Spatially and temporally controlled biomineralization is facilitated by interaction

between self-assembled dentin matrix protein 1 and calcium phosphate nuclei in

solution. Biochemistry. 2005; 44:16140-16148.

[42] H. Ozawa, K. Hoshi, N. Amizuka. Current concepts of bone biomineralization.

Journal of Oral Bioscience. 2008; 50:1-14.

[43] K. Hoshi, S. Ejiri, H. Ozawa. Organic components of crystal sheaths in bones.

Journal of Electron Microscopy (Tokyo). 2001; 50:33-40.

[44] T. Saito, A. Arsenault, M. Yamauchi, Y. Kuboki, M. Crenshaw. Mineral induction

by immobilized phosphoproteins. Bone. 1997; 21:305-311.

[45] B. Marelli, C. E. Ghezzi, Y. L. Zhang, I. Rouiller, J. E. Barralet, S. N. Nazhat.

Fibril formation pH controls intrafibrillar collagen biomineralization in vitro and in

vivo. Biomaterials. 2015; 37:252-259.

Page 115: Fabrication and Characterization of Collagen-Apatite

96

[46] Z. Xu, Y. Yang, W. Zhao, Z. Wang, W. J. Landis, Q. Cui, et al. Molecular

mechanisms for intrafibrillar collagen mineralization in skeletal tissues. Biomaterials.

2015; 39:59-66.

[47] G. Boivin, P. Meunier. The degree of mineralization of bone tissue measured by

computerized quantitative contact microradiography. Calcified Tissue International.

2002; 70:503-511.

[48] D. Toroian, J. E. Lim, P. A. Price. The Size Exclusion Characteristics of Type I

Collagen Implications for the Role of Noncollagenous Bone Constitutes in

Mineralization. Journal Biological Chemistry. 2007; 282:22437-22447.

[49] F. R. Tay, D. H. Pashley. Guided tissue remineralisation of partially

demineralised human dentine. Biomaterials. 2008; 29:1127-1137.

[50] C. Holt, S. Lenton, T. Nylander, E. S. Sørensen, S. C. Teixeira. Mineralisation of

soft and hard tissues and the stability of biofluids. Journal of Structural Biololy. 2014;

185:383-396.

Page 116: Fabrication and Characterization of Collagen-Apatite

97

5. Intrafibrillar silicification of collagen fibrils

Abstract

Inspired by the silicification process in nature, intrafibrillar silicified collagen

fibrils were successfully fabricated using a biomimetic one-step collagen

self-assembly/silicification approach, in which collagen self-assembly and

intrafibrillar silicification occurred simultaneously. To the best of our knowledge, it is

the first time that intrafibrillar silicified collagen fibrils were formed via a one-step

simultaneous approach, where collagen fibrils served as a templating matrix, and

poly(allylamine) hydrochloride and sodium tripolyphosphate acted as the respective

positive and negative analogs of zwitterionic proteins. By tailoring zwitterionic

proteins analogs, silicified collagen fibrils with different microstructures, including

core-shell, twisted and banded structures, were achieved. The intrafibrillar silicified

collagen fibrils demonstrated significant improvement of osteoblasts proliferation

compared with apatite mineralized collagen fibrils. These findings open a new avenue

for preparation of silicon-containing hierarchical biocomposites for biomedical needs.

Keywords: Biomimetic; intrafibrillar silicification; hierarchical structure; zwitterionic

proteins; one-step approach

Page 117: Fabrication and Characterization of Collagen-Apatite

98

5.1 Introduction

Silicon is an essential element contributing significantly to the health of bone and

cartilage. [1] It has been discovered that silicon-containing organic-inorganic

biocomposites with unique hierarchical structures exist in numerous natural biological

systems, such as bone, fish scales, diatom frustules, and spicules of glass sponges.

[2-5] For example, the long spicules of Hyalonema sieboldi demonstrate a perfect

hierarchical structure, remarkable optical properties, and excellent durability and

flexibility. [6] More importantly, silicon-containing biomaterials have been found to

be osteoinductive for new bone formation and stimulative for neovascularization

without supplements of growth factors. [7, 8] Due to the unique properties of the

silicon-containing biomaterials, great enthusiasm has been generated and extensive

effort has been made to replicate these materials to be used in different biomedical

applications. [6, 9, 10]

In recent years, effort has been directed to develop an in-depth understanding of

the biosilicification mechanism so as to replicate the process in vitro and thereby

fabricate organic-inorganic hybrid biomaterials with a hierarchical structure.

Chemical analyses of glass sponge and diatom have shown that fibrillar collagen can

serve as templates for silicification, while the highly zwitterionic proteins, such as

silaffins, silicateins, and long-chain polyamines, are actively involved in the interplay

with collagen and the mediation of silicification process. [6, 11] The zwitterionic

proteins exhibit both positive charges introduced by the quaternary ammonium groups

and negative charges by the phosphorylated serine residues. [12] The zwitterionic

Page 118: Fabrication and Characterization of Collagen-Apatite

99

nature of these proteins enables them to form aggregates, which controls the

silicification process and silica morphology. [13] Polyamines have been found to

induce silica precipitation in vitro in the presence of polyanions, where polyamines

and polyanions act as positive and negative analogs of zwitterionic proteins,

respectively. [3] Despite all the successes in fabrication of silicon-containing

biomaterials, an in-depth understanding of the mechanism of biomimetic silicification

process has not yet been achieved.

A variety of silicified collagen hybrids have been produced, such as

silica-collagen xerogels, and silicified collagen hydrogels with different silicon

sources (silica and silicates). [14-18] Typically, silica is formed on the surface of

collagen fibrils (extrafibrillar silicification), resulting in a weak bond between silica

and collagen. [14, 15, 19] However, the recent successes in intrafibrillar calcification

of collagen fibrils have shed light on the understanding of intrafibrillar silicification

(forming silica within collagen fibrils), as intrafibrillar calcification and silicification

of collagen fibrils may share a similar mechanism. In our previous work, intrafibrillar

calcified collagen fibrils have been achieved by a one-step collagen self-assembly and

calcification approach. [20, 21] In the present study, we also propose to produce

intrafibrillar silicified collagen fibrils via a one-step collagen

self-assembly/silicification (OCSS) approach, where collagen self-assembly and

silicification occur simultaneously with the aid of positive and negative analogs of

zwitterionic proteins. To test this hypothesis, a systemic study was conducted to

investigate the effect of zwitterionic proteins analogs on the OCSS approach, and an

Page 119: Fabrication and Characterization of Collagen-Apatite

100

in vitro cell culture study was also conducted to test the biocompatibility of the

intrafibrillar silicified collagen fibrils.

5.2 Materials and methods

5.2.1 Materials

Type I collagen was extracted from rat tails, purified based on the protocol

reported by Rajan et al.. [22] Poly(allylamine) hydrochloride (PAH, Mw = 15,000 Da)

and sodium tripolyphosphate (TPP) were purchased from Sigma-Aldrich. All

chemicals were of analytical chemical grade.

5.2.2 Preparation of hydrolyzed silicic acid

A 3 wt% hydrolyzed silicic acid stock solution was prepared as followed:

Silbond® 40 (40% hydrolyzed tetraethyl orthosilicate (TEOS); Silbond Corp., Weston,

MI, USA), absolute ethanol, deionized water (DIW) and 37% HCl were mixed at a

molar ratio of 1.03:199:6.11:0.03 for 1 hour at room temperature to hydrolyze TEOS

into orthosilicic acid and its oligomers. [23]

5.2.3 Preparation of PAH-stabilized silica (PAH-SA) precursors

PAH-stabilized silica (PAH-SA) precursors were prepared by the following

procedure: A solution with a PAH concentration of 0.267 mM was mixed with 3 wt%

silicic acid at a 1:1 volume ratio to obtain 1.5 wt% silicic acid stabilized by PAH.

After stirring for 1 min, the mixture was centrifuged at 3000 RPM and 25 °C for 10

min. The PAH-stabilized silica (PAH-SA) precursors were then collected. [24]

Page 120: Fabrication and Characterization of Collagen-Apatite

101

5.2.4 Controlled biomimetic silicification of collagen fibrils via an OCSS approach

Buffer solution was used to avoid sudden pH changes and prepared based on our

established protocol, [20, 25] where 184.8 mM NaCl, 6.3 mM K2HPO4·3H2O and

100.7 mM HEPES (4-(2-hydroxyethyl)piperazine-1-ethane-sulfonic acid) were added

to DIW in the order listed above.

*Collagen-silica (Col-SA) fibrils: Type I collagen stock solution (4.5 mg/mL)

was diluted to a given concentration (0.3 mg/mL) using buffer solution at 4 °C. Then

different amounts of silicic acid solution (3 wt%) were added dropwise to the diluted

collagen solution immediately with vigorous stirring. The amount of silicic acid added

was controlled at a SA/Col ratio of 1:2, 1:1, 2:1, 3:1 and 4:1. After mixing the diluted

collagen solution with the silicic acid solution and stirring for 5 min, the mixtures (pH

= 6.4) were incubated at 25 °C for 10 min, and then incubated in a water bath at 37 °C

for 24 h.

*Collagen-poly(allylamine) hydrochloride-silica (Col-PAH-SA) fibrils: Type I

collagen stock solution (4.5 mg/mL) was diluted to a given concentration (0.3 mg/mL)

using the buffer solution at 4 °C. Then different amounts of PAH-SA precursors were

added dropwise into the diluted collagen solution immediately with vigorous stirring.

The amount of PAH-SA precursors was controlled at a SA/Col ratio of 1:1, 4:3 and

2:1. Higher ratios of SA/Col were not used since precipitates formed rather quickly

when the ratios of SA/Col were 3:1 and 4:1. After mixing the diluted collagen

solution with PAH-SA precursors and stirring for 5 min, Col-PAH-SA mixtures (pH =

6.4) were incubated at 25 °C for 10 min, and then transferred to a water bath at 37 °C

Page 121: Fabrication and Characterization of Collagen-Apatite

102

for 24 h.

*Intrafibrillar silicified Collagen (Col-inSA) fibrils: Type I collagen stock

solution (4.5 mg/mL) was diluted to a given concentration (0.3 mg/mL) using the

buffer solution at 4 °C. TPP (1.2 wt%) was added to this diluted collagen solution and

stirred for 5 min. Different amounts of PAH-SA precursors were then added dropwise

into this collagen-TPP solution with vigorous stirring. The amount of PAH-SA

precursors was controlled at a SA/Col ratio of 1:1, 4:3 and 2:1. After stirring for 5 min,

Collagen containing mixtures (pH = 6.8) were incubated at 25 °C for 10 min, and then

transferred to a water bath at 37 °C for 24 h.

5.2.5 Preparation of silicified collagen fibrils via a two-step approach

Firstly, collagen fibrils without mineralization were produced by the following

procedure: type I collagen stock solution (4.5 mg/mL) was diluted to a given

concentration (0.3 mg/mL) using the buffer solution at 4 °C, followed by incubation

at 25 °C for 10 min and then in a water bath at 37 °C for another 24 h to precipitate

collagen fibrils. Secondly, the collagen fibrils without mineralization were silicified

by incubating in silica precursors, which was described below. *Col-SA fibrils via a

two-step approach (T-Col-SA): Silicic acid was added dropwise to the collagen fibrils

at a SA/Col ratio of 4:1, and then incubated at 37 °C for another 24 h. *Col-PAH-SA

fibrils via a two-step approach (T-Col-PAH-SA): PAH-SA precursors were added to

the self-assembled collagen fibrils at a SA/Col ratio of 4:3, and then incubated at

37 °C for 24 h. *Col-inSA fibrils via a two-step approach (T-Col-inSA): TPP was

dissolved in the self-assembled collagen fibrils at a concentration of 1.2 wt%, and

Page 122: Fabrication and Characterization of Collagen-Apatite

103

PAH-SA precursors (SA/Col = 4:3) were added after 5 min of stirring. The mixed

solution was then incubated at 37 °C for 24 h.

5.2.6 Materials characterization

Transmission electron microscopy (TEM): Silicified collagen fibrils and collagen

fibrils without mineralization were dropped onto Formvar carbon coated copper grids,

rinsed with DIW and then dried at room temperature. TEM (JEOL, JEM-2010) was

used to examine the silicification of collagen fibrils at an accelerating voltage of 80

kV. All samples were examined without staining.

Field emission scanning electron microscopy (FESEM): FESEM (JEOL

JSM-6335F, Japan) was used to characterize the surface morphology of the silicified

collagen fibrils. Different silicified collagen fibrils were sputter-coated with gold and

examined using FESEM operating at 8-10 kV.

Thermogravimetric analysis (TGA): TGA was performed using a Q500 TGA

analyzer (TGAQ-500, TA Instrument, New Castle, DE, USA). About 20 mg

freeze-dried silicified collagen fibrils were placed in a platinum pan and heated from

30 to 900 °C at a rate of 10 °C/min in air to determine the amount of minerals in the

silicified collagen fibrils.

Attenuated total reflection-Fourier transform infrared spectroscopy (ATR-FTIR):

The freeze dried silicified collagen fibrils were characterized using ATR-FTIR. A

Nicolet Magna 560 FT-IR spectrophotometer (Artisan Technology Group ®, Illinois,

USA) with an ATR setup was used to collect infrared spectra in the range 4,000-400

cm-1 at a 4 cm-1 resolution, 32 scans.

Page 123: Fabrication and Characterization of Collagen-Apatite

104

5.2.7 In vitro cell culture and proliferation

Mouse calvaria 3T3-E1 (MC3T3-E1, ATCC, USA) cell line, an osteoblast

precursor cell line, was used to determine the cytocompatibility of various samples.

MC3T3-E1 cells were plated in a culture flask (FALCON, USA) containing 10 mL of

α-minimum essential medium (α-MEM; Gibco, Invitrogen, Inc., USA), which was

supplemented with 10% fetal bovine serum (FBS; Gibco, Invitrogen, Inc., USA) and

1% penicillin–streptomycin (Gibco, Invitrogen, Inc., USA). In a humidified

atmosphere of 5% CO2 under 37 °C that maintained by an incubator (NAPCO, USA),

cells were further passaged when reaching 80-90% confluence. The fibrous

membranes of 5.5 mm in diameter were collected and cross-linked with 1 wt% EDC

(1-ethyl-3-(3-dimethylaminopropyl)-carbodiimide). Collagen-apatite membrane with

an apatite content of about 67 wt% was produced using a co-precipitation approach,

which serves as a control. The cross-linked fibrous membranes were sterilized and

placed in a 96-well plate.

To observe cell proliferation and viability using alamarBlue assay, cells were

seeded on various samples (four replicates) in 96-well plates at a cell density of 8,000

cells per well in 100 μL cell culture medium. The cells were cultured for 14 days and

the cell culture medium was refreshed every 2 days. After incubating for 1 day,

samples that attached with cells were transferred to a new 96-well plate to leave

behind the cells escaped from samples. Phenol-red-free cell culture medium with 10%

alamarBlue (Thermo Scientific, USA) was added to each well. After culturing for

another 4 h, 100 μL of the culture solution from each well was transferred to a new

Page 124: Fabrication and Characterization of Collagen-Apatite

105

96-well plate to measure the absorbance at wavelengths of 570 nm and 600 nm by a

mircoplate reader (BioTek, USA). The samples were then rinsed against PBS twice to

remove residue alamarBlue, and 100 μL fresh cell culture medium was added to each

well and refreshed every other day. After culturing for 4, 7, 10, and 14 days, cell

viabilities were investigated according to the procedure descripted above. Percentage

of reduction of alamarBlue was calculated according to the instruction provided by

the assay.

5.2.8 Statistical analysis

Statistically significant differences (p) between various groups were measured

using two-way RM ANOVA analysis of variance that carried out by GraphPad Prism

5.

5.3 Results and discussion

5.3.1 Col-SA fibrils

Silicified collagen fibrils with a core-shell structure are successfully produced

using a one-step approach. When silicic acid is mixed with collagen molecules with a

SA/Col ratio of 3:1 or 4:1, a thick layer of dense silica is coated on the collagen fibrils,

forming a core-shell structured Col-SA hybrid (Figures 5.1A1, 5.1A2, 5.1B1 and

5.1B2). This result is in accordance with the reports by Eglin et al. and Ono et al.,

where hollow silica fibrils were produced with silica coated onto the surface of

templating collagen fibrils. [26, 27] Some Col-SA fibrils also demonstrate a twisted

structure, which is due to the presence of negatively charged silanol groups of silica

Page 125: Fabrication and Characterization of Collagen-Apatite

106

coating on the surface of Col-SA hybrid (Figures 5.1B1, 5.1B2, and 5.1C). [9, 28]

When the SA/Col ratio is low (1:2, 1:1, or 2:1), silicified collagen fibrils with

core-shell structures are not formed (Data not shown). When the SA/Col ratio is 3:1,

the distance between two adjacent protrusions is 66.9±5.8 nm (n = 30, Figures 5.1A1

and 5.1B1), which is identical to the distance of the gap zone present in collagen

fibrils (67 nm), indicating the templating function of collagen fibrils. Negatively

charged silicic acid preferentially accumulates at sites of highly positively charged

gap zone of collagen fibrils, leading to the formation of protrusions with a periodicity

of about 67 nm (Figure 5.1D). [15, 29] When the SA/Col ratio is 4:1, the shell

thickness of the Col-SA core-shell structure increases, and the periodicity of the

distinguished protrusions increases to 82.7±21.0 nm (n = 40, Figure 5.1A2 and 5.1B2).

This is due to that with the increase of the SA/Col ratio, more and more silica

precipitates on the surface of collagen fibrils, resulting in an increase in shell

thickness and a decrease in electrostatic attraction between silicic acid and the gap

zone of collagen fibrils. As a result, the surface of the coating was smoothened, and

some adjacent protrusions fused together, leading to a longer distance between two

adjacent protrusions (Figures 5.1A2 and 5.1B2).

Page 126: Fabrication and Characterization of Collagen-Apatite

107

Figure 5.1. TEM images (A1 and A2), FESEM images (B1 and B2) of Col-SA fibrils

with different SA/Col ratios: 3:1 (A1, B1), 4:1 (A2, B2). ATR-FTIR spectra (C) of

Col-SA fibrils (Stars indicate the peaks for Si-O-Si, and triangle shows the peak for

Si-OH). Schematic illustration (D) of the process to form Col-SA fibrils, and the TEM

image shows Col-SA fibrils with a SA/Col ratio of 3:1.

5.3.2 Col-PAH-SA fibrils

By controlling the silica precursors, twisted silicified collagen fibrils with

bandings are produced (Figure 5.2). A long-chain polyamine (poly(allylamine)

hydrochloride (PAH, Mw = 15,000 Da)) was used to stabilize silica by the formation

of hydrogen bonds with silanol groups of silicic acid, which reduces the availability of

silicic acid from forming polysilicic acid, and thereby leads to the formation of fluidic

PAH-SA nanoprecursors. [4] Rope-like, twisted silicified collagen fibrils are formed

via a one-step approach by mixing PAH-SA precursors and collagen molecules

Page 127: Fabrication and Characterization of Collagen-Apatite

108

(Figures 5.2A-D). This result is in agreement with the reports by Desimone et al. and

Heinemann et al., [9, 28] where twisted fibrils were formed due to the presence of

negatively charged silanol groups of silica (Figure 5.2D). The average distances of

two adjacent bandings for Col-PAH-SA fibrils are 143.3±12.1 nm, 145.8±7.5 nm, and

145.0±11.8 nm for SA/Col ratio of 1:1, 4:3 and 2:1, respectively (Figures 5.2A-C, n =

30). All the banding distances are significantly larger than that of the typical

D-banding of collagen fibrils (67 nm). This may be due to that the amine groups of

PAH in PAH-SA precursors interact with carboxyl groups of collagen molecules

during the self-assembly of collagen fibrils, where PAH is inserted between two

adjacent collagen molecules as a space linker, leading to a longer distance between

adjacent gap zones (Figure 5.2E). In PAH-SA precursors, silica is stabilized by PAH

through ionic bonding and/or hydrogen bonding. [4] Subsequently, PAH-SA

precursors infiltrate into collagen fibrils along with PAH during the collagen fibril

formation process. PAH is then separated from silicic acid, leading to the precipitation

of silica at the C-terminus of collagen molecules, where the potential energy is the

lowest. [29] This explains why the banding in Figure 5.2 is larger than that of the

typical collagen D-banding because of the longer distance between adjacent gap zones

of the fibrils and the precipitation of electron-dense minerals at the C-terminus of

collagen molecules. The FTIR result also confirms the attachment of PAH onto

collagen fibrils (Figure 5.2D). The 528 cm-1 peak of FT-IR is assigned to the amine

hydrogen bonding of PAH, indicating the attachment of PAH onto the silicified

collagen fibrils (Figure 5.2D).

Page 128: Fabrication and Characterization of Collagen-Apatite

109

Figure 5.2. TEM images of Col-PAH-SA fibrils with different SA/Col ratios: 1:1 (A),

4: 3 (B), and 2:1 (C). ATR-FTIR spectra (D) of Col-PAH-SA fibrils (Stars indicate the

peaks for Si-O-Si, triangle shows the peak for Si-OH, and arrow indicates amine

hydrogen bonding). Schematic illustration (E) of the process to form Col-PAH-SA

fibrils, and the TEM image shows Col-PAH-SA fibrils with a SA/Col ratio of 4:3.

5.3.3 Col-inSA fibrils

As shown in Figure 5.3, intrafibrillar silicified collagen fibrils were produced

using PAH and TPP as the positive and negative analogs of zwitterionic proteins,

respectively. Compared to Col-PAH-SA fibrils, the morphology of collagen fibrils in

the Col-inSA group exhibits a normal structure instead of twisted fibrils (Figure 5.3).

[20] This may be due to the competition between TPP and PAH to interact with

collagen molecules during the collagen self-assembly process. As a penta-anion, TPP

Page 129: Fabrication and Characterization of Collagen-Apatite

110

has a stronger electrostatic attraction to the gap zone of collagen fibrils than PAH,

leading to prohibition of further bonding of collagen molecules with PAH. Instead,

fluidic PAH-SA precursors infiltrate into the collagen fibrils and precipitate within the

gap zone (Figure 5.5). Clear D-bandings are observed using TEM due to

electron-dense silica precipitating within specific sites of collagen fibrils (Figures 5.3

and 5.4). As shown in Figure 5.3, the D-banding becomes more and more obvious

with the increase of SA/Col ratios, and silica nanoparticles are also formed on the

surface of collagen fibrils. Similar to silaffins and long-chain polyamines in diatoms,

PAH and TPP have been used as respective positive and negative analogs of highly

zwitterionic proteins to promote silica formation in vitro. [30] In details, PAH mimics

the positively charged groups of lysine residues in silaffins, while TPP mimics the

negatively charged groups introduced by phosphorylation of serine residues. [3, 31]

Similar to the mechanism of calcium phosphate intrafibrillar mineralization, TPP

accumulates in the gap zone of collagen fibrils due to electrostatic attraction between

the highly negatively charged TPP and the positively charged gap zone of collagen

fibrils. [20, 29] TPP further attracts fluidic PAH-SA nanoprecursors to infiltrate into

gap zones of collagen fibrils via electrostatic attraction and capillary action (Figures

5.5 and 5.6). [4, 32, 33] PAH promotes silica precipitation in the presence of TPP in

the gap zone of collagen fibrils, leading to intrafibrillar silicification of collagen

fibrils (Figure 5.6). [31, 34]

Page 130: Fabrication and Characterization of Collagen-Apatite

111

Figure 5.3. TEM images (A1, B1, and C1) and FESEM images (A2, B2, and C2) of

Col-inSA fibrils with different SA/Col ratios: 1:1 (A1, A2), 4:3 (B1, B2), and 2:1 (C1,

C2).

Figure 5.4. ATR-FTIR spectra (A) and TGA spectra (B) of Col-inSA fibrils with

different SA/Col ratios. (Stars indicate the peaks for Si-O-Si, triangle shows the peak

for Si-OH, and arrow indicates amine hydrogen bonding.)

Page 131: Fabrication and Characterization of Collagen-Apatite

112

Figure 5.5. The initial stage of PAH-SA precursors interacted with collagen fibrils in

the presence of TPP as a regulator. Low (A) and high (B) magnification of the fluidic

silica precursors infiltrated into the collagen fibrils during the collagen self-assembly

process. White arrows in B: Vague D-banding can be observed.

Figure 5.6. Schematic illustration of the intrafibrillar mineralization process of

Col-inSA fibrils. TEM image shows Col-inSA fibrils with a SA/Col ratio of 4:3.

5.3.4 Silicified collagen fibrils via a two-step approach

Moreover, silicification of collagen fibrils via a one-step approach was compared

with a two-step approach. In a two-step approach, collagen fibrils were first

self-assembled, and then silica precursors were added to the self-assembled collagen

Page 132: Fabrication and Characterization of Collagen-Apatite

113

fibril suspension to enable the silicification of the precipitated collagen fibrils.

Compared T-Col-SA fibrils with Col-SA fibrils at the same SA/Col ratio

(SA/Col=4:1), silica particles rather than a continuous coating are deposited on the

surface of collagen fibrils (Figures 5.7A1 and 5.7A2). Silica particles may be

absorbed onto the collagen fibrils due to electrostatic attraction, or through hydrogen

bonding between silanol groups of silica and amine or hydroxyl groups of collagen. [6]

In the T-Col-PAH-SA fibrils, silica particles are only formed on the surface of

collagen fibrils, while twisted silicified collagen fibrils are formed in Col-PAH-SA

fibrils using one-step approach (Figures 5.7B1 and 5.7B2). The comparison between

the T-Col-PAH-SA and the Col-PAH-SA fibrils indicates that in the one-step approach

PAH and silica are well integrated into the collagen self-assembly process and that the

silicification and self-assembly of collagen fibrils occur simultaneously. In contrast,

intrafibrillar silicification was observed in the Col-inSA fibrils formed via both the

two-step and the one-step processes (Figures 5.7C1 and 5.7C2). In both the two-step

and the one-step approaches, highly negatively charged TPP accumulates in the

positively charged gap zone of collagen fibrils through electrostatic attraction, which

further attracts fluidic PAH-SA precursors to infiltrate into collagen fibrils. [4, 29, 31]

PAH and TPP function as the positive and negative analogs of zwitterionic proteins,

respectively, which mediate the biosilicification process in nature systems. Similarly,

the accumulation of PAH-SA and TPP in the gap zone of collagen fibrils enables the

PAH-SA precursors to aggregate and form silica within collagen fibrils. [3, 13]

Two kinds of polyanions were used: phosphate ions and TPP, but they have

Page 133: Fabrication and Characterization of Collagen-Apatite

114

different functions. TPP inhibits the incorporation of PAH during the collagen

self-assembly process (Figures 5.7B2 and 5.7C2), and it also functions as a regulator

to guide precise deposition of silica within collagen fibrils (Figures 5.5B, 5.7B1 and

5.7C1). Phosphate ions, being one kind of kosmotrope anions, were used to favor

collagen self-assembly. [16] The comparison between T-Col-PAH-SA and T-Col-inSA

fibrils indicates that, even though both phosphate ions and TPP induce phase

separation of PAH-SA precursors, [34, 35] phosphate ions do not function as a

regulator to attract PAH-SA to infiltrate into collagen fibrils (Figure 5.7B1), but TPP

does (Figure 5.7C1).

Figure 5.7. Col-SA fibrils produced using either a two-step process (T-Col-SA (A1),

T-Col-PAH-SA (B1), and T-Col-inSA (C1)) or a one-step approach (Col-SA (A2),

Col-PAH-SA (B2), and Col-inSA (C2)).

Page 134: Fabrication and Characterization of Collagen-Apatite

115

5.3.5 In vitro cell culture

Different mineralized collagen fibrous membranes were tested for their ability to

support cell growth in vitro. An osteoblastic cell line MC3T3-E1 was used.

Collagen-apatite (Col-Ap) matrices have been widely used as scaffolds for bone tissue

engineering, thus the Col-Ap fibrous membrane was used as the control. [36-42] Cell

proliferation and viability on different fibrous membranes were quantified using

alamarBlue assay. The Col-inSA group supports a higher proliferation rate of

MC3T3-E1 cells, especially at the late stage of incubation, compared with Col-Ap,

Col-SA and Col-PAH-SA (Figure 5.8). This result indicates that the intrafibrillar

silicified collagen fibrous membrane has a better biological response to osteoblast

cells compared with Col-Ap. It is well-established that silicon-induced cell responses

are dose-dependent. [43, 44] As shown in the TEM and SEM images of Col-inSA

fibrils (Figure 5.3), silica is wrapped up by collagen fibrils, demonstrating dense and

obvious D-banding structures, allowing gradual release of silicon. [45] Therefore,

intrafibrillar silicified collagen fibril can serve as a carrier and gradually release

silicon ions from the collagen fibrils to positively impact cell activities instead of

burst release causing negative cell responses. [46]

Page 135: Fabrication and Characterization of Collagen-Apatite

116

Figure 5.8. Proliferation of MC3T3-E1 cells on mineralized collagen membranes:

Collagen-apatite (Col-Ap, control), Col-SA (SA/Col=4:3), Col-PAH-SA

(SA/Col=4:3), and Col-inSA (SA/Col=4:3) over the course of 14 days. Data is

presented as mean ± standard error. Statistically significant differences (p) between

various groups were measured using two-way RM ANOVA analysis of variance, and

*: p<0.05, **: p<0.01, ***: p<0.001.

5.4 Conclusions

We have presented a one-step collagen self-assembly/silicification approach in

preparation of intrafibrillar silicified collagen fibrils. Highly negatively charged TPP

first accumulates in the positively charged gap zone of collagen fibrils due to

electrostatic attraction, which further attracts positively charged fluidic PAH-SA

precursors to enter the collagen fibrils. As the positive analog of the zwitterionic

proteins in silicification, PAH promotes silica precipitation in the presence of negative

analog, TPP. Thus the accumulation of PAH-SA and TPP in the gap zone of collagen

fibrils facilitates the formation of silica within gap zone of collagen fibrils, resulting

Page 136: Fabrication and Characterization of Collagen-Apatite

117

in intrafibrillar silicification. Moreover, the structure of the silicified collagen fibrils

can be manipulated by varying silica precursors, polyanions and SA/Col ratios to

produce silicified collagen fibrils with a core-shell, twisted, or banded structure. The

intrafibrillar silicified collagen fibrils possess better cell compatibility compared with

the collagen-apatite fibrils. Thus, this approach provides a facile method to produce

intrafibrillar silicified collagen fibrils for biomedical applications.

Page 137: Fabrication and Characterization of Collagen-Apatite

118

References

[1] C. C. Perry, T. Keeling-Tucker. Biosilicification: the role of the organic matrix in

structure control. JBIC Journal of Biological Inorganic Chemistry. 2000; 5:537-550.

[2] E. Beniash. Biominerals—hierarchical nanocomposites: the example of bone.

Wiley Interdisciplinary Reviews: Nanomedicine and Nanobiotechnology. 2011;

3:47-69.

[3] S. Wenzl, R. Hett, P. Richthammer, M. Sumper. Silacidins: highly acidic

phosphopeptides from diatom shells assist in silica precipitation in vitro. Angewandte

Chemie. 2008; 120:1753-1756.

[4] L.-n. Niu, K. Jiao, H. Ryou, A. Diogenes, C. K. Yiu, A. Mazzoni, et al.

Biomimetic silicification of demineralized hierarchical collagenous tissues.

Biomacromolecules. 2013; 14:1661-1668.

[5] A. Woesz, J. C. Weaver, M. Kazanci, Y. Dauphin, J. Aizenberg, D. E. Morse, et al.

Micromechanical properties of biological silica in skeletons of deep-sea sponges.

Journal of Materials Research. 2006; 21:2068-2078.

[6] H. Ehrlich, R. Deutzmann, E. Brunner, E. Cappellini, H. Koon, C. Solazzo, et al.

Mineralization of the metre-long biosilica structures of glass sponges is templated on

hydroxylated collagen. Nature Chemistry. 2010; 2:1084-1088.

[7] A. J. Mieszawska, N. Fourligas, I. Georgakoudi, N. M. Ouhib, D. J. Belton, C. C.

Perry, et al. Osteoinductive silk–silica composite biomaterials for bone regeneration.

Biomaterials. 2010; 31:8902-8910.

[8] W. Zhai, H. Lu, L. Chen, X. Lin, Y. Huang, K. Dai, et al. Silicate bioceramics

Page 138: Fabrication and Characterization of Collagen-Apatite

119

induce angiogenesis during bone regeneration. Acta Biomaterialia. 2012; 8:341-349.

[9] M. F. Desimone, C. Hélary, G. Mosser, M. M. Giraud-Guille, J. Livage, T. Coradin.

Fibroblast encapsulation in hybrid silica–collagen hydrogels. Journal of Materials

Chemistry. 2010; 20:666-668.

[10] T. Delclos, C. Aimé, E. Pouget, A. Brizard, I. Huc, M. H. Delville, et al.

Individualized silica nanohelices and nanotubes: Tuning inorganic nanostructures

using lipidic self-assemblies. Nano Letters. 2008; 8:1929-1935.

[11] S. Heinemann, H. Ehrlich, T. Douglas, C. Heinemann, H. Worch, W. Schatton, et

al. Ultrastructural studies on the collagen of the marine sponge Chondrosia reniformis

Nardo. Biomacromolecules. 2007; 8:3452-3457.

[12] N. Kröger, S. Lorenz, E. Brunner, M. Sumper. Self-assembly of highly

phosphorylated silaffins and their function in biosilica morphogenesis. Science. 2002;

298:584-586.

[13] C. C. Lechner, C. F. Becker. Silaffins in Silica Biomineralization and Biomimetic

Silica Precipitation. Marine Drugs. 2015; 13:5297-5333.

[14] M. F. Desimone, C. Hélary, I. B. Rietveld, I. Bataille, G. Mosser, M. M.

Giraud-Guille, et al. Silica–collagen bionanocomposites as three-dimensional

scaffolds for fibroblast immobilization. Acta Biomaterialia. 2010; 6:3998-4004.

[15] S. Heinemann, C. Heinemann, R. Bernhardt, A. Reinstorf, B. Nies, M. Meyer, et

al. Bioactive silica–collagen composite xerogels modified by calcium phosphate

phases with adjustable mechanical properties for bone replacement. Acta

Biomaterialia. 2009; 5:1979-1990.

Page 139: Fabrication and Characterization of Collagen-Apatite

120

[16] S. Quignard, G. J. Copello, C. Aimé, I. Bataille, C. Hélary, M. F. Desimone, et al.

Influence of Silicification on the Structural and Biological Properties of Buffer‐

Mediated Collagen Hydrogels. Advanced Engineering Materials. 2012; 14:B51-B55.

[17] S. Heinemann, C. Heinemann, H. Ehrlich, M. Meyer, H. Baltzer, H. Worch, et al.

A novel biomimetic hybrid material made of silicified collagen: perspectives for bone

replacement. Advanced Engineering Materials. 2007; 9:1061-1608.

[18] S. Heinemann, H. Ehrlich, C. Knieb, T. Hanke. Biomimetically inspired hybrid

materials based on silicified collagen. International Journal of Materials Research.

2007; 98:603-608.

[19] K. Spinde, M. Kammer, K. Freyer, H. Ehrlich, J. N. Vournakis, E. Brunner.

Biomimetic silicification of fibrous chitin from diatoms. Chemistry of Materials. 2011;

23:2973-2978.

[20] L. Zhang. C. Hu, Mei Wei. Development of biomimetic scaffold woth both

intrafibrilar and extrafibrillar mineralization. ACS Biomaterials Science &

Engineering. 2015; 1:669-676.

[21] C. Hu, M. Zilm, M. Wei. Fabrication of intrafibrillar and extrafibrillar

mineralized collagen/apatite scaffolds with a hierarchical structure. Journal of

Biomedical Materials Research Part A. 2016; 104:1153-1161.

[22] N. Rajan, J. Habermehl, M. F. Coté, C. J. Doillon, D. Mantovani. Preparation of

ready-to-use, storable and reconstituted type I collagen from rat tail tendon for tissue

engineering applications. Nature Protocol. 2006; 1:2753-2758.

[23] L. Niu, K. Jiao, Y. Qi, C. K. Yiu, H. Ryou, D. D. Arola, et al. Infiltration of silica

Page 140: Fabrication and Characterization of Collagen-Apatite

121

inside fibrillar collagen. Angewandte Chemie. 2011; 123:11892-11895.

[24] L. Niu, K. Jiao, H. Ryou, C. K. Yiu, J. Chen, L. Breschi, et al. Multiphase

intrafibrillar mineralization of collagen. Angewandte Chemie. 2013; 125:5874-5878.

[25] D. A. Cisneros, J. Friedrichs, A. Taubenberger, C. M. Franz, D. J. Muller.

Creating ultrathin nanoscopic collagen matrices for biological and biotechnological

applications. Small. 2007; 3:956-963.

[26] Y. Ono, Y. Kanekiyo, K. Inoue, J. Hojo, M. Nango, S. Shinkai. Preparation of

Novel Hollow Fiber Silica Using Collagen Fibers as a Template. Chemistry Letters.

1999:475-476.

[27] D. Eglin, G. Mosser, M. M. Giraud-Guille, J. Livage, T. Coradin. Type I collagen,

a versatile liquid crystal biological template for silica structuration from nano-to

microscopic scales. Soft Matter. 2005; 1:129-131.

[28] S. Heinemann, C. Heinemann, M. Jager, J. Neunzehn, H. Wiesmann, T. Hanke.

Effect of silica and hydroxyapatite mineralization on the mechanical properties and

the biocompatibility of nanocomposite collagen scaffolds. ACS Applied Materials &

Interfaces. 2011; 3:4323-4331.

[29] F. Nudelman, K. Pieterse, A. George, P. H. Bomans, H. Friedrich, L. J. Brylka, et

al. The role of collagen in bone apatite formation in the presence of hydroxyapatite

nucleation inhibitors. Nature Materials. 2010; 9:1004-1009.

[30] N. Kröger, R. Deutzmann, M. Sumper. Polycationic peptides from diatom

biosilica that direct silica nanosphere formation. Science. 1999; 286:1129-1132.

[31] N. Li, L. Niu, Y. Qi, C. K. Yiu, H. Ryou, D. D. Arola, et al. Subtleties of

Page 141: Fabrication and Characterization of Collagen-Apatite

122

biomineralisation revealed by manipulation of the eggshell membrane. Biomaterials.

2011; 32:8743-8752.

[32] Y. Liu, Y. K. Kim, L. Dai, N. Li, S. O. Khan, D. H. Pashley, et al. Hierarchical

and non-hierarchical mineralisation of collagen. Biomaterials. 2011; 32:1291-1300.

[33] Y. K. Kim, L.-s. Gu, T. E. Bryan, J. R. Kim, L. Chen, Y. Liu, et al. Mineralisation

of reconstituted collagen using polyvinylphosphonic acid/polyacrylic acid templating

matrix protein analogues in the presence of calcium, phosphate and hydroxyl ions.

Biomaterials. 2010; 31:6618-6627.

[34] E. Brunner, K. Lutz, M. Sumper. Biomimetic synthesis of silica nanospheres

depends on the aggregation and phase separation of polyamines in aqueous solution.

Phys Chem Chem Phys. 2004; 6:854-857.

[35] J. Fothergill, M. Li, S. A. Davis, J. A. Cunningham, S. Mann.

Nanoparticle-Based Membrane Assembly and Silicification in Coacervate

Microdroplets as a Route to Complex Colloidosomes. Langmuir. 2014; 30:14591-6.

[36] Z. Xia, M. Wei. Biomimetic fabrication of collagen-apatite scaffolds for bone

tissue regeneration. Journal of Biomaterials and Tissue Engineering. 2013; 3:369-384.

[37] Z. Xia, X. Yu, X. Jiang, H. D. Brody, D. W. Rowe, M. Wei. Fabrication and

characterization of biomimetic collagen–apatite scaffolds with tunable structures for

bone tissue engineering. Acta Biomaterialia. 2013; 9:7308-7319.

[38] Y. Pek, S. Gao, M. M. Arshad, K. J. Leck, J. Y. Ying. Porous collagen-apatite

nanocomposite foams as bone regeneration scaffolds. Biomaterials. 2008;

29:4300-4305.

Page 142: Fabrication and Characterization of Collagen-Apatite

123

[39] M. M. Villa, L. Wang, J. Huang, D. W. Rowe, M. Wei. Bone tissue engineering

with a collagen–hydroxyapatite scaffold and culture expanded bone marrow stromal

cells. Journal of Biomedical Materials Research Part B: Applied Biomaterials. 2015;

103:243-253.

[40] X. Yu, Z. Xia, L. Wang, F. Peng, X. Jiang, J. Huang, et al. Controlling the

structural organization of regenerated bone by tailoring tissue engineering scaffold

architecture. Journal of Materials Chemistry. 2012; 22:9721-9730.

[41] X. Yu, L. Wang, F. Peng, X. Jiang, Z. Xia, J. Huang, et al. The effect of fresh

bone marrow cells on reconstruction of mouse calvarial defect combined with

calvarial osteoprogenitor cells and collagen–apatite scaffold. Journal of Tissue

Engineering and Regenerative Medicine. 2013; 7:974-983.

[42] M. M. Villa, L. Wang, J. Huang, D. W. Rowe, M. Wei. Improving the

permeability of lyophilized collagen–hydroxyapatite scaffolds for cell‐based bone

regeneration with a gelatin porogen. Journal of Biomedical Materials Research Part B:

Applied Biomaterials. 2015; 104:1580-1590.

[43] H. A. Santos, J. Riikonen, J. Salonen, E. Mäkilä, T. Heikkilä, T. Laaksonen, et al.

In vitro cytotoxicity of porous silicon microparticles: effect of the particle

concentration, surface chemistry and size. Acta Biomaterialia. 2010; 6:2721-2731.

[44] E. A. Bortoluzzi, L. Niu, C. D. Palani, A. R. El-Awady, B. D. Hammond, D. Pei,

et al. Cytotoxicity and osteogenic potential of silicate calcium cements as potential

protective materials for pulpal revascularization. Dental Materials. 2015;

31:1510-1522.

Page 143: Fabrication and Characterization of Collagen-Apatite

124

[45] L. Niu, K. Jiao, Y. Qi, S. Nikonov, C. K. Yiu, D. D. Arola, et al. Intrafibrillar

silicification of collagen scaffolds for sustained release of stem cell homing

chemokine in hard tissue regeneration. The FASEB Journal. 2012; 26:4517-4529.

[46] S. Bhattacharjee, I. M. Rietjens, M. P. Singh, T. M. Atkins, T. K. Purkait, Z. Xu,

et al. Cytotoxicity of surface-functionalized silicon and germanium nanoparticles: the

dominant role of surface charges. Nanoscale. 2013; 5:4870-4883.

Page 144: Fabrication and Characterization of Collagen-Apatite

125

6. Intrafibrillar calcified collagen scaffolds

Abstract

A biomimetic collagen-hydroxyapatite (Col-HA) scaffold resembling the

composition and structure of natural bone from the nano- to macro-scale has been

successfully prepared for bone tissue engineering. We have developed a bottom-up

approach to fabricate hierarchically biomimetic Col-HA scaffolds with both

intrafibrillar and extrafibrillar mineralization. To achieve intrafibrillar mineralization,

polyacrylic acid (PAA) was used as a sequestrating analog of non-collagenous

proteins (NCPs) to form a fluidic amorphous calcium phosphate (ACP) nanoprecursor

through attraction of calcium and phosphate ions. Sodium tripolyphosphate (TPP) was

used as a templating analog to regulate orderly deposition of HA within collagen

fibrils. Both X-ray diffraction and Fourier transform infrared spectroscopy suggest

that the mineral phase was HA. Field emission scanning electron microscopy,

transmission electron microscopy and selected area electron diffraction indicate that

hierarchical collagen-HA scaffolds were formed with both intrafibrillar and

extrafibrillar mineralization. Biomimetic Col-HA scaffolds with both intrafibrillar and

extrafibrillar mineralization (IE-Col-Ap) were compared to Col-HA scaffolds with

extrafibrillar mineralization only (E-Col-Ap) as well as pure collagen scaffolds in

vitro for cellular proliferation using MC3T3-E1 cells. Pure collagen scaffolds had the

highest rate of proliferation, while there was no statistically significant difference

between IE-Col-Ap and E-Col-Ap scaffolds. This approach renders a promising

Page 145: Fabrication and Characterization of Collagen-Apatite

126

Col-HA scaffold, which bears high resemblance to natural bone in terms of

composition and structure.

Keywords: Intrafibrillar mineralization; bottom-up approach; hierarchical scaffolds;

amorphous calcium phosphate nanoprecursors

Page 146: Fabrication and Characterization of Collagen-Apatite

127

6.1 Introduction

In the United States, approximately one million bone fractures are treated

every year. [1-3] Autografts have long been considered as the gold standard for

bone grafting procedures. However, the amount of bone that can be harvested is

limited and there is a high potential for donor site morbidity. Allografts are

alternatives to autografts, but there are concerns of disease transmission. [4-7]

Thus a promising strategy to address these problems is bone tissue engineering,

where a combination of a supporting scaffold, cells and stimuli is used to

regenerate bone. [4, 8, 9]

There are many different types of scaffold materials, but biomimetic

scaffolds have attracted great attention from researchers. [10, 11] Natural bone is

a composite of hydroxyapatite (HA) and collagen fibrils. HA constitutes 65 wt% of

natural bone, exhibiting biocompatibility, osteoconductivity, and bone-bonding ability.

[3, 12] Collagen, the most abundant protein in bone, has good resorbability and high

affinity to cells. [13, 14] Inspired by the composition of natural bone, scaffolds for

bone repair consisting of collagen and apatite have attracted extensive attention of

researchers in recent years. [10, 11]

Recently, great efforts have been made to produce collagen-apatite composites

resembling the composition and structure of natural bone from the nano- to the

macro-scale. The conventional strategies to produce bone substitutes include coating a

collagen scaffold with biomimetic apatite, physically mixing apatite nanocrystals with

collagen fibrils, and biomimetic co-precipitation of collagen and apatite in which

Page 147: Fabrication and Characterization of Collagen-Apatite

128

apatite is deposited on the surface of collagen fibrils. [3, 12, 15-17] Collagen/apatite

scaffolds with improved mechanical properties were achieved by either coating

collagen scaffolds with HA or physically mixing pre-synthesized apatite with type I

collagen. [3, 18] Recently, our research group employed a self-assembly approach to

form extrafibrillar mineralized collagen fibrils in a collagen-containing modified

simulated body fluid (m-SBF). [13, 16] The materials characterization confirmed that

the scaffold has a composition closely mimicking that of natural bone, and our in vitro

cell culture and in vivo animal tests suggested that this structure has excellent

osteoconductivity and is suitable for bone repair. [1, 15, 16, 19] Although these

biomimetic collagen/apatite scaffolds successfully mimic the composition of natural

bone, they do not recapitulate the structure of bone at all levels, especially at the

nanostructural level. [20]

In natural bone, HA nanocrystals deposit within the gap zone of collagen fibrils

(intrafibrillar mineralization) at the early stage of mineralization and then on the

surface of collagen fibrils (extrafibrillar mineralization) at a later stage. [10, 21-25]

Bundles and arrays of mineralized collagen fibrils further arrange into a multilayered

structure rotating across concentric lamellae, which pack densely into compact bone

or loosely into spongy bone. [26-28] Both intrafibrillar and extrafibrillar

mineralization are essential for the functions of natural bone. Intrafibrillar and

extrafibrillar mineralization of collagen fibrils have been reported to contribute to the

high mineral content in bone and the improvement of mechanical and biological

properties in collagen matrix. [29-33] Thus, biomimetic mineralization must

Page 148: Fabrication and Characterization of Collagen-Apatite

129

recapitulate both intrafibrillar and extrafibrillar mineralization of collagen fibrils to

mimic natural bone.

Extensive research has been conducted on the mechanism of intrafibrillar

mineralization, and it has been proposed that non-collagenous proteins (NCPs) play

an important role in the mineralization process. Chen et al. investigated the effect of

NCPs (osteocalcin (OC) and bone sialoprotein (BSP)) on intrafibrillar and

extrafibrillar mineralization of collagen fibers. [21] Both BSP and OC are present on

the surface of collagen fibrils, but only OC is present within the intrafibrillar space of

collagen fibrils. [21] BSP is too large to be accommodated by the gap zone of

collagen fibrils due to the size exclusion rule for NCPs - molecules larger than a 40

kDa protein are completely excluded from collagen fibrils while those smaller than 6

kDa can freely diffuse into all spaces within collagen fibrils. [34] Moreover, OC

would potentially be located at the boundary between the gap and the overlap region

of collagen fibers, further confirming the observations by Nudelman et al..[25]

Nudelman et al. proposed that the net charges of each band play crucial roles in the

deposition and maturation of HA. Rodriguez et al. further proved that osteopontin, the

most abundant NCP in bone, can induce intrafibrillar mineralization of collagen and

modulate osteoclastic activity in vitro. [35]

With a better understanding of the mechanism for intrafibrillar mineralization of

collagen, biomimetic collagen/apatite scaffolds mimicking the natural structure of

bone from nano- to macro-scale have been investigated. Collagen scaffolds, turkey

tendons, eggshell membranes and demineralized human dentines have been subjected

Page 149: Fabrication and Characterization of Collagen-Apatite

130

to either intrafibrillar or extrafibrillar mineralization via a polymer-induced

liquid-precursor (PILP) process. [36-42] In the PILP process, an electrolyte is used to

form a fluidic amorphous calcium phosphate (ACP) precursor. Polyanionic

electrolytes, such as poly (aspartic acid) (PAsP), poly (acrylic acid) (PAA) and poly

(vinyl phosphonic acid) (PVPA), are often used to mimic the functional groups of

NCPs due to the limited availability and high cost of NCPs. [43-45] Collagen tapes

were immersed in PAsP-stabilized ACP nanoprecursors, and hydroxyapatite initially

precipitated within collagen fibrils, and then a mineral coating formed on the surface

of the collagen tape. [43, 44] In a separate study, human demineralized dentine was

incubated in ACP precursors, which were stabilized by poly(amidoamine) (PAMAM),

another candidate to mimic NCPs. [40] Intrafibrillar HA mineralization was achieved,

which mimicked the arrangement of natural minerals in human dentine at the

nanometer scale. Thus, a bioinspired intrafibrillar mineralization process can be

achieved via employing polyanionic electrolytes, which are analogs of NCPs. [36-40]

However, a two-step process is normally required to produce a collagen/HA

scaffold with intrafibrillar mineralization. Typically, a collagen scaffold is produced,

and then incubated in a precursor solution to induce mineral infiltration into collagen

fibrils. There are no reports on a bottom-up approach to fabricate scaffolds with a

bone-like composition and hierarchical structure. In the bottom-up approach,

materials assemble from the nanoscale, such as molecules, ions and atoms, to form a

macroscopic structure. [46] In this study, we employed a bottom-up approach to

produce intrafibrillar and extrafibrillar mineralized collagen-apatite scaffolds from the

Page 150: Fabrication and Characterization of Collagen-Apatite

131

collagen molecules and ACP nanoprecursors instead of the aforementioned two-step

process. The structure and biological performance of the intrafibrillar and

extrafibrillar mineralized collagen-apatite scaffolds were compared with that of

extrafibrillar mineralized scaffolds produced via our previous co-precipitation

method.

6.2 Materials and methods

6.2.1 Materials

Type I collagen was extracted from rat tails, purified based on the protocol

reported by Rajan et al.. [47] Polyacrylic acid (PAA, Mw 5000 Da) and sodium

tripolyphosphate (TPP) were purchased from Sigma-Aldrich. All chemicals were of

analytical chemical grade.

6.2.2 Fabrication of extrafibrillar mineralized collagen-apatite (E-Col-Ap) scaffolds

Based on our previously established protocol, E-Col-Ap scaffolds were produced

by the following procedure: A 6X simulated body fluid, called modified simulated

body fluid (m-SBF) was prepared as reported previously: [48-50] NaCl,

K2HPO4·3H2O, MgCl2·6H2O, CaCl2 were added in deionized water (DIW) in the

order as they are listed, and the ion concentrations were adjusted according to Table

6.1. A collagen stock solution (4.5 mg/mL) was diluted to 2.2 mg/mL by the addition

of m-SBF. The pH of the collagen containing m-SBF solution was adjusted to 7.2 by

addition of HEPES (4-(2-hydroxyapatiteethyl)-1-piperazineethanesulfonic acid),

NaHCO3 and NaOH. A two-temperature process was used to form a mineralized

Page 151: Fabrication and Characterization of Collagen-Apatite

132

collagen hydrogel: The solution was incubated at 25 °C for 1 h and then transferred to

a water bath at 37 °C for 24 hr. Subsequently, the collagen-apatite hydrogel was

allowed to undergo unconfined self-gravity driven compression (self-compression) at

room temperature for 20 min. The resulting self-compressed gel was transferred to a

custom-made mold and subjected to unidirectional freezing in a freeze-drier (Free

Zone®, Labconco, USA) precooled to -25 °C. The freeze-dried scaffolds were

subsequently cross-linked with 1 wt% N-(3-dimethylaminopropyl)-N’

-ethylcarbodiimide hydrochloride (EDC) for 24 h. Cross-linked scaffolds were then

rinsed thoroughly in DIW, followed by rinsing with a 5% glycine solution to react

with residual active carboxylic acid groups. [51, 52] The scaffolds were then rinsed

again with DIW, and freeze-dried for a second time to completely remove all the

moisture within the scaffold while maintaining the structure of the scaffold.

6.2.3 Fabrication of intrafibrillar and extrafibrillar mineralized collagen apatite (IE-Col-Ap)

scaffolds

Based on our previously established protocol for the biomimetic intrafibrillar

mineralized collagen fibrils, [16, 53] IE-Col-Ap scaffolds with a lamellar structure

were fabricated via the following procedure: PAA, the sequestration analog, was

added to a m-SBF solution to form stabilized ACP nanoprecursors (PAA-ACP). The

m-SBF was prepared as described above. [48-50] The concentration of PAA in

PAA-ACP nanoprecursor was 800 ug/mL. Type I collagen stock solution (4.5 mg/mL)

was diluted to a given concentration (2.2 mg/mL) using PAA-ACP nanoprecursors.

TPP (1.2 wt%), the templating analog, was added to the collagen stock solution to

Page 152: Fabrication and Characterization of Collagen-Apatite

133

template the orderly arrangement of hydroxyapatite within collagen fibrils. The pH of

the solution was then adjusted to 7.2 by addition of HEPES, NaHCO3 and NaOH. All

of the reactions were taken place at 4 °C, and then transferred to a water bath at 37 °C

for 24 h. The self-compression process, unidirectional freeze-drying and cross-linking

process were conducted following the same procedure as described above.

6.2.4 Materials characterization

A small piece of collagen-apatite hydrogel (IE-Col-Ap and E-Col-Ap hydrogel)

prepared by the two-temperature process was placed on a copper grid and observed

using transmission electron microscopy (TEM) at an accelerating voltage of 80 kV

and selected area electron diffraction (SAED) coupled to TEM (TEM, JEOL

JEM-2010). [54] The cross-linked dry scaffold (IE-Col-Ap and E-Col-Ap scaffold)

was grinded into tiny pieces in liquid nitrogen, dispersed in ethanol, deposited onto a

copper grid, dried in air, and then observed using both TEM and SAED coupled to the

TEM. The surface morphology of the freeze-dried collagen-apatite scaffold was

imaged using field emission scanning electron microscopy (FESEM, JEOL

JSM-6335F, Japan).

Freeze dried IE-Col-Ap and E-Col-Ap scaffolds were characterized using

attenuated total reflection-Fourier transform infrared spectroscopy (ATR-FTIR). A

Nicolet Magna 560 FT-IR spectrophotometer (Artisan Technology Group ®, Illinois,

USA) with an ATR setup was used to collect infrared spectra in the range of

4,000-400 cm-1 at a 4 cm-1 resolution, 32 scans. X-ray diffraction (BRUKER AXS D2

Phaser) was also performed on the mineralized collagen scaffolds from 10° to 50° at a

Page 153: Fabrication and Characterization of Collagen-Apatite

134

step size of 0.02° and a scan rate of 0.5 °/min with CuKα radiation.

Thermogravimetric analysis (TGAQ-500, TA Instrument, USA) was performed from

30 to 900 °C with a heating rate of 10 °C/min in air to determine the mineral content

in the collagen composite scaffolds.

6.2.5 In vitro cell culture

The ability of E-Col-Ap, and IE-Col-Ap scaffolds to support cell proliferation

was monitored through the reduction of a 3-(4,5-dimethylthiazol

-2-yl)-2,5-diphenyltetrazolium (MTT) dye. A pre-osteoblastic cell line, mouse

calvaria 3T3-E1 (MC3T3-E1), was used for the study. Cells were grown in alpha

modified eagles medium (α-MEM) supplemented with 10% fetal bovine serum and 1%

penicillin-streptomycin at 37°C under an atmosphere of 5% CO2. The medium was

changed every other day until the culture reached 90% confluence, at which point

cells were harvested and seeded onto pure collagen, E-Col-Ap, and IE-Col-Ap

scaffolds at 7,500 cells per scaffold (n ≥ 4). Scaffolds were cylindrical with a diameter

of 3.5 mm and a height of 1 mm.

For the proliferation study the medium was replaced every other day.

Proliferation was assessed by replacing the medium at 1, 3, and 7 days with a working

solution of α-MEM supplemented with 10% MTT solution (5 mg/ml). The treated

scaffolds were incubated for 4 hours at 37 °C. The medium was removed and

dimethyl sulfoxide was added to dissolve the formed formazan crystals. Aliquots of

150 μl were measured using a microplate reader (Biotek) at a wavelength of 560 nm.

All data are represented as the mean ± standard error. Statistical analysis was

Page 154: Fabrication and Characterization of Collagen-Apatite

135

conducted using a student t-test with two tails, and differences were considered

significant if p < 0.05.

6.3 Results and discussions

Figure 6.1 shows the TEM images of mineralized collagen fibrils in the

IE-Col-Ap and E-Col-Ap hydrogels. At the initial stage of IE-Col-Ap scaffold

preparation, a hydrogel consisting of intrafibrillar and extrafibrillar mineralized

collagen fibrils was formed using a two-temperature process (Figure 6.1A). In Figure

6.1A, a clear D-banding pattern is observed in the unstained collagen fibrils,

indicating electron-dense minerals have deposited within specific sites of collagen

fibrils, reproducing the D-banding pattern of naturally mineralized collagen fibrils

normally observed in bone. [22] The SAED result of intrafibrillar minerals does not

demonstrate a ring pattern, but only a diffraction spot (Figure 6.1A, inset), indicating

that the minerals deposited are nanocrystalline. This specific alignment of minerals

within collagen fibrils was regulated by TPP, a templating analog of NCPs. [37]

Highly negatively charged TPP accumulates within positively charged gap zones of

collagen fibrils through electrostatic attraction, which further attracts fluidic

PAA-ACP nanoprecursors to enter the gap zone of collagen fibrils via electrostatic

attraction. [25, 44, 53, 55] Thus, TPP functions as an ionic bridge between collagen

fibrils and PAA-ACP nanoprecursors due to the high affinity of TPP to collagen and

calcium. [53] A large fraction of the charged functional groups in the gap zone recruit

and bind with calcium and phosphate ions from the PAA-ACP nanoprecursors to

precipitate electron-dense minerals within the gap zone, which results in the observed

Page 155: Fabrication and Characterization of Collagen-Apatite

136

D-banding of collagen fibrils. [42, 55, 56] In extrafibrillar mineralized collagen fibrils,

D-banding is not observed. Instead, minerals are only formed on the surface of

collagen fibrils (Figure 6.1B). The SAED result of extrafibrillar mineralized collagen

fibrils exhibits a ring pattern of typical low crystalline hydroxyapatite (Figure 6.1B,

inset).

Figure 6.1. TEM images of a small piece of IE-Col-Ap (A) and E-Col-Ap (B)

hydrogel fabricated using the two-temperature process. Inset images in A and B:

SAED results. The mineralized collagen fibrils were unstained.

Then IE-Col-Ap and E-Col-Ap hydrogels were self-compressed and frozen

radially to form the hierarchical scaffold as shown in Figure 6.2. Both IE-Col-Ap and

E-Col-Ap scaffolds exhibit a multi-lamellar structure from the macrostructure to the

microstructure level. At the macrostructure level, unidirectional macro-pores are

aligned across the entire scaffold (Figures 6.2A and 6.2B). At the microstructure level,

each layer is comprised of aligned and well stacked lamellae with a thickness of

Page 156: Fabrication and Characterization of Collagen-Apatite

137

several hundred nanometers (Figures 6.2A1 and 6.2B1). In each lamella, mineralized

collagen fibers are partially aligned along the freezing direction of the scaffold, as

indicated by double headed arrows shown in Figures 6.2A2 and 6.2B2. In the

extrafibrillar mineralized scaffold, minerals are observed to be randomly dispersed on

the surface of collagen fibrils (Figure 6.2B2, black circle). The shear stress during the

gravity self-compression and unidirectional freezing processes leads to the alignment

of macro-pores, a multi-lamellar structure, and partially aligned mineralized collagen

fibrils of the Col-HA scaffold (Figure 6.2). [16]

Page 157: Fabrication and Characterization of Collagen-Apatite

138

Figure 6.2. SEM images of IE-Col-Ap (A, A1, A2) and E-Col-Ap (B, B1, B2)

scaffolds. Double headed arrows in A2 and B2 indicate the freezing direction of

self-compressed hydrogels. Black circle in B2 shows the randomly dispersed minerals

on the surface.

Freeze-dried IE-Col-Ap and E-Col-Ap scaffolds were grinded into tiny pieces,

and observed using TEM (Figure 6.3). At the nanostructure level of the IE-Col-Ap

Page 158: Fabrication and Characterization of Collagen-Apatite

139

scaffold, minerals are arranged within collagen fibrils and a clear D-banding pattern is

observed (Figure 6.3A). Minerals are also observed on the surface of the collagen

fibrils, as indicated by white arrow in Figures 6.3A. As demonstrated by SAED (inset

of Figure 6.3A), the typical ring pattern for HA is observed, indicating that the

minerals formed within collagen fibrils are HA. An arc-shape diffraction pattern of

the (002) growth face is observed in Figure 6.3A, implying that the mineral phase is

oriented along the long axis direction of collagen fibrils. [22, 57] In contrast, the

minerals within the E-Col-Ap scaffold are randomly dispersed on the surface of

collagen fibrils (Figures 6.2B2 and 6.3B). The ring pattern observed from SAED is

consistent with the typical ring pattern for HA which implies that the minerals formed

are HA (inset of Figure 6.3B).

Figure 6.3. Freeze-dried IE-Col-Ap (A) and E-Col-Ap (B) scaffolds were ground into

small pieces in liquid nitrogen, and one of these small pieces were observed using

TEM. Inset images in A and B: SAED results. White arrows in A and B indicate the

extrafibrillar minerals.

Page 159: Fabrication and Characterization of Collagen-Apatite

140

Figure 6.4 shows the XRD spectra of IE-Col-Ap and E-Col-Ap scaffolds.

IE-Col-Ap and E-Col-Ap scaffolds exhibit similar XRD spectra (Figures 6.4A and

6.4B). The broad peak at about 20° (2θ) is attributed to collagen. The peak at 26.2°

(2θ) corresponds to the (002) plane of HA. The peak at 29.4° is assigned to the (210)

plane. The broad peak from 30.7-35° is an overlap of three major reflections (112),

(211) and (300). The peak at about 39.6° corresponds to (310) plane of HA. The broad

peaks of 30.7-35° indicate that the apatite obtained is at the nanosized level and

poorly crystalline. [16]

Figure 6.4. XRD spectra of IE-Col-Ap (A) and E-Col-Ap (B) scaffolds.

Figure 6.5 shows the FT-IR spectra obtained from IE-Col-Ap and E-Col-Ap

scaffolds. Both IE-Col-Ap and E-Col-Ap scaffolds show similar spectra. The broad

peak around 3400 cm-1 is associated with the absorption of water in the scaffolds.

Both spectra show typical peaks of collagen: the peak around 1660 cm-1 is attributed

to the C=O stretching vibration (amide I), while those at 1549 cm-1 and 1188 cm-1

arise from angular deformation of N-H bond (amide II) and vibration of C-N bond

(amide III). All of these peaks are typical major peaks for collagen. Peaks between

Page 160: Fabrication and Characterization of Collagen-Apatite

141

500-600 cm-1 are assigned to anti-symmetric bending vibration of the PO43- groups,

and peaks at 950 cm-1 and 1036 cm-1 are identified as the stretching vibration of PO43-

groups. Moreover, the peaks at 1410 cm-1 and 875 cm-1 are typical vibration bands of

carbonate apatite, which is similar to the mineral phase in natural bone. [58]

Figure 6.5. FT-IR spectra of IE-Col-Ap (A) and E-Col-Ap (B) scaffolds.

Figure 6.6 shows TGA measurements of E-Col-Ap and IE-Col-Ap scaffolds.

Both E-Col-Ap and IE-Col-Ap scaffolds have a mineral content of approximately 30

wt%. In the E-Col-Ap scaffolds, the first 7 % of weight loss occurs between

30-150 °C (peak at 71.2 °C), which corresponds to the evaporation of physisorbed

water on the scaffold. Decomposition of collagen molecules occurs between

200-650 °C (peaks at 308.9 °C and 544.5 °C), leaving a residual 30 wt% which are

the minerals obtained (Figure 6.6B). In the IE-Col-Ap scaffolds, the initial weight loss

(10 wt%) occurs between 30-150 °C (peak at 50 °C), and the decomposition of

collagen molecules occurs between 150-800 °C (peaks at 227.7 °C, 306.6 °C and

Page 161: Fabrication and Characterization of Collagen-Apatite

142

576.2 °C), which yields approximately 30.6 wt% of minerals (Figure 6.6A). From

Figure 6.6 the decomposition of collagen molecules in E-Col-Ap scaffolds started at a

higher temperature than that of IE-Col-Ap scaffolds. This may be due to the

extrafibrillar HA minerals existing on the surface of collagen fibrils, which in turn

protects the collagen fibrils from decomposition. However, a higher temperature was

needed to fully decompose the collagen in the IE-Col-Ap scaffolds, which may be due

to the association of intrafibrillar minerals with collagen fibrils.

Figure 6.6. TGA of IE-Col-Ap (A) and E-Col-Ap (B) scaffolds.

Compared to E-Col-Ap scaffolds fabricated in our previous study using a

co-precipitation method, similar composition and crystallinity of HA minerals were

obtained. However, oriented HA minerals were formed within collagen fibrils via this

bottom-up approach, which closely mimics the organization of mineralized collagen

fibrils in natural bone.

Substantial research has been performed on the mechanism of intrafibrillar

mineralization of collagen fibrils. [25, 43, 44, 56, 59] Two functional groups in NCPs

Page 162: Fabrication and Characterization of Collagen-Apatite

143

have been found to play vital roles in regulation of nucleation, dimension, and order

of HA deposition within natural bone. Acidic sequestrating motifs of NCPs bind to

calcium and inhibit the precipitation of calcium phosphate, forming stabilized-ACP

nanoprecursors. [60] The templating groups of NCPs regulate the deposition of HA

within gap zone of collagen fibrils through its excellent binding capacity for calcium

and its strong affinity for collagen fibrils. [37] Fluidic ACP nanoprecursors infiltrate

and are retained within the gap zone of collagen fibrils through electrostatic attraction,

capillary action and size exclusion. [61] Then the infiltrated ACP minerals mature into

HA within collagen fibrils.

However, it is still not clear why the polymer-induced liquid precursor droplets

are more likely to infiltrate collagen fibrils instead of precipitating on the surface of

collagen fibrils. If only size exclusion and electrostatic attraction played a role in

mineralization, then more minerals should have deposited on the surface of collagen

fibrils since there is more room on the surface for nucleation than within the collagen

fibers. The electrostatic attraction is presumably the same, since charged groups are

located on both the interior and exterior of collagen fibers. [35, 56] Thus, more

minerals should have deposited on the surface of collagen fibrils. However, this was

not observed in our study. Herein, we propose that there are other driving forces

playing crucial roles in directing the deposition and maturation of HA minerals within

collagen fibrils: the stereochemical match between the interior of collagen fibrils and

ACP precursors. On one hand, the charged functional groups of collagen molecules

are located and oriented spatially within collagen fibrils to increase the likelihood of

Page 163: Fabrication and Characterization of Collagen-Apatite

144

calcium and phosphate ions binding to the fibrils, leading to the nucleation of ACP

within collagen fibrils. This has been confirmed by Xu et al. by applying all-atom

Hamiltonian replica exchange molecular dynamics simulation. [56] On the other hand,

calcium ions induce conformation changes of many proteins. [21, 62] Bound calcium

ions to dentin matrix protein 1 (DMP 1) induce a conformation change resulting in

self-assembly into elongated dimers. [63] Osteocalcin also undergoes a conformation

change when calcium binds. The conformation change resembles the interatomic

locations of calcium in the HA lattice. [21, 64] Thus, conformation changes of

proteins occur after binding with calcium to form ACP nanoprecursors, which leads to

the re-arrangement of calcium and phosphate ions into a structure resembling HA.

This re-arrangement of calcium and phosphate ions decreases the activation energy

for the transformation of ACP into HA. Moreover, the conformation change helps

NCPs interact specifically and spatially with certain domains of collagen sequences,

which leads to the spatial deposition of minerals within collagen fibrils. [65] Thus, we

propose that HA minerals form within collagen fibrils due to two factors: (1) The

stereochemical match of NCPs with the interior functional groups of collagen fibrils,

and (2) the conformation change of NCPs decreases the activation energy of

nucleation within collagen fibrils and regulates HA orientation. [65, 66] Meanwhile,

the work by other research groups also proposed the stereochemical template function

of collagen fibrils, which is conducive to nucleation, growth and development of

apatite crystals within collagen fibrils. [67, 68] A similar mechanism has been

proposed that specific molecular recognition between DMP1 and apatite surface aids

Page 164: Fabrication and Characterization of Collagen-Apatite

145

in the controlled growth of HA crystals with oriented crystallography. [63] Thus,

PAA-ACP nanoprecursors penetrate the gap zone of collagen fibrils via capillary

action and electrostatic attraction. Subsequently, both spatial interactions between

collagen molecules and ACP nanoprecursors along with preferential association of

NCPs on certain HA crystallographic planes play important roles in the orderly

alignment of HA within collagen fibrils. [65, 69]

Figure 6.7. Proliferation of MC3T3-E1 cells on collagen, E-Col-Ap, and IE-Col-Ap

scaffolds over the course of 7 days. Data is represented as the mean ± standard error.

# denotes statistical significance (p < 0.05) between collagen scaffolds and

experimental scaffolds.

The biocompatibility of the scaffolds was assessed through the proliferation of

MC3T3-E1 cells on E-Col-Ap and IE-Col-Ap scaffolds compared to cell proliferation

on pure collagen scaffolds which were used as a control, and the results are depicted

in Figure 6.7. All scaffolds supported cell growth over the course of 7 days. Cell

Page 165: Fabrication and Characterization of Collagen-Apatite

146

proliferation was the greatest on collagen scaffolds and statistically significant from

mineralized scaffolds. The difference in mineralization strategies did not result in a

statistically significant difference in proliferation between E-Col-Ap and IE-Col-Ap.

Despite the lower rate of cell proliferation on E-Col-Ap and IE-Col-Ap scaffolds

compared to collagen scaffolds, the mineralized scaffolds are a suitable material for

bone tissue engineering. The in vitro performance of a material is not always

indicative of its in vivo performance. Calcium phosphate materials are well known for

their excellent biocompatibility and in vivo performance, but these materials have

been shown to have inhibitory effects on osteoblasts in vitro. [70-73] Lyons et al.

observed inhibited cell proliferation on a biomimetic collagen-calcium phosphate

scaffold compared to a collagen-glycosaminoglycan scaffold in vitro but observed a

better response in vivo with the collagen-calcium phosphate scaffold. [73] Minardi et

al. also observed inhibited cell proliferation of hBM-MSC on a collagen HA

composite in vitro compared to a pure collagen scaffold. [71] The presence of the

biomimetic apatite phase was found to support the expression of osteogenic genes

better than the pure collagen scaffold. [74] E-Col-Ap scaffolds fabricated with the

described technique have previously been shown to have excellent in vivo

performance. [15] Future studies to compare the in vivo performance of IE-Col-Ap to

E-Col-Ap scaffolds will be conducted to determine the role of intrafibrillar

mineralization on the scaffolds bone forming ability.

6.4 Conclusions

A biomimetic Col-HA scaffold mimicking the composition and structure of

Page 166: Fabrication and Characterization of Collagen-Apatite

147

natural bone has been successfully prepared via a bottom-up approach. Compared

with E-Col-Ap scaffolds prepared using the co-precipitation approach in our previous

study, IE-Col-Ap scaffolds showed similar multi-lamellar structure and mineral

composition, which were confirmed by FESEM, XRD and FT-IR examinations.

However, the TEM and SAED results indicated that HA minerals in IE-Col-Ap

scaffolds were deposited within collagen fibrils and oriented along the long axis of

these fibrils, which bears higher resemblance to natural bone than the scaffold

fabricated using the co-precipitation approach. The biomimetic Col-HA scaffolds

demonstrated good biocompatibility in vitro with comparable cellular proliferation for

both E-Col-Ap and IE-Col-Ap scaffolds. This biomimetic IE-Col-AP scaffold can be

a promising biomaterial for bone tissue engineering applications.

Page 167: Fabrication and Characterization of Collagen-Apatite

148

References:

[1] Z. Xia, M. Wei. Biomimetic fabrication of collagen-apatite scaffolds for bone

tissue regeneration. Journal of Biomaterials and Tissue Engineering. 2013; 3:369-384.

[2] L. Meinel, V. Karageorgiou, R. Fajardo, B. Snyder, V. Shinde-Patil, L. Zichner, et

al. Bone tissue engineering using human mesenchymal stem cells: effects of scaffold

material and medium flow. Annal of Biomedical Engineering. 2004; 32:112-122.

[3] A. A. Al‐Munajjed, N. A. Plunkett, J. P. Gleeson, T. Weber, C. Jungreuthmayer, T.

Levingstone, et al. Development of a biomimetic collagen‐hydroxyapatite scaffold for

bone tissue engineering using a SBF immersion technique. Journal of Biomedical

Materials Research Part B: Applied Biomaterials. 2009; 90:584-591.

[4] J. R. Porter, T. T. Ruckh, K. C. Popat. Bone tissue engineering: a review in bone

biomimetics and drug delivery strategies. Biotechnology Progress. 2009;

25:1539-1560.

[5] C. Szpalski, M. Barbaro, F. Sagebin, S. M. Warren. Bone tissue engineering:

current strategies and techniques—part II: cell types. Tissue Engineering Part B:

Reviews. 2012; 18:258-269.

[6] J. O. Smith, A. Aarvold, E. R. Tayton, D. G. Dunlop, R. O. Oreffo. Skeletal tissue

regeneration: current approaches, challenges, and novel reconstructive strategies for

an aging population. Tissue Engineering Part B: Reviews. 2011; 17:307-320.

[7] S. Liao, F. Cui, W. Zhang, Q. Feng. Hierarchically biomimetic bone scaffold

materials: nano‐HA/collagen/PLA composite. Journal of Biomedical Materials

Research Part B: Applied Biomaterials. 2004; 69:158-165.

Page 168: Fabrication and Characterization of Collagen-Apatite

149

[8] L. H. Nguyen, N. Annabi, M. Nikkhah, H. Bae, L. Binan, S. Park, et al.

Vascularized bone tissue engineering: approaches for potential improvement. Tissue

Engineering Part B: Reviews. 2012; 18:363-382.

[9] I. Marcos-Campos, D. Marolt, P. Petridis, S. Bhumiratana, D. Schmidt, G.

Vunjak-Novakovic. Bone scaffold architecture modulates the development of

mineralized bone matrix by human embryonic stem cells. Biomaterials. 2012;

33:8329-8342.

[10] X. Su, K. Sun, F. Cui, W. Landis. Organization of apatite crystals in human

woven bone. Bone. 2003; 32:150-162.

[11] T. Dvir, B. P. Timko, D. S. Kohane, R. Langer. Nanotechnological strategies for

engineering complex tissues. Nature nanotechnology. 2011; 6:13-22.

[12] S. Yunoki, H. Sugiura, T. Ikoma, E. Kondo, K. Yasuda, J. Tanaka. Effects of

increased collagen-matrix density on the mechanical properties and in vivo

absorbability of hydroxyapatite–collagen composites as artificial bone materials.

Biomedical Materials. 2011; 6:015012.

[13] Z. Xia, X. Yu, X. Jiang, H. D. Brody, D. W. Rowe, M. Wei. Fabrication and

characterization of biomimetic collagen–apatite scaffolds with tunable structures for

bone tissue engineering. Acta Biomaterialia. 2013; 9:7308-7319.

[14] J. M. Holzwarth, P. X. Ma. Biomimetic nanofibrous scaffolds for bone tissue

engineering. Biomaterials. 2011; 32:9622-9629.

[15] X. Yu, Z. Xia, L. Wang, F. Peng, X. Jiang, J. Huang, et al. Controlling the

structural organization of regenerated bone by tailoring tissue engineering scaffold

Page 169: Fabrication and Characterization of Collagen-Apatite

150

architecture. Journal of Materials Chemistry. 2012; 22:9721-9730.

[16] Z. Xia, M. Villa, M. Wei. A biomimetic collagen–apatite scaffold with a

multi-level lamellar structure for bone tissue engineering. Journal of Materials

Chemistry B. 2014; 2:1998-2007.

[17] H. S. Yang, W. G. La, S. H. Bhang, T. J. Lee, M. Lee, B. S. Kim. Apatite-coated

collagen scaffold for bone morphogenetic protein-2 delivery. Tissue Engineering Part

A. 2011; 17:2153-2164.

[18] Y. Pek, S. Gao, M. M. Arshad, K. J. Leck, J. Y. Ying. Porous collagen-apatite

nanocomposite foams as bone regeneration scaffolds. Biomaterials. 2008;

29:4300-4305.

[19] M. M. Villa, L. Wang, D. W. Rowe, M. Wei. Effects of Cell-Attachment and

Extracellular Matrix on Bone Formation In Vivo in Collagen-Hydroxyapatite

Scaffolds. PLoS One. 2014; 9:e109568.

[20] S. Weiner, H. D. Wagner. The material bone: structure-mechanical function

relations. Annual Review of Materials Science. 1998; 28:271-298.

[21] L. Chen, R. Jacquet, E. Lowder, W. J. Landis. Refinement of collagen–mineral

interaction: A possible role for osteocalcin in apatite crystal nucleation, growth and

development. Bone. 2015; 71:7-16.

[22] Y. Wang, T. Azaïs, M. Robin, A. Vallée, C. Catania, P. Legriel, et al. The

predominant role of collagen in the nucleation, growth, structure and orientation of

bone apatite. Nature materials. 2012; 11:724-733.

[23] X. Wang, F. Cui, J. Ge, Y. Wang. Hierarchical structural comparisons of bones

Page 170: Fabrication and Characterization of Collagen-Apatite

151

from wild-type and liliput dtc232 gene-mutated Zebrafish. Journal of Structural

Biology. 2004; 145:236-245.

[24] F. Nudelman, P. H. Bomans, A. George, N. A. Sommerdijk. The role of the

amorphous phase on the biomimetic mineralization of collagen. Faraday Discuss.

2012; 159:357-370.

[25] F. Nudelman, K. Pieterse, A. George, P. H. Bomans, H. Friedrich, L. J. Brylka, et

al. The role of collagen in bone apatite formation in the presence of hydroxyapatite

nucleation inhibitors. Nature Materials. 2010; 9:1004-1009.

[26] W. Landis, M. Song, A. Leith, L. McEwen, B. McEwen. Mineral and organic

matrix interaction in normally calcifying tendon visualized in three dimensions by

high-voltage electron microscopic tomography and graphic image reconstruction.

Journal of Structural Biology. 1993; 110:39-54.

[27] T. T. Thula, D. E. Rodriguez, M. H. Lee, L. Pendi, J. Podschun, L. B. Gower. In

vitro mineralization of dense collagen substrates: a biomimetic approach toward the

development of bone-graft materials. Acta Biomaterialia. 2011; 7:3158-3169.

[28] W. Wagermaier, H. Gupta, A. Gourrier, M. Burghammer, P. Roschger, P. Fratzl.

Spiral twisting of fiber orientation inside bone lamellae. Biointerphases. 2006; 1:1-5.

[29] I. Jäger, P. Fratzl. Mineralized collagen fibrils: a mechanical model with a

staggered arrangement of mineral particles. Biophysical Journal. 2000; 79:1737-1746.

[30] M. J. Olszta, X. Cheng, S. S. Jee, R. Kumar, Y. Y. Kim, M. J. Kaufman, et al.

Bone structure and formation: a new perspective. Materials Science and Engineering:

R: Reports. 2007; 58:77-116.

Page 171: Fabrication and Characterization of Collagen-Apatite

152

[31] A. K. Nair, A. Gautieri, M. J. Buehler. Role of intrafibrillar collagen

mineralization in defining the compressive properties of nascent bone.

Biomacromolecules. 2014; 15:2494-2500.

[32] Y. Liu, D. Luo, X. X. Kou, X. D. Wang, F. R. Tay, Y. L. Sha, et al. Hierarchical

intrafibrillar nanocarbonated apatite assembly improves the nanomechanics and

cytocompatibility of mineralized collagen. Advanced Functional Materials. 2013;

23:1404-1411.

[33] R. B. Martin, D. B. Burr, N. A. Sharkey. Skeletal tissue mechanics. 1998.

[34] D. Toroian, J. E. Lim, P. A. Price. The Size Exclusion Characteristics of Type I

Collagen Implications for the Role of Noncollagenous Bone Constitutes in

Mineralization. Journal of Biological Chemistry. 2007; 282:22437-22447.

[35] D. E. Rodriguez, T. Thula-Mata, E. J. Toro, Y. W. Yeh, C. Holt, L. S. Holliday, et

al. Multifunctional role of osteopontin in directing intrafibrillar mineralization of

collagen and activation of osteoclasts. Acta Biomaterialia. 2014; 10:494-507.

[36] S. S. Jee, R. K. Kasinath, E. DiMasi, Y. Y. Kim, L. Gower. Oriented

hydroxyapatite in turkey tendon mineralized via the polymer-induced liquid-precursor

(PILP) process. CrystEngComm. 2011; 13:2077-2083.

[37] Y. K. Kim, L. Gu, T. E. Bryan, J. R. Kim, L. Chen, Y. Liu, et al. Mineralisation of

reconstituted collagen using polyvinylphosphonic acid/polyacrylic acid templating

matrix protein analogues in the presence of calcium, phosphate and hydroxyl ions.

Biomaterials. 2010; 31:6618-6627.

[38] J. Mahamid, B. Aichmayer, E. Shimoni, R. Ziblat, C. Li, S. Siegel, et al.

Page 172: Fabrication and Characterization of Collagen-Apatite

153

Mapping amorphous calcium phosphate transformation into crystalline mineral from

the cell to the bone in zebrafish fin rays. Proceedings of the National Academy of

Sciences. 2010; 107:6316-6321.

[39] B. Antebi, X. Cheng, J. N. Harris, L. B. Gower, X. D. Chen, J. Ling. Biomimetic

collagen–hydroxyapatite composite fabricated via a novel perfusion-flow

mineralization technique. Tissue Engineering Part C: Methods. 2013; 19:487-496.

[40] J. Li, J. Yang, J. Li, L. Chen, K. Liang, W. Wu, et al. Bioinspired intrafibrillar

mineralization of human dentine by PAMAM dendrimer. Biomaterials. 2013;

34:6738-6747.

[41] B. Marelli, C. E. Ghezzi, A. Alessandrino, J. E. Barralet, G. Freddi, S. N. Nazhat.

Silk fibroin derived polypeptide-induced biomineralization of collagen. Biomaterials.

2012; 33:102-108.

[42] N. Li, L. Niu, Y. Qi, C. K. Yiu, H. Ryou, D. D. Arola, et al. Subtleties of

biomineralisation revealed by manipulation of the eggshell membrane. Biomaterials.

2011; 32:8743-8752.

[43] S. S. Jee, T. T. Thula, L. B. Gower. Development of bone-like composites via the

polymer-induced liquid-precursor (PILP) process. Part 1: Influence of polymer

molecular weight. Acta Biomaterialia. 2010; 6:3676-3686.

[44] Y. Liu, Y. K. Kim, L. Dai, N. Li, S. O. Khan, D. H. Pashley, et al. Hierarchical

and non-hierarchical mineralisation of collagen. Biomaterials. 2011; 32:1291-1300.

[45] R. Q. Song, H. Colfen, A. W. Xu, J. Hartmann, M. Antonietti.

Polyelectrolyte-directed nanoparticle aggregation: systematic morphogenesis of

Page 173: Fabrication and Characterization of Collagen-Apatite

154

calcium carbonate by nonclassical crystallization. Acs Nano. 2009; 3:1966-1978.

[46] Y. Liu, S. Mai, N. Li, C. K. Yiu, J. Mao, D. H. Pashley, et al. Differences between

top-down and bottom-up approaches in mineralizing thick, partially demineralized

collagen scaffolds. Acta biomaterialia. 2011; 7:1742-1751.

[47] N. Rajan, J. Habermehl, M. F. Coté, C. J. Doillon, D. Mantovani. Preparation of

ready-to-use, storable and reconstituted type I collagen from rat tail tendon for tissue

engineering applications. Nature Protocol. 2006; 1:2753-2758.

[48] Z. Xia, X. Yu. Biomimetic collagen/apatite coating formation on Ti6Al4V

substrates. Journal of Biomedical Materials Research Part B: Applied Biomaterials.

2012; 100:871-881.

[49] H. Qu, Z. Xia, D. A. Knecht, M. Wei. Synthesis of dense collagen/apatite

composites using a biomimetic method. Journal of American Ceramic Society. 2008;

91:3211-3215.

[50] X. Yu, L. Wang, F. Peng, X. Jiang, Z. Xia, J. Huang, et al. The effect of fresh

bone marrow cells on reconstruction of mouse calvarial defect combined with

calvarial osteoprogenitor cells and collagen–apatite scaffold. Journal of Tissue

Engineering and Regenerative Medicine. 2013; 7:974-983.

[51] M. Gelinsky, P. Welzel, P. Simon, A. Bernhardt, U. König. Porous

three-dimensional scaffolds made of mineralised collagen: preparation and properties

of a biomimetic nanocomposite material for tissue engineering of bone. Chemical

Engineering Journal. 2008; 137:84-96.

[52] X. Duan, C. McLaughlin, M. Griffith, H. Sheardown. Biofunctionalization of

Page 174: Fabrication and Characterization of Collagen-Apatite

155

collagen for improved biological response: scaffolds for corneal tissue engineering.

Biomaterials. 2007; 28:78-88.

[53] L. Zhang. C. Hu, Mei Wei. Development of biomimetic scaffold woth both

intrafibrilar and extrafibrillar mineralization. ACS Biomaterials Science &

Engineering. 2015; 1:669-676.

[54] C. Hu, W. Cui. Microscopy of Biopolymer Systems. Handbook of

Biopolymer-Based Materials: From Blends and Composites to Gels and Complex

Networks. 2013; 611-644.

[55] Q. Wang, X. Wang, L. Tian, Z. Cheng, F. Cui. In situ remineralizaiton of partially

demineralized human dentine mediated by a biomimetic non-collagen peptide. Soft

Matter. 2011; 7:9673-9680.

[56] Z. Xu, Y. Yang, W. Zhao, Z. Wang, W. J. Landis, Q. Cui, et al. Molecular

mechanisms for intrafibrillar collagen mineralization in skeletal tissues. Biomaterials.

2015; 39:59-66.

[57] N. Nassif, F. Gobeaux, J. Seto, E. Belamie, P. Davidson, P. Panine, et al.

Self-assembled collagen− apatite matrix with bone-like hierarchy. Chemistry of

Materials. 2010; 22:3307-3309.

[58] G. Piga, A. Santos-Cubedo, A. Brunetti, M. Piccinini, A. Malgosa, E. Napolitano,

et al. A multi-technique approach by XRD, XRF, FT-IR to characterize the diagenesis

of dinosaur bones from Spain. Palaeogeography, Palaeoclimatology, Palaeoecology.

2011; 310:92-107.

[59] F. R. Tay, D. H. Pashley. Guided tissue remineralisation of partially

Page 175: Fabrication and Characterization of Collagen-Apatite

156

demineralised human dentine. Biomaterials. 2008; 29:1127-1137.

[60] H. Ozawa, K. Hoshi, N. Amizuka. Current concepts of bone biomineralization.

Journal of Oral Bioscience. 2008; 50:1-14.

[61] L. Gu, Y. K. Kim, Y. Liu, K. Takahashi, S. Arun, C. E. Wimmer, et al.

Immobilization of a phosphonated analog of matrix phosphoproteins within

cross-linked collagen as a templating mechanism for biomimetic mineralization. Acta

Biomaterialia. 2011; 7:268-277.

[62] S. R. Shi, R. J. Cote, D. Hawes, S. Thu, Y. Shi, L. L. Young, et al.

Calcium-induced modification of protein conformation demonstrated by

immunohistochemistry: What is the signal? J Histochem Cytochem. 1999;

47:463-469.

[63] G. He, S. Gajjeraman, D. Schultz, D. Cookson, C. Qin, W. T. Butler, et al.

Spatially and temporally controlled biomineralization is facilitated by interaction

between self-assembled dentin matrix protein 1 and calcium phosphate nuclei in

solution. Biochemistry. 2005; 44:16140-16148.

[64] P. V. Hauschka, S. A. Carr. Calcium-dependent. alpha.-helical structure in

osteocalcin. Biochemistry. 1982; 21:2538-2547.

[65] E. Beniash. Biominerals—hierarchical nanocomposites: the example of bone.

Wiley Interdisciplinary Reviews: Nanomedicine and Nanobiotechnology. 2011;

3:47-69.

[66] G. He, T. Dahl, A. Veis, A. George. Nucleation of apatite crystals in vitro by

self-assembled dentin matrix protein 1. Nature Materials. 2003; 2:552-558.

Page 176: Fabrication and Characterization of Collagen-Apatite

157

[67] W. Stetler-Stevenson, A. Veis. Bovine dentin phosphophoryn: calcium ion

binding properties of a high molecular weight preparation. Calcified Tissue

International. 1987; 40:97-102.

[68] W. J. Landis, F. H. Silver. Mineral deposition in the extracellular matrices of

vertebrate tissues: identification of possible apatite nucleation sites on type I collagen.

Cells Tissues Organs. 2009; 189:20-24.

[69] S. S. Jee, L. Culver, Y. Li, E. P. Douglas, L. B. Gower. Biomimetic

mineralization of collagen via an enzyme-aided PILP process. Journal of Crystal

Growth. 2010; 312:1249-1256.

[70] Y. F. Chou, W. Huang, J. C. Dunn, T. A. Miller, B. M. Wu. The effect of

biomimetic apatite structure on osteoblast viability, proliferation, and gene expression.

Biomaterials. 2005; 26:285-295.

[71] J. B. Nebe, L. Müller, F. Lüthen, A. Ewald, C. Bergemann, E. Conforto, et al.

Osteoblast response to biomimetically altered titanium surfaces. Acta Biomaterialia.

2008; 4:1985-1995.

[72] R. O. Oreffo, F. C. Driessens, J. A. Planell, J. T. Triffitt. Growth and

differentiation of human bone marrow osteoprogenitors on novel calcium phosphate

cements. Biomaterials. 1998; 19:1845-1854.

[73] F. G. Lyons, A. A. Al-Munajjed, S. M. Kieran, M. E. Toner, C. M. Murphy, G. P.

Duffy, et al. The healing of bony defects by cell-free collagen-based scaffolds

compared to stem cell-seeded tissue engineered constructs. Biomaterials. 2010;3

1:9232-9243.

Page 177: Fabrication and Characterization of Collagen-Apatite

158

[74] S. Minardi, B. Corradetti, F. Taraballi, M. Sandri, J. Van Eps, F. Cabrera, et al.

Evaluation of the osteoinductive potential of a bio-inspired scaffold mimicking the

osteogenic niche, for bone augmentation. Biomaterials. 2015; 62:128-137.

Page 178: Fabrication and Characterization of Collagen-Apatite

159

7. Conclusions and future work

7.1 Conclusions

The objective of this study was to prepare and characterize biomimetic

collagen-apatite composite coatings and scaffolds for enhanced bone repair and

regeneration. In Chapter 2, a crack-free apatite coating was produced using a

biomimetic approach. Biomimetic apatite was deposited on the surface-treated

Ti-6Al-4V substrates by soaking them in the modified simulated body fluid (m-SBF)

for 24 hours. The apatite morphology could be adjusted by controlling the pH of the

m-SBF solution. Dual beam FIB/SEM was used to mill the cross-sections and prepare

TEM foils. Both cross-sectional SEM images and TEM observations confirmed that a

well-bonded HA coating has been formed on the Ti-6Al-4V substrate, and that the

biomimetic apatite coating possessed a unique gradient structure, with a dense

structure adjacent to the substrate to facilitate a strong bonding, and a porous structure

at the surface to allow bone cell attachment.

Furthermore, building on the success of preparation of biomimetic apatite

coatings, biomimetic collagen-apatite composite coatings were successfully fabricated

on the Ti-6Al-4V substrates to better mimic the composition of natural bone. The

coating was also formed by simply immersing the surface-treated Ti-6Al-4V substrate

into a collagen-containing m-SBF. The cross-sectional microstructure of the

collagen-apatite coating has been investigated using dual beam FIB/SEM. It has been

revealed that the collagen-apatite coating also consists of a gradient structure with a

Page 179: Fabrication and Characterization of Collagen-Apatite

160

dense a dense layer adjacent to the interface and a porous structure towards the

surface. Compared with apatite coating, collagen-apatite composite coating formed is

thinner and less porous than the apatite coating.

The cross-sectional microstructure of both the apatite coating and the

collagen-apatite composite coating has been investigated using dual beam FIB/SEM

with either gallium (GFIB) or xenon plasma (PFIB) ion source, and the capabilities of

these two dual beam FIB/SEM systems were also compared. Although GFIB and

PFIB share a lot of similarities, there are also significant differences, including the

optimal voltage of ion-beam-induced Pt deposition, the TEM sample preparation

process, and the efficiency and the quality of the cross-sectional cuts produced.

Particularly, the GFIB is more beneficial for preparing the high-quality final finish of

the sample, while the PFIB is more efficient for creating a large cut in the sample.

Inspired by the composition and structures of natural bone, intrafibrillar

calcification of collagen fibrils have been successfully produced using poly (acrylic

acid) as a sequestration analog and sodium tripolyphosphate as a templating analog.

First, the effect of the sequestration analog on the intrafibrillar mineralization has

been investigated. Based on the success of intrafibrillar calcification of collagen

fibrils, intrafibrillar silicification of collagen fibrils has been successfully produced by

a one-step collagen self-assembly/silicification approach, in which both the collagen

fibril self-assembly and the intrafibrillar silicification occurred simultaneously. As the

positive analog of the zwitterionic proteins in silicification, poly (allylamine

hydrochloride) promotes silica precipitation in the presence of a negative analog, such

Page 180: Fabrication and Characterization of Collagen-Apatite

161

as sodium tripolyphosphate. The structure of the silicified collagen fibrils can be

manipulated by varying silica precursors, polyanions and silica/collagen ratios to

produce silicified collagen fibrils with a core-shell, twisted, or banded structure.

Further, the mechanism of the collagen silicification has also been explored. The

intrafibrillar silicified collagen fibrils possess better cell compatibility compared with

the collagen-apatite fibrils.

A collagen-apatite scaffold consisting of intrafibrillar mineralized collagen fibrils

has been successfully prepared via a bottom-up approach, closely mimicking the

composition and structure of natural bone. Such prepared scaffolds demonstrate a

hierarchical structure at nano-, micro- and macro-levels, bearing striking resemblance

to natural bone. The results indicated that at the nanoscale level, the apatite minerals

deposited within the gap zone of collagen fibrils and oriented along the long axis of

these fibrils; At the microscale level, the aligned thin multi-lamella formed a thick

layer; At the macroscale level, the scaffolds consisted of interconnected macropores

throughout the entire scaffold. This biomimetic collagen-apatite scaffold with high

resemblance to natural bone can be a promising biomaterial for future bone tissue

engineering applications.

7.2 Future work

Great progresses have been made to produce intrafibrillar calcification and

silicification of collagen fibrils, but there are still many challenges remaining, for

examples: (1) the apatite mineral content in the intrafibrillar mineralized

collagen-apatite scaffolds is much lower than that in the natural bone; (2) the lack of

Page 181: Fabrication and Characterization of Collagen-Apatite

162

osteoinductivity of the scaffolds.

Silicon-containing biomaterials have been found to be osteoinductive for new

bone formation and stimulative for neovascularization. Intrafibrillar silification of

collagen fibrils have been successfully developed in our study. In order to impart the

osteoinductivity to the collagen-apatite scaffolds, a scaffold with both intrafibrillar

apatite and silica have been produced. Figure 7.1 shows the TEM images of

intrafibrillar mineralized collagen fibrils in the collagen-apatite, collagen-silica and

collagen-apatite-silica hydrogels. At the initial stage of scaffold preparation, a

hydrogel consisting of apatite and/or silica mineralized collagen fibrils was formed

using a two-temperature process (Figure 7.1). In Figure 7.1, clear D-banding patternd

are observed in the unstained collagen fibrils, indicating electron-dense minerals have

deposited within the gap zone of collagen fibrils, reproducing the D-banding pattern

of naturally mineralized collagen fibrils normally observed in bone. As discussed

above, sodium tripolyphosphate was applied as the templating analog to direct the

deposit of calcium phosphate (Figure 7.1A) and silica (Figure 7.1B) within gap zone

of collagen fibrils, and then composite minerals of calcium phosphate and silica were

deposited within the collagen fibrils (Figure 7.1C).

Further characterization needs to be conducted to prove the application of the

intrafibrillar collagen-apatite-silica composite scaffolds in bone tissue engineering. In

vitro cell culture studies are essential for initial assessing the biocompatibility of the

material, because they help predict how the material will interact with cells in vivo.

Cell culture studies will be conducted using MC3T3-E1 cells, an osteoblast precursor.

Page 182: Fabrication and Characterization of Collagen-Apatite

163

Intrafibrillar silica, apatite and apatite/silica mineralized collagen scaffolds will be

prepared to assess the viability and proliferation of cells, which will be monitored

through the use of MTT dye.

Figure 7.1. TEM images of a small piece of hydrogel composed of intrafibrillar

apatite (A), silica (B) and apatite/silica (C) mineralized collagen fibrils. The

mineralized collagen fibrils were unstained.

Page 183: Fabrication and Characterization of Collagen-Apatite

164

Appendice A: List of publications

Publications

1. Changmin Hu, Le Yu, Mei Wei*. Sectioning studies of biomimetic

collagen-hydroxyapatite coatings on Ti-6Al-4V substrates using focused ion beam.

Paper in preparation.

2. Changmin Hu, Le Yu, Mei Wei*. Intrafibrillar silicification of collagen fibrils

through a one-step collagen self-assembly/silicification approach, RSC Advances.

2017: 55, 34624-34632.

3. Changmin Hu, Mark Aindow, Mei Wei*. Focused ion beam sectioning studies of

biomimetic hydroxyapatite coatings on Ti-6Al-4V substrates, Surface and Coatings

Technology. 2017: 313, 255-262.

4. Changmin Hu, Michael Zilm, Mei Wei*. Fabrication of intrafibrillar and

extrafibrillar mineralized collagen/apatite scaffolds with a hierarchical structure,

Journal of Biomedical Materials Research Part A. 2016: 5, 1153-1161.

5. Changmin Hu, Lichun Zhang, Mei Wei*. Development of biomimetic scaffolds

with both intrafibrillar and extrafibrillar mineralization, ACS Biomaterials Science &

Engineering. 2015: 8, 669-676.

Conference presentations

1. Changmin Hu, Mei Wei*. Sectioning biomimetic collagen-hydroxyapatite coatings

on Ti-6Al-4V substrates using focused ion beam. Society For Biomaterials 2017

Annual Meeting & Exposition. USA, MN, 2017.

2. Changmin Hu, Mark Aindow, Mei Wei*. A comparison of using plasma FIB and

Page 184: Fabrication and Characterization of Collagen-Apatite

165

Ga FIB to study porous hydroxyapatite coating on Ti6Al4V alloy. 2016 MRS Fall

Meeting & Exhibit, USA, MA, 2016.

3. Changmin Hu, Mei Wei*. Biomimetic silicification of collagen fibrils through a

one-step collagen self-assembly/silicification approach. 10th World Biomaterials

Congress, Canada, Montreal, 2016.

4. Changmin Hu, Lichun Zhang, Mei Wei*. Development of biomimetic scaffolds

with intrafibrillar and extrafibrillar mineralization. Society For Biomaterials 2015

Annual Meeting & Exposition. USA, NC, 2015.