Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

PEEK Biomaterials Handbook
PEEK Biomaterials Handbook
PEEK Biomaterials Handbook
Ebook799 pages10 hours

PEEK Biomaterials Handbook

Rating: 5 out of 5 stars

5/5

()

Read preview

About this ebook

PEEK biomaterials are currently used in thousands of spinal fusion patients around the world every year. Durability, biocompatibility and excellent resistance to aggressive sterilization procedures make PEEK a polymer of choice, replacing metal in orthopedic implants, from spinal implants and hip replacements to finger joints and dental implants.

This Handbook brings together experts in many different facets related to PEEK clinical performance as well as in the areas of materials science, tribology, and biology to provide a complete reference for specialists in the field of plastics, biomaterials, medical device design and surgical applications.

Steven Kurtz, author of the well respected UHMWPE Biomaterials Handbook and Director of the Implant Research Center at Drexel University, has developed a one-stop reference covering the processing and blending of PEEK, its properties and biotribology, and the expanding range of medical implants using PEEK: spinal implants, hip and knee replacement, etc.

  • Covering materials science, tribology and applications
  • Provides a complete reference for specialists in the field of plastics, biomaterials, biomedical engineering and medical device design and surgical applications
LanguageEnglish
Release dateOct 28, 2011
ISBN9781437744644
PEEK Biomaterials Handbook

Related to PEEK Biomaterials Handbook

Titles in the series (25)

View More

Related ebooks

Technology & Engineering For You

View More

Related articles

Reviews for PEEK Biomaterials Handbook

Rating: 5 out of 5 stars
5/5

1 rating0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    PEEK Biomaterials Handbook - Steven M. Kurtz

    Table of Contents

    Cover image

    Dedication

    Front Matter

    Copyright

    Foreword

    List of Contributors

    Chapter 1. An Overview of PEEK Biomaterials

    1.1. Introduction

    1.2. What Is a Polymer?

    1.3. What Is PEEK?

    1.4. Crystallinity and PEEK

    1.5. Thermal Transitions

    1.6. PEEK Composites

    1.7. Overview of This Handbook

    Chapter 2. Synthesis and Processing of PEEK for Surgical Implants

    2.1. Introduction

    2.2. Synthesis of PAEKs

    2.3. Nomenclature

    2.4. Quality Systems for Medical Grade Resin Production

    2.5. Processing of Medical Grade PEEK

    2.6. Machining

    2.7. Summary

    Chapter 3. Compounds and Composite Materials

    3.1. Introduction

    3.2. What Is a Composite Material?

    3.3. Additive Geometry, Volume, and Orientation Effects

    3.4. Preparation of Materials

    3.5. Processing to Make Parts

    3.6. Biocompatibility of CFR PEEK

    3.7. Summary and Conclusions

    Chapter 4. Morphology and Crystalline Architecture of Polyaryletherketones

    4.1. Introduction

    4.2. Chain Architecture and Packing

    4.3. Crystallization Behavior

    4.4. Characterization Techniques

    4.5. Structure Processing–Property Relationships

    4.6. Summary and Conclusions

    Chapter 5. Fracture, Fatigue, and Notch Behavior of PEEK

    5.1. Introduction

    5.2. Fracture and Fatigue of Materials

    5.3. PEEK Fracture Studies

    5.4. PEEK Notch Studies

    5.5. Summary

    Chapter 6. Chemical and Radiation Stability of PEEK

    6.1. Introduction to Chemical Stability

    6.2. Water Solubility

    6.3. Thermal Stability

    6.4. Steam Sterilization of PEEK

    6.5. Radiation Stability: Implications for Gamma Sterilization and Postirradiation Aging

    6.6. Summary

    Chapter 7. Biocompatibility of Polyaryletheretherketone Polymers

    7.1. Introduction

    7.2. Cell Culture and Toxicity Studies

    7.3. Mutagenesis (Genotoxicity)

    7.4. Immunogenesis

    7.5. Soft Tissue Response

    7.6. Osteocompatibility of PEEK Devices

    7.7. Biocompatibility of PEEK Particulate—X-STOP™ PEEK Explant Studies

    7.8. Summary and Conclusions

    Chapter 8. Bacterial Interactions with Polyaryletheretherketone

    8.1. Introduction

    8.2. Bacterial Adhesion to Biomaterials

    8.3. The Role of Surface Topography and Chemistry in Bacterial Adhesion

    8.4. Strategies to Reduce Bacterial Adhesion to PEEK

    8.5. Summary and Perspectives

    Chapter 9. Thermal Plasma Spray Deposition of Titanium and Hydroxyapatite on Polyaryletheretherketone Implants

    9.1. Introduction

    9.2. Coating Technology

    9.3. Biomedical Plasma-Sprayed Coatings

    9.4. Coating Analysis Methods

    9.5. Substrate Analysis Method

    9.6. Plasma-Sprayed Coatings on PEEK-Based Substrates

    9.7. Plasma-Sprayed Osteointegrative Surfaces for PEEK: The Eurocoating Experience

    9.8. Summary and Conclusions

    Chapter 10. Surface Modification Techniques of Polyetheretherketone, Including Plasma Surface Treatment

    10.1. PEEK–Tissue Interactions

    10.2. Surface Modification

    10.3. Surface Modification Techniques

    10.4. Applications of These Surface Modification Methods and the Translation to Industry

    10.5. Perspectives

    Chapter 11. Bioactive Polyaryletherketone Composites

    11.1. Introduction

    11.2. Processing–Structure Relationships

    11.3. Structure–Property Relationships

    11.4. Concluding Remarks

    Chapter 12. Porosity in Polyaryletheretherketone

    12.1. Introduction

    12.2. Porous Biomaterials in Existing Implants

    12.3. Porous Polymer Production for Industrial Applications

    12.4. Manufacturing of Porous PEEK Biomaterials

    12.5. Case Study 1—Porosity Through Porogen Leaching at Production Scale

    12.6. Case Study 2—Comparison of Small and Large Pore Sizes

    12.7. Case Study 3—Mid-Sized Porosity

    12.8. Conclusions

    Chapter 13. Applications of Polyaryletheretherketone in Spinal Implants

    13.1. Introduction

    13.2. Origins of Interbody Fusion and the Cage Rage of the Late 1990s

    13.3. CFR-PEEK Lumbar Cages: The Brantigan Cage

    13.4. Threaded PEEK Lumbar Fusion Cages

    13.5. Clinical Diagnostic Imaging of PEEK Spinal Cages and Transpedicular Screws

    13.6. Subsidence and Wear of PAEK Cages

    13.7. Posterior Dynamic Stabilization Devices

    13.8. Cervical and Lumbar Artificial Discs

    13.9. Summary

    Chapter 14. Isoelastic Polyaryletheretherketone Implants for Total Joint Replacement

    14.1. Introduction

    14.2. Incompatible Design Goals for an Uncemented Hip Stem

    14.3. Setbacks with Early Polymer–Metal Composite Hip Stems

    14.4. The Epoch Hip Stem

    14.5. Other PAEK Composite Hip Stems

    14.6. Stress Shielding in the Acetabulum

    14.7. PEEK in the Acetabulum

    14.8. Outlook for PEEK in Orthopedic Implants

    Chapter 15. Applications of Polyetheretherketone in Trauma, Arthroscopy, and Cranial Defect Repair

    15.1. Introduction

    15.2. Principles of Fracture Repair

    15.3. Principles of Arthroscopic Repair

    15.4. Principles of Craniofacial Defect Repair

    15.5. Summary

    Chapter 16. Arthroplasty Bearing Surfaces

    16.1. Introduction

    16.2. Total Hip and Knee Replacement

    16.3. Basic Biotribology Studies of PEEK Articulations

    16.4. Hip Resurfacing

    16.5. Mobile-Bearing, Unicondylar Knee Joint Replacements

    16.6. Other Total Joint Replacement Applications

    16.7. MOTIS: Medical Grade CFR-PEEK for Bearing Applications

    16.8. Summary and Concluding Remarks

    Chapter 17. FDA Regulation of Polyaryletheretherketone Implants

    17.1. Introduction

    17.2. What Is the FDA?

    17.3. Common Misconceptions About What the FDA Does

    17.4. Brief History of the FDA

    17.5. Medical Device Definition and Classification

    17.6. Regulatory Approval Process and Types of Applications

    17.7. Content of an FDA Application

    17.8. Material Considerations

    17.9. Current Uses of PEEK in FDA-Approved Spinal and Orthopedic Implants

    17.10. The Use of Master Files in Supplying Material Data for FDA Regulation

    17.11. The Use of Standards in FDA Regulation

    17.12. Summary and Conclusions

    Index

    Dedication

    For the patience and understanding of my wife Karen, and my children, Katie, Peter, Michael, Sophia, and Andrew, for allowing me to spend my increasingly scarce free time on this writing project.

    Front Matter

    PEEK Biomaterials Handbook

    Edited by

    Steven M. Kurtz

    Amsterdam • Boston • Heidelberg • London • New York • Oxford Paris • San Diego • San Francisco • Singapore • Sydney • Tokyo

    William Andrew is an imprint of Elsevier

    PLASTICS DESIGN LIBRARY (PDL)

    PDL HANDBOOK SERIES

    Series Editor: Sina Ebnesajjad, PhD

    President, FluoroConsultants Group, LLC

    Chadds Ford, PA, USA

    www.FluoroConsultants.com

    The PDL Handbook Series is aimed at a wide range of engineers and other professionals working in the plastics industry, and related sectors using plastics and adhesives.

    PDL is a series of data books, reference works and practical guides covering plastics engineering, applications, processing, and manufacturing, and applied aspects of polymer science, elastomers and adhesives.

    Recent titles in the series

    Sastri, Plastics in Medical Devices

    ISBN: 9780815520276

    McKeen, Fatigue and Tribological Properties of Plastics and Elastomers, Second Edition

    ISBN: 9780080964508

    Wagner, Multilayer Flexible Packaging

    ISBN: 9780815520214

    Chandrasekaran, Rubber Seals for Fluid and Hydraulic Systems

    ISBN: 9780815520757

    Tolinski, Additives for Polyolefins

    ISBN: 9780815520511

    McKeen, The Effect of Creep and Other Time Related Factors on Plastics and Elastomers, Second Edition

    ISBN: 9780815515852

    Ebnesajjad, Handbook of Adhesives and Surface Preparation

    ISBN: 9781437744613

    Grot, Fluorinated Ionomers, Second Edition

    ISBN: 9781437744576

    Mckeen: Permeability Properties of Plastics and Elastomers, Third Edition

    ISBN: 9781437734690

    To submit a new book proposal for the series, please contact

    Sina Ebnesajjad, Series Editor

    sina@FluoroConsultants.com

    or

    Matthew Deans, Senior Publisher

    m.deans@elsevier.com

    Copyright

    William Andrew is an imprint of Elsevier

    The Boulevard, Langford Lane, Kidlington, Oxford OX5 1GB, UK

    225 Wyman Street, Waltham, MA 02451, USA

    First edition 2012

    Copyright © 2012 Elsevier Inc. All rights reserved.

    No part of this publication may be reproduced, stored in a retrieval system or transmitted in any form or by any means electronic, mechanical, photocopying, recording or otherwise without the prior written permission of the publisher

    Permissions may be sought directly from Elsevier's Science & Technology Rights Department in Oxford, UK: phone (+44) (0) 1865 843830; fax (+44) (0) 1865 853333; email: permissions@elsevier.com. Alternatively you can submit your request online by visiting the Elsevier web site at http://elsevier.com/locate/permissions and selecting Obtaining permission to use Elsevier material

    Notice

    No responsibility is assumed by the publisher for any injury and/or damage to persons or property as a matter of products liability, negligence or otherwise, or from any use or operation of any methods, products, instructions or ideas contained in the material herein. Because of rapid advances in the medical sciences, in particular, independent verification of diagnoses and drug dosages should be made

    British Library Cataloguing in Publication Data

    A catalogue record for this book is available from the British Library

    Library of Congress Cataloging-in-Publication Data

    A catalog record for this book is availabe from the Library of Congress

    ISBN–13: 978-1-4377-4463-7

    For information on all William Andrew publications visit our web site at books.elsevier.com

    Printed and bound in USA

    12 13 14 15 1610 9 8 7 6 5 4 3 2 1

    Foreword

    Editor's Foreword

    The idea for researching the history and background of PEEK began in the late 1990s with my professional involvement with the development of medical devices, including pharmaceutical packaging, suture anchors, spine and hip implants. Throughout these projects, I was struck not only by the remarkable properties of this biomaterial, but also by the lack of a reference or clear guide for biomedical engineers and material scientists using PEEK in the design and development of medical devices. Information was available on PEEK in the literature, but it was scattered in polymer engineering, polymer science, materials science, clinical, and trade journals.

    In early 2006, I began the task of a comprehensive review of the scientific literature for PEEK. The goal of this review was to summarize the different silos of scientific information about PAEK polymers, and also to provide a historical context for PEEK biomaterials in clinical use today. The background reading for the review commenced in October 2006. For the next six months, my evenings and weekends were spent piecing together the milestones for PAEK polymers in the literature. The culmination in this immersion experience was the Biomaterials review article published in 2007, PEEK Biomaterials in Trauma, Orthopedic, and Spinal Implants. Although this review was rather extensive for a published article, there were many details we could not include due to space constraints. In addition, it was felt that the erudite style of the review might not be approachable for students or junior bioengineers who were interested in learning more about PEEK biomaterials.

    With this motivation, I began developing an online monograph about medical grade PEEK that would be accessible not only to scientists and medical device researchers, but also to students and bioengineers as well. The Medical PEEK Lexicon (www.medicalpeek.org) was born. Its online format was most appealing for its lack of space constraints and ease of expansion. The website was launched in the fall of 2007 with two distinct parts: the first part, an online reference, is updated on an as-needed basis; the second, a compendium of the latest scientific papers, conferences, and standards for medical grade PEEK, is regularly updated as new developments emerge.

    Feedback from the website was extremely encouraging, and site visitors began to inquire about the availability of a hardcopy book dedicated to PEEK biomaterials. The website is timely and dynamic, but it is also transient. A book, on the other hand, is a historical reference, establishing the state-of-the-art in the field of PEEK biomaterials at a specific point in time. I also wanted to cover a broader range of research topics for PEEK biomaterials in greater depth than could be accomplished in a website. Initial planning for the book started in late 2009, and work began in earnest in January 2010.

    The original review article, the website, and this Handbook were all made possible thanks to the strong collaboration with my friends from Invibio. For this book, I am likewise indebted to my fellow contributors, who are listed following the table of contents. Thank you for your efforts to make this Handbook a success.

    I also wish to thank many colleagues who helped flesh out the historical context for medical grade PEEK and agreed to be interviewed for the website and book. Thanks are due to Bill Christianson, for providing the regulatory background on the Brantigan cage; Robert Hastings and Stanley Brown, for helpful discussions about fracture fixation; Mike Hawkins, Hallie Brinkerhuff, and Michele Marcolongo for their feedback on composite stems; Michael Manley, Jim Nevelos, Eric Jones, Aiguo Wang, Martyn Elcocks, Richard Field, and Neil Rushton for collaboration on the MITCH PCR acetabular shell; and Jonathan Peck, Genevieve Hill, and Jove Graham from the US FDA for their patient tutelage regarding the regulatory pathway for PEEK biomaterials and related medical devices. Special thanks are also due to collaborators at Exponent and Drexel University, including David Jaekel, Maureen Reitman, Judd Day, Kevin Ong, Ryan Siskey, Chris Espinosa, and Marta Villarraga, who contributed to chapters, assisted with the preparation of figures, and otherwise provided editorial assistance.

    I am, as always, especially thankful for the patience and understanding of my wife Karen, and my children, Katie, Peter, Michael, Sophia, and Andrew, for allowing me to spend my increasingly scarce free time on this writing project.

    Steven M. Kurtz (email: skurtz@drexel.edu)

    On a train from Paris to Torino

    April 3, 2011

    List of Contributors

    Steven M. Kurtz, Ph.D.

    Corporate Vice President, Exponent, Inc., Research Professor, Drexel University, Philadelphia, PA, USA

    Mark Brady, Ph.D.

    Product Development Project Manager, Invibio Biomaterial Solutions, Thornton Cleveleys, United Kingdom

    Timothy L. Conrad, B.S.

    University of Notre Dame, Notre Dame, IN, USA

    Nicholas M. Cordaro, M.S.

    Director, Engineering, SeaSpine, San Diego, CA, USA

    Judd Day, Ph.D.

    Managing Scientist, Exponent, Inc., Philadelphia, PA, USA

    Jove Graham, Ph.D.

    Food and Drug Administration, Center for Devices and Radiological Health, Silver Spring, MD, USA

    Stuart Green, Ph.D.

    Global Technical Program Leader, Victrex Polymer Solutions, Thornton Cleveleys, United Kingdom

    Noreen J. Hickok, Ph.D.

    Associate Professor, Thomas Jefferson University, Philadelphia, PA, USA

    David Jaekel, Ph.D. Candidate

    Drexel University, Philadelphia, PA, USA

    Marcus Jarman-Smith, Ph.D.

    Technology Leader, Invibio Biomaterial Solutions, Thornton Cleveleys, United Kingdom

    Scott Lovald, Ph.D.

    Senior Associate, Exponent, Inc., Philadelphia, PA, USA

    T. Fintan Moriarty, Ph.D.

    Manager, Musculo-skeletal Infection, AO Research Institute Davos, Davos, Switzerland

    Jim Nevelos, Ph.D.

    Director, Hip Research, Stryker Orthopaedics, Mahwah, NJ

    Kevin Ong, Ph.D.

    Managing Engineer, Exponent, Inc., Philadelphia, PA, USA

    Jonathan Peck, M.E.

    Food and Drug Administration, Center for Devices and Radiological Health, Silver Spring, MD, USA

    Alexandra H.C. Poulsson, Ph.D.

    Post-Doctoral Research Associate, AO Research Institute Davos, Davos, Switzerland

    Maureen Reitman, Sc.D.

    Principal and Practice Director, Exponent, Inc., Bowie, MD, USA

    R. Geoff Richards, Ph.D.

    Professor and Director, AO Research Institute Davos, Davos, Switzerland

    Clare M. Rimnac, Ph.D.

    Wilbert J. Austin Professor of Engineering, Case Western Reserve University, Cleveland, OH, USA

    Pierfrancesco Robotti, MS

    Eurocoating S.p.A., Italy

    Edward T.J. Rochford

    AO Research Institute Davos, Davos, Switzerland

    Ryan K. Roeder, Ph.D.

    Associate Professor, University of Notre Dame, Notre Dame, IN, USA

    Ryan L. Siskey, M.S.

    Senior Manager, Exponent, Inc., Philadelphia, PA, USA

    Michael C. Sobieraj, M.D., Ph.D.

    NYU Hospital for Joint Diseases, New York, NY, USA

    Jeffrey M. Toth, BSE, Ph.D.

    Professor of Orthopaedic Surgery, The Medical College of Wisconsin, Milwaukee, WI, USA

    William R. Walsh, Ph.D.

    Professor & Director, Surgical & Orthopaedic Research Laboratories (SORL), University of New South Wales, Sydney, Australia

    Gianluca Zappini, MS

    Eurocoating S.p.A., Italy

    Chapter 1. An Overview of PEEK Biomaterials

    Steven M. Kurtz, Ph.D.

    1.1. Introduction

    Following confirmation of its biocompatibility two decades ago [1], polyaryletherketone polymers (PAEKs) have been increasingly employed as biomaterials for orthopedic, trauma, and spinal implants. Polyaryletheretherketone, commonly referred to as PEEK, is a member of the PAEK polymer family that has been used for orthopedic and spinal implants. Historically, the availability of PEEK arrived at a time when there was growing interest in the development of isoelastic hip stems and fracture fixation plates, with stiffnesses comparable with bone [2]. Although neat (unfilled) PEEK biomaterials can exhibit an elastic modulus ranging between 3 and 4 GPa, the modulus can be tailored to closely match cortical bone (18 GPa) or titanium alloy (110 GPa) by preparing carbon fiber-reinforced (CFR) composites with varying fiber length and orientation [2]. In the 1990s, researchers characterized the biocompatibility and in vivo stability of various PAEK materials, along with other high-performance engineering polymers, such as polysulfones and polybutylene terephthalate [3]. However, concerns were raised about the stress-induced cracking of polysulfones by lipids [4] and the use of these polymers in implants was subsequently abandoned.

    By the late 1990s, PEEK had emerged as the leading high-performance thermoplastic candidate for replacing metal implant components, especially in orthopedics [5] and [6] and trauma [7] and [8]. Not only was the material resistant to simulated in vivo degradation, including damage caused by lipid exposure, but starting in April 1998, PEEK was offered commercially as a biomaterial for implants (Invibio Ltd., Thornton Cleveleys, United Kingdom) [9]. Facilitated by a stable supply, research on PEEK biomaterials flourished and is expected to continue to advance in the future [10].

    Numerous studies documenting the successful clinical performance of PAEKs in orthopedic and spine patients continue to emerge in the literature [11], [12], [13], [14], [15] and [16]. Recent research has also investigated the biotribology of PEEK composites as bearing materials and flexible implants used for joint arthroplasty [17], [18], [19] and [20]. Because of the interest in further improving implant fixation, PEEK biomaterials research has also focused on compatibility of the polymer with bioactive materials, including hydroxyapatite, either as a composite filler or as a surface coating [21], [22], [23], [24] and [25]. As a result of ongoing biomaterials research, PEEK and related composites can be engineered today with a wide range of physical, mechanical, and surface properties, depending upon their implant application.

    The purpose of this Handbook is to introduce PEEK as an established member of the biomaterials armamentarium to students, engineers, and surgeons. Our aim is to cover the terminology, history, and recent advances related to its use in implantable devices for trauma, spine, and orthopedics. We hope that this monograph will serve two useful purposes. Our primary objective is to provide biomaterials researchers with a timely synthesis of the existing literature for PEEK, to help stimulate further studies to fill existing gaps in knowledge and experience. Our second goal is to provide the surgical community with state-of-the-art information about PEEK to facilitate accurate communications with patients.

    In this introductory chapter, we begin with the basics about polymers and PEEK. This chapter reviews basic information about polymers in general and describes the structure and composition of PEEK. The concepts of crystallinity and thermal transitions are introduced at a basic level. Readers familiar with these basic polymer concepts may want to consider skipping ahead to the next chapter.

    1.2. What Is a Polymer?

    PEEK belongs to a class of materials known as polymers or in lay terms more simply as plastics. More specifically, PEEK is classified as a linear homopolymer. Before proceeding to a definition of PEEK, it is helpful to first understand the significance of these italicized terms.

    The definition of polymer has its origins in the Greek, polumerēs, meaning having many parts. The repeating units, or monomer segments, of a polymer can all be the same. In such a case, we have a homopolymer (Fig. 1.1).

    When two or more different monomers are used, the resulting material is classified as a copolymer. However, PEEK is a homopolymer, and so throughout this chapter we will focus our attention on polymers having only a single monomer.

    Polymers may be linear or branched (Fig. 1.2). The tendency for branching in a homopolymer depends strongly on its synthesis conditions. The distinguishing feature of a polymer—as compared with a metal or ceramic—is its molecular size. In a polymer such as PEEK, the molecule is a linear chain of 100 monomer units with an average molecular weight of 80,000–120,000 g/mol.

    In general, the length and composition of the molecular chain result in many unique attributes for polymers, most notably the dependence of its properties on the temperature and rate at which deformations are applied. The rate and temperature sensitivity of polymers are strongly dependent on their chemical composition and structure. In other words, certain polymers are more rate and temperature sensitive than others.

    As we shall see in subsequent sections and chapters of this Handbook, when used for implants under clinically relevant conditions, PEEK is relatively insensitive to changes in rate and temperature. Further explanation of general polymer concepts can be found in the excellent textbook by Rodriguez [26].

    1.3. What Is PEEK?

    Commercialized for industry in the 1980s, PAEK is a family of high-performance thermoplastic polymers, consisting of an aromatic backbone molecular chain, interconnected by ketone and ether functional groups [27]. Thus, PEEK belongs to a larger family of PAEK polymers, sometimes referred to as polyetherketones (PEKs) or more simply as polyketones. The chemical formula of PEEK is shown in Fig. 1.3. Other members of the PAEK family that are considered for implants include PEK and polyetherketoneketone (PEKK), with chemical structures depicted in Fig. 1.4. PEEK is the dominant member of the PAEK polymer family for implant applications and is consequently the main focus of this Handbook.

    The chemical structure of PEEK, similar to its PAEK cousins, confers stability at high temperatures (exceeding 300 °C), resistance to chemical and radiation damage, compatibility with many reinforcing agents (such as glass and carbon fibers), and greater strength (on a per mass basis) than many metals, making it highly attractive in industrial applications, such as aircraft and turbine blades; for example, see Ref. [27] and [28]. Its stability, biocompatibility, radiolucency, and mechanical properties make PEEK a suitable biomaterial for orthopedic and spine implants.

    Unfilled neat PEEK is available as tan pellets or powder (Fig. 1.5), which can be converted into implant parts by standard polymer processing techniques, such as injection molding. PEEK implants are also fabricated by machining from extruded rods or compression molded sheets.

    Although neat PEEK has a tan appearance, when PEEK powder is blended with carbon fibers for added strength, the resulting material is black. For those interested in additional details about manufacturing, we explain the processing of PEEK for implants in Chapter 2 of this Handbook.

    1.4. Crystallinity and PEEK

    The molecular chain of PEEK may be visualized as a tangled strand of spaghetti that is hundreds of meters long. The molecular chain is not static but vibrates and rotates due to thermal energy or in response to an externally applied deformation. The PEEK molecule is relatively stiff because of the presence of the aromatic (benzene) rings along its backbone (Fig. 1.3); however, the molecule does have the freedom to rotate axially about the ether (–O–) bonds and ketone-carbon bonds (–CO–). When cooled slowly from the molten state, the molecular chain can rotate upon itself to form chain folds and to organize into ordered domains, known as crystals. PEEK crystals are embedded within amorphous (disordered) regions and form a two-phase microstructure (Fig. 1.6).

    PEEK conforms well to the conceptual model of a two-phase semicrystalline polymer, consisting of an amorphous phase and a crystalline phase. Similar to many semicrystalline polymers, including ultrahigh-molecular-weight polyethylene (UHMWPE), the crystalline content of PEEK varies depending upon its thermal processing history. The crystallinity of injection-molded PEEK in implants typically ranges from 30% to 35%[9]. By adjusting the cooling rate during fabrication of films, crystallization of PEEK can be greatly reduced, resulting in a nearly completely amorphous material. More details about the crystallinity, microstructure, and physical properties of PEEK can be found in Chapter 4 of this Handbook.

    1.5. Thermal Transitions

    As described in a previous section of this chapter, an important distinguishing feature of polymers is the temperature dependence of their properties. In general, upon heating many polymers undergo three major thermal transitions: the glass transition temperature (Tg), the melt temperature (Tm), and the flow temperature (Tf). As we shall see, PEEK components also exhibit a fourth transition, a recrystallization transition (Tc), depending upon how it was originally fabricated. In practical terms, all these melt transitions occur at temperatures far exceeding the boiling point of water and any clinical applications of the material. One of the characteristics of PEEK is its high temperature stability, and it is used in engine components for this reason. Although tangential to the clinical function of the material, some knowledge of its thermal behavior is crucial for materials scientists and engineers who are interested in producing PEEK implants, because thermal processing is a critical step in the manufacturing of PEEK components.

    The glass transition (Tg) is classically considered to be the temperature below which the polymer chains are supposed to behave like a brittle glass. Below Tg, the polymer chains have insufficient thermal energy to slide past one another, and the primary way for the material to respond to mechanical stress is by stretching (or rupture) of the covalent bonds constituting the molecular chain. With PEEK, the glass transition occurs around 143 °C. Ironically, although PEEK is below the glass transition at room and body temperatures, it is surprisingly ductile for a glassy polymer, as it is capable of elongations of up to 10–60%, depending upon the processing method and testing conditions.

    As we raise the temperature above Tg, the amorphous regions within the polymer gain increased mobility, and secondary intermolecular forces (e.g., van der Waals forces) can influence the flow and movement of the polymer chains. If the polymer sample was quickly cooled down from the melt during its previous history, when the temperature rises above Tg, there will be a thermodynamic tendency for the polymer to continue to form crystals or to recrystallize. The features of this transition provide clues to the materials scientist about how the material was previously processed; however, for implants, it has limited practical significance as the component will remain below Tg for its entire service life.

    When the temperature of PEEK rises above its recrystallization temperature, the smaller crystallites in the polymer begin to melt. The melting behavior of semicrystalline polymers, including PEEK, is typically measured using differential scanning calorimetry (DSC). DSC measures the amount of heat needed to increase the temperature of a polymer sample. Some representative DSC data for PEEK is shown in Fig. 1.7.

    The DSC trace for PEEK shows several key features. One feature of the DSC trace is its recrystallization peak, which for this annealed rod sample occurs around 150 °C, corresponding to the heat needed by the material to form crystals as it is heated above the glassy state. The glass transition temperature itself is difficult to discern from a conventional DSC trace; a specially modulated DSC analysis is usually needed to clearly demonstrate the presence of Tg.

    Another key feature of the DSC curve above Tc is the peak melting temperature (Tm), which for this sample occurs at 343 °C and corresponds to the point at which the majority of the crystalline regions have melted. The melt temperature reflects the thickness of the crystals, as well as their perfection. Thicker and more perfect PEEK crystals will tend to melt at a higher temperature than smaller crystals.

    As the temperature of a semicrystalline polymer is raised above the melt temperature (not shown on the DSC trace), it may undergo a flow transition and become liquid. PEEK undergoes a flow transition (Tf) around 390 °C and is typically processed at this temperature.

    1.6. PEEK Composites

    PEEK can be readily combined with certain additives to create a composite. A composite material is comprised of two or more distinct phases, each retaining unique physical, bioactive, and mechanical properties, bonded together by aninterface. The overall mechanical behavior of a composite is thus governed by the properties of the individual constituents and the interfaces between them. In the case of PEEK, the polymer is typically designed as the matrix of the composite and constitutes most of the volume in the polymer composite.

    As already alluded to previously in this chapter, carbon and glass fillers were among the first reinforcement additives for PEEK to increase its strength and stiffness [29]. PEEK forms a strong interface with carbon fibers, effectively transferring stress between the fibers and the polymer matrix (Fig. 1.8). The strength and modulus of carbon fiber-reinforced PEEK (CFR-PEEK) depend on the size, length, and orientation of the fibers. CFR-PEEK biomaterials are currently used in implants for spine fusion and joint replacement.

    PEEK biomaterials are also engineered for the biomedical, as well as their biomechanical, function. PEEK may be mixed with radiopacifiers, such as barium sulfate, to improve visualization and contrast in medical imaging. Image contrast grades of PEEK are commercially available for implant applications and are currently used in spinal implants.

    Researchers are also investigating the combination of PEEK and bioactive fillers, such as hydroxyapatite, to enhance bone growth around implants. Although structural and image contrast formulations of PEEK are relatively well understood, bioactive PEEK composites represent a novel field in biomaterials under active research and development. In Chapter 3 of this Handbook, we explore PEEK composites more thoroughly.

    1.7. Overview of This Handbook

    The primary goal of this Handbook is to provide a comprehensive, state-of-the-art assessment of PEEK and PEEK composites as a family of biomaterials. In recent years, advances in the processing and biomaterials applications of PEEK have been progressing steadily. Previously, much of the research on PEEK implants has been fragmented in the materials science, composites, biomaterials, and application-specific literature. Consequently, we have also sought to synthesize data from the materials science, polymers engineering, biomaterials, and clinical literature to make this information more readily available and to hopefully facilitate new research in this field.

    This Handbook is organized in three main sections. The first part of this Handbook provides the reader with a foundation in PEEK structure, properties, and behavior. As background for this monograph, we have provided in this introductory chapter an initial summary of polyaromatic ketones as the basis for understanding the chemical, physical, and mechanical properties of this family of polymeric biomaterials. Chapter 2 summarizes the techniques for processing PEEK and fabricating PEEK components, and Chapter 3 further covers the field of PEEK composites. Chapter 4, Chapter 5 and Chapter 6 describe the structure and morphology, fatigue and fracture behavior, and chemical and radiation stability of PEEK biomaterials.

    The second part of this Handbook summarizes the biocompatibility of PEEK and recent developments in the engineering of PEEK biocomposites. Chapter 7 provides an overview of studies in the literature analyzing the biocompatibility of PEEK. In Chapter 8, the interaction between PEEK and microbiological organisms is summarized. Surface modification of PEEK can be achieved using bioactive coatings (Chapter 9) or by plasma treatment (Chapter 10). Chapter 11 covers advances in the field of hydroxyapatite-PEEK biocomposites, and in Chapter 12, we describe efforts to introduce porosity into PEEK biomaterials for creating tissue scaffolds.

    The third part of this Handbook provides an overview of current applications of PEEK implants. We provide an overview of the clinical applications of PEEK and related polyaromatic ketones in fusion and motion preserving spine implants (Chapter 13), PEEK isoelastic hip stems (Chapter 14), trauma implants and suture anchors (Chapter 15), and orthopedic bearings (Chapter 16). The final chapter in this part of this Handbook summarizes the regulatory pathway for PEEK implants.

    In short, this Handbook is designed as a tour de force for the field of PEEK biomaterials, starting with the basics of device fabrication, engineering of the implant/host tissue interface, and culminating in development of new implants. In the chapter that follows, we explore the materials processing, as well as the blending and conversion practices that underlie the manufacture of PEEK components, and continue to build upon the conceptual foundation established in this introduction.

    References

    [1] Williams, D.F.; McNamara, A.; Turner, R.M., Potential of polyetheretherketone (PEEK) and carbon-fibre-reinforced PEEK in medical applications, J. Mater. Sci. Lett. 6 (1987) 188–190.

    [2] Skinner, H.B., Composite technology for total hip arthroplasty, Clin. Orthop. Relat. Res. 235 (1988) 224–236.

    [3] Brown, S.A.; Hastings, R.S.; Mason, J.J.; Moet, A., Characterization of short-fibre reinforced thermoplastics for fracture fixation devices, Biomaterials 11 (8) (1990) 541–547.

    [4] Trentacosta, J.D.; Cheban, J.C., In: Lipid sensitivity of poly-aryl-ether-ketones and polysulfone (1995) Transactions of the 41st Orthopedic Research Society, Orlando (FL), p. 783.

    [5] Liao, K., Performance characterization and modeling of a composite hip prosthesis, Exp. Tech. 18 (5) (1994) 33–38.

    [6] Maharaj, G.R.; Jamison, R.D., Intraoperative impact: characterization and laboratory simulation on composite hip prostheses, In: (Editors: Jamison, R.D.; Gilbertson, L.N.) STP 1178: Composite Materials for Implant Applications in the Human Body: Characterization and Testing (1993) ASTM, Philadelphia, pp. 98–108.

    [7] Kelsey, D.J.; Springer, G.S.; Goodman, S.B., Composite implant for bone replacement, J. Compos. Mater. 31 (16) (1997) 1593–1632.

    [8] Corvelli, A.A.; Biermann, P.J.; Roberts, J.C., Design, analysis, and fabrication of a composite segmental bone replacement implant, J. Adv. Mater. 28 (3) (1997) 2–8.

    [9] Green, S.M.; Schlegel, J., A polyaryletherketone biomaterial for use in medical implant applications, In: Polymers for the Medical Industry, Proceedings of a Conference held in Brussels (2001), pp. 1–7.

    [10] Williams, D., New horizons for thermoplastic polymers, Med. Device. Technol. 12 (4) (2001) 8–9.

    [11] Toth, J.M.; Wang, M.; Estes, B.T.; Scifert, J.L.; Seim 3rd, H.B.; Turner, A.S., Polyetheretherketone as a biomaterial for spinal applications, Biomaterials 27 (3) (2006) 324–334.

    [12] Brantigan, J.W.; Neidre, A.; Toohey, J.S., The Lumbar I/F Cage for posterior lumbar interbody fusion with the variable screw placement system: 10-year results of a Food and Drug Administration clinical trial, Spine J. 4 (6) (2004) 681–688.

    [13] Brantigan, J.W.; Steffee, A.D.; Lewis, M.L.; Quinn, L.M.; Persenaire, J.M., Lumbar interbody fusion using the Brantigan I/F cage for posterior lumbar interbody fusion and the variable pedicle screw placement system: two-year results from a Food and Drug Administration investigational device exemption clinical trial, Spine 25 (11) (2000) 1437–1446.

    [14] Akhavan, S.; Matthiesen, M.M.; Schulte, L.; Penoyar, T.; Kraay, M.J.; Rimnac, C.M.; et al., Clinical and histologic results related to a low-modulus composite total hip replacement stem, J. Bone Jt. Surg. 88 (6) (2006) 1308–1314.

    [15] Glassman, A.H.; Crowninshield, R.D.; Schenck, R.; Herberts, P., A low stiffness composite biologically fixed prosthesis, Clin. Orthop. Relat. Res. 393 (2001) 128–136.

    [16] Karrholm, J.; Anderberg, C.; Snorrason, F.; Thanner, J.; Langeland, N.; Malchau, H.; et al., Evaluation of a femoral stem with reduced stiffness. A randomized study with use of radiostereometry and bone densitometry, J. Bone Jt. Surg. Am. 84 (9) (2002) 1651–1658.

    [17] Wang, A.; Lin, R.; Stark, C.; Dumbleton, J.H., Suitability and limitations of carbon fiber reinforced PEEK composites as bearing surfaces for total joint replacements, Wear 225–229 (1999) 724–727.

    [18] Jones, E.; Wang, A.; Streicher, R., Validating the limits for a PEEK composite as an acetabular wear surface, In: Society for Biomaterials 27th Annual Meeting, St Paul, MN, April 24–29 (2001), p. 27.

    [19] Joyce, T.J.; Rieker, C.; Unsworth, A., Comparative in vitro wear testing of PEEK and UHMWPE capped metacarpophalangeal prostheses, Bio-Med. Mater. Eng. 16 (1) (2006) 1–10.

    [20] Manley, M.; Ong, K.; Kurtz, S.M.; Rushton, N.; Field, R.E., Biomechanics of a PEEK horseshoe-shaped cup: comparisons with a predicate deformable cup, In: 53rd Annual Meeting of the Orthopedic Research Society, San Diego, CA, February 11–14 (2007).

    [21] Yu, S.; Hariram, K.P.; Kumar, R.; Cheang, P.; Aik, K.K., In vitro apatite formation and its growth kinetics on hydroxyapatite/polyetheretherketone biocomposites, Biomaterials 26 (15) (2005) 2343–2352.

    [22] Fan, J.P.; Tsui, C.P.; Tang, C.Y.; Chow, C.L., Influence of interphase layer on the overall elasto-plastic behaviors of HA/PEEK biocomposite, Biomaterials 25 (23) (2004) 5363–5373.

    [23] Tan, K.H.; Chua, C.K.; Leong, K.F.; Cheah, C.M.; Cheang, P.; Abu Bakar, M.S.; et al., Scaffold development using selective laser sintering of polyetheretherketone-hydroxyapatite biocomposite blends, Biomaterials 24 (18) (2003) 3115–3123.

    [24] Abu Bakar, M.S.; Cheng, M.H.W.; Tang, S.M.; Yu, S.C.; Liao, K.; Tan, C.T.; et al., Tensile properties, tension-tension fatigue and biological response of polyetheretherketone-hydroxyapatite composites for load-bearing orthopedic implants, Biomaterials 24 (13) (2003) 2245–2250.

    [25] Ha, S.W.; Kirch, M.; Birchler, F.; Eckert, K.L.; Mayer, J.; Wintermantel, E.; et al., Surface activation of polyetheretherketone (PEEK) and formation of calcium phosphate coatings by precipitation, J. Mater. Sci. Mater. Med. 8 (11) (1997) 683–690.

    [26] Rodriguez, F., Principles of Polymer Systems. fifth ed. (2003) Taylor & Francis, London.

    [27] May, R., Polyetheretherketones, In: (Editors: Mark, H.F.; Bikales, N.M.; Overberger, C.G.; Menges, G.; Kroschiwitz, J.I.) Encyclopedia of Polymer Science and Engineering (1988) John Wiley and Sons, New York, pp. 313–320.

    [28] Rigby, R.B., Polyetheretherketone, In: (Editor: Margolis, J.M.) Engineering Thermoplastics: Properties and Applications (1985) Marcel Dekker, Inc., New York, pp. 299–314.

    [29] Jones, D.P.; Leach, D.C.; Moore, D.R., Mechanical properties of poly(ether-ether-ketone) for engineering applications, Polymer 26 (1985) 1385–1393.

    Chapter 2. Synthesis and Processing of PEEK for Surgical Implants

    Steven M. Kurtz, Ph.D.

    2.1. Introduction

    Polyaryletheretherketone (PEEK) is a challenge to synthesize and convert into surgical implants. The polymer is chemically inert and insoluble in all conventional solvents at room temperature. Indeed, PEEK can only be completely dissolved using fairly esoteric solvents, such as diaryl sulfones [1]. Although inertness and insolubility are desirable for a biomaterial, these attributes constrain the synthesis and manufacture of PEEK.

    Fabrication techniques for polyaryletherketone (PAEK) polymers have undergone constant refinement since the preparation of polyetherketoneketone (PEKK) was first described in the 1960s [1], [2] and [3]. Although many of the details associated with synthesis and processing of PAEKs are proprietary to resin and stock material suppliers, it is important to understand the steps used in the manufacture of raw materials, because these techniques can substantially impact the properties and quality of the stock shapes and molded implant components [4]. Consequently, this chapter summarizes the principal steps used to synthesize and fabricate PEEK implant components.

    Early studies on PEEK processing tend to emphasize the fabrication of PEEK composites, using carbon and glass fibers [4]. This chapter is focused on the synthesis and processing of neat, unfilled PEEK polymer. We begin by outlining the two main synthesis routes for contemporary PEEK. As a high-temperature thermoplastic, PEEK can be processed using a variety of commercial techniques, including injection molding, extrusion, and compression molding. This chapter provides an overview of the methods for processing unfilled PEEK used in biomedical applications. Readers with an interest in PEEK composites may wish to skip ahead to Chapter 3, which covers blending of PEEK with additives.

    2.2. Synthesis of PAEKs

    As noted previously, the polymerization of aryletherketones is a complex and challenging process due to the insolubility of PAEKs in typical solvents. Furthermore, the solvents and high temperatures necessary to carry out successful polymerization of PEEK, such as benzophenone or diphenylsulfone above 300°C, necessitate dedicated plant facilities with rigorous safety procedures (Fig. 2.1). Because of the precautions necessary to carry out safe polymerization, the reactions are typically carried out in batches, as opposed to in a continuous process. All these challenges contribute to the higher cost in producing PAEK polymers, when compared with other thermoplastics.

    Historically there are two main routes involved in the production of PAEKs. The first method involves linking aromatic ether species through ketone groups, whereas a second method involves linking aromatic ketones by an ether bond. The first method involves an electrophilic reaction and Friedel Crafts acylation chemistry, and the second route involves a nucleophilic displacement reaction.

    2.2.1. Electrophilic Routes to PAEK Polymers

    The inherent solvent resistance and propensity to reach high crystallinity levels prevents PAEK polymers from being synthesized in common organic solvents. Early attempts to synthesize PEEK in methylene chloride or nitrobenzene produced only low-molecular-weight variants.

    Work by DuPont using a combination of anhydrous hydrogen fluoride/boron trifluoride succeeded in protonating the carbonyl groups and meant that high-molecular-weight polyetherketone (PEK) became a possibility (Scheme 2.1) [5].

    Raychem also reported the synthesis of PAEK polymers using similar reaction conditions in the presence of alkylthiochloroformates.

    Another electrophilic process exemplified by Ueda and Oda uses methanesulfonic acid (MSA)/phosphorus pentoxide (P2O5) at low temperatures [6]. Although PEEK produced by this method has a less branched structure than AlCl3-catalyzed systems, it also suffers from high temperature instability and hence cannot be molded or extruded without extensive cross-linking and degradation.

    Colquhoun and Lewis [7] have described the Friedel Crafts polycondensation of 4-(4′-phenoxyphenoxybenzoic acid) in trifluoromethanesulfonic acid to form PEEK. This route has only remained of academic interest due to the extremely high cost and corrosive nature of the solvent used (Scheme 2.2).

    The electrophilic synthesis of PAEK polymers produces materials with reactive end groups such as benzoic acids. Such polymers cannot be processed, without endcapping, due to their high thermal instability. When the reactive end group materials are subjected to high-temperature processing, the polymer immediately cross-links, producing gels, which cannot be shaped into desired articles. Therefore, PEEK production by electrophilic processes as described earlier has historically had limited commercial success.

    More recently, a modification to the electrophilic process for manufacturing PAEK polymers has been described. This again involves the polycondensation of 4-(4′-phenoxyphenoxybenzoic acid). However, methanesulfonic acid was used as the reaction solvent in the absence of phosphorus pentoxide, and 1,4′-diphenoxybenzene was used as an endcapping agent [8]. This route permits the manufacture of thermally stable PAEK polymers and has been used in industrial processes (Scheme 2.3).

    It should be noted that to ensure thermal stability, significant quantities of the endcapping agent are used and as a result may be present in significant quantities in the finished polymer. The choice of the endcapping agent may therefore significantly alter the leachable and biocompatibility profile of the material.

    2.2.2. Nucleophilic Routes to PAEK Polymers

    The nucleophilic route to PAEK polymers provides a straightforward pathway to polymers such as PEEK. Initial attempts to form high-molecular-weight PAEKs from the reaction of a dihalobenzophenone and an equivalent bisphenate failed due to the polymer product crystallizing from the sulfolane solvent (Scheme 2.4).

    Owing to the poor solubility of PEEK, the selection of the synthesis solvent is crucial. Suitable solvents should be thermally stable and inert to phenoxide species. It became apparent that solvents such as benzophenone or diphenylsulfone could be used in the synthesis of PAEK polymers [3]. The inherent instability of bisphenates to oxidation was overcome by the use of hydroquinone and sodium or potassium carbonate to form the bisphenate in situ.

    Very high temperatures (>300°C) are required to reach high molecular masses, the molecular weight being controlled by a slight excess of difluorobenzophenone, leading to fluorine-terminated chains (Scheme 2.5).

    This process was patented in 1977 by ICI and sold under the brand Victrex PEEK, and this route provided the majority of PEEK polymer used in industrial applications.

    The establishment of the nucleophilic route to PAEK polymers permitted the investigation of polymer variants by the use of different bisphenols to produce PAEK polymers with various properties, as reported by Attwood et al. [1]. The family of PAEK polymers grew to contain variants such as PEK, PEEK, PEKK, PEKEKK, and so on, with a range of glass transition temperatures (143–160°C) and high crystalline melt temperatures (335–441°C). As the dominant member of the PAEK family of polymers, PEEK is in its glassy state at room temperature, as its glass transition temperature occurs about 143°C, whereas the crystalline melt transition temperature (Tm) occurs around 343°C.

    2.3. Nomenclature

    The literature on PAEK resin is a maze of trade names and producers, which have changed over the years, complicating interpretation of reported data for today's materials. For researchers interested in deciphering the historical polymer science literature, we provide here a brief primer on the nomenclature of PAEK resins used for industrial purposes as well as for biomaterials (Table 2.1). Resin, when used in this context, refers to the neat, unfilled powder that is created by polymerization, whereas grades are typically characterized by flow characteristics (e.g., for injection molding or compression molding) or based on their filler content (e.g., glass fiber or carbon fiber). Because PAEK polymers are converted using standard thermoplastic processing techniques, such as injection molding, they are generally available as pellets, although powder resin is also available. Stock shapes, such as rods, are also available from producers.

    Enjoying the preview?
    Page 1 of 1