Encyclopedia of Nanoscience and Nanotechnology 9781588830623, 1588830624

v. 1. A-Ch -- v. 2. Cl-E -- v. 3. El-H -- v. 4. Hy-M -- v. 5. Mag-Mu -- v. 6. Nano A-M -- v. 7. Nano Me-T -- v. 8. Ne-P

476 73 14MB

English Pages 909 Year 2004

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
ToC......Page 1
p001......Page 4
p023......Page 26
p053......Page 56
p061......Page 64
p077......Page 80
p161......Page 164
p179......Page 182
p197......Page 200
p215......Page 218
p235......Page 238
p249......Page 252
p301......Page 304
p321......Page 324
p341......Page 344
p365......Page 368
p389......Page 392
p417......Page 420
p435......Page 438
p465......Page 468
p477......Page 480
p495......Page 498
p505......Page 508
p537......Page 540
p555......Page 558
p587......Page 590
p605......Page 608
p623......Page 626
p643......Page 646
p665......Page 668
p683......Page 686
p703......Page 706
p727......Page 730
p739......Page 742
p757......Page 760
p775......Page 778
p795......Page 798
p815......Page 818
p823......Page 826
p843......Page 846
p857......Page 860
p879......Page 882
Recommend Papers

Encyclopedia of Nanoscience and Nanotechnology
 9781588830623, 1588830624

  • 0 0 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

Ingenta Select: Electronic Publishing Solutions

1 of 3

file:///C:/06/06/06.files/m_cp1-1.htm

Encyclopedia of Nanoscience and Nanotechnology Volume 6 Number 1 2004 Nanoanalysis of Biomaterials

1

Matthias Mondon; Steffen Berger; Hartmut Stadler; Christiane Ziegler Nanoassembly for Polymer Electronics

23

Tianhong Cui; Yuri Lvov; Jingshi Shi; Feng Hua Nanobiosensors

53

Tuan Vo-Dinh Nanocables and Nanojunctions

61

Yuegang Zhang; Weiqiang Han; Gang Gu Nanocapsules

77

Zhi-dong Zhang Nanocatalysis

161

S. Abbet; U. Heiz Nanochemistry

179

Sebastian Polarz Nanocluster Characterization by EXAFS Spectroscopy

197

Giuseppe Faraci Nanoclusters and Nanofilaments in Porous Media

215

Parasuraman Selvam Nanocomposites of Polymers and Inorganic Particles

235

Walter Caseri Nanocomputers: Theoretical Models

249

Michael P. Frank Nanocontainers

301

Samantha M. Benito; Marc Sauer; Wolfgang Meier Nanocrystal Memories

321

E. Kapetanakis; P. Normand; K. Beltsios; D. Tsoukalas Nanocrystalline Aluminum Alloys

341

Livio Battezzati; Simone Pozzovivo; Paola Rizzi Nanocrystalline and Amorphous Magnetic Microwires

365

A. Zhukov; J. González; M. Vázquez; V. Larin; A. Torcunov

4/25/2007 9:16 PM

Ingenta Select: Electronic Publishing Solutions

2 of 3

Nanocrystalline Barium Titanate

file:///C:/06/06/06.files/m_cp1-1.htm

389

Jian Yu; Junhao Chu Nanocrystalline Ceramics by Mechanical Activation

417

J. M. Xue; Z. H. Zhou; J. Wang Nanocrystalline Diamond

435

Narendra B. Dahotre; Padmakar D. Kichambare Nanocrystalline Phosphors

465

Guangshun Yi; Baoquan Sun; Depu Chen Nanocrystalline Silicon Superlattices

477

David J. Lockwood; Leonid Tsybeskov Nanocrystalline Silicon: Electron Spin Resonance

495

Takashi Ehara Nanocrystalline TiO2 for Photocatalysis

505

Hubert Gnaser; Bernd Huber; Christiane Ziegler Nanocrystals Assembled from the Bottom Up

537

Edson Roberto Leite Nanocrystals from Solutions and Gels

555

Marc Henry Nanocrystals in Organic/Inorganic Materials

587

Peter Reiss; Adam Pron Nanodeposition of Soft Materials

605

Seunghun Hong; Ling Huang Nanodielectrophoresis: Electronic Nanotweezers

623

P. J. Burke Nanoelectromechanical Systems

643

Josep Samitier; Abdelhamid Errachid; Gabriel Gomila Nanoelectronics with Single Molecules

665

Karl Sohlberg; Nikita Matsunaga Nanoembossing Techniques

683

Yong Chen Nanofabrication by Glancing Angle Deposition

703

Brian Dick; Michael J. Brett Nanofibers

727

Jason Lyons; Frank K. Ko

4/25/2007 9:16 PM

Ingenta Select: Electronic Publishing Solutions

3 of 3

file:///C:/06/06/06.files/m_cp1-1.htm

Nanofluidics

739

P. Mela; N. R. Tas; A. van den Berg; J. E. ten Elshof Nanofluids

757

Stephen U. S. Choi; Z. George Zhang; Pawel Keblinski Nanohelical/Spiral Materials

775

S. Motojima; X. Chen Nanoicosahedral Quasicrystal

795

J. Saida; A. Inoue Nanomagnetics for Biomedical Applications

815

Chong H. Ahn; Jin-Woo Choi; Hyoung J. Cho Nanomagnets for Biomedical Applications

823

Pedro Tartaj Nanomaterials by Severe Plastic Deformation

843

Yuntian T. Zhu; Darryl P. Butt Nanomaterials for Cell Engineering

857

Xiaoyue Zhu; Tommaso F. Bersano-Begey; Shuichi Takayama Nanomaterials from Discotic Liquid Crystals

879

Gregory P. Crawford; Robert H. Hurt Copyright © 2004 American Scientific Publishers

4/25/2007 9:16 PM

Encyclopedia of Nanoscience and Nanotechnology

www.aspbs.com/enn

Nanoanalysis of Biomaterials Matthias Mondon, Steffen Berger, Hartmut Stadler, Christiane Ziegler University of Kaiserslautern, Kaiserslautern, Germany

CONTENTS 1. Introduction 2. Biomaterials 3. Nanoanalytical Tools 4. Nanoanalysis of Biomaterials 5. Conclusions Glossary References

1. INTRODUCTION Nanotechnology has been a fast-developing field over the last decade. Features of dimensions below 100 nm are an interesting and important field of studies in a world of miniaturization in engineering but as well in the already small world of biology. New methods of analysis have to be developed along the way, as conventional light microscopy techniques are unable to resolve features of these dimensions. In this chapter recent progress in nanoanalytics in biomaterial respectively biointerface research is reviewed, focusing on publications that are relevant for the characterization of the surfaces of biomaterials rather than the characterization of biofilms on biomaterials. Scanning probe microscopy (SPM) methods, that is, scanning force microscopy (SFM), scanning tunneling microscopy (STM), and related techniques are unique to investigate on a molecular and even on an atomic level surfaces and molecules interacting with them. Morphology and roughness are important parameters that can be characterized with them. Further fields are the spatial arrangement of chemical functional groups and their interaction with the surrounding medium. The techniques based on the electron-sample interaction like scanning electron microscopy (SEM) and transmission electron microscopy (TEM) are also able to achieve material- or morphology-dependent information with nanometer resolution and are therefore important tools in biological and materials science. ISBN: 1-58883-062-4/$35.00 Copyright © 2004 by American Scientific Publishers All rights of reproduction in any form reserved.

The issue of nanotechnology has become also part of the world of biomaterials for production as well as for analysis to investigate phenomena occurring on this scale. Biomaterials include a wide range of metals and alloys, glasses and ceramics, natural and synthetic polymers, biomimetics, composites, and natural materials, including combinations of synthetic and living tissue components. Materials in contact with biological systems are of great relevance not only in medicine but for many applications in engineering, food processing [1], marine environments [2], and biosensing [3]. The use of biomaterials is therefore part of the larger area of biological surface science (BioSS) that stretches as well into areas like tissue engineering, biosensors and biochips for diagnostics, artificial photosynthesis, bioelectronics, and biomimetic materials [4]. The market volume for medical biomaterials exceeds a billion Euro per year. In the United States alone the current value is over 700 million Euro [5]. More than 500,000 arthroplastic procedures and total joint replacements are performed annually in the United States [6]. Clinical implants like artificial tooth replica, knee and hip joints, blood vessels, heart valves, and intraocular lenses are in use for many years now and applied with hundreds of thousands of patients improving their quality of life. A lot of materials consist of metals and ceramics but the use of polymers for new medical products and materials has increased strongly over the last decades to keep improving current health care procedures. These biomaterials must meet the demands of materials science on various length scales as well as clinical requirement. The mechanical specifications range from high loads in implantology for hip and knee replacements to high elasticity of artificial blood vessels. They have to be accompanied by others, for example, high transmittance in intraocular or contact lenses. In many cases blood compatibility with suppression of blood coagulation between the physiological environment and the biomaterial surface is necessary. A good understanding of the interaction with the biological environment, namely with biomolecules like proteins, cells and tissue, is crucial in order to be able to improve the functionality of biointerfaces. This interaction is essential for many applications like lubrication, adhesion, and recognition Encyclopedia of Nanoscience and Nanotechnology Edited by H. S. Nalwa Volume 6: Pages (1–22)

Nanoanalysis of Biomaterials

2 in biological systems and for biocompatibility of the implant interface [7]. The interface is important for biomaterials, as it defines the interaction with the environment. The knowledge of the surface structure, the surrounding medium in direct contact (e.g., water structure [8]), and the chemical composition is essential to understand further reactions taking place at the biomaterial’s surface. The interface structures are in dimensions of a few nanometers for proteins up to the micrometer range for cells in the field of biomaterials. In Figure 1 an overview of the interfaces on different scales is presented to show the parameters that can be of interest. Macroscopically, the shape of an implant, the structure, and its mechanical stability are important. The microscopic level is determined by the morphology (i.e., the domain structure, the presence of ionic groups, and the chemical composition respectively modification), the topography (i.e., the surface roughness, planarity, and feature dimensions), and the hardness respectively elasticity (Young’s modulus). These characteristics determine other features like wetting behavior and interaction forces (inter- and intramolecular) including cell-surface or cell-cell interactions. Various degrees of information about these properties can be

obtained using different analysis methods including microscopic and spectroscopic methods. The field of investigation is extending further as the material properties have to be characterized after manufacturing, sterilization, before and after clinical insertion, and before, after, and especially during the immersion in simulated natural environments. The resulting changes of the interaction, namely corrosion and the buildup of a biofilm, have to be characterized. The surface properties as well should be characterized thoroughly as it is essential for correlating any surface modifications with changes in biological performance. In the following, techniques that are able to resolve features of materials on the nanometer scale (Section 3) as well as applications for biomaterial research (Section 4) are presented. The whole topic is introduced by a more general discussion of functionality, properties, and types of biomaterials (see Section 2). The focus in experimental work will be put on methods investigating materials on the nanometer scale, that apply to nanometer resolution in x-, y-, and z-direction. Methods leading to a resolution in the nanometer regime in only one or two dimensions will be described only shortly. Investigations of the above introduced groups of biomaterials including metals, ceramics, polymers as well as biological specimens are presented. In the context of this review biomaterials are defined as materials used in medicine which are in contact with the tissue of the patient. Therefore special emphasis will be put on medical implant materials and the characterization with scanning probe techniques. Experiments dealing with hard tissues like teeth and bone and the influences of substrates on biofilm formation will be briefly mentioned.

2. BIOMATERIALS Biomaterials can be classified with respect to their chemical composition and their use. Generally, the following criteria have to be fulfilled by biomaterials: • Functionality: The biomaterial must replace at least in the important parts natural body functions. This means that, for example, for bone substitution a suitable mechanical stability has to be achieved. • Biocompatibility: The biomaterial must not cause negative body reactions resulting, for example, in inflammation. Furthermore, a good integration into the body environment is favorable. • Aesthetical aspects: Biomaterials used for external implants should not deface the patient. Preferably, they should look like the replaced tissue.

Figure 1. Different relevant length scales in biomaterials research demonstrated at hand of a bone implant. The biomaterial-tissue interaction is based on molecular events, which affect meso- and macroscopic material properties. On larger scales additional (e.g., cellular, mechanical) effects resulting from combining individual units (molecules, cells, crystallites) to a more extended ensemble arise.

Surface properties influence strongly the mechanical function, the biochemical interaction, and the optical appearance of the biomaterial. Therefore the interface of the materials deserves special interest in investigation, control, and modification of properties to reach the criteria mentioned above. Mechanical aspects are relevant for applications which require force transfer (e.g., bone substitution), which impose dynamic loads on, for example, artificial joints, and which, for example, in the case of substitutes for blood vessels, meet certain hydrodynamic respectively fluid-mechanical properties. The interaction of the biomaterial with the body

Nanoanalysis of Biomaterials

environment concerns the tissue contact (cell adhesion and fixation), the tissue organization (cell-to-cell linkage and communication), and the exchange of substances across the interface. Optical appearance is determined by light transmission (e.g., of contact lenses) or light reflection (e.g., of natural looking tooth replacements). Mechanical, biochemical, and optical properties depend mainly on the topography and the chemistry of the surface. Topography may be defined as size, shape, distribution, and hierarchy of surface features. These features are either discrete (holes and peaks) or continuous (furrows and ridges) in random (statistical), fractal (self-similar on different length scales), or periodic distribution across the surface. Rough surfaces exhibit a larger surface area and more contact points for biological molecules and cells. The reason for larger entities like cells to adhere to the surface and how they arrange their layer growth is dependent on the size and distribution of the surface features: Cavities which are smaller than single cells cannot be colonized, while larger structures regulate the direction, connectivity, and differentiation of the growing cell layer. A hierarchical surface consists of structure elements with increasing complexity, for example, two-dimensional planes scrolled up to threedimensional fibers which are organized in three-dimensional fiber composites. Examples of those materials were synthesized with carbon nanotubes with increased mechanical stability and electrical conductivity and decreased thickness compared to common wires [9–11]. In principle such materials are also interesting for biomaterials as on one hand they show a very large surface area and on the other hand they enable combinations of surface sections with different properties and therefore different cell types within a single superstructure. The chemical properties of the interface between biomaterial and body environment determine the interaction of the surface with water molecules, ions, biological macromolecules, and cells. The surface reactivity depends on the chemical composition, the production process, and the pretreatment before use of the materials and is essential for fixation, growth, and proliferation of tissue and bone cells. Cell adherence via membrane receptors and adhesion proteins is influenced by active surface groups (generated, e.g., by oxidation of metals or coating of materials with specially designed polymer layers) which regulate the adsorption of the anchoring protein layer. The functionalization of the substrate additionally controls the wetting behavior where particularly hydrophilic surfaces show enhanced cell adhesion. The latter is due to the fact that systems like cells and hydrophilic surfaces reach a thermodynamically favorable state by combining their high-energy surfaces. In addition, even hydrophobic interfaces like gold can get more hydrophilic by protein adsorption [12] which may help to increase biocompatibility. Polarizability and charging of the interface affect electrostatic interactions between charged species and the surface. Charge screening and complexation by multivalent ions present in all types of body fluids allow attraction between molecular and surface groups of like charges and stabilize the biological layers. At conductive interfaces electrochemical reactions with charge transfer between the electrolyte solution and the substrate may occur, which interfere with the cellular metabolism and the

3 conformation of adsorbed adhesion proteins. This can set free toxic substances and cause allergic and inflammatory body reactions. Topographical as well as chemical effects play a role for tribological properties of biomaterials. Tribology describes the behavior of interfaces in motion and gets important when implants are designed to support body movements. In this case friction, wear, and lubrication of implant and body respectively different moving implant elements influence function and longevity of the applied biomaterial. Interfacial friction depends on kind and strength of interaction forces, the clasping of surface uprisings and troughs, and force transfer properties of intermediate fluids. Hardness, cohesion, and adhesion of the individual surfaces in contact determine to what extent the materials are worn off by abrasion (e.g., scratch and particle formation), adhesion (e.g., welding processes), and surface fatigue (e.g., crack formation). Liquid films in the crevice between two hard surfaces can reduce friction and wear. The effect of such lubricants is influenced by surface chemistry and separation as well as by the viscoelastic and hence force transferring properties of the fluids themselves. This comprehensive discussion of biomaterials in view of surface properties shows that the interplay between implant and body environment can be very complex and includes a large variety of parameters from materials science, biology, and medicine explained in more detail in [7]. The need of surface analysis with nanometer resolution arises on one hand from the importance of the interface between natural and artificial material and on the other hand from the high dependence of the macroscopic behavior on the microscopic appearance of the biomaterial. In the following the main different chemical substance classes for biomaterials will be described shortly. Biomaterials can be divided into native materials like teeth and bone, and artificial biomaterials which are in contact with biological systems like implants in medicine. Within these fields the materials in their actual state of occurrence and model surfaces like thin layers on supports are under investigation. The three metal systems mainly utilized in implantology are stainless steel, cobalt-chromium alloys, and titanium. Such metals are mainly used in bone replacement and here titanium is the most important metal because of its good biocompatibility and nearly bonelike mechanical properties. That is why titanium and its alloys TiAl6V4 and TiAl6Nb7 are frequently applied as biomaterials for hard tissue replacement such as dental and orthopedic (e.g., hip and knee joint prosthesis) implants, in the audiological field, and for cardiovascular applications such as mechanical heart valves and as material for surgical instruments such as vascular clamps, needle holders, and forceps [13]. Titanium surfaces are covered by a 2- to 6-nm-thick oxide layer that is formed spontaneously in the presence of oxygen. This layer contributes to the biocompatibility of titanium by preventing corrosion and the release of ions from the metal surface [7]. The oxide layer thickness and surface morphology can be altered by according treatment or coatings applied to change the physicochemical properties and the biological response [13].

4 Stainless steel is also used for orthopedic implants. Its lower biocompatibility and its mechanical properties, which are less bonelike than titanium, reduce the use as human implant material, but it is cheaper than other metals and therefore is still used in animal medicine and for surgical instruments. Other applications of steel are cardiovascular stents. Cobalt-chromium-molybdenum alloys are very hard and in combination with other smooth surfaces very abrasionproof. They are therefore an ideal material for joint implants. In many patients Co-Cr-Mo, that is sintered from Co-Cr-Mo beads, is used as a femoral hip implant. Gold as one of the oldest implant materials is used not only as dental restorative material but also as electrode material for implantable biosensors or to increase the contrast in electron microscopy. An overview on studies concerning metals as biomaterials is given in Section 4.1. Bone, tooth enamel, and dentin have a composition which is very similar to that of special ceramic materials like hydroxyapatite, calcium phosphate, and calcium carbonate. These materials in their natural or manmade origin are used as coatings to increase the biocompatibility of implant materials. On the other hand they can replace parts of bones; for example, porous hydroxyapatite is a good basis for new bone growth. Biologically active glass or bioglass is similar to ceramics and assists actively new bone formation. It is therefore used in every place where bone has to be built up. This is necessary after tumor surgery or if natural bone is destroyed by an inflammation. Ceramic and glassy materials are further considered in Section 4.2. Another kind of biomaterials are the polymers. They can be tailor-made for many different applications as their properties and their composition are variable in a large range. Polymers can be used as manufactured or treated with different techniques to yield special functional groups at the surface for specific interaction, increased surface energy, hydrophilicity respectively hydrophobicity, or chemical inertness. Furthermore, such surface layers may induce crosslinking to increase hardness, remove weak boundary layers or contaminants, modify the surface morphology (increase or decrease of surface crystallinity and roughness), or increase surface electrical conductivity or lubricity. A review of these modifications, examples, and characterization is given in [14]. Polymers like high molecular weight polyethylene (HMWPE) are used as counterpart for the titanium ball head in artificial hip joints. Poly(tetrafluoro ethylene) (PTFE) and other fluorinated polymers are used for blood vessel replacements and catheters, poly(methyl methacrylate) (PMMA) for intraocular lenses, silicone acrylate for contact lenses, polyurethane for pacemakers and left ventricular assist devices, cellulose for renal dialyzers, and silicone for catheters and breast implants [15]. Biodegradable polymers like poly(sebacic anhydride) (PSA) and poly(DLlactic acid) (PLA) are naturally removed from the body after a definitive time span and can be used for controlled drug delivery [16]. Generally, carbon in different forms (e.g., diamond, graphite, or activated carbon) is an important biomaterial particularly because of its high blood compatibility. Glassy polymeric carbon (GPC) made from phenolic resins by pyrolization, is mostly used as coating, for example, in prosthetic heart valves and other prosthetic devices [17, 18]. Some selected studies on polymers are discussed in Section 4.3.

Nanoanalysis of Biomaterials

A newer group of materials used in implantology are composites, especially composites of inorganic material and polymers which combine the advantages of both. Examples can be found in Section 4.4.

3. NANOANALYTICAL TOOLS A wide range of applications use the methods described in this section as nanoanalytical tools. This field is extending rapidly and new applications and methods are developed constantly. A short description will be given of the methods and the parameters investigated with them. The possible areas of application and examples are given in Section 4. Particularly, scanning force microscopy (SFM) is stressed which has proven to be a valuable and easy to handle method for studies on biomaterials. Additional advantages can be found in the nondestroying operation which is applicable under ambient conditions. Experimental setups revealing nanometer resolution in only one dimension are discussed at the end of this section.

3.1. Scanning Probe Techniques Scanning probe techniques have been a valuable tool since their invention in the 1980s [19, 20]. In these experimental methods distance-dependent interactions like tunneling current, force, or light transmission between a sharp needle (“tip” or “probe”) in close proximity to a surface (“sample”) are utilized to produce an image of the sample. Two principal measurement modes were implemented: (i) to maintain a constant vertical probe position while measuring the interaction change often due to surface topography (“constant distance mode”), and (ii) to maintain constant interaction while adjusting the distance with the feedback signal reflecting surface topography (“constant interaction mode”). Both modes offer the possibility to characterize surfaces down to the atomic scale in a great variety of environments from ultrahigh vacuum to aqueous solutions. It is as well possible to characterize time-dependent reactions like crystallization or corrosion processes, as it can be done continuously and hence on-line. To scan a probe over a surface in the desired way while reacting to the topography of the sample, positioning tools with spatial resolution in the 0.1-nm regime are necessary. The latter is fulfilled by piezoelectric actuators made of ceramics like PZT (lead zirconate titanate) or PMN (lead magnesium niobate), which in different directions can be extended by less than the size of one crystal unit cell. With a suitable detection unit to measure small interaction changes connected to a distance feedback circuit and digital data representation, an image of the surface can be portrayed. Experimental adaptations and extensions based on the interaction mechanisms originally used for imaging purposes lead the way to monitor more complex features than the topography as described below. In scanning tunneling microscopy (STM) the current of electrons tunneling between a conductive wire (preferentially heavy metals like tungsten, platinum, or iridium) with an atomically sharp tip and a (semi-) conductive surface across vacuum is measured (Fig. 2). The tunneling current decays exponentially with increasing distance between tip

Nanoanalysis of Biomaterials

5

controller piezo-crystal current amplifier tunneling-current

Figure 2. Schematic drawing of the STM principle.

and surface and additionally depends on the local densities of electronic states of the tunneling partners and the work function of the sample. Because of its strong distance behavior only the atom at the very end of the tip and the nearest surface atom are involved within the tunneling event. A slight change in distance according to progression along rows of surface atoms alters the tunneling current in a measurable manner so that true atomic resolution can be achieved. Experimentally the onset of a tunneling current is obtained by approaching tip and surface to less than a few tenths of a nanometer under a constant bias voltage. The sign of the bias determines the direction of the tunneling current; this means unoccupied electronic surface states are probed with occupied electronic tip states or vice versa. Scanning the surface row by row either at constant height or constant current (see above) reveals the surface topography. Additionally, STM can be applied to probe electronic structures. Modulating the tip-surface distance and measuring the change of the tunneling current at constant bias allows to extract the local work function of the sample. Modulating the bias and measuring the corresponding current changes at a constant tip-surface distance allows to extract the electronic states around the Fermi level of the sample (STS, scanning tunneling spectroscopy). This can lead to some chemical information about the surface, but for more detailed information electronic core levels have to be sensed, which is not possible with the STM (see [21] and especially references therein). Scanning electrochemical microscopy (SECM) measures highly localized electrochemical currents associated with charge transfer reactions on metallic sample surfaces under liquid environment [22, 23]. In macroscopic measurements it can be compared with cyclovoltammetry. The reactions occur in a four-electrode electrochemical cell under bipotentiostatic control. There are two mechanisms respectively pathways of image production. Electron tunneling and electrochemical reactions via a water bridge occur according to the applied voltage. It can be used for the detection of localized electrochemical reactions at surfaces. It can also be used for microstructuring biomaterials like titanium [24]. In addition SECM is also capable of probing the kinetics of solution reactions and adsorption phenomena and monitoring heterogeneous electron transfer kinetics associated with processes on conducting surfaces [25].

The invention of the scanning force microscopy (SFM) was a breakthrough for these techniques as it is possible to image nonconducting substrates with a resolution of 0.2 nm laterally and 0.001 nm vertically. It does not require specimens to be metal coated or stained. Noninvasive imaging can be performed on surfaces in their native states and under near physiological conditions. It has been proven to be particularly successful for imaging biological samples, such as proteins, nucleic acids, and whole cells. By scanning, dynamic processes can be imaged, such as erosion, hydration, physicochemical changes, and adsorption at interfaces. Therefore the SFM is currently the SPM technique with the widest applicability for biomaterial research. In SFM a small tip attached to a micro beam (cantilever) is scanned across the surface of the specimen and deflected by topographic features. The force of interaction may be repulsive or attractive, giving rise to different modes of operation. Moving the cantilever from the interaction free zone far above the surface, it snaps into contact (Fig. 3) due to the attractive force between the tip and the sample, which can be described in a simple way by the Lennard Jones potential. The piezo pushes the tip further towards the sample and the positive repulsive force reaches a maximum. As the piezo is retracted the repulsive force is reduced and the force changes sign. If the bending force of the cantilever gets larger than the attractive force towards the surface, the tip loses contact. The tip can be held in the repulsive regime of the Lennard Jones potential or oscillated in the attractive or repulsive regime, resulting in different interactions [26, 27]. These differences are important as biomolecules are deformed by applying a load of some nanonewtons as present in contact mode [28]. The deflection is usually monitored by a laser beam that is reflected and detected with a four split photodiode. This signal is used to maintain a constant force via a feedback loop and to monitor the height data. The SFM can be operated in a variety of modes that can provide different information about the sample. Usually the z-deflection is monitored and interpreted according to the parameters under investigation. An easy to read overview covering many scanning probe

return from surface towards surface

Cantilever deflection [nm]

monitor

linear deflection on stiff surfaces 0

interaction free region -90

point of contact to sample "jump in" adhesion 0

200

400

Piezo-z-position [nm] Figure 3. Cantilever deflection versus sample-z-position curve on a stiff surface monitored via SFM.

Nanoanalysis of Biomaterials

6 techniques is given in [29] while in [30] results obtained by SFM in the field of nanotribology are reviewed. Sensing of the z-position is usually done via the monitoring of the z-piezo voltage. This can cause ambiguities as piezo crystals exhibit hysteresis. This is overcome by monitoring the distance separately via inductive or fiber-optical sensors [31]. For dynamic modes [32] the application of an additional oscillation to the cantilever by a piezo crystal has to be performed. Magnetically driven cantilevers are used as well for actuation [33]. The quality of SFM data is essentially determined by the cantilever and the tip, influencing the resolution of topography and force measurements. Micromechanical properties of the cantilever and the shape and chemical composition of the tip, which comes into direct contact with the sample, are essential. High aspect ratios and small tip radii are desirable for imaging steep slopes and deep crevices. Depending on the mode of operation, different parameters have to be optimized, which can be realized according to the methods described in the following. Silicon and silicon nitride cantilever fabrication based on photolithographic techniques is well established. Metalbased (Ni) cantilevers [34] and cantilevers made of piezoelectric material (lead zirconate titanate) are produced as well for independent actuation and sensing [35]. Cantilevers of various shapes (e.g., rectangular or V-shaped) and dimensions (usually 100 to 200 m in length) are available. New approaches to a further miniaturization of the system have been made on microfabricated aluminum probes with length scales of 9 m [36–39]. Coatings (Cr-Au, Pt, Al, TiN, W2 C, TiO2 , Co, Ni, Fe, Au with biological coating) can be applied to the cantilever to modify it against corrosion in the liquid phase or for different applications when conductive, magnetic, or biological properties are necessary. The resonance frequency of the cantilever ranges between several kHz to several hundred kHz. Cantilevers with high resonance frequencies are used in dynamic modes as the tip oscillates with several hundred kHz above the surface. The stiffness of the cantilever is defined by the force constant k and ranges between 0.01 and 100 N/m for the vertical deflection. The soft (k < 01 N/m) cantilevers are used in contact modes to minimize the disturbance of the sample. Rigid cantilevers with a force constant larger than 1 N/m are used in noncontact or dynamic modes since they exhibit high resonant frequencies and small oscillation amplitudes of about several nanometers. The force constant for the lateral twisting can be determined for friction measurements [40] and the movement of the cantilever has been modeled accordingly with finite element analysis [41]. The mechanical behavior and the determination of the spring constant are well described in literature [42–49]. The tip itself can consist of different materials or is coated according to the application. For some applications a hard surface is necessary and a diamond-like coating (DLC) is applied [34]. The production of DLC coatings is described in [50]. Different functionalities [51–53] can be applied. Cantilever tips can be modified by chemical [54] and biochemical [55] functionalization. Via silanization [56, 57] or via thiols, self-assembled monolayers (SAMs) are produced for

further modification [53]. Proteins and bacteria [58] can be attached to the tip. Additional materials which vary the shape of the tip can be deposited, such as polystyrene, borosilicate and silica spheres, C60 molecules [59], carbon-nanotubes in general [60–63], or single-wall nanotubes with diameters of 3 Pas had the best properties, especially when they were heat treated at 175  C. Controlled release glass (CRG) based on calcium and sodium phosphates differs substantially from commercial glass in that it dissolves completely in aqueous environment. The solution rate can be predetermined and adjusted by altering the chemical composition. A SEM study of the hemocompatibility [189] of such glass indicated that selected and controlled compositions of these materials might provide good blood-contacting surfaces.

4.3. Polymers Many different ways to use polymers as biomaterial are reported. This field spreads from vascular implants [193– 196] and bone replacement [197–200] to intraocular lenses [201].

4.3.1. Structure and Mechanical Properties High- and low-density polyethylene (HDPE and LDPE), isotactic and atactic polypropylene (iPP and aPP), and polymer blends were investigated regarding the friction coefficient and the elastic modulus via SFM. Structural changes and the mechanical property changes around the glass transition temperature have been monitored during temperature runs. The enhanced ordering of the backbone correlates to the increased surface modulus. Additionally, the elastic modulus, the time-dependent relaxation process, and the friction properties were measured as a function of pressure. Loads between 1 and 1000 nN were applied. At low pressure the deformation of the polymer is elastic. With increasing pressure there is a phase transition to a plastic behavior attributed to a polymer alignment effect. Friction properties were investigated concerning the contact pressure and contact area, revealing an increased elastic modulus with increased density and crystallinity and a linear increase with contact pressure. Stretching of LDPE was shown to lead to surface roughening and alteration of surface mechanical properties. These experiments show that a material subject

Nanoanalysis of Biomaterials

16 to complex mechanical stresses will change continuously its surface mechanical properties as the nature of the stress changes [102]. Polyethylene and the influence of low molecular weight additives that are added to prevent oxidation were investigated via friction versus load curves. For loads of 1000 nm, the lower friction of the polymer with additives gives way to friction properties of the different samples that are identical [102]. Polyurethane copolymers with hydrophobic and hydrophilic side chains were investigated regarding the mechanical properties changing with the hydrogen bonding between soft (SS) and hard segments (HS). The adhesion force showed a maximum at 57% HS correlated with the highest number of nonassociated urethane groups. The friction mechanisms were investigated as well, showing different mechanisms for the different compositions [102]. The surface structure of six different polycarbonate polyurethane copolymers was investigated with SFM in topography and phase modes. The stoichiometry of the reagents and the chemistry of the hard segment were changed, varying the contribution of diisocyanate, poly(hexyl, ethyl) carbonate, and butane diol. The 1,6-hexane diisocyanate showed a stronger phase separation and the highest values of mean square roughness (RMS) with 3.3 to 10.0 nm with regard to methylene bis(p-phenyl isocyanate) and methylene bis(p-cyclohexyl isocyanate) that have RMS values of 0.8 respectively 0.25 nm. These values have to be viewed with respect to the image processing of a third-order flatten. This is only obvious in this article as the image processing is usually not mentioned, and therefore the values of the other investigations must be examined closely as well. The polymers presented relatively stiff rodlike structures with dimensions of 12 nm in width and 170 nm in length [202]. Investigations on polyurethane and rubber were done with force indentation curves [28]. Differences between polystyrene (PS) and poly(methylmethacrylate) (PMMA) could be determined by measuring the adhesion differences of the different polymers by PFM [95]. Polymer blends of poly(styrene)-blockpoly(ethene-co-but-1-ene)-block-poly(styrene) with isotactic and atactic polypropylenes were characterized observing the morphology according to different thermal treatments. It was possible to monitor microphase separations with tapping mode in the phase image [203]. Interfaces between PMMA and polystyrene (PS) could be identified with HM-LFM [84] as well as interfaces between poly(acrylonitrile-co-styrene) and polybutadiene respectively between polypropylene and poly(propylene-co-ethylene) [85]. Fractographic investigations looking at the surfaces after a fracture of dental composites consisting of silicon dioxide fillers in a matrix of dimethacrylate resins are presented in [107]. The pristine surfaces and the results of different preparation methods for poly(ethylene glycol) (PEGMA) grafted poly(tetrafluoro ethylene) (PTFE) surfaces were investigated with SFM. The PEGMA treated surfaces exhibited a larger RMS of 144 nm with regard to 129 nm RMS of the pristine PTFE [204]. Hexafluoroethane and tetrafluoroethylene films produced by glow discharge plasma deposition show with SFM a pinhole free and smooth surface

with a RMS roughness of 0.4 to 2 nm. Scanning in contact mode with moderate applied loads yields rippled films. At small scan sizes, and high velocities and load, the film disrupts. The surface modulus was estimated between 1.2 and 5.5 GPa [205]. Poly(vinylidene difluoride) (PVDF) and poly(vinylidene fluoride/hexafluoropropylene) [P(VDFHFP)] polymers were radiation grafted with polystyrene (PS) yielding an increase in surface roughness upon irradiation and a smoothing of the valleys by further functionalization by chlorosulfonation and sulfonamidation [206].

4.3.2. Response to Proteins In the field of ultrafiltration of biological fluids such as blood, biofouling respectively membrane fouling is an important issue resulting in pore plugging, pore narrowing, and cake deposition. Hydrophobic polysulfone membrane interactions with hen egg lysozyme were investigated with the surface force apparatus and the topography with SFM [207]. Forces between cellulose acetate films and human serum albumin, HSA, respectively ribonuclease A, RNase, were investigated in the same way [208]. The comparison of forces between the proteins fibronectin, fibrinogen, and albumin and a hydrophilic cellulosic membrane, a hydrophilic glass surface, and a hydrophobic polystyrene surface showed that the forces on hydrophobic surfaces are an order of magnitude larger than the ones on hydrophilic surfaces. Additionally, the influence of surface structure on hydrophilic surfaces was described [209]. Collagen on native and oxidized poly (ethylene terephthalate) (PET) was investigated with the aid of X-ray photoelectron spectroscopy (XPS). Time-dependent differences were observed in stability at times between 5 s and 5 min with an easy-to-disturb surface structure (scan ranges from 2 to 5 m) as monitored via SFM. The investigations were done in air, leaving the results ambiguous, as the topographical information can be altered by drying leading to different structures and orientations on the surface [210].

4.3.3. Tissue Response The tissue response to polymers is investigated in a variety of studies [200, 211–214]. There are different attributes, which complicate the use of polymer vascular implants. Especially for small-diameter vascular grafts the blocking is an important problem. For example, calcification decreases the vascular diameter. So calcification of subdermally implanted polyurethane is enhanced by calciphylaxis as shown by SEM [194]. Another aim is to decrease the thrombogenicity [195] of polymeric biomaterials. These “heparin-like” polymers have a specific affinity for antithrombin III and thus are able to catalyze the inhibition of thrombin. In this study sulfur, sodium, fluorine, and carbon are determined by SEM combined with energy dispersive X-ray analysis. The objective of another project [196] was to define a series of assays for the evaluation of hemocompatibility of cardiovascular devices. In [193] a polyamino-acid urethane copolymer coated vascular prosthesis (1.5 mm diameter) was developed to enhance endothelialization. SEM showed the homogeneous coating of clinically available synthetic PTFE grafts. Prosthetic heart valves made of isotropic pyrolytic carbon (LTIC) are

Nanoanalysis of Biomaterials

a successful biomaterial application which is investigated together with other difficult-to-image biomaterials in [17]. Low-voltage, high-resolution scanning electron microscopy shows that LTIC valve leaflets are significantly rougher than previously described, and that LTIC induces extensive platelet spreading in vitro, even in the presence of considerable albumin. Calcification is also crucial in the use of intraocular lenses [201]. To assess this, intraocular lens optic materials were implanted intramuscularly and/or subcutaneously in rabbits for up to 90 days. SEM and energy dispersive X-ray spectroscopy (EDX) were used to detect discrete nodules containing both calcium and phosphate. Calcification only was noted on intramuscularly, subcutaneously, and intraocularly implanted experimental acrylic and intramuscularly implanted hydrogel material. In contrast, the intramuscularly or subcutaneously implanted silicon, PMMA, and acrylic optic materials showed no calcification. Glassy polymeric carbon was treated by bombardment with energetic ions in different curing states. The roughness that leads to a better thromboresistance was monitored with SFM [18]. An approach to manage chronic osteomyelitis (inflammation of the periosteum) utilizes the implementation of antibiotic-impregnated PMMA beads for local delivery of antibiotics. The study of the getamicin sulphate release from a commercial acrylic bone cement was presented in [197]. A very interesting feature is the biodegradability of some polymers either natural or synthetic. The surface structure of blends with different composition of poly lactic acid (PLA) and poly sebaic acid (PSA) and their degradation properties were investigated. Single component PLA exhibits a smooth surface whereas PSA films possess spherulites at the surface. Blends of 70%, 50%, and 30% PLA exhibited granular respectively pitted surface structures but no fibrous structures visible in SFM. The monitoring of the erosion within 10 minutes showed that PLA is present in granules [16]. Phase imaging with the SFM yields phase shifts at different compositions ranging between 45 and 52 . The presence of PSA microdomains in blend of 50% PSA could be confirmed [215]. Screws, made of poly-L-lactic acid, were inserted axially into the right distal femur in 18 rabbits. The degradation and phagocytosis were assessed histologically and by TEM [198]. In the TEM specimens, polymeric particles of an average area of 2 m2 were seen to be located intercellularly with phagocytic cells. In 4.5-year specimens, the size of the polymeric particles, measured as area and perimeter, was significantly smaller than that of the 3-year specimens. The findings indicate that the ultimate degradation process of PLA is much longer than it was previously thought. Porous structured polymers to mimic natural extracellular matrix were investigated in [216] and in [217]. A biodegradable blend of starch with ethylene-vinyl alcohol copolymer (SEVA-C) with hydroxyapatite as filler was investigated, monitoring the degradation and the buildup of calcium phosphate crystals over a time span of up to 30 days in simulated body fluid. Only with the filler the buildup of the calcium phosphate layer could be observed. Within 8 hours

17 the roughness increased, after 24 hours calcium phosphate nuclei covered the surface, and after 126 hours a dense and uniform layer was present [218].

4.4. Composites Composites of different material systems can combine the advantages of both. Therefore a variety of composite materials have been developed. One distinguishes composites which work as bone cements [172, 219, 220] and those that include living cells and were used for tissue engineering [176, 221]. To quantify the bone-implant interface of high-strength HA/poly(L-lactide) (PLA) composite rods an affinity index was calculated, which was the length of bone directly joined to the rods, expressed as a percentage of the total length of the rod’s surface [219]. Calcined and uncalcined HA particles which amount to 30 or 40% by weight of the composite were implanted in the distal femora of 50 rabbits and after 2, 4, 8, and 25 weeks they were examined by SEM, TEM, and light microscopy. In all composites new bone formation could be examined after 2 weeks. SEM showed direct bone contact with the composites without intervening fibrous tissue. The affinity indices of all the composite rods were significantly higher than those of unfilled control PLA rods. The maximum affinity index (41%) was attained at 4 weeks in 40% calcinated HA-containing rods. Another approach to use composites is to increase the compressive strength of calcium phosphate cement (CPC) [220], which limits their use to non-load-bearing applications, by the means of water-soluble polymers. Composites formulated with the polycations poly(ethylenimine) and poly(allylamine hydrochloride) exhibited compressive strengths up to six times greater than that of pure CPC material. SEM results indicate a denser, more interdigitated microstructure. The increased strength was attributed to the polymer’s capacity to bridge between multiple crystallites and to absorb energy through plastic flow. Glass-ceramic, as mentioned previously, is also used in composites with bisphenol-a-glycidyl dimethacrylate (Bis-GMA)-based resin. These were compared with composites containing HA or -tricalcium phosphate (TCP) as the inorganic filler [172]. After 10 weeks of implantation into tibial metaphyses of rabbits, the ceramic-containing sample was in direct contact with bone and ceramic particles were partially adsorbed at the surface. The HA-containing cement was in contact with partially mineralized extracellular matrix. In 25-week specimens, ceramic particles were completely absorbed and replaced by new bone, and there was no intervening soft tissue. The tissue response of nano-hydroxyapatite/collagen composite was investigated in [176]. At the interface of the implant and marrow tissue, solution-mediated dissolution and giant-cell-mediated resorption led to the degradation of the composite. Interfacial bone formation by osteoblasts was also evident. The composite can be incorporated into bone metabolism instead of being a permanent implant. Nano-HA was also used in a porous collagen composite, to build up a three-dimensional osteogenic cell/nano-HA-collagen construct [221]. SEM and histological examination have demonstrated the development of the cell/material complex. Other biodegradable composites are

Nanoanalysis of Biomaterials

18 based on polyhydroxybutyrate-polyhydroxyvalerate (PHBPHV) [222]. SEM showed that the intended compositions of composites were achieved and bioceramic particles were well distributed in the polymer.

4.5. Biological Samples In the dental field enamel and dentin are materials under investigation. Effects of demineralization, for example, by soft drinks on the natural tooth structure, and improvements of resin adhesion important in dental therapy were studied. Topographical changes of human teeth in various liquids were characterized in [223]. The initially smooth and finely grained surfaces were changed to a rougher and more coarsely grained surface. Dissolution rates for soda pops with pH values between 4 and 5 were calculated to be 3 mm/