Sie sind auf Seite 1von 13

From Alloying to Nanocomposites Improved Performance of Hard Coatings**

By Roland Hauert* and Jrg Patscheider*


Nowadays a variety of different hard coatings are commercially available, the most widely used ones are TiN, TiC, TiCN, TiAlN, CrN, Al2O3, and combinations thereof, as well as some coatings with lubricating properties such as diamond-like carbon (DLC), WC/C or MoS2. To fulfil the industrial demands for improved coatings, a lower friction, a longer lifetime, a desired biological behavior or a better thermal stability in different environments, improved and application adapted coatings are developed. The different properties of a coating can be tuned to a desired value by alloying with suitable elements. Composite materials such as multilayer coatings and isotropic nanocomposite coatings, having structures in the nanometer range, can even show properties which can not be obtained by a single coating material alone. The authors review research and development work on the improvement of the overall coating performance. It mainly addresses alloying, the development of multilayer systems and the recently emerged field of nanocomposite coatings.

REVIEWS

1. Introduction
Attempts to change the properties of materials by alloying or structuring are as old as the earliest human activities in metal working; since several thousand years this technique has been used to improve the performance of bulk materials. The ongoing development of materials has been the basis of new products and industries since ancient times. During the early bronze-age copper was the main metal used. The alloying of copper with tin lowered the melting point, making casting much easier, and conferred considerable strength on the cast objects. During the iron age, iron and steel (carbon containing iron) could be produced. However, for heavy duty applications like tools (hammers, chisel) and weapons, iron was too soft and high carbon steel too brittle. Alloying of iron and steel did not result in improved properties, also the temperatures necessary for alloying could not be reached. By welding of iron and steel and subsequent repeated folding and forging, a laminated material named Damascus steel could be produced. The structure of Damascus steel consisted of iron and steel lamellas in the micrometer range, the material showing high hardness as well as a high ductility and was

capable of producing superior tools and weapons.[1,2] It may be regarded as the first multilayer material. From the Middle Ages to present, the increasing analytical possibilities allowed the creation of several thousand different types of steel. They differ by their alloying elements and their structure, each type of steel being optimized for a class of specific applications. Steel technology also provides examples of how additional phases can strengthen an material: PH-steels (precipitation hardening) shows high yield and tensile strength in combination with a high fracture toughness due to the precipitations formed by a suitable heat treatment. Nowadays, composite

[*] Dr. R. Hauert, Dr. J. Patscheider Swiss Federal Laboratories for Materials Testing and Research (EMPA) CH-8600 Dbendorf (Switzerland)

[**] Financial support by the presidents' funds of EMPA and EPFL and the Swiss Priority Program on Materials (PPM) is gratefully acknowledged. The authors thank M. Diserens for the TEM micrograph.

ADVANCED ENGINEERING MATERIALS 2000, 2, No. 5

1438-1656/00/00505-0247 $ 17.50+.50/0

247

Hauert, Patscheider/From Alloying to Nanocomposites materials are developed to synergistically combine desired material properties for many different applications. A large variety of functional properties (mechanical, wear, corrosion, fretting, weight, cost, bioreaction, etc.), can be optimized separately for the bulk material and the surface by applying an appropriate coating. Therefore, coated parts usually show a superior performance compared to uncoated workpieces. In the second half of the 20th century, surface coatings have emerged as an important industrial branch. In the beginning, coatings have been mainly deposited electrochemically. However, coatings exceeding the hardness of hard chrome (HV about 1000) like transition metal nitrides and carbides cannot be deposited from solution. In the late 1960s, TiC films deposited by CVD (chemical vapor deposition) on hard metal cutting tools have been introduced to the market. The beginning of the eighties saw the first PVD (physical vapor deposition) hard coatings appearing on the market, and allowed the coating of other materials than hard metals. Around 1980 TiN coatings deposited by CVD became commercially available (Figure 1). At the same time drills and cutting inserts with a TiN and TiC overlayer, coated by PVD (ion plating) were introduced. Around this time low friction coatings such as DLC (diamond-like carbon) deposited by PACVD (plasma activated chemical vapor deposition) and MoS2 deposited by PVD also became available (Figure 2). An introduction to the different deposition techniques can be found in the literature.[3] The ever-growing need for superior coatings to withstand severe operating conditions even at

REVIEWS

Fig. 1. TiN-cated cutting tools (Photo Berna-Bernex AG)

high temperatures, to have low friction, to last longer or to cause a desired bioreaction is the driving force for further development of new hard coatings. It should be highlighted, that it is crucial to regard the combination coating/substrate as a system. Differences in material properties (i.e., elastic modulus, thermal expansion coefficient, fatigue behavior, chemical compatibility) play a decisive role in the final behavior of the workpiece. Additionally the chemical situation at the interface coating/substrate will influence coating adhesion, i.e., a single atomic layer can decide if a coating will delaminate or adhere to the sub-

Roland Hauert studied Physics at the ETH in Zrich, Switzerland from 19791982. He worked as a Ph.D. student at the University of Basel in the research group of Prof. P. Oelhafen where he completed his thesis on electron spectroscopy of chemisorbed carbon monoxide on different alloys in 1986. After one year as a postdoctoral fellow at the IBM Almaden Research Center in San Jose, USA, he joined the Swiss Federal Laboratories for Materials Testing and Research (EMPA) in Switzerland. As head of the Surface Technology Laboratory, he is responsible for several research projects in bioreactions on modified diamond-like carbon coatings, nanostructured surfaces for tribological applications, nano-sender development, interface analysis, and adhesion studies. Additionally his surface analysis research group is a service center for solving surface-related problems for customers from industry and universities. In 1993 he worked during his sabbatical in the IBM Almaden Research Center in San Jose on diamond-like carbon coatings.

Jrg Patscheider received his diploma in inorganic chemistry from the University of Zrich, Switzerland for his work on plasma restoration and conservation of historic iron objects. During his Ph.D. time at ek he worked on PACVD of alumina and titanium nitride the University of Zrich with Prof. S. Vepr hard coatings. In 1990 he spent eight months as a postdoctoral fellow at Texas A&M University in College Station, TX on surface characterization of metal catalysts. In 1991 Patscheider joined the Swiss Federal Institute for Materials Testing and Research (EMPA) to establish a thin film deposition technology laboratory. He carried out application-oriented R+D work on deposition and adhesion of diamond-like carbon films and nucleation of diamond films. At present he is concerned with the development of nanocomposite coating systems. Besides this he is active in surface analysis as a service for external customers and quality assurance issues in surface analysis.

248

ADVANCED ENGINEERING MATERIALS 2000, 2, No. 5

Hauert, Patscheider/From Alloying to Nanocomposites physical properties such as hardness, lattice constant, thermal expansion coefficient, optical properties, electrical properties, grain size, texture, chemical reactivity, diffusion coefficient, elasticity, oxidation behavior, internal stress, grain size, texture, defects, and impurities, etc. altering the final performance. Although some basic correlations between the basic parameters and the final performance of the coating are known[5] there is still a lack of knowledge of how the basic properties and microstructure of a coating correlate with the final performance in a technical application. It is possible to deliberately change single parameters such as the lattice constant,[6] the electrical resistivity,[7] etc. by alloying en existing coating with a certain amount of an appropriate element. However, it is usually difficult or impossible to theoretically predict the influence of an alloyed element on the final performance in an application, especially in cases where the final properties are controlled by a nonlinear interaction of several different parameters as, e.g., in the case of the tribological behavior. Additionally, the introduction of an additional element into the deposition process (CVD or PVD) may alter the deposition conditions and therefore also the properties of the coating. When alloying an additional element with an existing single phase coating the additional element can either be diluted in the matrix and a new single phase coating is produced or the additional element forms a second phase resulting in a two phase coating. The case of two-phase coatings and one phase being present in the form of small grains with dimensions in the nanometer range is discussed separately in section 4 nanocomposite hard coatings. Since nearly all of the different properties of a coating can be influenced by alloying and there are thousands of publications on this subject, it is not possible to discuss them all within this article. We will restrict ourselves to the discussion of some selected properties. 2.1. Hardness Hardening of a material by alloying may be obtained in two ways. First, the chemical bonding between the atoms involved is changed by the element added. It has been demonstrated by Holleck[8] for nitrides and carbides that at a valence electron concentration of about 8.4 a maximum in the hardness in reached. A systematic approach to the different alloying possibilities of coatings is described by Lffler.[9] Second, the hardness of a material is also increased by hindering the dislocation movement by lattice distortions (grain boundaries, defects, dislocations, precipitations) as known from classical metallurgy. By alloying an additional element, defects (solidsolution hardening) as well as additional precipitations (precipitation hardening) and grain boundaries may be introduced into the coating, resulting in increased hardness. For example, on adding aluminum to TiN increased hardness is observed.[10,11] The incorporation of smaller aluminum atoms into the lattice of TiN leads to a local tensile stress (lattice distortion) which contributes to the higher hardness. Additionally a 1:1 mixture of TiN (9 valence electrons)

REVIEWS

Fig. 2. SEM cross-section of a 10 mm thick diamond-like carbon coating on a steel substrate showing a glass-like fracture morphology (Photo Berna-Bernex AG)

strate.[4] In this article we will only present the different possibilities to improve the properties of hard coatings. To change the properties of a coating various possibilities are at hand. Adequate elements can be alloyed to an existing coating to change its properties to a desired value as described in the section 2, modified coatings by alloying. In many cases the requirements on a coating cannot be fulfilled by a single coating material alone. Therefore, composite materials are developed, where different material properties are combined in a way that new properties are created. The final properties of such new materials are often controlled by the mutual interaction of the materials forming the composite, as described in the sections 3 and 4, multilayer coatings and nanocomposite hard coatings. When the dimensions of the layers or grains are in the nanometer range, superlattice effects may further improve the material properties. The different possibilities are displayed schematically in Figure 3. The deposition of gradient coatings, especially to adapt for differences in material properties between the substrate and the coating will not be discussed here.

2. Modified Coatings by Alloying


Adding different elements to an existing coating is one possibility to adapt some of its properties to a value desired for specific applications. This allows modification of the basic

Fig. 3. Schematic representation of the structure of the different coatings.

ADVANCED ENGINEERING MATERIALS 2000, 2, No. 5

249

Hauert, Patscheider/From Alloying to Nanocomposites and AlN (8 valence electrons) results in a valence electron concentration of 8.5 which is close to the favorable value of 8.4 described above. However, hard coatings can be produced in a way that they have already a very good defect and grain boundary density (usually a small grain size) as described for example in the reviews by Sundgren[5,12] and Holleck.[8] This is also the main reason why a deposited coating usually exhibits a higher hardness than a single crystal made of the same material, beside the internal stress in the films. Additionally, the internal stress in the films also contributes to the hardness of the film, but on the other hand large internal stress cause or favor delamination of the film. combinations thereof) into the film and still to maintain the amorphous phase of the coating. By this technique the different film properties such as thermal stability, hardness, internal stress, tribological properties, electrical conductivity, surface energy, and biological reactions of cells in contact with the surface (cell growth, cell differentiation) can be continuously adapted to a desired value. In the last two decades several hundred papers have been published on the changes of different properties of a-C:H when alloyed or doped with different elements. For an introduction to a-C:H see the review articles by Robertson.[20,21] Due to its bio-homeocompatible nature,[22,23] there is growing interest in the application of a-C:H on orthopedic and other implants.[24] Especially the manipulation of surface bioreactions by adding adequate elements into the amorphous aC:H matrix is expected to result in new fields of application (Figure 4). For example, Sicontaining a-C:H coatings are under evaluation as a special coating for cardiovascular prostheses. In the case of artificial heart valves the properties of aC:H could be improved by the incorporation of Si. The Si in the a-C:H matrix renders the surface antithrombogenic by inhibiting the fibrinogen activation.

REVIEWS

2.2. High Temperature Oxidation Surface oxidation at high temperatures is an issue not only for applications in hot environments, but also in cutting and drilling operations, where locally elevated temperatures of up to 800 C can be reached at cutting edges, and even on planes exposed to severe friction. Efforts have therefore been undertaken to improve the high temperature oxidation behavior of coatings. TiN, for example, has, despite its wide range of applicability, several major disadvantages. One is that it starts to oxidize severely at temperatures above 500 C. The concept of adding oxidation-resistant elements such as Al, Zr, and Cr during the deposition process of TiN led to the development of several new single-phased materials derived from TiN such as TiAlN,[10,11] TiZrN,[11,13,14] and CrZrN.[13] The phase diagrams and selected thermophysical properties of ternary transition metal nitride and carbide systems, which are of potential interest as hard coatings, have been described in extenso by Holleck.[15] In TiAlN, which is commercially available since the first half of the nineties, the introduction of aluminum in TiN increased the oxidation resistance from approximately 550 C to 800 C and additionally an increased hardness is observed.[10,11] The enhanced oxidation resistance is a consequence of the formation of an aluminum-rich protective Al2O3 passive layer at the surface.[1618]

Fig. 4. SEM image of mouse fibroblast cells just after mitosis on N-containing DLC (Photo ETH-BWB).

2.3. Surface Chemistry, Bioreactions The surface chemistry, which can be easily modified by introducing additional elements into a coating, plays an important role in bioreactions as well as tribology, surface energy, catalysis, etc. If a surface has to be created for a desired bioreaction, a promising approach is to start from an existing biocompatible coating (to prevent inflammatory reaction or rejection) and to alloy it with suitable elements. Additionally, it has to be considered that the bioreactions and the in-vivo behavior of an implant are influenced also by the surface texture as described by Boyan.[19] Amorphous DLC (a-C:H) is an excellent base coating to be alloyed with different elements. The amorphous nature of aC:H opens the possibility to introduce certain amounts of additional elements such as Si, F, N, O, W, V, Co, Mo, Ti, (and

A desirable achievement is to obtain a bioreactive surface enhancing bone.[25] As an example, when Ti containing a-C:H (a-C:H/Ti) is exposed to a biological environment, the adsorption of different proteins could be altered as a function of the Ti content in the film.[26] The adsorbed proteins will subsequently influence cell attachment, cell proliferation, and cell differentiation. Bone marrow cell culture experiments on these a-C:H/Ti coatings demonstrated that the differentiation of bone marrow cells into bone resorbing cells, i.e., osteoclasts, is inhibited on all titanium containing a-C:H coatings[25] making a-C:H/Ti a valuable coating for implants. 2.4. Friction and Wear The tribological behavior of a coating can be influenced by adding different elements to the coating, however, a prediction of the influence of the elements added is usually not possible. The real response of a coating in a specific application has to be determined by conducting field tests.[27,28] In a tribo-

250

ADVANCED ENGINEERING MATERIALS 2000, 2, No. 5

Hauert, Patscheider/From Alloying to Nanocomposites logical system many properties and parameters (the different material properties of the friction counterparts, environment, test conditions) interact with eachother in a nonlinear and locally different way which prevents a theoretical prediction of the final tribological behavior. Additionally, when a coating is improved by alloying for a certain tribological application, it may perform worse under some other tribological condition, as shown by Lffler[9] for the wear of TiN, TiAlN, TiCN during face milling for different cutting speeds. While the classical hard coatings such as TiN, TiAlN, TiC, TiCN have friction coefficients of 0.50.8 against steel, a-C:H exhibitsbeside a low wearalso low friction coefficients between 0.05 and 0.2 against most materials except some polymers. These properties are the main reasons for its application as a low friction coating. Amorphous DLC (a-C:H), modified by alloying with different elements has, due to its outstanding properties, been studied extensively as a tribological coating as described in the review articles by Grill, Donnet, and Gangopadhyay[2932] and is also established in several industrial applications.[3336] For example, the addition of N to the a-C:H matrix in an application as a protective overcoat for magnetic thin film rigid disks resulted in a 3 to 4 times improved performance (wear, start stop, stiction) compared to plain a-C-H.[34] Also wear-resistant F containing a-C:H films (F is introduced into the film to render the surface more hydrophobic to prevent stiction of the read/write head in magnetic rigid disks), have been deposited.[30,37] The dependence of the coefficient of friction on the ambient humidity could be improved by the addition of Si (Figure 5).[30,38,39] A review of the influence of different concentrations of Ti, W, Nb, Si, and Ta in a-C:H films on the wear and other properties is given by Klages.[40]
0.20
Counterface 100Cr6 steel

mechanical and tribological properties by the metal content in a-C:H is given by Klages.[40] As an example, the surface energy of most coatings can be modified by suitable alloying element as in the case of a-C:H where the wetting angle was continuously lowered by the incorporation of N or O and was increased to more hydrophobic values by Si, SiO, and F.[41] As an example of continuous variable material properties by alloying, the advancing and receding wetting angle, hardness, and electrical resistivity are displayed as a function of the N content of a-C:H:N coating in Figure 6.[7] Implantation of different elements can improve the surface properties, either by changing the chemical situation due to the implanted elements or by introducing intrinsic stress into the surface. An introduction to ion implantation can be found in the literature.[42]

REVIEWS

Fig. 6. The advancing and receding wetting angle, hardness, and electrical resistivity are displayed as a function of the N content of a-C:H:N coatings.

3. Multilayer Coatings
0.15 Friction coefficient
Pure DLC

0.10
Si-containing DLC

0.05

Relative humidity R.H. 65%

R.H. 5%

R.H. 85%

0.00 0 2 4 6 Sliding distance (km) 8

Fig. 5. The friction coefficient for different relative humidities of pure DLC and SiDLC containing 5 atomic percent Si against 100Cr6 steel as af function of the sliding distance in a pin-on-disk experiment.

2.5. Other Selected Properties There are several thousand papers presenting different results on how one or more parameters have been changed by adding one or more elements to an existing coating. A survey of a continuous variation of the electrical, magnetic, optical,

A composite coating usually consists of two or more phases combined either as different layers (multilayer) or as a homogeneous isotropic mixture of different phases (multiphase). The aim of a multicomponent coating is to combine desired properties from different components as well as the creation of new properties generated by the combination of suitable materials. Since coatings are usually deposited by CVD or PVD processes, multilayers can easily be obtained by changing either the feed gas in CVD process or by the alternate use of different material sources (magnetron, evaporator, etc.) in PVD deposition processes. When dealing with multilayer coatings two cases have to be distinguished. Classical multiple layer coatings and nanoscale multilayer coatings. Classical multilayer or multiple layer coatings consists of a few (usually 3 to 20) different layers and having a total thickness of 18 mm. These types of multilayer coatings are commercially available since the early 1970s. They usually consist of one or more Al2O3 or TiAlN layers (0.051 mm thick) between layers of TiN, TiC, and TiCN.[43,44] The Al2O3 or TiAlN layers are introduced to ob-

ADVANCED ENGINEERING MATERIALS 2000, 2, No. 5

251

Hauert, Patscheider/From Alloying to Nanocomposites tain a better high temperature oxidation behavior (stable Al2O3 surface oxide layer) compared to single layers of TiN or TiC. Additionally, multilayer coatings consisting of materials with different mechanical properties may show a better performance since crack propagation may be deflected or stopped at the interfaces or in the more ductile material.[45] The tribological behavior of the classical single layer coatings TiN and TiC could be improved by building multilayer structures. TiN/TiC, TiN/TiC/BN, TiN/TiC/B4C, TiN/ TiVC/AlN, and TIN/TiC/SiC multilayer structures composed of 3150 layers, exhibit a lower coefficient of friction as well as a longer edge life when applied on cutting tools, as shown by Holleck.[46] Reviews of the different multilayer and multicomponent coatings and their improved tribological performance are given in the articles of Submaranian and Holleck.[45,46] Improved wear and hardness of different layered systems containing up to 20 layers of Ti/TiC/a-C:H can be found in the review article on pulsed laser deposition by Voevodin.[47] An additional benefit was obtained by introducing a gradient TixC1x interlayer, where the concentration (x) is adapted to gradually increase hardness and thermal expansion coefficient from the steel substrate to the hard DLC top layer.[4749] Forces caused by the internal stress and by differences in the thermal expansion can be very high at the interface coating/substrate and, exceeding the adhesion, result in delamination. By the introduction of an adapted gradient layer the forces are dispersed over the gradient coating and therefore lowered at the interface.

REVIEWS

Fig. 7. Hardness as a function of the lattice period for TiN/VN, TiN/NbN, TiN/ V0.6Nb0.4N (no difference lattice mismatch), and Nb/VN (no difference in elastic shear modulus) multilayers. Data taken from [51,52,55,57].

3.1. Nanoscale Multilayers In nanoscale multilayer coatings, the thickness of each single layer being in the nanometer range, superlattice effects may additionally increase the hardness of the coating. Since the late 1980s single crystal (epitaxial) nanoscaled superlattice films with increased hardness have been grown and analyzed. An introduction to nanoscaled multilayer coatings can be found in several review articles.[46,5053] It was shown by Helmersson and Shinn that, when the superlattice period (thickness of the bilayer packet) is around 5 nm the hardness of TiN/VN[54] and TiN/NbN[50] epitaxial nanoscale multilayer films exceeds 50 GPa, which is an enhancement of more than a factor of two compared to the hardness of the base materials. Similar hardness values have also been obtained for polycrystalline TiN/VN and TiN/NbN nanoscale coatings by Chu[55] and Sproul[51,52] for a superlattice period of 48 nm. These measurements are displayed in Figure 7. The reason for this hardness enhancement is a hindering of the dislocation movement across a sharp interface, when the two materials have a large difference in their dislocation line energies.[56] Since the dislocation line energy is proportional to the shear modulus, the main requirement for a hardness enhancement in these films is a difference in the elastic moduli of the two layer materials, as described by Sproul,[51] Shinn,[57] and

Chu.[58] To verify this concept, multilayer structures have been produced with two materials having nearly the same share moduli, namely with NbN and VN.[55] These hardness values, where no increase in hardness was observed, are also displayed in Figure 7. Analogous V0.6Nb0.4N/NbN nitride multilayer films have been produced by Shinn[57] and are also included in Figure 7. V0.6Nb0.4N and NbN have nearly the same shear moduli, but their lattice mismatch is about 3.6 % the same as that of TiN/NbN.[57] No hardness increase was observed which can be taken as a proof that coherency strains play a minor role in the hardening of multilayer systems, but a difference in the shear modulus is a precondition for a hardness increase in nanoscale multilayers. A convincing model explaining the hardness enhancement of nanoscale multilayer films is given by Chu.[58] The model takes into account dislocation movement across an interface, dislocation glide within individual layers as well as an interface with interface composition gradient. The calculations confirmed that a difference in the shear modulus is a precondition for a hardness increase in nanoscale multilayers. If each single layer is thinner than 35 nm the strain field around a dislocation is mainly outside the particular layer and the hardness increase induced by the periodical shear modulus variation will disappear in the multilayer structure.[50] Additionally, depending on the deposition conditions and the thermal load applied, interdiffusion between the layers will decrease the amplitude of the shear modulus modulation and therefore also the hardness. The interface sharpness plays therefore a decisive role in the hardness enhancement of multilayers. At an interface between two materials, the lattice mismatch is producing a strain in the first few atomic rows of the material, the so called coherency strain. The influence of the

252

ADVANCED ENGINEERING MATERIALS 2000, 2, No. 5

Hauert, Patscheider/From Alloying to Nanocomposites coherency strain on the hardness of multilayers was investigated by Mirkarimi.[59] He deposited TiN/V0.6Nb0.4N multilayer films in which no lattice mismatch exists between the two materials (see Fig. 7). An enhanced hardness could also be obtained in this lattice matched multilayer but the sharp hardness maximum observed for TiN/NbN and TiN/VN in the range of 48 nm is not present. This indicates that the coherency strain may be responsible only for the additional hardness increase between 48 nm multilayer period[50] which is named super modulus effect.[53] At multilayer periods above 10 nm the coherency strain disappears due to lattice relaxation as also observed by XRD (X-ray diffraction) analysis.[50] The magnitude of a possible hardness increase caused by coherency strain is not clear. The V0.6Nb0.4N/NbN nitride multilayer has the same coherency strain as TiN/ NbN, but no additional hardness increase is observed in the range of a 310 nm multilayer period (Fig. 7). Possibly, a change in material properties near an interface, induced by strain or by electronic interaction with the neighboring material has to be considered as well. In the literature[60] it was demonstrated, that the hardness as well as the oxidation resistance of TiAlN at elevated temperatures can be improved additionally by building a TiAlN/ CrN multilayer structure having a multilayer period (thickness of bilayer) of 3.6 nm. 3.2. Templates in Nanoscale Multilayers When building a superlattice coating from two materials with different crystallographic structures, one material may act as a template and force the first few atomic layers of the other material in its own crystallographic structure and therefore create a material with new properties. For example in the TiN/Cr2N superlattice system the usually hexagonal Cr2N can be forced into a fcc structure,[57] as long as the Cr2N layer is thin enough, so that the energy difference from the forced cubic phase to the hexagonal phase is smaller than the coherency strain energy. TiN/AlN multilayer coatings show maximum hardness enhancement at a multilayer period of about 2.5 nm. In these superlattice films, TiN (NaCl-structure) forces the AlN (usually Wurzite structure) to be present in the NaCl-structure, a structure that usually exists only at high pressure and which is supposed to be very hard.[61] 3.3. Tribology of Nanoscale Multilayers When wear-resistant coatings are applied to chipping tools (drills, cutting inserts, hobs, etc.), very high temperatures up to 800 C may occur at the cutting edge. The relevant parameters for a good coating are therefore not the mechanical parameters alone, but also the chemical stability towards oxidation. Especially the toughness at the high temperatures during the application is a more relevant parameter than the hardness.[46] Multilayer structures may also be produced simply by a periodical variation of the deposition conditions. Since the internal stress on TiN films produced by reactive sputtering (ion plating) depends on the applied sample bias voltage, a periodical modulation of the internal stress in the film can be introduced by modulating the self bias voltage during deposition. By this technique TiN/TiN multilayer films with increased hardness and decreased wear have been obtained for a multilayer period of around 8 nm.[62] Analogously, a-C:H multilayer structures have been deposited by periodical variation of the sample self bias voltage between 100 and 600 V during film growth. These a-C:H multilayer structures show periodical variation of their physical properties (hardness, density, internal stress, etc.). Although these films do not show a hardness increase for the different multilayer periods, they show a low wear down to 4 1018 m3 N m for multilayer periods of 20 and 30 nm which could not be reached by the homogeneous films deposited at 100 V and 600 V self bias.[63] Multilayer coatings with multilayer periods in the nanometer range are commercially available, for example WC/C coatings by the tradename Balinit C (Figure 8).

REVIEWS

Fig. 8. TEM micrograph of the WC/C multilayer structure of Balint C (Photo Balzers AG)

4. Nanocomposite Hard Coatings


Analogous to the nanoscale multilayer coatings presented above, it is possible to deposit isotropic nanocomposite coatings consisting of crystallites, embedded in an amorphous matrix, with grain sizes in the nanometer range. Two different materials, namely the crystalline and the amorphous phase, are deposited simultaneously and the nanocomposite material forms by a phase separation. A prerequisite for the phase separation is complete immiscibility of the two phases. In contrast to the multilayer structures, where any material combination can be deposited at any multilayer period, nanocomposites can only be obtained for certain material combinations. Additionally, the size of the crystallites can not be independently controlled by the deposition process, because it is essentially determined both by the properties of the materials and by the deposition conditions (temperature, plasma conditions, elemental concentrations, etc.). In the last decade some nanocomposite thin film systems, which show promising results for applications, have been deposited and investigated.

ADVANCED ENGINEERING MATERIALS 2000, 2, No. 5

253

Hauert, Patscheider/From Alloying to Nanocomposites

REVIEWS

4.1. The TiN/Si3N4 System Based on the concept of incorporating stable oxide-forming elements (Al, Si, Hf, Cr, Zr, Nb) into TiN, efforts have been undertaken to codeposit silicon and titanium nitride. In contrast to TiAlN, TiZrN, and other single-phased hard materials, silicon cannot be substitutionally built in the lattice of TiN. In accordance with the TiSiN phase diagram which does not present any stable ternary phase under equilibrium conditions,[64] two-phase TiN/Si3N4 coatings form when silicon is added during deposition of TiN. The first coatings consisting of TiSiN have been produced by CVD in 1982 by Hirai et al.[65] Posadowski used triode sputtering from a Ti/ Si-component target to deposit cermet (ceramic metal composite) coatings with a temperature-independent electrical resistance.[66] Nicolet produced TiSiN coatings by reactive sputtering from TiSix-targets as electrically conductive diffusion barriers for copper metallization.[67,68] The first successful attempts to improve the hardness of TiN by addition of silicon have been reported by Li et al.[69] ek et al.[7072] They deposited TiNSi3N4 films by and Vepr plasma-enhanced CVD at deposition temperatures of 550 C 600 C. The process of Li used TiCl4, SiCl4 and H2,[69] while ek used SiH4 instead of SiCl4. These films showed an unVepr usual high hardness of about 60 GPa at 15 at.-% silicon in the TiSiN film while only X-ray signals from TiN were observed.[69] In further experiments these coatings were identified as nanocomposites consisting of TiN crystallites of about 47 nm (nc-TiN) surrounded by an amorphous Si3N4 (aSi3N4) matrix.[70] Such a nanocomposite is schematically illustrated in Figure 9 with TiN crystals embedded in an amorek and coworkers were phous matrix of Si3N4. Li et al., Vepr the first to report hardness values of nanocomposite TiN Si3N4 coatings. The nc-TiN/a-Si3N4 films showed very high hardness values of above 50 GPa at a Si content of 8 at.-%. At this value, which corresponds to approximately 19 at.-% Si3N4, a minimum of the crystallite size of the TiN grains and a hardness maximum of 50 GPa was found.[70,73,74] This finding suggests a decisive influence of the interface between the grains and the amorphous phase, since the highest hardness is observed at the highest fraction of crystallite surface per volume. The hardness (and the crystallite size) is a distinct function of the silicon concentration in the film as can be seen in Figure 10, which compares results from various groups.

Fig. 10. Hardness of nc-TiN/a-Si3N4 nanocomposites as a function of the silicon content. Solid lines are for reactive PVD experiments, dashed lines are for PACVD films. The hardness maxima between 5 and 12 at.-% Si are obvious.

Fig. 9. Schematic representation of a nanocomposite consisting of a nanocrystalline phase embedded in an amorphous matrix.

The variation of the data is probably due to different procedures for hardness measurements and for composition determination. The advantage of using chlorinated species in deposition processes is the easy introduction of all reactants into the deposition system. However, gas phase nucleation processes, which occur at appreciable rates at typical PACVD pressures (10100 Pa), are a threat to the homogeneity of the growing film. Moreover, the incorporation of chlorine, originating from unreacted educts, is an inherent problem of PACVD with chlorinated educts. Intense discharges are required to shift the chemical equilibrium in the PACVD process sufficiently to the product side, so that the chlorine content of the films is less than fractions of 1 %.[75] Such discharges, however, lead to substrate temperatures of 500 C and higher, which is too high for applications as protective coatings for most steels. The gases involved in the PACVD process of TiN/Si3N4, (TiCl4, SiH4, H2, HCl) pose additional drawbacks for process engineering. A way to circumvent these problems is the use of PVD. The process has to be reactive PVD in order to form the nitrides from the sputtered Ti and Si targets using nitrogen as ek, there has to be reactive gas. As pointed out earlier by Vepr a sufficiently high activity of nitrogen to suppress the formation of titanium silicides at the expense of titanium nitride and silicon nitride.[73] Unbalanced magnetron sputtering (UBM), a magnet arrangement in a conventional planar magnetron where the central magnets are stronger than the outer ones, is a technique which provides sufficiently high degrees of dissociation and ionization at the substrate.[76] Superimposing a glow discharge on the substrates placed on the radiofrequency-powered (RF-powered) sample holder increases the reactivity of the gas mixture even more as well as the surface mobility of adsorbed species on the growing film. Vaz et al.[77,78] as well as Diserens et al.[79] were the first to report on hardness enhancement of TiN/Si3N4 coatings by PVD, which were codeposited from titanium and silicon targets using re-

254

ADVANCED ENGINEERING MATERIALS 2000, 2, No. 5

Hauert, Patscheider/From Alloying to Nanocomposites active UBM sputtering. Both observed the increase of hardness at relatively low silicon concentrations and in this way ek's group. Hardness measureconfirmed the results of Vepr ments performed with a Nanoindenter using the continuous stiffness method (a method providing a depth profile of the nanohardness within only one indentation cycle[80]) showed that at approximately 10 % to 12 % Si3N4 in nc-TiN/a-Si3N4 the hardness peaks are close to 40 GPa.[7779,81] Transmission electron micrographs clearly show the presence of diffracting (200) planes of TiN nanocrystallites (see Fig. 11). Fourier transformation of the micrograph yield a mean crystallite size of 2.5 nm. However, much higher values are reported for materials deposited by PACVD, where hardness values exceeding that of diamond are postulated.[73,74,82] These hardness findings relate to the remaining plastic deformation according to the definition of the Vickers hardness. dislocation movement. The sharp interfaces encountered in nc-TiN/a-Si3N4 form due to the immiscibility of the two phases. There are chemical reasons which control the formation of sharp interfaces during deposition of nanocrystalline/ amorphous systems while in multilayer systems these interfaces have to be generated by a proper process technology. In order to achieve the sharp phase transition, a perfect encapsulation of the crystalline material by the second phase is necessary. This is the case for superlattices with atomically sharp interfaces. For nanocrystalline isotropic materials such a close coverage of nanocrystals can best be obtained by amorphous materials which, due to their structural flexibility, can adapt to the shape and orientation of the nanocrystals. Furthermore, incoherence stresses which originate from the random orientation of the crystallites, can best be accommodated by an amorphous material with its possibility to adopt any orientation within a few atomic distances. Hardening of materials due to finite grain size effects has been observed in metals since long time. Such phenomena are described, e.g., by the HallPetch relation, which ascribe the higher hardness of metals at smaller grain size to a pileup of dislocations at grain boundaries or, in other words, to a hindered dislocation motion across grain boundaries.[83] However, classical theories, like the HallPetch theory, cannot account for the observed increase of hardness of nc/a systems, exceeding the hardness of the base materials by a factor of 24. Like in multilayer systems, the highest hardness is observed at very small grain sizes of the order of a few nanometers. This distinct dependence highlights the crucial role of the grain size for the hardness of nc/a systems. Hindered dislocation movement alone cannot account for these superior properties. A dislocation has to be at least of several nanometers to form. In an isotropic nanocrystal of e.g., 2 nm diameter, the distances are of the order of five unit cells. The small size of such a nanocrystal precludes the development of dislocations and hence no dislocations may agglomerate at grain boundaries. Instead, at least for carbide-based materials, plastic deformation as a consequence of mechanical load is supposed to occur via pseudoplastic deformation, where the nanocrystals are moved against each other.[84] This requires more work per unit volume because a substantial rearrangement of material is necessary to produce a remaining deformation. On a macroscopic scale this is equivalent to a higher resistance against deformation or, in other words, is synonymous to an increased hardness. An additional improvement of nc-TiN/aSi3N4 coatings as compared to TiN is their improved oxidation resistance. While TiN already shows significant rates of oxidation above 550 C to 600 C, nanocomposite nc-TiN/a-Si3N4 coatings are clearly more resistant to oxidation.[70,73,81] In contrast to singlephased TiN-based coatings like TiAlN, no silicon enriched protective layer was found at the outer surface of oxidized TiN/a-Si3N4; this is indicative of a complex oxidation mechanism. At a Si3N4 fraction of 12 % (which corresponds to the hardness maximum), the oxidation rate at 800 C is lowered

REVIEWS

Fig. 11. Transmission electron microscopy (TEM) cross section micrograph of a ncTiN/a-Si3N4 nanocomposite coating with 11 at.-% silicon where the nanocrystalline structure is clearly apparent. The (200)-planes of TiN with its spacing of 0.212 nm are resolved. The mean crystallite size of TiN is 2.5 nm.

The observed increase of hardness with decreasing crystalline size resembles strikingly the behavior of the hardness evolution in multilayer superlattices which are described above in Section 3. The resistance of a crystalline material against deformation by dislocation movement is described by the shear modulus. It basically gives the necessary work to induce shearing deformation by dislocation movement of the material as it occurs upon plastic deformation of a crystalline solid. Analogous to the situation in multilayer structures described above, a hardness increase can be expected in any twophase material, as long as they have sharp interfaces and a large difference in their shear modulus. One striking difference to the multilayer systems, however, is the fact that nanocrystalline/amorphous systems are isotropic, or, in other words, that the orientation and sequence of both phases (crystalline and amorphous) is random. It is reasonable to assume that the shear moduli of these two phases are different (both materials differ in their structure and in their electronic properties); following the argumentation of Chu[58] they will form barriers for

ADVANCED ENGINEERING MATERIALS 2000, 2, No. 5

255

Hauert, Patscheider/From Alloying to Nanocomposites by a factor of about ten as compared to TiN. A further increase of the silicon nitride content lowers the oxidation rate even more, however, at the expense of the hardness, which decreases down to the hardness of a-Si3N4 at Si3N4 fractions of 30 % and higher. At these values the oxidation rate is lower than the one of TiAlN.[16] A two-step oxidation process is likely to govern the oxidation in the regime between 600 C and 1000 C. The first process, being dominant at low temperatures up to about 820 C, is characterized by the slow diffusion of oxygen through the Si3N4 barriers into the TiN nanocrystals. An increase of the Si3N4 fraction in the coatings means a thicker Si3N4 layer encapsulating the TiN grains, and hence a lower oxidation rate is observed. This is accompanied by the formation of an amorphous silica passivation layer which further reduces the oxygen diffusion rate into TiN. Above 820 C the oxidation is dominated by the recrystallization and growth of TiO2 (rutile) crystals which leads to wellfacetted grains and, caused by the discontinued protective a-SiOx and a-Si3N4 layers, to an accelerated oxidation.

REVIEWS

4.2. Carbon-Based Nanocrystalline/Amorphous Materials Nanocomposite thin films with outstanding properties are not restricted to the nitride systems alone. The incorporation of carbide particles into amorphous carbon (a-C:H, DLC) was reported by Sundgren already since the beginning of DLC as a low-friction hard coating.[85] Work by Klages et al.[40] and Dimigen[86,87] on doping of a-C:H films with metals like Ti, Ta, Nb, and others also supposed the presence of carbide nanocrystals in the amorphous a-C:H matrix. Knotek et al.[88] observed a phase segregation of titanium carbide and hydrogenated carbon when deposited by reactive PVD at conditions overstoichiometric in carbon (Figure 12). Films consisting of TiC crystallites in a hydrogen-free a-C matrix have been deposited by a combined pulsed laser desorption and reactive magnetron sputtering process. The ablation of carbon in the laser plume and its subsequent ionization lead to the formation of a dense amorphous network of a-C.[89,90] In contrast to the TiN/Si3N4 nanocomposites, this type contains significantly larger grains of some tens of nanometers with a thicker separating amorphous phase between the nanocrystallites of about 5 nm. The crystallite size of 1050 nm is large enough to allow the formation of dislocations, but is too small for self-propagation of cracks. The larger grain separation permits adaptation of incoherency strains and enables formation of nanocracks between the crystallites to allow pseudoplastic behavior. In this way, this material exhibits four times higher toughness than single-crystalline TiC and hardness values of about 32 GPa.[84] During mechanical load, the pseudoplasticity (instead of the highly elastic behavior of superhard TiN/Si3N4) allows the distribution of load peaks originating from asperities of the friction counterpart, instead of crack formation and hence the lifetime of the coating can be prolonged or its destruction can even be avoided. These TiC/ a-C films were therefore called load-adaptive coatings.

Fig. 12. Hardness a), crystallite size and mean distance between the crystallites b) of TiC/a-C:H nanocomposites as a function of the composition. Maximum hardness occurs when the crystallite size and the separation by the amorphous phase are minimal.

4.3. Other Selected Properties Many other material properties can be changed by introducing an additional element in the form of nanocrystals into an existing coating. The optical properties of a-C:H coatings have been adapted by the introduction of W and Cr as nanosize carbidic inclusions in the film. These coatings are applied, e.g., as selective absorber coatings for thermal solar energy conversion.[91] An additional benefit can also be obtained by introducing materials with lubricating properties such as MoS2, C or DLC to a coating, either as top layer or as a composite coating as shown for example by Gilmore for the TiN/ MoS2, TiB2/MoS2, and TiB2/C nanocomposite systems.[92,93] From the work on nanocomposites consisting of ncTiN/aSi3N4 and nc-TiC/a-C some generalized building principles for nanocomposites with enhanced mechanical properties can be derived: l One phase must be sufficiently hard to bear the load. This is usually fulfilled for transition metal nitrides and carbides as well as for some main group oxides. l The other phase should provide structural flexibility to act as a binder for the nanocrystallites. Amorphous materials like Si3N4, a-C, a-C:H, and possibly others are best suited for this purpose. l Immiscibility of the phases is a prerequisite to ensure the sharp transition of the elastic properties from one phase to the other.

256

ADVANCED ENGINEERING MATERIALS 2000, 2, No. 5

Hauert, Patscheider/From Alloying to Nanocomposites

5. Conclusions and Outlook


A variety of possibilities to deliberately improve the properties of hard coating have been presented. Alloying of a coating during deposition, while maintaining the deposit as a single phase, is a possibility of changing most of the properties of a coating. Especially the hardness, toughness and the chemical properties are of prime interest for applications. The most important characteristics for the successful use of a hard coating are the tribological properties, hardness and toughness, its oxidation behavior and in some cases, the biological behavior. The dilemma of increased brittleness at high hardness seems to have been relieved by the development of composite coatings. We have shown for the multiphase coatings (e.g., nc-TiN/ a-Si3N4, nc-TiC/a-C, multilayers) that the combination of different materials which have either good oxidation resistance, ductility, hardness or chemical behavior can lead to a synergistic combination of these properties to give coatings with improved performance. These composite coatings are prepared and used as multilayers or as nanocomposites. The increased hardness of these biphased materials is a consequence of additional interfaces between different materials which hinder dislocation movement and which are places of energy dissipation and crack deflection. Multilayers can be regarded as the highly anisotropic case and nanocomposites as the isotropic case of biphased materials. When the dimensions of the layer thickness in multilayer coatings or the dimensions of the grains in nanocomposite coatings are in the range of a few nanometers (dimension of the strain field around dislocations, of the coherency strain and of altered electronic structure), superlattice effects can further increase the hardness of the coating. It has been shown that a main requirement for improved hardness by hindered dislocation movement is a difference in the shear modulus between the two materials. In order to have a material for practical use, the adhesion between the two materials at each interface of the composite material as well as the one between the coating and the substrate has to be at least in the range of the cohesion in the weaker of the two. If this is not the case, the interfaces will be the weakest part of the whole composite system and on increased mechanical load one of the interfaces will be the primary source of cracks, fatigue or delamination. A prime area of applications for new hard coatings is the use of coated tools for dry machining. Moreover, new coatings have to show improved wear resistance to extend their lifetime. Coatings with increased hardness are expected to fulfil this demand. High load bearing and rolling applications do not necessarily need extremely hard surfaces, but toughness and low friction are key requirements which may be matched by nanocomposites or multilayers with lubricating solid phases. Hard coatings development in the seventies and eighties led to the appearance of a limited choice of coatings which were used for a wide palette of applications. This holds especially true for the single-phase materials TiN, TiCN, and TiC.

Future new coatings will have to compete against their successful predecessors and therefore new hard coatings with synergetic ensembles of properties will occupy niches where no existing product can satisfy the requirements. Another incentive for the development of new performance-improved coatings is the scientific curiosity of materials scientists. Based on prediction rules and theoretical models, which emerged over the last two decades, the behavior of materials, the interaction between different phases, especially in the nanometer range, is fairly well understood. This knowledge will lead to the appearance of new coatings, which may have superior properties as compared to what is available today. However, the resistance of a coated surface to wear or even catastrophic failure cannot be predicted from such first principles. If and when a coating will fail, is not determined by its properties alone, but is to a large extent influenced by the nature of the friction counterpart and, in most cases, by lubricant phases (oils, transfer films, etc.) separating the two bodies. When the counterpart or the force on the counterpart or a lubricating phase changes, the overall situation is modified and a different wear behavior will occur. Friction and wear problem engineering should therefore be regarded as a whole system where hard coatings play an important, but not the only role. [1] R. F. Tylecote, A History of Metallurgy, 2nd ed., Institute of Materials, London, 1992. [2] C. S. Smith, A Search for Structure, Reprint ed., MIT Press, Cambridge, MA May 1983. [3] Thin Film Processes (Eds: J. L. Vossen, W. Kern), Academic Press, New York 1978. [4] K. H. Ernst, J. Patscheider, R. Hauert, M. Tobler, Surf. Interface Anal. 1994, 21, 32. [5] J.-E. Sundgren, H. T. G. Hentzell, J. Vac. Sci. Technol. A 1986, 4, 2259. [6] P. Hones, R. Sanjins, F. Lvi, Thin Solid Films 1998, 332, 240. [7] R. Hauert, A. Glisenti, S. Metin, J. Goitia. J. H. Kaufman, P. H. M. van Loosdrecht, A. J. Kellock, P. W. Hoffmann, Thin Solid Films 1995, 268, 22. [8] H. Holleck, J. Vac. Sci. Technol. A 1986, 4, 2661. [9] F. H. W. Lffler, Surf. Coat. Technol. 1994, 68/69, 729. [10] W.-D. Mnz, J. Vac. Sci. Technol. A 1986, 4, 2717. [11] O. Knotek, M. Bhmer, T. Leyendecker, J. Vac. Sci. Technol. A 1986, 4, 2695. [12] J.-E. Sundgren, Thin Solid Films 1985, 128 , 21. [13] D. P. Monaghan, D. G. Teer, K. C. Laing, I. Efeoglu, R. D. Arnell, Surf. Coat. Technol. 1993, 59, 21. [14] W. D. Sproul, Surf. Coat. Technol. 1996, 81, 1. [15] H. Holleck, Binre und Ternre Carbid- und Nitridsysteme der bergangsmetalle, Borntrger, Berlin 1984. [16] D. McIntyre, J. E. Greene, G. Hakansson, J.-E. Sundgren, W.-D. Mnz, J. Appl. Phys. 1990, 67, 1542. [17] C. W. Kim, K. H. Kim, Thin Solid Films 1997, 307, 113. [18] S. Hofmann, H. A. Jehn, Surf. Interface Anal. 1988, 12, 329.

REVIEWS

ADVANCED ENGINEERING MATERIALS 2000, 2, No. 5

257

Hauert, Patscheider/From Alloying to Nanocomposites [19] B. D. Boyan, T. W. Hummert, D. D. Dean, Z. Schwartz, Biomaterials 1996, 17, 137. [20] J. Robertson, Prog. Solid State Chem. 1991, 21, 199. [21] J. Robertsom, Surf. Coat. Technol. 1992, 50, 185. [22] M. Allen, F. Law, R. Rushton, Clin. Mater. 1994, 17, 1. [23] R. Butter, M. Allen, L. Chandra, A. H. Lettington, R. Rushton, Diamond Relat. Mater. 1995, 4, 857. [24] D. P. Dowling, P. V. Kola, K. Donelly, Diamond Relat. Mater. 1997, 6, 390. [25] G. Francz, A. Schroeder, R. Hauert, Surf. Interface Anal. 1999, 28, 3. [26] R. Hauert, L. Knoblauch-Meyer, G. Francz, A. Schroeder, E. Wintermantel, Surf. Coat. Technol., in press. [27] S. Hogmark, P. Hedenqvist, S. Jacobson, Surf. Coat. Technol. 1997, 90, 247. [28] S. Hogmark, S. Jacobson, J. STLE Lubrication Eng. 1992, May 1992, 401. [29] A. Grill, Surf. Coat. Technol. 1997, 9495, 507. [30] C. Donnet, Surf. Coat. Technol. 1998, 100101, 180. [31] A. Gangopadhyay, Tribol. Lett. 1998, 5, 25. [32] A. Grill, Diamond Relat. Mater. 1999, 8, 428. [33] R. Hauert, U. Mller, M. Tobler, Conference proceeding of the 17th international SAMPE EURPE Conference (Ed: U. Meier), 2830. Mai 1996, Basel, Switzerland, p. 367. [34] E. C. Cutiongco, D. Li, Y. W. Chung, C. S. Bhatia, Trans. ASME 1996, 118, 543. [35] K.-R. Lee K. Y. Eun, Mater. Sci.Eng. A 1996, 209, 264. [36] J. Gttler, J. Reschke, Surf. Coat. Technol. 1993, 60, 531. [37] C. Donnet, J. Fontaine, A. Grill, V. Patel, C. Jahnes, M. Belin, Surf. Coat. Technol. 1997, 94/95, 531. [38] D. Neerinck, P. Persoone, M. Sercu, A. Goel, D. Kester, D. Bray, Diamond Relat. Mater. 1998, 7, 468. [39] A. K. Gangopadhyay, P. A. Willermet, M. A. Tamor, W. C. Vassell, Tribol. Int. 1997, 30(1), 9. [40] C. P. Klages, R. Memming, Mater. Sci. Forum 1989, 52/ 53, 609. [41] M. Grischke, K. Bewilogua, K. Trojan, H. Dimigen, Surf. Coat. Technol. 1995, 74/75, 739. [42] G. A. Collins, R. Hutchings, J. Tendys, M. Samandi, Surf. Coat. Technol. 1994, 68/69, 285. [43] J. R. Coleman, Manuf. Eng. 1990, 104, 38. [44] W. Schnitlmeister, O. Pacher, T. Krall, W. Wallgram, T. Raine, Powder Metall. 1981, 13(1), 26. [45] C. Subramanian, K. N. Strafford, Wear 1993, 165, 85. [46] H. Holleck, V. Schier, Surf. Coat. Technol. 1995, 76/77, 328. [47] A. A Voevodin, M. S. Donley, J. S. Zabinski, Surf. Coat. Technol. 1997, 92, 42. [48] A. A Voevodin, M. A. Capano, S. J. P. Laube, M. S. Donley, J. S. Zabinski, Thin Solid Films 1997, 298, 107. [49] A. A Voevodin, S. D. Walk, J. S. Zabinski, Wear 1997, 203/204, 516. [50] M. Shinn, L. Hultman, S. A. Barnett, J. Mater. 1992, 7, 901. [51] W. D. Sproul, Surf. Coat. Technol. 1996, 8687, 170. [52] W. D. Sproul, Science 1996, 273, 889. [53] J.-E. Sundgren, J. Birch, G. Hkansson, L. Hultman, U. Helmerson, Thin Solid Films 1990, 193/194, 818. [54] U. Helmersson, S. Todorova, S. A. Barnett, J. Sundgren, L. C. Markert, J. E. Greene, J. Appl. Phys. 1987, 62, 481. [55] X. Chu, PhD Thesis, Northwestern Univeristy, Evanstown, IL 1993. [56] J. C. Koehler, Phys. Rev. B 1970, 2, 547. [57] M. Shinn, S. A. Barnett, Appl. Phys. Lett. 1994, 64, 61. [58] X. Chu, S. A. Barnett, J. Appl. Phys 1995, 77, 4403. [59] P. B. Mirkarimi, L. Hultman, S. A. Barnett, Appl. Phys. Lett. 1990, 57, 2654. [60] I. Wadsworth, I. J. Smith, L. A. Donohue, W.-D. Mnz, Surf. Coat. Technol. 1997, 94/95, 315. [61] M. Setoyama, A. Nakayama, M. Tanaka, N. Kitagawa, T. Nomura, Surf. Coat. Technol. 1996, 86/87, 225. [62] S. J. Bull, A. M. Jones, Surf. Coat. Technol. 1996, 78, 173. [63] L. Knoblauch-Meyer, R. Hauert, Thin Solid Films 1999, 338, 172. [64] P. Rogl, J. C. Schuster, Phase diagrams of ternary boron nitride and silicon nitride systems, ASM Int., Materials Park, OH 1992. [65] T. Hirai, S. Hayashi, J. Mater. Sci. 1982, 17, 1320. [66] W. Posadowski, Thin Solid Films 1988, 126, 111. [67] X. Sun, J. S. Reid, E. Kolawa, M.-A. Nicolet, J. Appl. Phys. 1997, 81, 656. [68] Y. Tsuji, S. M. Gasser, E. Kolawa, M.-A. Nicolet, Thin Solid Films 1999, 350, 1. [69] L. Shizhi, S. Yulong, P. Hongrui, Plasma Chem. Plasma Process. 1992, 12(3), 287. ek, S. Reiprich, L. Shizhi, Appl. Phys. Lett. 1995, [70] S. Vepr 66, 2640. ek, Pure Appl. Chem. 1996, 68, 1023. [71] S. Vepr ek, S. Reiprich, Thin Solid Films 1995, 268, 64. [72] S. Vepr ek, M. Haussmann, L. Shizhi, Proc. Electrochem. [73] S. Vepr Soc. 1996, 96(5), 619. ek, P. Nesladek, A. Niederhofer, F. Glatz, M. Ji[74] S. Vepr lek, M. Sima, Surf. Coatings Technol. 1998, 109/(13), 138. ek, Curr. Top. Mater. Sci.1980, 4, 151. [75] S. Vepr [76] B. Window, N. Savvides, J. Vac. Sci. Technol. A 1986, 4/2, 196. [77] F. Vaz, L. M. Rebouta, S. Ramos, M. F. da Silva, J. C. Soares, Surf. Coat. Technol. 1998, 108109, 236. [78] F. Vaz, L. M. Rebouta, S. Ramos, A. Cavaleiro, M. F. da Silva, J. C. Soares, Surf. Coat. Technol.1998, 101, 110. [79] M. Diserens, J. Patscheider, F. Lvy, Surf. Coat. Technol. 1998, 108109, 241. [80] W. C. Oliver, G. M. Pharr, J. Mater. Res. 1992, 7(6), 1564. [81] M. Diserens, J. Patscheider, F. Lvy, Surf. Coat. Technol., in press. [82] A. Niederhofer, P. Nesladek, H.-D. Mnnling, K. Moto, ek, M. Jilek, Surf. Coat. Technol., in press. S. Vepr [83] R. W. K. Honeycombe, The Plastic Deformation of Metals, Edward Arnold,London 1975, p. 234. [84] A. A. Voevodin, J. S. Zabinski, J. Mater. Sci. 1998, 33, 319.

REVIEWS

258

ADVANCED ENGINEERING MATERIALS 2000, 2, No. 5

Hauert, Patscheider/From Alloying to Nanocomposites [85] J.-E. Sundgren, B.-O. Johansson, S.-E. Karlsson, Thin Solid Films 1983, 105, 353. [86] H. Dimigen, C. P. Klages, Surf. Coat. Technol. 1991, 49, 543. [87] K. Bewilogua, H. Dimigen, Surf. Coat. Technol. 1993, 61, 144. [88] O. Knotek, E. Lugscheider, F. Lffler, B. Bosserhoff, S. Schmitz, Mater. Sci. Eng. A 1996, 209, 394. [89] A. A. Voevodin, M. A. Capano, A. J. Safriet, S. Donley, J. S. Zabinski, Appl. Phys. Lett. 1996, 69, 188. [90] A. A. Voevodin, S. V. Prasad, J. S. Zabinski, J. Appl. Phys. 1997, 82, 855. [91] R. Gampp, P. Gantenbein, Y. Kuster, P. Reimann, R. Steiner, P. Oelhafen, S. Brunold, U. Frei, A. Gombert, R. Joerger, W. Graf, M. Khl, Proc. of Optical Materials Technology for Energy Efficiency and Solar Energy Conversion XIII (Eds: V. Wittwer, C. G. Granqvist, C. M. Lampert), Proc. SPIE 1994, 2255, pp. 92106. [92] R. Gilmore, M. A. Baker, P. N. Gibson, W. Gissler, M. Stoiber, P. Losbichler, C. Mitterer, Surf. Coat. Technol. 1998, 108/109, 345. [93] R. Gilmore, M. A. Baker, P. N. Gibson, W. Gissler, Surf. Coat. Technol. 1998, 105, 45.

REVIEWS

______________________

ADVANCED ENGINEERING MATERIALS 2000, 2, No. 5

259

Das könnte Ihnen auch gefallen