389 Nanoindentat 16. Nanoindentation: Localized Probes of Mechanical Behavior of Materials David F. Bahr, Dylan J. Morris 16.1 Hardness Testing: Macroscopic Beginnings ........................ 16.1.1 Spherical Impression Tests: Brinell and Meyers ....................... 16.1.2 Measurements of Depth to Extract Rockwell Hardness ......... 16.1.3 Pyramidal Geometries for Smaller Scales: Vickers Hardness 16.2 Extraction of Basic Materials Properties from Instrumented Indentation ............. 16.2.1 General Behavior of Depth Sensing Indentation ........ 16.2.2 Area Functions ............................. 16.2.3 Assessment of Properties During the Entire Loading Sequence 389 390 390 391 392 392 394 395 16.3 Plastic Deformation at Indentations ....... 396 16.3.1 The Spherical Cavity Model ............ 397 16.3.2 Analysis of Slip Around Indentations ............................... 398 16.4 Measurement of Fracture Using Indentation ................................ 399 16.4.1 Fracture Around Vickers Impressions................................. 399 16.4.2 Fracture Observations During Instrumented Indentation .. 400 16.5 Probing Small Volumes to Determine Fundamental Deformation Mechanisms.. 402 16.6 Summary ............................................. 404 References .................................................. 404 16.1 Hardness Testing: Macroscopic Beginnings Out of the many ways in which to measure the mechanical properties of a material, one of the most common is the hardness test. While more basic materials properties can be determined from testing materials in uniaxial tension and compression (along with shear testing), the hardness test remains useful in both industrial and laboratory settings for several reasons. First, the testing method is very quick and easy, and allows many measurements on a given sample. Secondly, the equipment for the test methods and required sample preparation Part B 16 This chapter focuses on mechanical probes of small volumes of materials using contact based testing methods performed on instruments designed to measure mechanical properties, rather than those which are inherently developed for scanning probe microscopy. Section 16.1 consists of basic information used in the measurement of localized mechanical properties and provides a brief review of engineering hardness testing, followed in Sect. 16.2 by a discussion of the basis for instrumented indentation testing to determine elastic and plastic properties of the material being tested: the common modulus and hardness techniques. More advanced methods for using localized probes beyond determining basic elastic and plastic properties, including developments in the analysis of the deformation zone around indentations and the resulting information that can be garnered from this information, are covered in Sect. 16.3. Methodologies that are used to estimate other mechanical properties such as fracture, are covered in Sect. 16.4, which focuses on bringing instrumented indentation testing into realms similar to current macroscopic mechanical testing. Finally, the chapter concludes in Sect. 16.5 with methods that use nanoscale indentation testing to determine more fundamental properties, such as the onset of dislocation activity, size-dependent mechanisms of deformation, and activation volumes. 390 Part B Contact Methods Part B 16.1 is relatively inexpensive, making the tests attractive for small laboratories with limited budgets. Finally, and most importantly, the tests can be made on the actual parts and is sensitive to lateral position. For large parts, a small indentation is usually nondestructive, and allows for quality control checks of each part. Sectioning the sample allows testing of parts which have varying properties through the thickness of the part, such as case hardened gears. While hardness may not be as pure a measurement of a material’s properties as uniaxial testing, the advantages of these methods ensure that the test methods will continue to be used in both industrial and research settings for many applications. Traditional hardness tests have been used for decades to measure the resistance of a material to deformation. In the late 18th and early 19th century several attempts were made to rank the deformation of materials by scratching one material with another. If the scratching material marked the tested material, the scratching material was considered harder than the scratched material. While this was useful for comparative purposes (and is still referred to in the Mohs hardness scale), there remained the problem of quantifying hardness. In the late 19th century work by Hertz [16.1] and Auerbach [16.2] brought about early examples of static indentation testing. In these cases, the samples (either balls or flat plates) were probed with a ball of a given material, and the deformation was measured by considering the contact area between the ball and sample. 16.1.1 Spherical Impression Tests: Brinell and Meyers It was not until the early 1900s that Brinell [16.3] presented a standard method of evaluating hardness, based on applying a fixed load to a hard spherical indenter tip into a flat plate. After 15–30 s, the load was removed and the diameter of the impression was measured using optical microscopy. The Brinell hardness number (BHN) is defined as BHN = π D2 2P , d 2 1− 1− D (16.1) where P is the applied load, D the diameter of the indenter ball, and d is the chordal diameter of the residual impression. Note that this method evaluates the load applied to the surface area of the residual impression. However, the Brinell test has been shown to be affected by both the applied load and diameter of the ball used for the indenter. The idea that hardness should be a materials property, and not dependent on the test method, led to the observations that the parameter which remains constant for large indentations appears to be the mean pressure, defined as the load divided by the projected contact area of the surface. Meyer suggested in 1908 [16.4] that the hardness should therefore be defined as 4P , (16.2) H= πd 2 where H is the hardness. For a confirmation of the work of Meyer (and a report of the ideas proposed in English) the reader is referred to the study of Hoyt [16.5] in 1924. This type of hardness measurement is alternately referred to as the mean pressure of an indentation, and denoted by either p0 or H in the literature. Tabor describes in detail one of the direct benefits of utilizing spherical indentation for determining the hardness of a material[16.6]. As the indentation of a spherical indentation progresses, the angle of contact of the indentation changes. From extensive experiments in the 1920s to 1950s, it was found that the effective strain of the indentation εi imposed by a spherical tip could be approximated by d (16.3) . D Similarly, for materials such as copper and steel, H is approximately 2.8 times the yield stress. This allowed the creation of indentation stress–strain curves; by applying different loads to spherical tips and measuring the impression diameters one found that the hardness of a material which work hardens increased with increasing indentation area for spherical tips. εi ≈ 0.2 16.1.2 Measurements of Depth to Extract Rockwell Hardness Both the Brinell and Meyer hardness methods require each indentation to be observed with an optical microscope to measure the diameter of the indentation. This makes the testing slow and requires a skilled operator to carry out the test, with the result that the method is more appropriate to a laboratory setting than industrial settings. The Rockwell hardness test, designed originally by Stanley P. Rockwell, focused on measuring the depth an indenter penetrated into a sample under load. As the depth would correspond to an area of contact if the geometry of the indenter was know, it was thought that measuring the depth would provide an Nanoindentation: Localized Probes of Mechanical Behavior of Materials accurate and automatic measurement of hardness. For a detailed description of the history of the development of hardness testing, the reader is referred to the work of Lysaght [16.7]. In standardizing the Rockwell test methods, it became common to use a conical diamond indenter tip which was ground to an included angle of 120◦ when testing very hard materials. This indenter tip, referred to as a Brale indenter, is the basis for the common Rockwell C, A, and D scale tests. During a Rockwell test, the sample is loaded to a preload of 10 kg to alleviate effects of surface roughness. After this preload is applied, a larger load is applied, referred to as the major load. The depth measured during the test is the change in distance between the penetration of the tip at preload and after the application of the major load. Effectively, this is the first time that depth sensing was instrumented during indentation testing, and eventually leads to the development of instrumented indentation methods. 16.1.3 Pyramidal Geometries for Smaller Scales: Vickers Hardness where DPH is the diamond pyramid hardness number, P is the load applied in grams, and now d is the length of the diagonals in mm (usually the average of the two diagonals). This is analogous to the Brinell test, in that the DPH is the ratio of the load to the surface area of the indentation. The main advantages of this method is that there is a continuous scale between very soft and very hard materials, and that the DPH is constant over a wide range of loads until very low loads (less than 50–100 g) are reached. However, the test still required each indentation to be measured optically to determine the hardness. The DPH can be converted to a mean pressure by simply using (16.2), by substituting the projected area for the ratio. The DPH tests were then adapted to loads of less than 1000 g, which are now commonly referred to as microindentation techniques. This allowed testing of very small areas around the indentation. In addition, as the tests began to gain use in probing cross sections of case hardened materials, the lateral resolution was altered by adding the Knoop indenter tip, which is a four-sided diamond with an aspect ratio of 7:1 between the diagonals [16.6]. The small areas deformed by the DPH and Knoop methods allowed testing of thin coatings and of different microstructural features, such as areas of ferrite and martensite in a steel alloy. Other indenter geometries have been developed, in particularly the Berkovich tip, which is a three-sided pyramid with the same projected contact area as the Vickers tip [16.9]. This results in a face angle between an edge and the opposing face of approximately 65.3◦ . The driving force towards using the Berkovich geometry over conical or Vickers pyramids is that fabricating diamond tips with extremely small root radii is challenging. As three planes intersect at a point, experimentally it is easier to fabricate the Berkovich geometry. One should be aware that even the most finely fabricated diamond tips do not reach the pyramidal geometries at their apex, and commonly are approximated by an effective tip radius [16.10]. Experimental measurements of commercially available tips [16.11] show that the tip radius of Berkovich indenters is commonly less than 200 nm. If a simple geometrical model of an axisymmetric indenter such as a cone radiating from a spherical cap is applied, one finds that the depth at which the indenter tip transitions from a spherical region of radius R to the conical aspect of the tip, h tip occurs at h tip = 1 − sin θ R (16.5) for a conical tip with an included half angle of θ. As there are many analytical mechanics solutions developed for the contact of axisymmetric solids, one 391 Part B 16.1 While depth sensing indentation methods were being developed, a change in imaging-based indention methods was also occurring. The development of the Vickers (or diamond pyramid hardness) indentation test, and determination of the angles used during pyramidal indentations [16.8], followed the basic method of the Brinell test, but replaced the steel ball with a diamond pyramid ground to an included angle between faces of 136◦ . It is interesting to note that the angle chosen was based on the Brinell test. It had become common to make indentations with the Brinell method to residual indent diameters corresponding to 0.25–0.5 of the ball diameter, and measurements of BHN as a function of applied load showed an effective constant hardness region between d/D ratios of 0.25–0.5. The average of these is 0.375, and if a four-sided pyramid is formed around a spherical cap such that the ratio of the chordal diameter to ball diameter is 0.375, the angle must be 136◦ . After applying a load, the indent is imaged, and the diagonals of the indentation are measured. The Vickers hardness is then defined as ◦ 2P sin 136 2 (16.4) DPH = d2 16.1 Hardness Testing: Macroscopic Beginnings 392 Part B Contact Methods common technique is to apply an equivalent included half angle for a cone that would provide the same relation between depth of contact and the projected contact area of the solid. For both the Vickers and Berkovich geometries, the ideal equivalent conical included half angle is approximately 70.3◦ . For a tip radius of 200 nm, the effective depth of transition is 12 nm. A microhardness testing machine was modified to continuously monitor the load and depth during an indentation by Loubet et al. [16.12] among others [16.13]. This led to the development of what is commonly referred to as nanoindentation, though the term instrumented indentation is more accurate. Any procedure which is capable of monitoring the load and depth during an indentation can be referred to as instrumented indentation, continuous indentation or nanoindentation, however inappropriate the scale may be when discussing indentations which penetrate depths of microns and contact areas over 10 μm2 . The following section details the processes by which one experimentally determines the properties of a material using these instrumented indentation techniques. 16.2 Extraction of Basic Materials Properties from Instrumented Indentation Part B 16.2 There are many recent reviews and resources which describe the general methods of determining the elastic and plastic properties of materials using nanoindentation. This section will provide a brief review of the developments in this area; readers that wish for a more extensive treatment of the technique or specific aspects are referred to several reviews [16.14–18]. In particular, this section will address the calculation of elastic properties from an indentation load–depth curve, the methods to determine hardness from the same test, methods which have improved the technique, and address some of the limitations of the overall methodology of nanoindentation. 16.2.1 General Behavior of Depth Sensing Indentation When a hard tip is impressed into a solid, the resulting load to the sample and resulting penetration of the tip relative to the initial surface can be recorded during the process. This is typically referred to as the load– depth curve for a particular tip–solid system, and is shown schematically in Fig. 16.1. The load–depth curve has several important features which will be used in the following discussion. The maximum load Pmax , the maximum penetration depth h max , and the final depth h f will be required to describe the indentation process. In addition, the load–depth relationship can be determined throughout the loading process, and the stiffness is given by dP/ dh. As shown by Doerner and Nix [16.19], and further developed by Oliver and Pharr [16.20], an analysis of the unloading portion of the load–depth curve allows the modulus and hardness of the material to be determined; the following treatment is based on that work. This type of analysis is based on the work of Sneddon for the indentation of an elastic half space with a rigid shape or column [16.21]. While Sneddon originally used the analysis for the impression of an axisymmetric punch, the current nanoindentation approach is often used for relatively shallow indenters of pyramidal geometries such as the Berkovich geometry. Under load, one assumes that the profile of the indentation is given by Fig. 16.2, where the contact radius (if a conical indentation were being used) would be defined by a, and the maximum penetration is given by h max . Contact mechanics since the time of Hertz has recognized that the elastic deformation of two solids can be Load P Creep Pmax S = dP/dh Loading hf Unloading hmax Displacement h Fig. 16.1 Schematic load–depth curve for the penetration of a indentation tip into a flat solid Nanoindentation: Localized Probes of Mechanical Behavior of Materials described by composite terms. Extensive treatments of contact mechanics for the advanced reader can be found in Johnson’s text [16.22]. For the elastic compression of two solids in contact, the compliances add in series, and it is convenient to define a reduced elastic modulus E ∗ 1 − νi2 1 − νs2 1 = + , ∗ E Ei Es (16.6) √ 2 dP = β √ E∗ A , dh π Initial surface (16.7) a Indenter At Pmax hc hmax Surface profile After unloading Initial surface hf Fig. 16.2 Schematic of the surface of a solid while in con- tact with an indentation tip and after removing the loaded tip where A is the projected contact area of the solid and β is a constant that attempts to account for the differences between the axisymmetric contact models and the experimental variations in using pyramidal geometries as well as possible variations due to plastic deformation, and the fact that the contact area is beyond the small-strain conditions assumed in many contact mechanics problems. This correction factor β is still a subject of debate and recent research and reported values range from the commonly references King’s 1.034 [16.23] to values of over 1.1, as reviewed by Oliver and Pharr [16.17] and Fischer-Cripps [16.24]. A recent study [16.25] describes these effects in detail. The currently accepted recommendation is that using β = 1.034 [16.17] or 1.05 [16.25] will allow reasonably accurate values of elastic modulus to be determined, given the other possible inaccuracies in the calibration process. There remain two system characteristics which must be calibrated if an accurate assessment of the elastic modulus and hardness of the sample are to be determined. Of course, the load–depth curve provides measurements of P and h, but not A, which is needed to calculate both the elastic properties and hardness of the solid. First, one must estimate the contact depth from the load–depth curve. The common method of assessing this requires the initial unloading slope, dP/ dh, to be determined. If the indenter were a flat punch, then the elastic unloading would be linear in load–depth space. This was the initial assumption made by Doener and Nix [16.19], that the initial unloading slope was linear. However, if the contact was the elastic contact of a blunt axisymmetric cone in the manner of Sneddon [16.26], the relationship would be parabolic with depth. If the contact were that of a sphere of radius R, rather than a flat punch or cone, the original Hertzian elastic contact solution could be used, in which the relationship would be to the 1.5 power. The loading conditions for an axisymmetric elastic contact between an indenter and a flat surface, such that a contact radius of a can be described as P = 2aE ∗ h for a flat punch indenter of constant radius a, 2E ∗ 2 P= h tan α π for a conical indenter of included half angle α, 4 P = E ∗ Rh 3 3 for a spherical indenter of radius R. (16.8) 393 Part B 16.2 where E and v are the elastic modulus and Poisson’s ratio respectively, and the subscripts ‘i’ and ‘s’ refer to the indenter and sample, respectively. For the conventional indentation technique, diamond is selected for the indenter tip, and so 1141 GPa is commonly used for E i , while vi is 0.07. During nanoindentation, one standard assumption is that both elastic and plastic deformation occur during loading, beginning at the lowest measurable loads. Section 16.5 will describe cases in which this is not a logical assumption, but for many tests this assumption will be valid. In this case, and with a shallow indenter (i. e., one with an included angle similar to that of the Berkovich geometry), the unloading of the indentation will be used to determine the elastic properties of the solid, since the common assumption is that the unloading is purely described by the elastic response of the system. In many metallic and ceramic systems this is indeed experimentally observed (i. e., after partially unloading, subsequent reloading will follow the load–depth curve until the previous maximum load is achieved). If this is the case, then the unloading of the system can be described by Extraction of Basic Materials Properties 394 Part B Contact Methods Therefore, if the unloading slope, dP/ dh, is to be determined, it would at first glance appear to be possible to select the appropriate relationship, perform a leastsquares fit on experimental data, and determine the elastic modulus of the material. However, due to plastic deformation, none of these methods are exactly accurate. In practice the easiest manner found to determine the unloading slope is to perform a least-squares fit to the unloading data using a power-law function which can be described by P = A(h − h f )m , (16.9) Part B 16.2 where m will be between 1 and 2, and h f can be estimated from the load–depth curve. This form is convenient in that the slope of this form will never be negative during the unloading segment. In practice various portions of the unloading segment are selected for fitting due to efforts to minimize drift or noise in the system; hold periods are often inserted in the loading schedule. Any curve fitting procedure should ignore regions of the unloading curve which contain extended holding segments. With the proper analysis of the unloading stiffness, dP/ dh at the maximum depth and load, the contact depth (Figs. 16.1 and 16.2) can be determined. For an axisymmetric indenter the common method is to determine the contact depth h c as hc = hf − ε Pmax , dP/ dh The frame compliance, commonly referred to in the indentation literature as Cf , relates the displacement in the system at a given applied load, wherein the system consists of the springs which suspend the indenter tip, the shaft or screw which holds the indenter tip, and the mounting mechanism for the sample. The displacement carried by the total system will be dependent on each given sample due to mounting, but this is often accounted for by determining the compliance for a typically stiff sample and then assuming that this frame compliance is constant for other similarly mounted samples. The total load–displacement relationship can be described by dh dh = + Cf . (16.11) dP measured dP sample In practice, if the unloading stiffness is measured at a variety of contact depths for a variety of stiff materials (sometimes fused silica, sapphire, and tungsten are recommended), and once these values are inverted this measured compliance is plotted as a function of inverse contact depth, the intercept is Cf . This also allows the definition of the contact stiffness of the sample, rather than of the system. The sample stiffness, S, is just ( dP/ dh)sample and will be used to determine the elastic modulus from a given indentation. 16.2.2 Area Functions (16.10) where ε is a constant equal to 0.72 for an axisymmetric indenter. Commonly ε is chosen to be 0.75, which describes the data well for pyramidal indenter geometries. One last parameter must be determined experimentally to accurately determine the stiffness at unloading: the frame compliance of the system itself. When any mechanical test is performed, there will be a portion of the displacement carried by the elastic deformation of the loading frame. In bulk mechanical testing this is well documented; if one wishes to utilize crosshead displacement and load data the compliance of the system is subtracted at a given load, assuming a linear spring constant. In the case of bulk testing, the suggested method to carry out mechanical testing is to utilize strain gauges and extensometers rather than relying on the cross head displacement, because the displacement given by the frame may be a substantial part of the overall displacement. However, during nanoindentation testing there is no direct analog to the extensometer, and therefore the frame compliance must be accounted for during the test. With all these parameters now defined, the final step is to determine the area function of the tip. The initial point of analysis for this method relied upon the contact area, not the depth of penetration, and therefore, in a manner reminiscent of the work of Rockwell, the depth of penetration can be related to an area for a given tip geometry. Many authors have developed a series of techniques to relate the penetration depth to the projected contact area of the tip with the sample. Two popular methods are described in the literature; one chooses an empirical functional dependence of the projected contact area to contact depth, and the other models the tip as a spherical cap on an effectively conical contact. In both cases, the calibration begins by performing indentations to various depths in materials of known elastic modulus (hence both are sometimes referred to as constant-modulus methods). The traditional materials chosen include fused silica and single-crystal aluminum; however other materials such as sapphire and tungsten are sometimes used as calibration standards. Rearranging (16.7), it is possible to Nanoindentation: Localized Probes of Mechanical Behavior of Materials describe A= π 4 S β E∗ 2 (16.12) and if E ∗ is constant, then A can be determined for any given S at a contact depth h c . A can either be described empirically [16.20] as A = C1 h 2c + C2 h c + C3 h 0.5 c + C4 h 0.25 + C51 h 0.125 +... , c c (16.13) 16.2.3 Assessment of Properties During the Entire Loading Sequence Analysis of the unloading slope provides information from an individual point during an indentation, and as much more data during an instrumented indentation tests is collected, there has been an effort to develop methods to extract more information about the properties of materials from the entire load–depth curve. These fall into three categories, those which approach the loading curve from an analytical standpoint, those which impose a small oscillation upon the loading curve 395 to create a series of unloading segments, and those which utilize finite element solutions to solve inverse problems that provide more information about materials properties including strain hardening coefficients and rate sensitivity of the test. The first two portions, as they are still being used to primarily calculate E and H, will be discussed here, while the third topic will be covered in Sect. 16.3. The first method to extract materials properties from the entire loading segment focuses on the fact that for a given material–tip combination the loading curve tends to describe the balance between elastic and plastic deformation, referred to by Page as the mechanical fingerprint of the material. Hainsworth and coworkers [16.27] developed a novel method initially proposed by Loubet et al. [16.28] in which the shape of the loading curve instead of the unloading curve is considered. This model is particularly applicable for evaluating materials where unloading curves exhibit a considerable amount of elastic recovery. For instance, some coated materials, such as CNx on silicon, show a well-behaved loading curve but a highly curved unloading segment. The relationship between loading displacement and applied load was described by [16.27] P = Kmh2 , (16.15) where P is the applied load, h is the corresponding displacement, and K m is an empirical constant which is a function of Young’s modulus, hardness as well as indenter geometry. For a Berkovich tip, K m is expressed as −2 E H . (16.16) K m = E 0.194 + 0.930 H E Therefore, if either E or H is known, the other may be calculated by curve fitting the experimental loading curve using the combination of (16.16) and (16.15). A second method to determine the properties of materials during indentation builds upon the unloading aspect of the test. If a sample is repeatedly loaded and unloaded to greater depths, one effectively determines the unloading slope at different contact depths. As the initial unloading slope can be determined from a partially unloaded segment, the ability to sample in one position the properties as a function of depth is particularly appealing in two cases: firstly where either the material is layered or has some positional variations in properties (or size-dependent properties) and secondly when a spherical indentation is used for the experiment. The instrumented indentation of a material with a sphere would allow sampling at different contact-area- Part B 16.2 where in this case C1 is often chosen to fit the ideal tip shape (24.5 for a Berkovich tip). Subsequent constants are empirically selected to provide a best fit to the elastic modulus from the measured data. The other method assumes that the tip is described using a spherical cap on an effective cone of included half angle α with a tip radius of R [16.10] h c C1 2 + A= C1 C2 πh 2c + 4Rπh c + 4R2 π cot2 α , = (16.14) cot2 α where the constants C1 and C2 are related to α and R, respectively. The other methodology used for tip calibration relies upon actual measurement of the tip as a function of position using scanning probe microscopy (or some other quantitative profilometer method) [16.11], and is less common due to the difficulty in performing these experiments; most users will likely utilize the constant-modulus methods. Having defined a method to determine the projected contact area as a function of contact depth, the hardness H as determined in (16.2) is easily determined. The hardness measured by nanoindentation is expected to deviate from the hardness measured by post-indentation inspection if the surface does not remain flat during the entire test. Extraction of Basic Materials Properties 396 Part B Contact Methods to-depth ratios, which is the same effect as measuring hardness at different d/D ratios (see the previous section). Field and Swain demonstrated this for nanoindentation using a cyclic loading–partial unloading sequence [16.29]. By this time nanoindentation testing was becoming more standardized, and they noted that spherical indentations would probe different strains, while conical or pyramidal indentation would probe a fixed strain value; effectively (just like (16.3) but under load) a (16.17) εi ≈ 0.2 R for spheres while for pyramids and cones εi ≈ 0.2 tan α , (16.18) Part B 16.3 allowing them to calculate hardness as a function of penetration depth. This paper also deals with the prediction of the load–depth graphs knowing materials properties such as modulus and yield strength. The last method of determining materials properties during nanoindentation is commonly referred to as the continuous stiffness method, though other terms are utilized by a variety of commercial manufacturers of systems. In all cases, the process relies upon an imposed oscillation in the force (or displacement) applied to the tip. This is carried out at a specific frequency. The technique imposes a sinusoidal forcing function at a fixed frequency ω, and uses this signal to calculate the contact stiffness from POS −1 2 S−1 + Cf + K s − mω2 + ω2 D2 h(ω) = (16.19) or from the phase difference between the displacement and force signals from ωD , (16.20) tan(φ) = −1 −1 S + Cf + K s − mω2 where Cf is the compliance of the load frame, K S is the stiffness of the column support springs, D is the damping coefficient, POS is the magnitude of the forcing oscillation, h(ω) is the magnitude of the resulting displacement oscillation, φ is the phase angle between the force and displacement signals, and m is the mass. Material properties can then be calculated at each of these tiny unloading sequences. Further extension of this work by Joslin and Oliver [16.30] led to the realization that, as both H and E ∗ use the area of contact and applied load, at any point during the loading cycle if the stiffness were measured using a lock-in amplifier the quantity P 4 (16.21) S2 π (neglecting correction factors such as β) provides a measurement of the resistance to deformation which does not require the use of an area function. For assessing the properties of materials which may be expected to have depth-dependent hardnesses (for instance, in ionimplanted solids) this technique promised the ability to assess these properties without the need to calibrate tips to extremely shallow depths. Of course, if one knows the area function of the tip through a calibration technique, then it is possible to determine the modulus directly if the stiffness is known at all points during loading. The continuous stiffness method and its variations are now common methods of assessing materials properties as a function of depth during a single indentation. The obvious benefits of this method include the ability to sample small volumes (i. e., probe laterally distinct features) which may not be able to be the subject of many tests, to perform tip calibrations and property measurements in thin-film systems rapidly, and to provide extremely sensitive methods of system constants. For instance, a new method of determining frame compliance uses this technique[16.17]; when fused quartz is indented using the continuous stiffness method, P/S2 is a constant 0.0015 GPa−1 . This method is quite sensitive to a system’s frame compliance, and obviously not to the area function. Therefore, users of indentation systems can rapidly calibrate the frame compliance to account for the variations in the systems with time and ensure more accurate area functions for use in materials properties assessment. H 2 E∗ = 16.3 Plastic Deformation at Indentations The importance of imaging and examining indentations after they are made is often underestimated. While there certainly is valuable data contained in load–depth curves, there is also much information contained in the plastically deformed region around the indentation, be it in localized probes of bulk materials or in thin-film coated systems [16.31]. Specifically, information about the dislocation mechanisms responsible for deformation Nanoindentation: Localized Probes of Mechanical Behavior of Materials can be obtained by examining atomic force microscopy (AFM) images of the slip steps which develop on the surface. 16.3.1 The Spherical Cavity Model Θ a w hf c Fig. 16.3 Definition of variables for indentations. Note that the included angle Θ is only used here to represent the half angle of the indentation. After the indenter tip has been removed from the sample, the remaining angle in the residual cavity will increase due to the elastic recovery of the indentation included angle of the indenter is 2θ, the applied load is P, and the extent of plastic deformation or plastic zone radius is c. For this discussion the plastic zone will be defined in the manner of Samuels and Mulhearn [16.33] given as the radius of the edge of the vertical displacements of the surface. Lockett [16.36] used a model of indentation which relied on slip lines for included indenter angles between 105◦ and 160◦ . These calculations predict an h/a ratio for a 105◦ indenter of 0.185, for a 120◦ indenter of 0.137, and for a 140◦ indenter of 0.085. Dugdale [16.35] measured the pile up as a function of indenter angle normalized by 2c, providing experimental data for steel, aluminum, and copper. When converted to w/a, these values fall between approximately 0.2 and 0.1 for cone angles between 105◦ and 160◦ . The trend for Dugdale’s experimental data is that the ratio c/a decreases as the indenter angle increases, as do the calculations of Lockett. Mulhearn [16.37] proposed that the maximum pile up depended on the indenter angle, and as a first approximation the surface of the sample being indented would remain planar during indentation. Upon removing the load, elastic recovery would occur, and as an upper bound the maximum pile up could be approximated as 1/(6 tan θ). Bower et al.’s model can be modified into a similar ratio, and has a 1/ tan θ dependence, just as with the model presented by Mulhearn. In addition, Bower’s model suggests that the w/a ratio for a given indenter should be constant, independent of the applied load. The extent of plastic deformation around an indentation, often assumed to be linked to the lateral dimensions of the out-of-plane deformation around the indentation, is not the only way in which plastic zones around indentations can be defined. In addition to the upheaval around the indenter, there is a region centered on the indenter tip which experiences radial compression from the indenter tip. The ratio of the plastic zone size to the contact radius, c/a, will quantify the extent of plastic deformation. According to Johnson’s conical spherical cavity model [16.32] of an elastic–plastic material, the plastic zone is determined by ∗ 1 E tan β 2 1 − 2ν 3 c , = + a 6σ y (1 − ν) 3 1 − ν (16.22) where α is the angle between the face of the cone and the indented surface, 90 − θ and σy is the yield strength of an elastic–plastic material. When the ratio E ∗ tan α/σy is greater than 40, Johnson suggests that 397 Part B 16.3 The plastic deformation caused by an indentation has two components, radial and tangential. The spherical cavity model which has been developed extensively by Johnson relies on this mode of deformation [16.32]. However, the spherical cavity model does not address the amount of upheaval around an indentation, and instead focuses on the radial expansion of the plastic zone. It has been shown [16.33] that materials which exhibit perfectly plastic behavior tend to pile up around an indenter, while annealed materials which have high work hardening exponents tend to sink in around an indenter. This behavior has been modeled using finite element methods by Bower et al. [16.34]. Previous work on pile up around indentations has shown the amount of pile up to vary with the indenter angle for a given load, with sharper indenters providing more pile up [16.35]. This discussion will only consider cones between the angles of 105◦ and 137◦ , where the spherical cavity model has been used to successfully determine plastic zones around indentations. Cones sharper than this may begin to approach cutting mechanisms modeled by slip line fields, and may be inappropriate for using a spherical cavity model. Figure 16.3 defines the variables which will be used in this discussion. The maximum pile up is referred to as w, the indentation depth relative to the nominal surface is h final , the contact radius of the indenter is a, the 16.3 Plastic Deformation at Indentations 398 Part B Contact Methods the analysis of elastic–plastic indentation is not valid, and that the rigid–plastic case has been reached. Once the fully plastic state has been reached, the ratio c/a becomes about 2.33. Another model of the plastic zone size [16.38] based upon Johnson’s analysis suggests that the plastic zone boundary c is determined by 3P . (16.23) c= 2πσ y With this model, and as noted by Tabor, the hardness is approximately three times the yield strength of a fully plastic indentation and the ratio c/a should be 2.12 (similar to the suggested values of Johnson). In either case, this provides a mechanistic reason for the conventional rules of indentation which require lateral spacing of at least five times the indenter diameter to ensure no influence of a previous indentation interferes with the hardness measurement. Experimental measurements of pile up around macroscopic indentations [16.39] have shown the pile up models described in this section to be reasonable, and both on the macroscopic and nanoscale [16.40] the plastic zone size approximation for the yield strength is reasonable. Part B 16.3 16.3.2 Analysis of Slip Around Indentations From what is known about the dislocation mechanisms taking place beneath an indenter tip during indentation experiments, it seems a valid assumption that dislocation cross-slip must take place for out-of-plane deformation, or material pile up, on free surfaces around indentations to occur [16.41–49]. Several studies on the structure of residual impressions have been reported in the literature. Kadijk et al. [16.50] has performed indentations in MnZn ferrite single crystals using spherical tips and identified slip systems responsible for the patterns which resulted in the residual depression. Using a combination of controlled etch pitting, chemomechanical polishing, and AFM, Gaillard et al. [16.46] described dislocation structures beneath indentations in MgO single crystals. Other work has been performed on body-centered cubic (BCC) materials to identify changes in surface topography with crystal orientation [16.48] as well as transmission electron microscopy (TEM) images of the subsurface dislocation structure [16.49, 51]. Many of these results have noted that the extent of plastic deformation may be dependent on the overall size of the indentation, as well as exceeding the c/a ratio of 2.33 for nanoscale testing of relatively defect-free materials. The face-centered cubic (FCC) crystal structure contains only four unique slip planes. Chang et al. [16.52] describe a method for determining the surface orientation of a particular grain in an FCC material by measuring the angles of the slip step lines on the surface around indentations and calculating the orientation from the combination of angles. With the current availability of orientation imaging microscopy (OIM), it is possible to carry out a similar process in reverse and use the known orientation of a grain to determine the slip plane responsible for each slip step. Each slip plane can be indexed with respect to a reference direction taken from the OIM. Several studies of surface topography have been performed [16.53, 54] using this technique, and effects such as the include angle of the tip are shown to influence the extent of the plastic zone around materials. Figure 16.4 demonstrates the extent of surface deformation in a stainless-steel alloy indented with a spherical tip with a tip radius of approximately 10 μm, and a 90◦ conical indenter with a tip radius of approximately 1 μm. For the second indentation the extent of plastic deformation surpasses the expected c/a ratio of 2.33. These results are not unique to surface measurements; dislocation arrays identified by etching after indentation in MgO have demonstrated similar effects [16.44]. Clearly the extent of deformation in the lateral direction which is transferred to the surface is influenced by the effective strain of the indentation. Of particular interest in recent studies using dislocation etch pitting in conjunction with nanoinden- 20 μm 20 μm Fig. 16.4 Slip steps from only the positively inclined slip planes around blunt tip indentations in a grain with surface orientation of (1 2 30). When the effective strain is increased by using a sharp tip, steps form from both the positively and negatively inclined slip planes Nanoindentation: Localized Probes of Mechanical Behavior of Materials tation experiments [16.55] is that the lattice friction stress for individual screw and edge dislocation components can be measured; Gaillard and co-workers 16.4 Measurement of Fracture Using Indentation 399 have determined lattice friction stresses of 65 MPa for edge dislocations and 86 MPa for screw dislocations in MgO. 16.4 Measurement of Fracture Using Indentation If nanoindentation is to prove as useful as bulk mechanical tests, properties other than hardness (which is poorly defined in a mechanics sense) and elastic modulus must be considered. Some of these properties, such as yield strength and strain hardening, were covered in the previous section. Two other main areas which provide data to be utilized in assessing a material’s response to an applied load are time-dependent deformation and mechanisms of fracture. In the case of fracture measurements observation of the surface of the material will generally be required. 16.4.1 Fracture Around Vickers Impressions where c is the crack radius, P is the peak indentation load, and χr is an dimensionless constant dependent on the specific indenter–material system that can be found by applying indentations to various maximum loads and measuring crack lengths. If the material system has a constant toughness, then a fitting procedure can be used to determine the dimensionless constant. 15 μm Fig. 16.5 Vickers indentation in silicon, showing radical cracks Part B 16.4 The indentation fracture technique is widely used to characterize the mechanical behavior of brittle materials due to its simplicity and economy in data collection [16.56]. The fracture patterns strongly depend on loading conditions, which fall into two basic categories: blunt or sharp contact. While blunt indenters (for instance, spherical tips) are associated with a Hertzian elastic stress field, sharp indenters (for instance Vickers or Knoop) lead to an elastic–plastic field underneath the contact region [16.57]. In the later case, radial cracks may be generated from the corners of the contact impression, and median cracks propagate parallel to the load axis beneath the plastic deformation zone in the form of circular segments truncated by the material surface. At higher peak loads, lateral cracks may also be formed in the manner of a saucer-shaped surface centered near the impression base [16.57]. As the driving force for lateral crack growth is rather complex compared to the centerloaded symmetric radial/median stress field, for the sake of simplicity, only radial/median crack system is considered in terms of indentation fracture mechanics herein. Lawn and coworkers developed a model to quantitatively represent the relationship of the radial crack size and the fracture toughness for a ceramic material [16.58]. The key idea was to divide the indentation stress field into elastic and residual components where the reversible elastic field enhances median extension during the loading half cycle and the irreversible plastic field primarily provides the driving force for radial crack evolution in the unloading stage. In a later study, they further the discussion by considering the effect of a uniform biaxial applied field on the crack evolution [16.59, 60] based on experimental observation that an imposed uniform stress field could alter the fracture behavior by either increasing or decreasing the crack size [16.61, 62]. In general, the toughness of a material is related to the extent of radial cracks that emanate from the corners of an indentation, such as those shown in Fig. 16.5. An extensive review of the modes of fracture can be found in the literature [16.57]. For the commonly reported fracture tests using Vickers indentation, the stress intensity factor due to the residual impression is 0.5 E P P = χr 1.5 , (16.24) K r = ζr H c1.5 c 400 Part B Contact Methods tests have arrested, the assumption that K r is equal to the fracture toughness is used to provide a measurement of the toughness of the system. Load P (mN) 600 500 16.4.2 Fracture Observations During Instrumented Indentation 400 300 Pop-in 200 hf 100 0 0 1000 2000 3000 4000 5000 Displacement h (nm) Fig. 16.6 Nanoindentation in fused silica using a 42◦ in- denter Of course, if a different material is tested one must determine the constant again and ζr is often used instead, where ζr is given as 0.016 [16.61] or sometimes 0.022 [16.63]. Since the cracks measured during these Part B 16.4 a) Load P (mN) There are drawbacks to using static indentation for fracture. First, many materials which are brittle do not exhibit fracture during nanoindentation using conventional indentation probes. Secondly, imaging the indentation can become problematic if very small features or indentations are made in the sample. Therefore, it would be convenient to determine the occurrence of fracture during an instrumented indentation. Some authors have chosen to identify fracture mechanisms using acoustic emission sensors during nanoindentation [16.64–66], but this does require specialized equipment and analysis to separate fracture from plasticity events. Morris et al. [16.67] have utilized sharper indenter probes to determine the effects of included angle on the resulting load–displacement curves, as shown in the typical result in Fig. 16.6. More sensitive measurements of these pop-in events are found if one measures the stiffness continually during the ex- b) Load P (mN) 600 600 500 500 52° 400 400 65° 300 65° 52° 42° 35° 300 42° 200 200 35° 100 0 c) 100 0 100 200 300 400 500 0 Displacement h (nm) 0 100 200 300 400 500 Displacement h (nm) d) Fig. 16.7a–d Unloading curves and 10 μm respective images of residual indentation impressions showing cracking an systems ((a) and (b)), while uniform unloading curves occur in materials which undergo only plastic deformation ((c) and (d)) Nanoindentation: Localized Probes of Mechanical Behavior of Materials 16.4 Measurement of Fracture Using Indentation 401 Load (μN) 3000 Lower than the excursion load Higher than the excursion load fim thickness: 92 nm 2500 2000 1500 Cracks in oxide film 1000 1000 nm 500 Fig. 16.8 AFM image of circumferential cracks around indentations into anodized titanium. The large ring corresponds to a through-thickness crack, while the smaller impression in the center is the contact area defined by the indenter tip [16.68] 0 20 40 60 80 100 120 140 Depth (nm) Fig. 16.9 Indentation into a 92 nm-thick TiO2 film grown anodically on a Ti substrate. The distinct excursion in depth, in this case corresponds to a through-thickness fracture in the film [16.68] to produce cracks in the film [16.75–80]. Malzbender and de With [16.75] demonstrated that the dissipated energy was related to the fracture toughness of the coating and interface by performing simple integrations of the loading and unloading portions from indentation to determine the amount of energy needed to damage the film. Other studies have suggested that the fracture energy could be estimated as the energy consumed during the first circumferential crack during the load drop or plateau on the load–displacement curve [16.74, 76, 78, 81]. A recent study has demonstrated one technique which relies upon examination of the load–depth curve, knowing the general behavior of hard film–soft substrate systems [16.68, 82]. In this case, the load–depth curve of a coated system, shown in Fig. 16.9 for a 90 nm TiO2 on Ti, also demonstrates a pop-in or excursion in the load–depth curve. These pop-in events, present only because the indentation testing system is a load-controlled device, are in this case indicative of film fracture because permanent deformation is observed prior to the pop-in. The plastic zone c from Sect. 16.3 is used to estimate the radius of the fracture, and the toughness of the film can be estimated from the energy difference between that which goes into the indentation of the film–substrate system and indentations to similar depths for the substrate only. Part B 16.4 periment. While it may be tempting to identify the presence of a pop-in to a fracture event, these authors demonstrate conditions under which no pop-in event occurs and yet the resulting indentation impression instead shows clear evidence of cracking. Probes of increasing acuity (moving from the conventional Berkovich geometry to sharper cube-corner-type systems) are more likely to generate fracture. A comparison of the unloading slopes of probes of various acutities demonstrated that samples which exhibit indentationinduced fracture demonstrate differences in unloading (while having the same final displacement), while materials which only exhibit plastic deformation do not show differences in unloading curves, as shown in Fig. 16.7. Therefore, the authors suggest utilizing a shallow (Berkovich) and acute (cube corner or similar) indenter pair, and that comparing the ratio of h max /h f for the different tips will enable the user to identify materials which are susceptible to indentation-induced fracture. Another case for fracture measurements particularly suited to continuous indentation is the application of hard films on compliant or soft substrates, which results in circumferential fracture around the indentation, as shown in Fig. 16.8. Thin films have been shown to fracture during indentation processes at critical loads and depths [16.69]. Some studies have analyzed these fracture events by calculating an applied radial stress at fracture [16.68, 70, 71], an applied stress intensity at fracture [16.71–74], and the amount of energy required 0 402 Part B Contact Methods 16.5 Probing Small Volumes to Determine Fundamental Deformation Mechanisms Part B 16.5 It has been well documented that small volumes of materials regularly exhibit mechanical strengths greatly in excess of more macroscopic volumes. Observations in whiskers of metals were among the first studies to demonstrate the ability of metals to sustain stresses approaching the theoretical shear strength of the material. The often described indentation size effect in metals [16.83–86] has been the subject of many studies with nanoindentation due to the ability to neglect elastic recovery (i. e., the indentation hardness is measured under load) and to measure very small sizes. In ceramics this effect has been explained by variations in elastic recovery during the indentation [16.87]. Results on nanocrystalline metals have continued that trend, until the literature has reached a point where models suggest that grain sizes too small to support the nucleation and growth of stable dislocation loops will likely deform via other mechanisms. Similarly, nanostructured metallic laminates have been demonstrated to exhibit particularly unique deformation behavior. Other recent experiments in which a flat punch in a nanoindenter has been used to carry out compression tests on machined structures fabricated via focus ion beam machining have demonstrated that just having a smaller volume of material, with the concomitant increase in surface area and decrease in sample size, may indeed impact plasticity in ways not immediately obvious through scaling arguments [16.88, 89]. Recent work has also demonstrated the ability to quantify the stress required to cause the onset of plasticity dislocations in relatively dislocation-free solids [16.90–94]. This method is only possible using small-scale mechanical testing; large-volume mechanical testing will generally measure the motion of pre-existing dislocations under applied stresses. Nanoindentation couples well with small-scale mechanical modeling, as it approaches volumes which can be simulated using molecular dynamics [16.95, 96] and the embedded atom technique [16.97]. In materials with low dislocation densities, the sudden onset of plasticity occurs at stresses approaching the theoretical strength of a material. Recent studies have proposed two primary models of these effects: homogenous nucleation and thermally activated processes. Gane and Bowden were the first to observe the excursion phenomena on an electropolished surface of gold [16.98]. A fine tip was pressed on the gold surface while observing the deformation in a scan- ning electron microscope. No permanent penetration was observed initially, but at some point during the process a sudden jump in displacement occurred, which corresponded to the onset of permanent deformation. Similar observations of nonuniform loading during indentation were made by Pethica and Tabor [16.99], who applied a load to bring a clean tungsten tip into contact with a nickel (111) surface in an ultrahighvacuum environment while monitoring the electrical resistance between the tip and surface. The nickel had been annealed, sputtered to remove contaminants, and then annealed again to remove surface damage from the sputtering process. When an oxide was grown on the nickel to thicknesses of greater than 50 Å, loading was largely elastic while the electrical resistance remained extremely high, with only minimal evidence of plastic deformation. However, past some critical load the resistance between the tip and sample dropped dramatically, with continued loading suggestive of largely plastic deformation. With the advent of instrumented indentation testing it was possible to monitor the penetration of an indenter tip into a sample. The pop-in or excursion in depth during indentation was observed by many researchers during this period, and is represented in Fig. 16.10. Load (mN) 6 5 Elastic loading prediction 4 3 2 1 0 0 50 100 150 200 Depth (nm) Fig. 16.10 Nanoindentation into electropolished tungsten, with an initial elastic loading followed by plastic deformation. Elastic loading prediction by [16.26] Nanoindentation: Localized Probes of Mechanical Behavior of Materials Some researchers have utilized experimental methods that are either depth controlled [16.100,101] or a system in which the overall displacement (and not penetration) was held constant [16.91, 102]. Post-indentation microscopy demonstrated that, after these discontinuous events during loading in single-crystal ceramics, dislocation structures were present [16.49, 55]. Using the Hertzian elastic model, 4 (16.25) P = E ∗ Rδ3 , 3 the maximum shear stress under the indenter tip τmax is related to the mean pressure pm and is given by 1 6(E ∗ )2 3 1/3 P (16.26) τmax = 0.31 π 3 R2 at a position underneath the indenter tip at a depth of 0.48a, where a is the contact area, and is determined by solving ∗2 1/3 6E 3P = P 1/3 . (16.27) 2πa2 π 3 R2 Recently, Schuh and co-workers [16.106] carried out indentation experiments that probed the onset of plastic deformation at elevated temperatures to determine an effective activation energy and volume for the initiation of plasticity. Using a method describing the probability of an excursion occurring underneath the indenter tip they use an Arrhenius model for dislocation nucleation. They suggest that the activation energy is similar to that of vacancy motion, and that the activation volume is on the order of a cubic Burgers vector. This suggests that point defects have a direct relationship with dislocation nucleation. Another recent study of incipient plasticity in Ni3 Al [16.107] suggests that the growth of a Frank dislocation source at subcritical stresses controls plasticity at lower stresses during indentation, and that self-diffusion along the indenter– sample surface dominates the onset of plasticity in these cases of nanoindentation. Bahr and Vasquez [16.108] have shown that, for solid solutions of Ni and Cu, the shear stress under the indenter tip at the point of dislocation nucleation can be correlated to changes in the elastic modulus of the material. Using a diffusion couple of Ni and Cu to create a region of material with a range of compositions ranging from pure Ni to pure Cu, it was shown that the changes in shear stress required to nucleate dislocations are on the order of the changes in elastic modulus, between 1/30 and 1/20 of the shear modulus of this alloy. The implication of these data is that overall dislocation nucleation during nanoindentation is strongly related to shear modulus in this system, but does not preclude the diffusion mechanism suggested by Ngan and Po [16.107]. Another example of these methods is nanoindentation being used to probe the effects of hydrogen on dislocation nucleation and motion in a stainless steel [16.109]. These pop-in events were used to demonstrate that, while dissolved hydrogen lowers the indentation load at which dislocations are nucleated, this was likely a result of hydrogen decreasing the shear modulus and not directly related to hydrogen creating or modifying nucleation sites. The pop-ins occurred more slowly in the presence of hydrogen, most likely due to hydrogen inhibiting the mobility of fast-moving dislocations as they are rapidly emitted from a source. Finally, a second excursion occurs after charging with hydrogen, but is uncommon in uncharged material. When coupled to analysis of slip steps around the indentations that show the existence of increased slip planarity, this suggests that dislocations emitted during the first excursion are inhibited in their ability to cross-slip and there- 403 Part B 16.5 The shear stress field does not drop off that rapidly, and there is a region under the indenter tip extending to approximately 2a which has an applied shear stress of approximately 0.25 pm . Many research groups have noted that this stress approaches the theoretical shear stress in a crystal. The similarity between the applied shear stress and shear modulus is close to the classical models of homogeneous dislocation nucleation espoused by the early theoretical work on dislocation mechanisms. For instance, Cotrell noted that the classical theoretical shear strength of a material is often approximated by μ/30 to μ/100 [16.103], where μ is the shear modulus. A new set of experiments utilizing in situ nanoindentation in a transmission electron microscope has been used to support this model of homogeneous dislocation nucleation. Minor et al. [16.92] observed dislocations to pop out underneath an indenter tip after some initial elastic loading during contact between the tip and aluminum sample. This direct observation is strong evidence that dislocations nucleate underneath the indenter tip. However, the most recent studies from in situ microscopy have shown that even the very initial contact can generate dislocations that nucleate and propagate to grain boundaries, after which an elastic loading behavior is again demonstrated followed by a second strain burst [16.104]. Multiple yield points during nanoindentation are often referred to as staircase yielding [16.93, 101, 105]. Probing Small Volumes 404 Part B Contact Methods fore cannot accommodate a fully evolved plastic zone. A second nucleation event is then needed on a differ- ent slip system to accommodate the plasticity from the indenter. 16.6 Summary This chapter has provided a summary of the basic methodologies for assessing the mechanical properties of materials on a small scale via instrumented indentation, often more commonly referred to as nanoindentation. The first two sections described the history of indentation testing and the concept of hardness, beginning with the quantification carried out by Brinnel. This aimed to provide a background for the reader, who may feel at times that many of the conventions in nanoindentation appear to be based on somewhat arbitrary rules. Moving to instrumented techniques, the development of contact mechanics approaches for determining elastic and plastic properties of solids via small-scale indentations was outlined, including the analysis of the unloading slope and the imposed oscillation techniques currently available on commercial instruments. The later sections of the chapter focused on more advanced uses of small-scale indentations. The deformation surrounding an indentation provides a researcher with more information than just the elastic properties; sophisticated analysis methods allow for determination of more complex mechanical properties such as strain hardening. Indentation-induced fracture, while still a subject of debate in the literature, is often used in brittle materials, so this topic is covered herein to provide the reader with a current status of the methods. Finally, the last section covers more fundamental dislocation behavior (such as nucleation or multiplication issues) that is often examined by very low-load indentations. Part B 16 References 16.1 16.2 16.3 16.4 16.5 16.6 16.7 16.8 16.9 16.10 16.11 H. Hertz: Miscellaneous Papers. English translation by D.E. Jones, G.E. Schott (MacMillan, New York, 1896) pp. 163-183 F. Auerbach: Smithsonian Report for 1891 (Government Printing Office, Washington 1893) pp. 207–236 J.A. Brinnnel: II.Cong. Int. Methodes d’Essai, J. Iron Steel Inst. 59, 243 (1901), translated to English by A. Wahlberg E. Meyer: Untersuchungen über Härteprüfung und Härte, Z. Ver. Dtsch. Ing. 52, 645–654 (1908) S.L. Hoyt: The ball indentation hardness test, Trans. Am. Soc. Steel Treating, 6, 396 (1924) D. Tabor: The Hardness of Metals (Oxford Univ. Press, Oxford 1951) V.E. Lysaght: Indentation Hardness Testing (Reinhold Pub., New York 1949) R.L. Smith, G.E. Sandland: Some notes on the use of a diamond pyramid for hardness testing, J. Iron. Steel Inst. 3, 285–301 (1925) E.S. Berkovich: Three faceted diamond pyramid for micro hardness testing, Industrial Diamond Rev. 11, 129–133 (1951) J. Thurn, R.F. Cook: Simplified area function for sharp indenter tips in depth-sensing indentation, J. Mater. Res. 17, 1143–1146 (2002) M.R. Vanlandingham, T.F. Juliano, M.J. Hagon: Measuring tip shape for instrumented indentation 16.12 16.13 16.14 16.15 16.16 16.17 16.18 16.19 using atomic force microscopy, Measurement Sci. Tech. 16, 2173–2185 (2005) J.L. Loubet, J.M. Georges, O. Marchesini, G. Meille: Vickers indentation curves of magnesium oxide (MgO), J. Tribology, 106, 43–48 (1984) D. Neweyt, M.A. Wilkins, H.M. Pollock: An ultralow-load penetration hardness tester, J. Phys. E: Sci. Instrum., 15, 119–122 (1982) A.C. Fischer-Cripps: Nanoindentation, 2nd edn. (Springer, Berlin Heidelberg 2004) B. Bhushan (Ed.): Springer Handbook of Nanotechnology, 2nd edn. (Springer, Berlin Heidelberg 2004) X. Li, B. Bhushan: A review of nanoindentation continuous stiffness measurement technique and its applications, Mater.Char. 48, 11–36 (2002) W.C. Oliver, G.M. Pharr: Review: Measurement of hardness and elastic modulus by instrumented indentation: Advances in understanding and refinements to methodology, J. Mater. Res. 19, 3–20 (2004) M.R. VanLandingham: Review of instrumented indentation, J. Res. Natl. Inst. Stand. Technol. 108, 249–265 (2003) M.F. Doerner, W.D. Nix: A method for interpreting the data from depth-sensing indentation instruments, J. Mater. Res. 1, 601–609 (1986) Nanoindentation: Localized Probes of Mechanical Behavior of Materials 16.20 16.21 16.22 16.23 16.24 16.25 16.26 16.27 16.28 16.30 16.31 16.32 16.33 16.34 16.35 16.36 16.37 16.38 16.39 16.40 16.41 16.42 16.43 16.44 16.45 16.46 16.47 16.48 16.49 16.50 16.51 S. Harvey, H. Huang, S. Vennkataraman, W.W. Gerberich: Microscopy and microindentation mechanics of single crystal Fe-3 wt.% Si: Part I. Atomic force microscopy of a small indentation, J. Mater. Res. 8, 1291–1299 (1993) D.F. Bahr, W.W. Gerberich: Pile up and plastic zone size around indentations, Metall. Mater. Trans. A 27A, 3793–3800 (1996) D. Kramer, H. Huang, M. Kriese, J. Robach, J. Nelson, A. Wright, D. Bahr, W.W. Gerberich: Yield strength predictions from the plastic zone around nanocontacts, Acta Mater. 47, 333–343 (1998) Y. Gaillard, C. Tromas, J. Woirgard: Pop-in phenomenon in MgO and LiF: observation of dislocation structures, Phil. Mag. Leters 83, 553–561 (2003) E. Carrasco, M.A. Gonzalez, O. Rodriguez de la Fuente, J.M. Rojo: Analysis at atomic level of dislocation emission and motion around nanoindentations in gold, Surf. Sci. 572, 467–475 (2004) M. Oden, H. Ljungcrantz, L. Hultman: Characterization of the induced plastic zone in a single crystal TiN(001) film by nanoindentation and transmission electron microscopy, J. Mater. Res. 12, 2134–2142 (1997) C. Tromas, J. C. Girard, V. Audurier, J. Woirgard, Study of the low stress plasticity in single-crystal MgO by nanoindentation and atomic force microscopy, J. Mater. Sci. 34, 5337–5342 (1999) W. Zielinski, H. Huang, W.W. Gerberich: Microscopy and microindentation mechanics of single crystal Fe-3 wt.%Si: Part II. TEM of the indentation plastic zone, J. Mater. Res. 8, 1300–1310 (1993) Y. Gaillard, C. Tromas, J. Woirgard: Study of the dislocation structure involved in a nanoindentation test by atomic force microscopy and controlled chemical etching, Acta Mater. 51, 1059– 1065 (2003) P. Peralta, R. Ledoux, R. Dickerson, M. Hakik, P. Dickerson: Characterization of surface deformation around Vickers indents in monocrystalline materials, Metall. Mater. Trans. A 35A, 2247–2255 (2004) N.A. Stelmashenko, M.G. Walls, L.M. Brown, Y.U.V. Milman: Microindentations on W and Mo oriented single crystals: An STM study, Acta Metal. Mater. 41, 2855–2865 (1993) T.F. Page, W.C. Oliver, C.J. McHargue: Deformation behavior of ceramic crystals subjected to very low load (nano)indentations, J. Mater. Res. 7, 450–473 (1992) S.E. Kadijk, A. Broese van Groenou: Cross-slip patterns by sphere indentations on single crystal MnZn ferrite, Acta Metall. 37, 2625–2634 (1989) W. Zielinski, H. Huang, S. Venkataraman, W.W. Gerberich: Dislocation distribution under a microindentation into an iron-silicon single crystal, Phil. Mag. A 72, 1221–1237 (1995) 405 Part B 16 16.29 W.C. Oliver, G.M. Pharr: An improved technique for determining hardness and elastic modulus using load and displacement sensing indentation experiments, J. Mater. Res. 7, 1564–1583 (1992) I.N. Sneddon: The relation between load and penetration in the axisymmetric Boussinesq problem for a punch of Arbitrary Profile, Int. J. Engng. Sci. 3, 47–56 (1965) K.L. Johnson: Contact Mechanics (Cambridge Univ. Press, Cambrige 1985) R.B. King: Elastic analysis of some punch problems for a layered medium, Int. J. Solids Structures 23, 1657–1664 (1987) A.C. Fischer-Cripps: Critical review of analysis and interpretation of nanoindentation test data, Surf. Coatings Tech. 200, 4153–4165 (2006) J.H. Strader, S. Shim, H. Bei, W.C. Oliver, G.M. Pharr: An experimental evaluation of the constant β relating the contact stiffness to the contact area in nanoindentation, Phil. Mag. 86, 5285–5298 (2006) I.N. Sneddon: Boussinesq’s problem for a rigid cone, Proc. Cambridge Phil. Soc. 44, 492–507 (1948) S.V. Hainsworth, H.W. Chandler, T.F. Page: Analysis of nanoindentation load-displacement loading curves, J. Materials Research 11, 1987–1995 (1996) J.L. Loubet, J.M. Georges, J. Meille: Microindentation Techniques in Materials Science and Engineering (American Society for Testing and Materials, Philadelphia 1986), ed. by P. J. Blau, B. R. Lawn, Vol. 72 J.S. Field, M.V. Swain: A simple predictive model for spherical indentation, J. Mater. Res. 8, 297–306 (1993) D.L. Joslin, W.C. Oliver: A new method for analyzing data from continuous depth-sensing microindentation tests, J. Mater. Res. 5, 123–126 (1990) T.F. Page, S.V. Hainsworth: Using nanoindentation techniques for the characterization of coated systems: a critique, Surf. Coating Technol. 61, 201–208 (1993) K.L. Johnson: The correlation of indentation experiments, J. Mech. Phys. Solids 18, 115–126 (1970) L.E. Samuels, T.O. Mulhearn: An experimental investigation of the deformed zone associated with indentation hardness impressions, J. Mech. Phys. Solids 5, 125–134 (1957) A.F. Bower, N.A. Fleck, A. Needleman, N. Ogbonna: Indentation of a Power Law Creeping Solid, Proc. Royal. Soc. 441A, 97–124 (1993) D.S. Dugdale: Cone indentation experiments, J. Mech. Phys. Solids, 2, 265–277 (1954) F.J. Lockett: Indentation of a rigid/plastic material by a conical indenter, J. Mech. Phys. Solids 11, 345– 355 (1963) T.O. Mulhearn: The Deformation of Metals by Vickers-Type Pyramidal Indenters, J. Mech. Phys. Solids 7, 85–96 (1959) References 406 Part B Contact Methods 16.52 16.53 16.54 16.55 16.56 16.57 16.58 16.59 16.60 Part B 16 16.61 16.62 16.63 16.64 16.65 16.66 16.67 S.C. Chang, H.C. Chen: The determination of F.C.C. crystal orientation by indentation, Acta Metall Mater. 43, 2501–2505 (1995) K.A. Nibur, D.F. Bahr: Identifying slip systems around indentations in FCC metals, Scripta Materialia 49, 1055–1060 (2003) K.A. Nibur, F. Akasheh, D.F. Bahr: Analysis of dislocation mechanisms around indentations through slip step observations, J. Mater. Sci. 42, 889–900 (2007) Y. Gaillard, C. Tromas, J. Woirgard: Quantitative analysis of dislocation pile-ups nucleated during nanoindentation in MgO, Acta Mater. 54, 1409–1417 (2006) B.R. Lawn, T.R. Wilshaw: Fracture of Brittle Solids (Cambridge Univ. Press, Cambridge 1975) R.F. Cook, G.M. Pharr: Direct observation and analysis of indentation cracking in glasses and ceramics, J. Am. Cer. Soc. 73, 787–817 (1990) B.R. Lawn, A.G. Evans, D.B. Marshall: Elastic/Plastic Indentation Damage In Ceramics: The Median/Radial Crack System, J. Am. Cer. Soc. 63, 574–581 (1980) D.B. Marshall, B.R. Lawn: J. Mater. Sci., Residual stress effects in sharp contact cracking Part 1, Indentation fracture mechanics 14, 2001–2012 (1979) D.B. Marshall, B.R. Lawn, P. Chantikul: Residual stress effects in sharp contact cracking, Part 2. Strength degradation, J. Mater. Sci. 14, 2225–2235 (1979) G.R. Anstis, P. Chantikul, B.R. Lawn, D.B. Marshall: A critical evaluation of indentation techniques for measuring fracture toughness I, Direct crack measurements, J. Am. Ceram. Soc. 64, 533–542 (1981) P. Chantikul, G.R. Anstis, B.R. Lawn, D.B. Marshall: A critical evaluation of indentation techniques for measuring fracture toughness II, Strength method, J. Am. Ceram. Soc. 64, 539–543 (1981) R.F. Cook: Strength Characterization of Ceramics using Controlled Indentation Flaws. Ph.D. Thesis (Univ. New South Wales, Australia 1986) W.W. Gerberich, W.M. Mook, M.J. Cordill, J.M. Jungk, B. Boyce, T. Friedmann, N.R. Moody, D. Yang: Nanoprobing fracture length scales, Int. J. Fract. 138, 75–100 (2006) D.F. Bahr, J.W. Hoehn, N.R. Moody, W.W. Gerberich: Adhesion and acoustic emission analysis of failures in nitride films with a metal interlayer, Acta Mater. 45, 5163–5175 (1997) A. Daugela, H. Kutomi, T.J. Wyrobeck: Nanoindentation induced acoustic emission monitoring of native oxide fracture and phase transformations, Z. Metallkunde 92, 1052–1056 (2001) D.J. Morris, S.B. Myers, R.F. Cook: Sharp probes of varying acuity: Instrumented indentation and fracture behavior, J. Mater. Res. 19, 165–175 (2004) 16.68 16.69 16.70 16.71 16.72 16.73 16.74 16.75 16.76 16.77 16.78 16.79 16.80 16.81 16.82 16.83 M. Pang, D.F. Bahr: Thin film fracture during nanoindentation of a titanium oxide film – titanium system, J. Mater. Res. 16, 2634–2643 (2001) N.G. Chechenin, J. Bottiger, J.P. Krog: Nanoindentation of amorphous aluminum oxide films II. Critical parameters for the breakthrough and a membrane effect in thin hard films on soft substrates, Thin Solid Films 261, 228–235 (1995) D.F. Bahr, C.L. Woodcock, M. Pang, K.D. Weaver, N.R. Moody: Indentation induced film fracture in hard film – soft substrate systems, Inter. J. Frac. 119/120, 339–349 (2003) E. Wepplemann, M.V. Swain: Investigation of the stresses and stress intensity factors responsible for fracture of thin protective films during ultra-micro indentation tests with spherical indenters, Thin Solid Films 286, 111–121 (1996) J. Thurn, R.F. Cook: Mechanical and thermal properties of physical vapour deposited alumina films Part II Elastic, plastic, fracture, and adhesive behaviour, J. Mater. Sci. 39, 4809–4819 (2004) C. Herzl: Fracture mechanics analysis of thin coatings under plane-strain indentation, Int. J. Sol. Struct. 40, 591–610 (2003) J. Malzbender, G. de With: Elastic modulus, indentation pressure and fracture toughness of hybrid coatings on glass, Thin Solid Films 366, 139–149 (2000) J. Malzbender, G. de With: Energy dissipation, fracture toughness and the indentation load– displacement curve of coated materials, Surf. Coat. Tech. 135, 60–68 (2000) P.G.T. van der Varst, G. de With: Energy based approach to the failure of brittle coatings on metallic substrates, Thin Solid Films 384, 85–91 (2001) A. Abdul-Baqi, E. Van der Giessen: Numerical analysis of indentation-induced cracking of brittle coatings on ductile substrates, Int. J. of Sol. Struct. 39, 1427–1442 (2002) X. Li, D. Diao, B. Bhushan: Fracture mechanisms of thin amorphous carbon films in nanoindentation, Acta Mater. 45, 4453–4461 (1997) K. Sriram, R. Narasimhan, S. Biswas: A numerical fracture analysis of indentation into thin hard films on soft substrates, Eng. Fract. Mech. 70, 1323–1338 (2003) Z. Xia, W. Curtin, B. Sheldon: A new method to evaluate the fracture toughness of thin films, Acta. Mater. 52, 3507–3517 (2004) X. Li, B. Bhushan: Measurement of fracture toughness of ultra-thin amorphous carbon films, Thin Solid Films 315, 214–221 (1998) K.R. Morasch, D.F. Bahr: An energy method to analyze through thickness thin film fracture during indentation, Thin Solid Films 515, 3298–3304 (2007) N.A. Fleck, G.M. Muller, M.F. Ashby, J.W. Hutchinson: Strain gradient plasticity: Theory and experiment, Acta Metall. 42, 475–487 (1994) Nanoindentation: Localized Probes of Mechanical Behavior of Materials 16.84 16.85 16.86 16.87 16.88 16.89 16.90 16.91 16.92 16.94 16.95 16.96 16.97 16.98 16.99 16.100 16.101 16.102 16.103 16.104 16.105 16.106 16.107 16.108 16.109 W.W. Gerberich, N.I. Tymiak, J.C. Grunlan, M.F. Horstemeyer, M.I. Baskes: Interpretations of indentation size effects, J. Appl. Mech. 69, 433–442 (2002) N. Gane, F.P. Bowden: Microdeformation of solids, J. Appl. Phys. 39, 1432–1435 (1968) J.B. Pethica, D. Tabor: Contact of characterized metal surfaces at very low loads: Deformation and adhesion, Surf. Sci. 89, 182–190 (1979) T.A. Michalske, J.E. Houston: Dislocation nucleation at nano-scale mechanical contacts, Acta Mater. 46, 391–396 (1998) J.D. Kiely, J.E. Houston: Nanomechanical properties of Au(111), (001), and (110) surfaces, Phys. Rev. B 57, 12588–12594 (1998) D.F. Bahr, D.E. Wilson, D.A. Crowson: Energy considerations regarding yield points during indentation, J. Mater. Res. 14, 2269–2275 (1999) A.H. Cotrell: Dislocations and Plastic Flow in Crystals (Oxford Press, Oxford 1953) pp. 11–12 A.M. Minor, S.A. Syed Asif, Z. Shan, E.A. Stach, E. Cyrankowski, T.J. Wyrobek, O.L. Warren: A new view of the onset of plasticity during the nanoindentation of aluminium, Nature Mater. 5, 697–702 (2006) S.G. Corcoran, R.J. Colton, E.T. Lilleodden, W.W. Gerberich: Anomalous plastic deformation at surfaces: Nanoindentation of gold single crystals, Phys. Rev. B 55, R16057–16060 (1997) C.A. Schuh, J.K. Mason, A.C. Lund: Quantitative insight into dislocation nucleation from high-temperature nanoindentation experiments, Nature Mater. 4, 617–621 (2005) A.H.W. Ngan, P.C. Wo: Delayed plasticity in nanoindentation of annealed crystals, Phil. Mag. 86, 1287–1304 (2006) D.F. Bahr, G. Vasquez: The effects of solid solution impurities on dislocation nucleation during nanoindentation, J. Mater. Res. 20, 1947–1951 (2005) K.A. Nibur, D.F. Bahr, B.P. Somerday: Hydrogen effects on dislocation activity in austenitic stainless steel, Acta Materialia 54, 2677–2684 (2006) 407 Part B 16 16.93 Q. Ma, D.R. Clarke: Size dependent hardness of silver single crystals, J. Mater. Res. 10, 853–863 (1995) A.A. Elmustafa, J.A. Eastman, M.N. Rittner, J.R. Weertman, D.S. Stone: Indentation size effect: large grained aluminum versus nanocrystalline aluminum-zirconium alloys, Scripta Mater. 43, 951–955 (2000) W.D. Nix, H. Gao: Indentation size effects in crystalline materials: A law for strain gradient plasticity, J. Mech. Phys. Solids 46, 411–425 (1998) S.J. Bull, T.F. Page, E.H. Yoffe: An explanation of the indentation size effect in ceramics, Phil. Mag. Lett. 59, 281–288 (1989) M.D. Uchic, D.M. Dimiduk, J.N. Florando, W.D. Nix: Sample dimensions influence strength and crystal plasticity, Science 305, 986–989 (2004) C.A. Volkert, E.T. Lilleodden: Size effects in the deformation of sub-micron Au columns, Phil. Mag. 86, 5567–5579 (2006) C.A. Schuh, A.C. Lund: Application of nucleation theory to the rate dependence of incipient plasticity during nanoindentation, J. Mater. Res. 19, 2152–2158 (2004) S.A. Syed-Asif, J.B. Pethica: Nanoindentation creep of single-crystal tungsten and gallium arsenide, Philos. Mag. A 76, 1105–1118 (1997) A.M. Minor, E.T. Lilleodden, E.A. Stach, J.W. Morris: Direct observations of incipient plasticity during nanoindentation of Al, J. Mater. Res. 19, 176–182 (2004) D.F. Bahr, D.E. Kramer, W.W. Gerberich: Non-linear deformation mechanisms during nanoindentation, Acta Mater. 46, 3605–3617 (1998) A.B. Mann, J.B. Pethica: The role of atomic size asperities in the mechanical deformation of nanocontacts, Appl. Phys. Lett. 69, 907–909 (1996) Y. Choi, K.J. V.Vliet, J. Li, S. Suresh: Size effects on the onset of plastic deformation during nanoindentation of thin films and patterned lines, J. Appl. Phys. 94, 6050–6058 (2003) C.L. Kelchner, S.J. Plimpton, J.C. Hamilton: Dislocation nucleation and defect structure during surface indentation, Phys. Rev. B 58, 11085–11088 (1998) References
© Copyright 2025 Paperzz