Sei sulla pagina 1di 10

Indentation Power-Law Creep of High-Purity Indium

B.N. LUCAS and W.C. OLIVER Using a variety of depth-sensing indentation techniques, the creep response of high-purity indium, from room temperature to 75 C, was measured. The dependence of the hardness on the variables of indentation strain rate (stress exponent for creep (n)) and temperature (apparent activation energy for creep (Q)) and the existence of a steady-state behavior in an indentation test with a Berkovich indenter were investigated. It was shown for the rst time that the indentation strain rate (h/h) could be held constant during an experiment using a Berkovich indenter, by maintaining the loading rate divided by the load (P/P) constant. The apparent activation energy for indentation creep was found to be 78 kJ/mol, in accord with the activation energy for self-diffusion in the material. Finally, by performing P/P change experiments, it was shown that a steady-state path independent of hardness could be reached in an indentation test with a geometrically similar indenter.

I.

INTRODUCTION

A. Effect of Stress (Stress Exponent for Creep) There are four types of tests that have been employed using depth-sensing indentation systems to gain insight into the relationship between indentation strain rate and hardness: indentation load relaxation (ILR) tests,[2,3] constant rate of loading (CRL) tests,[4,5] constant-load indentation creep tests,[611] and impression creep tests.[12,13] All of these tests have drawn analogies between hardness and ow strength, as well as between the uniaxial strain rate and the indentation strain rate, i.e., H and
u

THE techniques for the measurement of time-dependent phenomena can be conveniently divided into two classes: (1) broad-band, quasistatic, or creep techniques, where the load, stress, or strain rate is held constant for a period of time while the response of the material is measured; and (2) frequency-specic dynamic techniques, where the load or stress is varied at a single frequency and the response of the material is measured. Although standard bulk-testing techniques exist for these types of measurements, in many of todays technologies, the volume of material of interest may be on such a scale that these techniques become impractical. While, in some instances, specimens may be prepared in a form that allows testing with modied uniaxial techniques, these preparation processes are often tedious or may very well alter those properties that are of interest. It then becomes necessary to nd an alternate means for mechanical characterization. In this work, a variation of the broad-band, quasistatic technique described earlier will be utilized to measure time-dependent plasticity or indentation creep at several temperatures. The goal of this study of indentation creep is to explore the dependency of the indentation hardness of high-purity indium on the variables of strain rate and temperature, as well as to investigate the existence of a steady-state behavior in indentation creep using a Berkovich indenter.
II. REVIEW OF DEPTH-SENSING INDENTATION CREEP LITERATURE

C1

[1]

C2

[2]

Since the ow stress of a material is a function of strain rate and temperature, the hardness of a material should be expected to vary in an analogous way.[1] This fact has led to numerous studies attempting to correlate hardness with the variables of time (or indentation strain rate) and temperature. A brief survey of this body of work follows.

B.N. LUCAS, Director of Analytical Services, and W.C. OLIVER, Vice President, are with the Nano Instruments Innovation Center, MTS Systems Corporation, Oak Ridge, TN 37830. Manuscript submitted April 27, 1998.

In an indentation test, the dynamics of deformation are very different than those occurring in the previously described uniaxial creep test. The deformed volume of material under the indenter is continually expanding to encompass previously undeformed material. As the material strains under the indenter, the material underneath the indenter is very often likened to an expanding cavity with a hydrostatic core, where no deformation is occurring, and an expanding elastic/plastic boundary. The creep process is believed to depend upon the rate at which the elastic/plastic boundary can proceed into the material. As the radius of the elastic/plastic boundary is often related to the radius of the indentation, the instantaneous change in contact area divided by the instantaneous contact area (A /A) may be the most appropriate denition for the indentation strain rate, as it is a direct measure of the progression of the elastic/plastic boundary into the material. However, for a geometrically similar indenter (h/h), the instantaneous displacement rate of the indenter divided by the instantaneous displacement is simply related to A/A, and the indentation strain rate has typically been dened as h/h. The ILR technique was rst presented by Hannula et al.[2] In this type of experiment, an indenter is pushed into a sample surface with the aid of a crosshead. When the desired load or displacement is achieved, the position of the indenter is xed. Load relaxation occurs by conversion of the elastic strain in the machine and the specimen into plas-

METALLURGICAL AND MATERIALS TRANSACTIONS A

VOLUME 30A, MARCH 1999601

tic strain in the specimen. This technique has been used[3] to study time-dependent deformation in small volumes, using a Vickers indenter with a 1 m2 at area on the apex. The CRL test is a multi-indentation test that was developed for determining stress exponents using depth-sensing indentation systems.[4,5] This type of test utilized the ability of the NANO INDENTER II* to incrementally apply the
*NANO INDENTER II is a trademark of MTS Systems Corp., Eden Prairie, MN.

load on the indenter at rates specied by the experimenter. In a CRL test, the indenter is loaded at a constant loading rate until the indenter has reached a prescribed depth in the material. A complete series of experiments would involve this procedure utilizing a different loading rate for each indent made and calculating a hardness for that loading rate from the applied load and the achieved depth. The stress exponent for creep (n) is then dened as d ln i/d ln H. The signicant problem associated with performing experiments in this fashion is that the indentation strain rate is continually changing throughout the loading segment. The idea of measuring creep properties of materials with depth-sensing indentation systems under conditions of constant load was rst introduced by Pollock et al.[6] and subsequently rened by Mayo et al.[7,8] Their experimental procedure involved loading the indenter at a high loading rate and then holding the load on the indenter constant while monitoring the displacement of the indenter as a function of time. Since the contact area increases as the indenter penetrates the material while the load is held constant, the hardness decreases as the test proceeds, with a corresponding decrease in the indentation strain rate. This allows the tabulation of several hardness-indentation strainrate pairs from one indentation experiment. This technique was subsequently used by a number of authors to investigate the time-dependent response of a variety of thin lms and bulk materials.[9,10,1416] One study of particular note is that of Poisl et al., who utilized this technique to empirically determine the relationship between the indentation strain rate and the uniaxial strain rate as measured for amorphous Se. They found that, at temperatures above the glass transition temperature, Newtonian viscous (power-law 1) ow was observed. Assuming that the coefcient relating the hardness and the ow stress ( ) of Se was equal to 3, i.e., H 3 , the relationship between the indentation strain rate and the uniaxial strain rate was determined. Poisl et al. found that the coefcient relating the indentation strain rate to the effective strain rate is equal to 0.09, or 0.09 I. B. Effect of Temperature (Activation Energy for Creep) The temperature dependence of indentation creep data is an area that has received limited attention with depth-sensing indentation, due mainly to the inherent difculties in measuring the small displacements associated with the technique at elevated temperatures. The work that has been accomplished has typically been conducted at around room temperature or using techniques that incorporate large displacements, such that the thermal effects become small in comparison to the creep displacements. The earliest investigations into the temperature dependence of indentation creep were carried out using an impression creep test. Yu and Li[17] and Chu and Li[12,13]
602VOLUME 30A, MARCH 1999

investigated the temperature dependence of the impression creep process in LiF single crystals, succinonitrile crystals, and -tin single crystals. They reported activation energies (Q) for high-temperature impression creep that were in good agreement with the activation energies reported for high-temperature uniaxial creep as well as the activation energies for self-diffusion. Poisl et al.[14] investigated the temperature dependence of the indentation creep process using a Berkovich indenter, by changing the ambient temperature of the laboratory, near room temperature in their investigation, into the relationship between indentation creep and uniaxial creep in amorphous Se. From the Newtonian portions of the curves of log hardness vs log inden/ tation strain rate, they were able to calculate an activation energy for the indentation creep process. Using the indentation strain rates at a constant hardness, they obtained an activation energy of 495 kJ/mol. This value also compares favorably with the activation energy for viscous ow of 535 kJ/mol. Stone and Yoder[18] also used an indentation system, capable of operating in the 160 to 298 K temperature range, to investigate the rate-dependent and rate-independent components of the hardness of Mo. They found activation volumes for Mo in agreement with those reported in the literature, given the 50 pct typical scatter in these type of data. C. Steady-State Indentation Test The nal concept associated with power-law creep behavior is that of a steady-state behavior. One of the problems associated with the current state of indentation creep testing with sharp indenters is that when the hardness decreases with time, the stress driving the creep process also decreases and no steady state can be achieved. This is true of all of the indentation tests previously discussed. To avoid these difculties, Chu and Li[12,13] proposed a test for bulk materials in which the shape of the indenter is changed from a sphere or pyramid to a circular cylinder with a at end. As the contact area remains constant during the test, the punching or indenting stress is constant at a constant load. Using a power-law constitutive equation, Chu and Li derive an expression relating the punch velocity to the effective stress. Following a short transient period, they observe a steady-state velocity during which the descent rate of the indenter is found to have the same stress dependence as that observed during conventional unidirectional creep tests using bulk specimens for LiF single crystals,[17] succinonitrile crystals,[12] and -tin single crystals.[13] The cause for the transient stage is believed to be due to the development of a steady-state plastic zone under the punch, for example, a subgrain or cell structure not very different from that developed during conventional steadystate creep. This hypothesis is supported by the subgrain formation beneath the cleaved surface of a LiF single crystal after high-temperature impression creep testing. The only drawback of this technique is that the volume of material being deformed is essentially dened by the radius of the punch.[17] To perform these experiments in such a way as to constrain the plastically deformed volume to shallow depths, it becomes necessary to use punches of smaller and smaller diameters. As the size of the punch is decreased, the errors associated with the contact area of the at punch become more pronounced, eventually limiting the useful
METALLURGICAL AND MATERIALS TRANSACTIONS A

range of the technique. These problems can be overcome by using geometrically similar indenters such as a pyramid or cone, if a technique can be developed that allows a similar steady state to be reached. III. EXPERIMENTAL EQUIPMENT

The room-temperature experiments conducted for this work were performed on a NANO INDENTER II. The intricacies of the system have been detailed elsewhere[19] and will not be discussed here. The resolutions of the loading system and displacement sensing system are 50 nN and 0.04 nm, respectively. The elevated-temperature experiments discussed in this work were performed with the High-Temperature Mechanical Properties Microprobe (HTMPM) at the Oak Ridge National Laboratory. The HTMPM is a low-load, ultrahigh-vacuum, high-temperature, depth-sensing indentation system. The system combines much of the technology found in the current ambient-temperature mechanical properties microprobes with the ability to perform experiments at temperatures ranging from 100 C to approximately 300 C. The details of the design can be found in Reference.[20] The resolutions of the loading system and displacement sensing system are 1 N and 0.125 nm, respectively. The physical structure of the HTMPM system is very similar to that of the NANO INDENTER II. The primary difference is that displacement measurements in the HTMPM are achieved through the use of an AXIOM 2/20*
*AXIOM 2/20 is a trademark of the Zygo Corporation, Middleeld, CT.

The indium specimens used for all indentation tests were prepared from commercially available In of greater than 99.999 pct purity. Specimens were cast, using standard techniques, to a thickness of 0.3885 in. The as-cast material was then rolled to a nal thickness of 0.0825 in. before being cut to the nal dimensions and subsequently annealed at 100 C for 1 hour. The surface was prepared for indentation by electropolishing in a 3:1 solution of methanol and nitric acid maintained at 20 C. This process resulted in a mirror-nish surface with a nal grain size of approximately 2 to 3 mm.

V.

LOAD-TIME HISTORIES

A. Percent Step/Hold A variation of a step-load experiment was utilized in the HTMPM to take advantage of the high data acquisition rates in determining the effect of the percentage of the maximum load that was applied to the specimen in a step fashion. All of the experiments were conducted to a maximum load of 10 mN. The load was applied by rst ramping the load to a percentage of Pmax in 10 seconds and then immediately applying the remainder of the load in the form of a step function. The load was then held constant for a period of time, such that the displacement vs time response under constant load could be monitored. Experiments were conducted in which 90, 80, 70, and 30 pct of the maximum load was applied in a step fashion. Note that the denition of t 0, which will be used for the displacement vs time response to this experiment type, is just prior to the application of the step load. B. Constant Rate of Loading/Hold In these experiments, the load on the indenter was ramped to a specied maximum at a constant loading rate and then held constant for a period of time. Elevated-temperature experiments on In were conducted using a 1-second ramp to maximum load of 10 mN. The load ramp was immediately followed by a hold period under constant load to monitor the displacement vs time (indentation creep). A typical hold time under constant load was 180 seconds. Room-temperature experiments were conducted on In at four different loading rates: 0.1, 0.3, 1, and 10 mN/s, each to a maximum load of 10 mN. This type of experiment allowed the indentation strain rate and hardness to be determined during a segment when both were expected to be changing. C. Constant Loading Rate Divided by the Load In most types of indentation creep tests, the indentation strain rate and, therefore, hardness continuously change during the experiment. For instance, in a constant-load creep test, the indenter is loaded at a specied rate and the load is then held constant for a period of time while the displacement is monitored. The indentation strain rate changes both during loading, due to the imposed loading rate, and also during the subsequent hold segment under constant load, due to the decreasing pressure (hardness) as the contact area increases. It is, therefore, desirable to perVOLUME 30A, MARCH 1999603

He-Ne laser interferometric displacement sensing system. Two of the beams from the interferometer strike a mirror that is clamped behind the indenter, while the other two beams strike a mirror that is rigidly coupled to the specimen. This conguration allows for a direct measure of the displacement of the indenter into the specimen with very little possibility for adverse contributions from dimensional instabilities in the load frame. The AXIOM 2/20 system is capable of producing displacement measurements with a time constant as short as 0.05 s. This very exciting feature of the measurement system gives the HTMPM unique capabilities not found in other mechanical properties microprobes. Temperature monitoring in the HTMPM is achieved through the use of one thermocouple located directly on the indenter tip and another located just below the specimen. Digital controllers independently monitor the temperature of the indenter and the specimen and control two independent power supplies for heating. The resolution of the temperature measurement is 0.2 C. Heating is achieved through the use of tungsten and tantalum resistance heaters, one located just above the indenter, the other located in the specimen stage. IV. MATERIALS/SAMPLE PREPARATION

All experiments in this work were performed on highpurity, electropolished indium (In), which has a melting point of 429.8 K (156.6 C). Its creep properties are well known over a wide range of homologous temperatures.[21]
METALLURGICAL AND MATERIALS TRANSACTIONS A

form an indentation experiment during which the indentation strain rate and, therefore, hardness remain constant. The proposed technique for conducting constant indentation strain-rate experiments can be developed from the equation for the hardness of a material, H P A P ch 2 [3]

where hc is the contact depth of the indentation. The Berkovich geometry is typically chosen over a Vickers indenter, as it is much easier to produce near-perfect tips due to the natural intersection of three planes at a point. Unless otherwise specied, the area function used for calculating the hardness will be that of Eq. [5]. VII. THERMAL DRIFT

where H is the hardness, P is the load, A is the projected contact area, h is the contact depth (or total depth in this work), and c is a constant that depends upon the geometry of the indenter (24.56 for the perfect Berkovich geometry). Equation [3] can be differentiated and simplied to obtain h h 1 P 2 P H H [4]

Given the time required to conduct the vast majority of the indentation creep experiments, any assumptions about the constancy of thermal drift during the entire experiment are questionable. With this in mind, every effort was made to minimize the effects of thermal drift in all of the experiments by allowing several hours for thermal equilibrium to be reached at each temperature. With this in mind, no correction was made to the data for thermal displacements. VIII. DEFINITION OF HARDNESS

where i is the indentation strain rate. Equation [4] suggests that an indentation experiment performed with a pyramidshaped indenter, during which the loading rate is controlled so that the loading rate divided by the load (P/P) is constant, can result in a constant value of the indentation strain rate if a steady-state value of the hardness can be reached and h 0. This type of test is very appealing, as the loading rate is a directly controllable parameter in the NANO INDENTER II. To test the hypothesis that controlling the loading rate in such a manner as to maintain P/P constant would result in a constant indentation strain rate, a series of experiments at a variety of P/P values were conducted on pure In at room temperature. The loading rate was controlled in such a way as to maintain the instantaneous value of P/P constant up to a maximum load of 10 mN. The load was then held constant for a period of time, and the displacement vs time was monitored. The P/P values of 0.2, 0.1, 0.02, 0.01, and 0.005 s 1 were used. Data were acquired at depths of 2 m and greater to avoid any surface effects, where the hardness of the material might be inuenced by surface layers or environmental exposure. D. P/P Change In order to investigate the response of pure In to an abrupt change in the indentation strain rate, a series of experiments were conducted in which the value of P/P was changed during the experiment at some specied depth. This type of experiment allowed the investigation of the existence of a path-independent indentation steady-state hardness. Experiments were conducted in which P/P was abruptly changed from a value of 0.02 to 0.1, as well as from 0.1 to 0.02, at a depth of 3 m.

In its simplest denition, hardness is thought of as describing the athermal strength of a material. However, as the ow stress of a material is a function of strain rate and temperature, the hardness of a material should be expected to vary in an analogous way. For this work, hardness will be dened as the instantaneous mean pressure that the surface will support as calculated from the load divided by the projected contact area. The contact area is calculated from the total depth of the indent and the area function for the perfect Berkovich indenter, i.e., Eq. [5]. The total depth was used rather than an elastically corrected contact depth, because elastic recovery of the indentation depth was small ( 1 pct of the depth). The accuracy of dening the hardness based on the total depth of the indent rather than an elastically corrected depth is discussed. The hardness will be considered to be a function of the indentation strain rate and temperature in a manner analogous to the ow stress in uniaxial creep. IX. RESULTS AND DISCUSSION

A. Effects of Stress The central theme of this section is to investigate fully the relationship between indentation strain rate and hardness. Data from a variety of experimental techniques are presented. The results obtained from the different experimental techniques are compared and evaluated. B. Percent Step/Hold Results Figure 1 is a semilog plot of the displacement vs time response of the indenter obtained for indium for the percent step/hold loading conditions, to a maximum load of 10 mN. The upper four curves show the displacement response to an applied step load, with varying amounts of the total load being applied during the ramp prior to the step to maximum load. The percentage of Pmax stepped on at t 0 is indicated for each curve. There are several important features to note from these four curves. First, note the wide range of descent rates experienced by the indenter, as indicated by the span of eight orders of
METALLURGICAL AND MATERIALS TRANSACTIONS A

VI.

INDENTER GEOMETRY

All of the experiments conducted in this work were performed using a Berkovich indenter, a three-sided pyramidal indenter with the same depth-to-area ratio as a Vickers indenter. The projected area of contact for a perfect Berkovich geometry is given by A
604VOLUME 30A, MARCH 1999

24.56 h 2 c

[5]

Fig. 1A semilog plot of displacement vs time for In showing the response of the material to a variety of step loads.

interferometric displacement sensing technique being used in the HTMPM. Second, the depth achieved immediately following the step load is largest for the largest percentage Pmax stepped on. This is believed to be due to the large inertial or dynamic forces that are being experienced by the specimen, as are evident from the oscillatory response of the indenter following the applied step, as shown in A in Figure 1. Finally, immediately following the step load, an incubation period is observed where the displacement rate of the indenter is observed to be smallest for the largest step fraction, and the four curves are seen to reverse their order in terms of overall displacement. Since, under normal data acquisition conditions, the rst data point on these curves is obtained at approximately 1 second, the region on the time axis where these four curves intersect is not normally observed. This initially reduced creep rate is believed to be due to (1) the stronger structure resulting from the higher strain rate experienced by the material when a larger fraction of the maximum load is applied in a step fashion, and (2) the larger contact area and, therefore, the lower stress that results from the dynamic overload. After the initial incubation period, the displacement rates for the four curves become similar. C. CRL/Hold Results Figure 2 is a log-log plot of the room-temperature indentation strain rate vs the hardness data from the CRL/hold experiments. Each curve represents an average of ve experiments performed under each loading condition. The strain rates were calculated by taking the time derivative of the displacement both during the load ramp and during the subsequent hold period under constant load and dividing the instantaneous rates by the displacement at that point in time. The indentation strain rate and hardness are observed to decrease continuously throughout the constant-loadingrate loading regime and, even more rapidly, during the period under constant load. It is also observed that the data immediately following the load application undergo a transient period immediately upon completion of the load ramp. These data are then observed to converge to a common curve over a period of time similar to the loading time. A power-law t to the linear portion of the data in this region yields an average slope of 6.1. In order to verify the calculated hardnesses and the validity of neglecting the elastic displacements, optical images were taken of indentations made under the three different loading conditions to measure the actual contact areas and to look for trends in the error as a function of the loading time. Since, for metals, the lengths of the diagonals of the indentation have been observed to remain relatively unchanged once the load is removed,[22] the residual contact area is expected to yield a good representation of the contact area under load. While a slight trend in the scatter of the data was observed, the scatter decreasing with increasing loading times, all of the data are observed to fall within an acceptable error window of 5 pct. Figure 3 is a log-log plot of indentation strain rate vs hardness data for experiments conducted at 28 C, 50 C, and 75 C. All of the data exhibit similar trends, again showing a transient region in progressing from the load ramp to the hold period under constant load. An interesting
VOLUME 30A, MARCH 1999605

Fig. 2A log-log plot of the average indentation strain rate vs the average hardness for series of CLR/hold experiments on In.

Fig. 3A log-log plot of indentation strain rate vs hardness for indentation creep experiments on In at temperatures of 28 C, 50 C, and 75 C.

magnitude on the time axis. This reveals one of the useful aspects of the high data acquisition rates available with the
METALLURGICAL AND MATERIALS TRANSACTIONS A

Fig. 4A plot of indentation strain rate vs displacement for three different constant P/P experiments conducted on In showing constant indentation strain rates equal to 0.5 P/P.

Fig. 7A log-log plot of indentation strain rate vs hardness comparing the results of the CRL/hold experiments to the results of the constant P/P experiments.

stress exponent of approximately 5 at the higher temperatures, the slope increasing slightly toward a value of 6 to 7 as the temperature is decreased. D. Constant P/P Figure 4 is a plot of the indentation strain rates vs displacement, obtained using the constant P/P loading schemes as well as the data from the 1 mN/s CRL/hold experiment for comparison. The strain rates were again calculated by taking the time derivative of the displacement, both during the loading segment and during the subsequent hold segment, and dividing the instantaneous rates by the displacement at that point in time. As suggested by Eq. [4], the indentation strain rates are constant during the loading segment and equal to 0.5 P/P, suggesting that an indentation steady state has, in fact, been reached and that h 0. The indentation strain rate is again observed to decrease once the load ramp terminates. Figure 5 is a plot of the calculated hardness vs displacement for the constant P/P experiments as well as for the 1 mN/s CRL/hold experiment, again, for comparative purposes. The calculated hardness is observed to remain constant during the constant P/P loading segment and to systematically decrease as P/P and, thus, h/h is decreased. Since, for each P/P experiment, a constant h/h and H is achieved, it is possible to tabulate indentation steadystate strain ratestress pairs analogous to uniaxial data. Figure 6 is a log-log plot of indentation strain rate vs hardness for the data from the constant P/P experiments. The data were generated by averaging the constant hardness/constant h/h data between 2 m and the termination of the loading segment. Each data point represents the average of ve experiments at each P/P. A power-law t to the data yields a slope of 7.3 (R 0.995). Over the range of strain rates investigated, the power-law relationship was observed to remain constant. The stress exponent for indentation creep of 7.3 is very near the value of 7.6 obtained for indium tested uniaxially by Weertman[21] near room temperature.
METALLURGICAL AND MATERIALS TRANSACTIONS A

Fig. 5A plot of hardness vs displacement for ve different constantP/P experiments conducted on In showing indentation steady, state behavior in response to constant P/P load ramp.

Fig. 6A log-log plot of average indentation strain rate vs average hardness for constant P/P experiments conducted on In.

note is that the length of the transient period seems to be reduced as the temperature is raised. All of the data are seen to be linear near the end of the hold period, with a
606VOLUME 30A, MARCH 1999

Table I.

Summary of Stress Exponents for Different Loading Schemes Technique CRL/hold CRL/hold (high T ) P/P Stress Exponent, n 6 to 7 5 to 7 7.3

creep. While the two sets of data were acquired under very different conditions, the appropriate stress exponent seems to come out of both sets of data. Second, the data from the CRL/hold experiments appears to have a stronger microstructure, as indicated by the higher hardness determined at the same strain rate. This is reasonable since the material has undergone a wide range of strain rates, beginning at very high values and decreasing throughout the experiment. F. Effects of Temperature (Determination of Apparent Activation Energy for Creep) Figure 8 is a plot of the natural log of the indentation strain rate at constant hardness vs 1/T for experiments conducted between 28 C and 75 C. The apparent activation energy for indentation creep obtained by a linear t to the data is 77.9 kJ/mol. This is in agreement with the activation energy for self-diffusion in pure indium, which is reported to be in the range from 75 to 78 kJ/mol[23,24] and is in accord with the observation that the activation energy for creep in pure metals is close to the activation energy for self-diffusion.[25] By compensating the indentation strain rate for the temperature of the experiment, a normalized plot of indentation strain rate vs hardness can be obtained. Figure 9 is a loglog plot of the temperature-compensated indentation strain rate vs hardness for the three sets of data shown in Figure 3. The fact that all three sets of data, when compensated for the test temperature, lie on a single curve supports the premise that a single thermally active process is controlling the deformation. The slope of 5 agrees with the observed values for stress exponents in pure metals obtained from conventional means, where the rate-controlling process is dislocation climb.[26] In calculating the temperature-compensated creep rate, it is often necessary to take into account the temperature dependence of the shear modulus (G) in order arrive at an appropriate activation energy.[27,28] It has been shown for a number of materials that taking values at a constant stress rather than at a constant value of /G(T) can lead to overestimation of the activation energy for creep.[21] Normalizing the hardness for the temperature dependence of the modulus and determining indentation strain rates at constant values of H/G or H/E yields values for the apparent activation energy for indentation creep of 59.3 and 58.0 kJ/mol, respectively. One possible explanation for the lower activation energy is that the mechanism is not self-diffusion, but perhaps some coupled diffusion mechanism such as a combination of lattice diffusion and dislocation core diffusion, as has been observed in Al at intermediate temperatures. However, at such high homologous temperatures, this explanation seems suspect. A more likely explanation is that the differences in the activation energies determined from the two analyses, which are roughly 20 pct different, are simply within the scatter of the data being used for determining the apparent activation energy for creep. G. Steady-State Path-Independent Hardness (P/P Change Tests) In order to test the hypothesis that a steady-state value of hardness could be reached that depends only on the imposed strain rate and temperature, a set of P/P change exVOLUME 30A, MARCH 1999607

Fig. 8A plot of natural log of indentation strain rate at constant values of hardness vs the reciprocal absolute temperature for In.

Fig. 9A log-log plot of temperature-compensated indentation strain rate vs hardness for 28 C, 50 C, and 75 C experiments on In.

E. Comparison of Techniques for Obtaining the Stress Exponent Figure 7 is a log-log plot of indentation strain rate vs hardness for the data from both the constant P/P experiments as well as the CRL/hold experiments. Strain rate hardness pairs were calculated for both the loading segment and hold segment from the CRL/hold experiments. Recall that each set of data represents the average of ve separate indentation experiments performed for each experiment type. Several interesting observations can be made about these data. The similarity of the stress exponents obtained from the two types of experiments and the offset in the hardness leads to two conclusions. First, the dominant factor in both types of experiments seems to be the stress exponent for
METALLURGICAL AND MATERIALS TRANSACTIONS A

Fig. 10A plot of indentation strain rate vs displacement for P/P change experiments on In.

In the indentation test, the deformed volume of material is continually expanding to encompass previously undeformed material. The material underneath the indenter is very often likened to an expanding cavity with a hydrostatic core, where no deformation is occurring, and an expanding elastic/plastic boundary. The creep process is believed to depend upon the rate at which the elastic/plastic boundary can proceed into the material. In contrast to the uniaxial test, the indentation creep process is, therefore, seen to entail an ever-increasing volume of material as deformation proceeds. The requirement of additional displacement rather than a characteristic time to reach a new indentation steady-state value of hardness can perhaps be understood in this light. As the plastically deformed volume in the indentation test is continually expanding into previously undeformed material, the indentation steady state is always being approached from a hardening direction rather than a softening direction, even when P P has been decreased. The necessity for additional deformation to reach a new steady state, therefore, seems to be appropriate. H. Comparison to Uniaxial Data While indentation creep testing is a viable technique for comparing the time-dependent properties of materials that cannot be characterized by standard bulk techniques, how the indentation results can be compared to uniaxial creep data is a subject of major importance. This section presents and discusses on an empirical level how the indentation creep data and the uniaxial data compare. Since the constant PP indentation experiments yield steady-state behavior, this type of experiment is believed to most closely approximate the results from the constant-stress tensile experiments and will, therefore, be compared to the steady-state uniaxial data. Before beginning, it is important to note that a central difference between uniaxial testing and indentation testing lies in the geometry of the two tests. In any mechanical tests, forces and displacements are controlled and/or measured depending upon the type of apparatus being used. However, the displacement response of a sample to an applied force, on its own, tells nothing of the properties of the material without incorporating the geometry of the specimen. In uniaxial testing, the geometry of the test is dened by the user. Specimens are typically geometrically simple, e.g., a cylindrical compression specimen or a dog-bone tensile specimen. In an indentation test, the geometry of the test is actually being controlled by the very properties of the material that are of interest, e.g., the hardness, Youngs modulus, or strain-rate sensitivity. The test itself is typically designed to examine the properties that are of interest. Figure 12 is a log-log plot of the temperature-compensated uniaxial strain rate vs uniaxial stress, plotted along with the temperature-compensated indentation strain rate vs the hardness. The data have been compensated for temperature using the activation energy for self-diffusion, 75 kJ/mol.[21] A slightly lower value for the stress exponent and a nite offset between the two data sets is observed for the indentation data, as compared to the tensile data at room temperature, when compared on a temperature-compensated strain-rate basis. Also included in Figure 12 is the indentation strain rate vs hardness/3.3, based on the experimental results of Tabor[29] as well as various modeling efMETALLURGICAL AND MATERIALS TRANSACTIONS A

Fig. 11A plot of hardness vs displacement for P/P change tests on In.

periments were conducted in which the value of P/P was abruptly changed from an initial value to a new value at a predetermined displacement. This type of experiment allowed the response of the material to changes in the imposed indentation strain rate to be determined. Figure 10 is a plot of the resulting indentation strain rate vs displacement for the PP change experiments. The initial indentation strain rate is again observed to be constant and equal to 0.5 P/P. Upon a change in P/P, the indentation strain rate is observed to go through a transient period and then approach a new constant h/h after a delta displacement of approximately 500 nm. Figure 11 is the resulting hardness vs displacement for the P/P change experiments. The measured hardness resulting from the P/P change is also observed to undergo a transient period and then approach the indentation steadystate value from the constant PP experiments. The displacement required seems to be somewhat a function of whether P/P was increased or decreased. The delta displacement required to reach the indentation steady-state value at a P/P of 0.1 is approximately 1 m after the P/P increase from 0.02, while the additional displacement required on going from a PP of 0.1 to 0.02 requires nearly 2 m of displacement.
608VOLUME 30A, MARCH 1999

While the offset between the two sets of data cannot be completely accounted for by either constraint factors relating the hardness to a ow stress of the material[33] or by predictions from time-dependent nite-element calculations,[32] the utility of the constant P/P indentation data is evident. The most often sought-after parameter in these types of experiments is how the rate of deformation of the material changes in response to a change in the applied stress, i.e., the stress exponent for creep. From the comparison of the two sets of data, it is obvious that the constant P/P indentation data yield an accurate value for the stress exponent for creep. X.
Fig. 12A log-log plot of temperature-compensated indentation strain rate vs hardness for constant P/P experiments on In plotted with temperature-compensated tensile strain rate vs stress data from Weertman.

CONCLUSIONS

Fig. 13A log-log plot of indentation strain rate/Bower factor vs hardness for constantP/P experiments on In plotted with tensile strain rate vs stress data from Weertman.

forts.[30,31] The constraint factor does bring the two sets of data closer together, but does not completely account for the observed difference. A nal method for relating the indentation data of a power-law creeping solid to its uniaxial counterpart is that of Bower et al.,[32] who used nite-element modeling to arrive at an expression for the effective stress and effective strain rate under the indenter. Their effective stress is taken to be the contact pressure or hardness, and the shape of the indenter and the displacement rate determine the effective strain rate. The effective stress is assumed to be related to the effective strain rate by the uniaxial stressstrain response. Assuming that the hardness is equal to the uniaxial stress, the indentation strain rate divided by the Bower factor should yield the uniaxial strain rate. Figure 13 is a plot of the indentation data modied according to the appropriate Bower factor. The Bower factor serves to shift the indentation strain rate nearly four orders of magnitude on the strain-rate axis. While this modication technique also serves to bring the two sets of data closer together, it also fails to completely reconcile the observed differences.
METALLURGICAL AND MATERIALS TRANSACTIONS A

It was shown for the rst time that the indentation strain rate could be held constant during an indentation experiment using a Berkovich indenter, by controlling the loading rate such that the loading rate divided by the load remained constant. The constant P/P technique seems to most closely approximate the steady-state results from uniaxial testing. Comparison of the results from the constant P/P experiments to the results from conventional CRL/hold indentation creep experiments shows that the transition from the load ramp to the hold period under constant load shows a brief period with an apparently higher stress exponent for creep at similar strain rates. The indentation strain rate hardness data for In at high homologous temperatures display a stress exponent of 5 once the apparent transient behavior is exhausted. This is in agreement with the stress exponent obtained for high-temperature uniaxial creep, where dislocation climb is the rate-controlling mechanism.[26] The apparent activation energy for indentation creep in indium was found to be approximately 78 kJ/mol. Temperature compensating the indentation strain rate using this experimentally determined activation energy is seen to bring all of the data at the various temperatures together onto a single master curve, supporting the hypothesis of a single thermally activated, rate-controlling mechanism. Finally, by performing P/P change experiments, it was shown that a steady-state path-independent hardness could be reached in an indentation test with a geometrically similar indenter. The arrival at a new steady-state value of hardness seems to depend on the accumulation of strain rather than a relaxation time. In closing, the application of these techniques to materials with moderate-to-high melting points is a subject of ongoing development and research. Current hardware limitations prevent temperatures higher than 300 C from being obtained. This effectively prevents creep studies at the high homologous temperatures of many important engineering materials. However, as the hardware evolves past the prototype stage, temperatures up to 1000 C are expected to be attainable. This temperature range will allow these techniques to be applied to a signicantly wider range of pure metals as well as alloy systems.

REFERENCES
1. A.G. Atkins, A. Silverio, and D. Tabor: J. Inst. Met., 1966, vol. 94, pp. 369-78. 2. S. Hannula, D.S. Stone, and C.Y. Li: Electronic Packaging Materials
VOLUME 30A, MARCH 1999609

3. 4. 5. 6. 7. 8. 9.

10. 11.

12. 13. 14. 15.

Science, Materials Research Society Symposia Proceedings, E.A. Geiss, K.N. Tu, and D.R. Uhlmann, eds., Materials Research Society, Pittsburgh, PA, 1985, vol. 40, pp. 217-24. W.R. LaFontaine, B. Yost, R.D. Black, and C.-Y. Li: J. Mater. Res., 1990, vol. 5, pp. 2100-06. M.J. Mayo and W.D. Nix: Acta Metall., 1988, vol. 36, pp. 2183-92. M.J. Mayo and W.D. Nix: in Strength of Metals and Alloys, P.O. Kettunen, T.K. Lepisto, and M.E. Lehtonen, eds., Pergamon Press, Elms ford, NY, 1988, pp. 1415-20. H.M. Pollock, D. Maugis, and M. Barquins: in Microindentation Techniques in Materials Science and Engineering, P.J. Blau and B.R. Lawn, eds., ASTM, Philadelphia, PA, 1986, pp. 47-71. M.J. Mayo, R.W. Siegel, Y.X. Liao, and W.D. Nix: J. Mater. Res., 1992, vol. 7, pp. 973-79. M.J. Mayo, R.W. Siegel, A. Narayanasamy, and W.D. Nix: J. Mater. Res., 1990, vol. 5, pp. 1073-82. V. Raman and R. Berriche: Thin Films: Stresses and Mechanical Properties II, Materials Research Society Symposia Proceedings, M.F. Doerner, W.C. Oliver, G.M. Pharr, and F.R. Brotzen, eds., Materials Research Society, Pittsburgh, PA, 1990, vol. 188, pp. 171-76. V. Raman and R. Berriche: J. Mater. Res., 1992, vol. 7, pp. 627-38. B.N. Lucas and W.C. Oliver: Thin Films: Stresses and Mechanical Properties III, Materials Research Society Symposia Proceedings, W.D. Nix, J.C. Bravman, E. Arzt, and L.B. Freund, eds., Materials Research Society, Pittsburgh, PA, 1992, vol. 239, pp. 337-41. S.N.G. Chu and J.C.M. Li: J. Mater. Sci., 1977, vol. 12, pp. 2200-08. S.N.G. Chu and J.C.M. Li: Mater. Sci. Eng., 1979, vol. 39, pp. 1-10. W.H. Poisl, W.C. Oliver, and B.D. Fabes: J. Mater. Res., 1995, vol. 10, pp. 2024-32. S.P. Baker, J.T.W. Barbee, and W.D. Nix: Thin Films: Stresses and Mechanical Properties III, Materials Research Society Symposia Proceedings, W.D. Nix, J.C. Bravman, E. Arzt, and L.B. Freund, eds., Materials Research Society, Pittsburgh, PA, 1992, vol. 239, pp. 33741.

16. K.B. Yoder, D.S. Stone, J.C. Lin, and R.A. Hoffman: Thin Films: Stresses and Mechanical Properties V, Materials Research Society Symposia Proceedings, S.P. Baker, C.A. Ross, P.H. Townsend, C.A. Volkert, and P. Borgesen, eds., Materials Research Society, 1994, vol. 356, pp. 651-56. 17. E.C. Yu and J.C.M. Li: Phil. Mag., 1977, vol. 36, pp. 811-25. 18. D.S. Stone and K.B. Yoder: J. Mater. Res., 1994, vol. 9, pp. 252433. 19. M.F. Doerner and W.D. Nix: J. Mater. Res., 1986, vol. 1, pp. 60109. 20. B.N. Lucas: Ph.D. Dissertation, The University of Tennessee, Knoxville, TN, 1997. 21. J. Weertman: Trans. AIME, 1960, vol. 218, pp. 207-18. 22. N.A. Stilwell and D. Tabor: Proc. Phys. Soc. London, 1961, vol. 78, pp. 169-79. 23. R.E. Eckert and H.G. Drickamer: J. Chem. Phys., 1952, vol. 20, pp. 13-17. 24. G.W. Powell and J.D. Braun: Trans. AIME, 1960, vol. 230, pp. 69499. 25. O.D. Sherby and A.K. Miller: ASME J. Mater. Technol., 1979, vol. 101, pp. 387-95. 26. O.D. Sherby: Acta Metall., 1962, vol. 10, pp. 135-47. 27. C.R. Barrett, A.J. Ardell, and O.D. Sherby: Trans. TMS-AIME, 1964, vol. 230, pp. 200-04. 28. J. Weertman, W.V. Green, and E.G. Zukas: J. Mater. Sci. Eng., 1970, vol. 6, pp. 199-211. 29. D.S. Tabor: Proc. R. Soc. A, 1948, vol. 192, pp. 247-74. 30. K.L. Johnson: J. Mech. Phys. Solids, 1970, vol. 18, pp. 115-26. 31. S.S. Chiang, D.B. Marshall, and A.G. Evans: J. Appl. Phys., 1982, vol. 53, pp. 298-311. 32. A.F. Bower, N.A. Fleck, A. Needleman, and N. Ogbonna: Proc. R. Soc. London A, 1993, vol. 441, pp. 97-124. 33. D.S. Tabor: Hardness of Metals, Clarendon Press, Oxford, United Kingdom, 1951, pp. 103-07.

610VOLUME 30A, MARCH 1999

METALLURGICAL AND MATERIALS TRANSACTIONS A

Potrebbero piacerti anche