Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Materials Science & Engineering A 751 (2019) 35–41

Contents lists available at ScienceDirect

Materials Science & Engineering A


journal homepage: www.elsevier.com/locate/msea

Nanoindentation creep behavior and its relation to activation volume and T


strain rate sensitivity of nanocrystalline Cu
⁎ ⁎
Weiming Suna, Yue Jianga, Guixun Suna, Jiangjiang Hub, , Tianyi Zhouc, Zhonghao Jianga, ,
Jianshe Liana
a
Key Laboratory of Automobile Materials, College of Materials Science and Engineering, Jilin University, Nanling Campus, Changchun 130025, China
b
College of Mechanical Engineering, Zhejiang University of Technology, Hangzhou 310014, China
c
IMI Precision Engineering, Shanghai 201108, China

A R T I C LE I N FO A B S T R A C T

Keywords: The creep behavior of nanocrystalline Cu with an average grain size of 25 nm was investigated by na-
Nanocrystalline Cu noindentation test at room temperature. Using the creep strain rate versus creep stress data obtained at different
Nanoindentation creep behavior loading rates, the activation volume and strain rate sensitivity were determined obtained by cooperating the
Activation volume continuous stiffness measurement (CSM) technique. The results showed that the activation volume first increases
Strain rate sensitivity
and then decreases, and the strain rate sensitivity first decreases and then increases with increasing the creep
Grain boundary
stress. The experimental activation volume and strain rate sensitivity versus the creep stress data exhibit very
Dislocation
good agreements with the theoretical values calculated by the previous models, respectively. The analysis based
on the data of the activation volume and strain rate sensitivity revealed that at lower stress, the grain boundary
activities dominate and lead to the lower creep strain rates; at higher stress, the dislocation activities dominate
and lead to the higher creep strain rates. The analysis based on the data of the nanoindentation test also revealed
that the use of the CSM technique can lead to the continuous creep strain rate versus creep stress data, which
allows us to uncover the creep mechanisms over a wide range of the creep stress from the initial to steady stage.

1. Introduction thermally activated to move the dislocations over localized obstacles in


their slip planes [14]. m is a sensitivity parameter of stress to strain rate
For coarse grained (CG) metals, plastic deformation is mediated by during plastic deformation and it can reflect a time or rate-dependent
nucleation and movement of intragranular dislocations. Interactions of deformation process [8]. V * and m can shed light on the thermally
dislocations on intersecting slip planes create permanent dislocation activated transition of plastic deformation mechanism with strain rate
network that hinders further the movement of dislocations and thus [15,16]. It is generally believed that a smaller V * and a larger m can be
hardens materials. However, such dislocation activity is no longer expected when the GB deformation process is activated at lower strain
possible in nanocrystalline (NC) metals and is replaced by emission of rates and, in contrast, a larger V * and a smaller m can be expected when
dislocations from grain boundary (GB) sources and their propagation the dislocation activity dominates at higher strain rates. Therefore, the
across the grains and eventual absorption in the opposite GBs [1–5]. On change of V * and m suggests a transition in plastic deformation me-
the other hand, GB-mediated process such as GB sliding and GB diffu- chanism.
sion can take place concurrently with the dislocation activity [6–8]. By using different testing methods, the stress dependences of the V *
Complicated interactions between the GB and dislocation mechan- and m values were obtained on different NC and ultrafine grained
isms also result in difficulty in identifying which of them dominates in a (UFG) materials, e.g., compressive test on UFG Cu [17], strain-rate
mechanical experiment with given experimental condition. The acti- jump test on UFG Cu [18], stress relaxation tests on NC Ni [19] and UFG
vation volume (V *) and strain rate sensitivity (m) are two important Ta [20], and nanoindentation creep tests on NC Ta [21], NC Ni [22],
parameters in the description of the kinetic and thermodynamic char- UFG Ti [23] and CoCrFeCuNi high-entropy alloy [24]. It is noted from
acteristics of a plastic deformation process and have been widely used these tests that the resultant V * and m values exhibit the either
to uncover the plastic deformation mechanisms of different metals and monotonous increase or monotonous decrease except for the UFG Cu
alloys [9–13]. V * is the number of atoms that have to be coherently [17] whose V * increases and then decreases with stress. Duhamel et al.


Corresponding authors.
E-mail addresses: jiangjiangwho@gmail.com (J. Hu), jzh@jlu.edu.cn (Z. Jiang).

https://doi.org/10.1016/j.msea.2019.02.027
Received 18 October 2018; Received in revised form 8 February 2019; Accepted 8 February 2019
Available online 10 February 2019
0921-5093/ © 2019 Elsevier B.V. All rights reserved.
W. Sun, et al. Materials Science & Engineering A 751 (2019) 35–41

[18] proposed a theoretical model to investigate the dependence of V * The nanoindentation tests were repeated at least ten times under the
on opening plastic deformation mechanisms. This model predicts that same conditions.
when the plastic deformation of a material is mainly controlled by GB
activities, V * will be very small and increase with increasing the stress 3. Results and discussion
and, in contrast, when the plastic deformation is mainly controlled by
dislocation activities, V * will decrease with increasing the stress. This 3.1. Creep behavior
means that V * should be a continuous and non-monotonous function of
the stress. Similar dependence of m on the stress was also suggested by Fig. 2(a) shows the representative load versus displacement (P − h )
Champion [17]. However, as mentioned above, few of the existing curves obtained at the P ̇ of 0.05 mN/s to 500 mN/s. In the loading
experimental data of V * and m [17–24] showed such stress dependence regime, h at given P increases sharply with decreasing P ̇ . In the holding
even GB and dislocation mechanisms are effectively activated. Possible regime, the P − h curves exhibit wide load plateaus (large creep dis-
reasons for the lack of such dependence are the limitation in the testing placement (hcreep )). Fig. 2(b) shows the corresponding creep displace-
method and the insufficient range of the stress tested. ment versus holding time (hcreep − tH ) curves in the holding regime. The
In this paper, the dependence of V * and m on creep stress of a na- h − t data in the holding regime can be well fitted by an empirical
nocrystalline Cu with an average grain size of 25 nm was investigated equation
by nanoindentation creep test cooperating the CSM technique in a wide r
range of loading rates at room temperature. The creep strain rate versus h = p + q ln ⎛ − 1⎞
⎝t − s ⎠ (1)
creep stress data were obtained. Using these data, the two relationships
between the V * and m values and creep stress were determined. These where p , q , r and s are the fitting parameters. The representative fitting
two relationships were compared with the theoretical models [17,18]. h − t curves and the experimental data in the holding regime at
Based on the V * and m data, the creep mechanisms were analyzed. P = 500 mN/s, 5 mN/s and 0.05 mN/s are shown in Fig. 2(c), (d)
Finally, some important features of the current nanoindentation testing and (e), respectively. It can be seen that the fitting curves show good
methods were discussed. agreements with the experimental data. The creep strain rate (εḢ ) can
be expressed as [26]
1 ∂h
2. Experiments εḢ =
h ∂t (2)

The NC Cu used in this text was prepared by electric brush-plating Fig. 2(f) shows the creep strain rate versus holding time (εḢ − tH )
technique and the synthetic process was described in Ref. [9]. This NC curves obtained from the h − t curves in the holding regime. It is seen
Cu consists of roughly equiaxed grains with random crystalline or- that εḢ drops rapidly with increasing tH , and then decreases slowly in
ientations and predominant high-angle GBs. The purity and density of the subsequent steady-state creep regime. At given tH , εḢ increases with
the NC Cu are ~99.73 wt% and ~8.93 g/cm3, respectively. The main increasing P ̇, which indicates that the creep behavior depends strongly
impurities, in wt%, are 0.060C, 0.054O, 0.023H, 0.056N, 0.001S and on P ̇ .
0.001P. Fig. 1(a) and (b) shows the TEM bright-field image of the NC Cu Fig. 3(a) and (b) show the representative load versus displacement
and the grain size distributions, respectively. As shown, this NC Cu has (P − h ) curves and the corresponding contact stiffness versus dis-
a narrow grain size distribution with an average grain size (d) of 25 nm. placement (S − h ) curves obtained by the CSM technique at the loading
Nanoindentation tests were performed on a nanoindenter (KLA- strain rates of 0.5, 0.05 and 0.005 s−1, respectively. It can be seen that S
Tencor Nanoindenter G200) with a Berkovich diamond indenter at is independent of loading strain rate. Fig. 4 shows the creep stress
room temperature. The surface of NC Cu was mechanically polished to versus time (σH − tH ) curves in the holding regime obtained by using
mirror finishing to acquire reliable data. The area function of the in- the following equation [26,27]
denter had been calibrated on the standard fused silicon before na- P
noindentation tests. The contact stiffness (S ) as a function of indenter σH =
3Ac (3)
displacement was obtained by the CSM technique, via loading strain
rates of 0.5, 0.05 and 0.005 s−1 and a maximum displacement of where Ac is the contact area. Ac can be expressed as [26,27]
2000 nm. Nanoindentation creep tests were performed at a wide range Ac = C0 hc2 + C1 hc + C2 hc1/2 + C3 hc1/4 (4)
of loading rates (P ̇ ) from 0.05 mN/s to 500 mN/s to a peak load of 300
mN with a holding time of 1000 s. The indenter was unloaded to 10% of where C0 , C1, C2 , C3 are the fitting parameters and hc is the contact
the maximum load and held for 500 s for thermal drift correction to depth. In this work, the fitting parameters are C0 = 24.3, C1 = 5115.9,
minimize its negative influence on the displacement measurement [25]. C2 = 10088.2 and C3 = −38910.5 which were obtained by the area

Fig. 1. The TEM bright-field image of the NC Cu (a) and the grain size distributions (b).

36
W. Sun, et al. Materials Science & Engineering A 751 (2019) 35–41

Fig. 2. The load versus displacement (P − h ) curves (a) obtained at a maximum load of 300 mN and holding time of 1000 s at different P ̇ . The corresponding creep
displacement versus holding time (hcreep − tH ) curves (b), the experimental displacement versus time (h − t ) curve and the fitted curve obtained at Ṗ = 500 mN/s (c),
P ̇ = 5 mN/s (d) and P ̇ = 0.05 mN/s (e) and the creep strain rate versus holding time (εḢ − tH ) curves (f) in the holding regime.

function calibration procedure. hc can be expressed as [26,27] 3.2. Determinations of activation volume and strain rate sensitivity

εP According to the results in Figs. 2(f) and 4, the εḢ and σH at the same
hc = h −
S (5) tH in the whole creep process can be obtained, so the εḢ versus σH data
were obtained at different P ̇ and the resultant εḢ − σH curves are shown
where ε is a constant. For the Berkovich indenter in this work, ε = 0.75 in Fig. 5(a). It is interestingly noted that the εḢ − σH data at different P ̇
[27]. In the calculations of σH (Eq. (3)), the S − h data obtained by CSM mutually superimpose and εḢ increases with increasing σH . Based on
technique in Fig. 3(b) were used to determine hc and Ac . As shown in these εḢ − σH curves, the true activation volume V * can be expressed as
Fig. 4, σH decreases rapidly with increasing tH and then decreases stably [28–30]
with increasing tH . σH has a larger dropping rate at higher P ̇ than that at
lower P ̇ , which indicates a more rapid stress relaxation at higher P ̇ than ∂ ln εḢ ⎞
that at lower P ̇ . V * = MkT ⎛
⎝ ∂σ * ⎠ (6)

37
W. Sun, et al. Materials Science & Engineering A 751 (2019) 35–41

Fig. 3. The load versus displacement (P − h ) curves (a) and the corresponding contact stiffness versus displacement (S − h ) curves (b) obtained by CSM technology
at the loading strain rates of 0.5, 0.05 and 0.005 s−1.

between V * and σH .

3.3. Stress dependence of activation volume and strain rate sensitivity

Duhamel et al. [18] proposed a model in which V * is connected with


σ*. In this model, a critical stress σ0*, which can be expressed as
σ0* = Mβ Gb/ d , is proposed, where β is the material constant, G is the
shear modulus. When σ * ≤ σ0*, the plastic deformation is controlled
only by GB activities and the corresponding V * can be written as [18]
V * = VGB (9)
where VGB = 1b3 is the activation volume when the plastic deformation
is controlled only by GBs activities [28]. When σ * > σ0* , the plastic
deformation is dominated by both GB and dislocation activities and the
corresponding V * can be given by [18]
−1
σ * − σ0* δ C (σ0*)2 ⎞
V * = VGB + 2MkT ⎜⎛ + ⎟

Fig. 4. The creep stress versus holding time (σH − tH ) curves in the holding ⎝ B b σ * − σ0* ⎠ (10)
regime.
where δ = 2b is the thickness of GB, B and C are the fitting parameters.
The first and second terms in the bracket of the right of Eq. (10) re-
where M is the Taylor factor, k is the Boltzmann constant, T is the present respectively the contributions of the dislocation and GB activ-
absolute temperature and σ* is the effective stress, which can be written ities to V *. B and C can be expressed as
as σ * = σH − σi , where σi is the athermal long-range internal stress and
δ ∂vGB
is usually approximated as a constant. So, Eq. (6) can be written as B=1+ >1
bd 2ve ∂ρ (11)
∂ ln εḢ ⎞
V * ≈ MkT ⎛ ⎜ ⎟
C=
νGB
⎝ ∂σH ⎠ (7) Bνe (12)
By taking M = 3 [16,18,31], k = 1.38 × 10−23 J/K and T = 298 K where νe is the frequency of dislocation emission from GB, νGB is the
and using the εḢ versus σH data in Fig. 5(a), the variation of V * with σH frequency of an elementary shear at GB companied with dislocation
was obtained and the result is shown in Fig. 5(b). As shown, there are absorption and ρ is the dislocation density. According to Duhamel et al.
two distinctive regions. In region I, V * increases with increasing σH [18], B and C can be approximately regarded as the constants, and C
from 6b3 to a maximum value of 10.5b3 at 530 MPa, where b is the can be determined by a condition in which the plastic deformation is
Burgers vector of the dislocation and b = 0.256 nm for Cu [32]. In re- controlled only by GB activities. According to this condition, the term
gion II, V * decreases with increasing σH from the maximum value to (σ * − σ0*)/ B in Eq. (10) will be far smaller than the corresponding term
2.5b3 . (δ C / b) σ0*2/(σ * − σ0*) and Eq. (10) can be written as
Using the εḢ versus σH data in Fig. 5(a), the variation of the strain
rate sensitivity m with σH , defined as [33,34] 2MkTb
V * ≈ VGB + (σ * − σ0*)
δ C (σ0*)2 (13)
∂ ln σH
m= Eq. (13) indicates that C can be determined from the slope of the
∂ ln εH˙ (8)
experimental V * versus σH data when the relation of σ * − σ0* to σH is
can be determined and the result is shown in Fig. 5(c). As shown, there determined. According to σ * = σH − σi and σi = σ0 − σ0* [18], one has
are also two distinctive regions. In region I, m decreases with increasing σ * − σ0* = σH − σ0 , where σ0 is the value of σH at which V * = 1b3 .
σH from 1.04 to a minimum value of 0.8 at 530 MPa. In region II, m Therefore, σ0 can be determined by finding the crossover point of the
increases with increasing σH from the minimum value to 1.2. It is noted tangent of the V * versus σH curve in region I with the σH axis (see
that the relationship between m and σH is contrary to the relationship Fig. 5(b)). From Fig. 5(b), σ0 is determined to be σ0 = 450 MPa. By

38
W. Sun, et al. Materials Science & Engineering A 751 (2019) 35–41

According to the relationship between the m and V *, i.e.,


m = MkT /(σH V *) [35], and the expressions for V * (Eqs. (9) and (10)),
the corresponding m can be written as [17]
MkT
m= (σH ≤ σ0)
σH VGB (14)

−1 −1
⎡σ V σ − σ0 δC σ02 ⎞ ⎤
m = ⎢ H GB + 2σH ⎛⎜ H + (σH σ0)
b σH − σ0 ⎠ ⎥

MkT ⎝ B (15)
⎣ ⎦
Inserting σ0 = 450 MPa and C = 0.02 and the values of other para-
meters, the m versus σH curves were calculated for different B values
and compared with the experimental curve. The comparison results in
Fig. 6(b) show that the calculated m versus σH curve with B = 2.5 ex-
hibits the best agreement with the experimental curve.

3.4. Explanations of activation volume and strain rate sensitivity

According to the work of Becker [35–37], V * can be written as


V * = b × ξ × l*, where ξ is the distance swept out by the mobile dis-
location during an activation event, which can be considered to be a
constant [35,38], and l* is the length of dislocation segment or the
distance between two pinning sites on GBs. Furthermore, as suggested
in Ref. [35], l* has two limits, i.e., when ρ is low, l* can be written as
l1* = χ d ; when ρ is high, l* can be expressed as l 2* = aρ−1/2 , where a and
χ are proportionality factors. There is the grain size across which me-
chanisms controlling plastic is different. This grain range is 10–500 nm.
This suggests that for the present NC Cu, the variation of l* depends on
the stress level and thus results in different dependence of V * on σH .
For NC metals, dislocation activities, including the nucleation from
GB sources, their propagation across the grains and eventual absorption
in the opposite GBs occurs when a critical stress is reached [1–5]. At
low stress, this critical stress cannot be satisfied, so the deformation
must be mediated by GB activities. In this case, l* is a very small con-
stant which are associated with the GB activities or some dislocation
activities within GBs or near GBs (see Fig. 7(a)), and a small V * value is
thus expected at low stress. As σH increases, GB dislocation sources
would be activated, and dislocations begin to nucleate from GBs and
propagate across the grains with their two ends depinning and shifting
along the GBs (see Fig. 7(b)). In this case, ρ is lower and l* is controlled
by l1* = χ d , where the χ value should be a stress dependent parameter
and 0 < χ ≤ 1. An increased l* or an increased V * would be expected
when σH varies in a range from 460 to 530 MPa. When σH reaches a
value larger than 530 MPa, would be effectively activated, the number
of the dislocations nucleated GBs increase significantly. An effective
dislocation storage inside the grain interior would occur due to the
enhanced nucleation of dislocations from GBs and the increased diffi-
culties for the dislocation absorption in the opposite GBs (see Fig. 7(c)).
In this case, ρ increases significantly and l* is controlled by l 2* = aρ−1/2 .
As a result, a decreased l* or a decreased V * would be expected when σH
is larger than 530 MPa. The above analysis clearly indicate that the
Fig. 5. Creep strain rate versus creep stress (εḢ − σH ) curves (a), activation steady creep process at lower stress level is mediated mainly by the GB
volume versus creep stress (V * − σH ) curves (b) and strain rate sensitivity activities, which is gradually replaced by the high speed creep process
versus creep stress (m − σH ) curves (c) in the holding regime. mediated mainly by the dislocation activities.
According to the relation m = MkT /(σH V *) mentioned above, there
taking β = 0.5 and d = 25 nm, σ0* is calculated to be σ0* = 374 MPa . is an approximate relationship between m and 1/ V * since σH is not a
Inserting σ * = σH − 76 MPa , σ0* = 374 MPa and the values of other constant. Therefore, the variation in the m value with σH (Fig. 5(c)) and
parameters into Eq. (13) and comparing with the slope value of the its relation to the creep mechanisms can be similarly explained to those
experimental V * versus σH curve in region I (0.07b3 MPa−1), the C value for the V * value.
can be determined to be C = 0.02 .
Furthermore, inserting σ * = σH − 76 MPa , σ0* = 374 MPa , C = 0.02 3.5. Some features of present nanoindentation creep tests
and the values of other parameters, the V * versus σH curves were cal-
culated for different B values and compared with the experimental An important result of our nanoindentation creep test is that the
curve. The comparison results in Fig. 6(a) show that the calculated V * εḢ − σH curves obtained at different P ̇ mutually superimpose, which
versus σH curves with B = 2.5 exhibits the best agreement with the allows us to obtain the continuous εḢ − σH data and hence the con-
experimental curve. tinuous V * and m versus σH data over a wide range of σH from 450 MPa

39
W. Sun, et al. Materials Science & Engineering A 751 (2019) 35–41

Fig. 6. Comparison of the activation volume versus creep stress (V * − σH ) curves obtained by experiment and Duhamel et al.’s model (a), and comparison of the
strain rate sensitivity versus creep stress (m − σH ) curves obtained by experiment and Champion's model with different B (b).

to 1350 MPa. It is noted that such V * and m versus σH data were less data. It is noted that in these tests [21–24], CSM technique was not used
reported. In the previous nanoindentation creep tests on NC and UFG and the unloading rates were not mentioned. In addition, it can be seen
metals [21–24], the εḢ − σH curves obtained at different loading strain from Fig. 3(b) that S increases linearly with h , while the S value
rates or loading rates were discrete and did not mutually superimpose. measured by the nanoindentation creep test is a single value. This is
Although the m values obtained from the discrete εḢ − σH curves in also an important reason responsible for producing the discrete εḢ − σH
these studies are accurate as indicated by Eq. (7) and were used to curves.
analyze the creep behavior in the steady stages, the range of σH covered Because the continuous V * or m versus σH data can be obtained over
are narrow. The resultant V * and m values either increase or decrease a very wide range of σH , the present nanoindentation creep test focuses
monotonously with increasing σH and the non-monotonous variation in on the entire creep process and the variation in the relevant creep
the V * and m value like that found in the present nanoindentation creep mechanisms from the initial to steady stage can be more clearly un-
test were less observed [17]. covered. This is obviously better than the previous nanoindentation
In general, for the creep deformation of a plastic material, there creep tests [21–24] in which only the creep process in the steady stages
should exist a certain relationship between εḢ and σH and the mutually were emphasized. In addition, it should be pointed out that the con-
superimposed εḢ − σH curves should be thus expected. Because εḢ is tinuous strain rate versus stress data also can be similarly achieved in
only related to the h − t data, the discrete εḢ − σH curves obtained in the loading region when the load-controlling mode is used. Under this
the previous nanoindentation creep tests [21–24] indicate that the σH mode, the strain rate varies with the loading stress. If such data can be
data might be not measured accurately. In fact, for plastic materials, achieved, the corresponding activation volume or strain rate sensitivity
especially NC and UFG metals, the creep deformation can occur not versus stress data can be obtained, which can be used to analyze the
only during the holding stage, but also in the unloading stage. A direct variation of plastic deformation mechanisms during loading process,
consequence of the creep deformation in the unloading stage is that it which will be the future work of the present authors.
leads to a nose-shaped P − h curve in the initial unloading stage
especially when a slow unloading rate was used [39]. In this case, it is 4. Conclusion
difficult to obtain the accurate S data. An effective method for over-
coming this inaccuracy in the S data is to suppress the creep de- Nanoindentation tests were carried out to investigate the creep
formation by using the higher unloading rate as possible. Another so- behaviors of NC Cu with a grain size of 25 nm after deforming under a
lution is to use CSM technique. It is believed that CSM technique can wide range of loading rate at room temperature. Based on the re-
obtain the accurate S − h data and such data is independent on the lationship between the creep strain rate and creep stress obtained at
unloading process [40,41]. Therefore, a possible reason for the discrete different loading rates, the activation volume and strain rate sensitivity
εḢ − σH curves observed in the previous nanoindentation creep tests versus creep stress data were determined by cooperating CSM tech-
[21–24] is the inaccurate σH data caused by the errors in the resultant S nique. The results shows that the activation volume increases and then

Fig. 7. Schematic illustrations of the pinning-depinning mechanism of creep behavior.

40
W. Sun, et al. Materials Science & Engineering A 751 (2019) 35–41

decreases, and the strain rate sensitivity decreases and then increases understanding of mechanical behavior of nanocrystalline metals, Acta Mater. 55
with the increasing creep stress. The experimental activation volume (2007) 4041–4065.
[16] Y. Wang, A. Hamza, E. Ma, Temperature-dependent strain rate sensitivity and ac-
and strain rate sensitivity versus the creep stress data are in good tivation volume of nanocrystalline Ni, Acta Mater. 54 (2006) 2715–2726.
agreements with the theoretical values calculated by the previous [17] Y. Champion, Competing regimes of rate dependent plastic flow in ultrafine grained
models, respectively. The analysis based on the data of the activation metals, Mater. Sci. Eng. A 560 (2013) 315–320.
[18] C. Duhamel, Y. Brechet, Y. Champion, Activation volume and deviation from
volume and strain rate sensitivity revealed that at lower stress, the grain Cottrell–Stokes law at small grain size, Int. J. Plast. 26 (2010) 747–757.
boundary activities dominate and lead to the lower creep strain rates, at [19] F. Dalla Torre, P. Spätig, R. Schäublin, M. Victoria, Deformation behaviour and
higher stress, the dislocation activities dominate and lead to the higher microstructure of nanocrystalline electrodeposited and high pressure torsioned
nickel, Acta Mater. 53 (2005) 2337–2349.
creep strain rates. The analysis based on the data of the nanoindenta- [20] Y. Wang, Y. Liu, J. Wang, Investigation on activation volume and strain-rate sen-
tion test also revealed that the use of the CSM technique can lead to the sitivity in ultrafine-grained tantalum, Mater. Sci. Eng. A 635 (2015) 86–93.
continuous creep strain rate versus creep stress data, which allows us to [21] Z. Cao, P. Li, H. Lu, Y. Huang, Y. Zhou, X. Meng, Indentation size effects on the
creep behavior of nanocrystalline tetragonal Ta films, Scr. Mater. 60 (2009)
uncover the creep mechanisms over a wide range of the creep stress
415–418.
from the initial to steady stage. Similar method of nanoindentation test [22] Z. Ma, S. Long, Y. Zhou, Y. Pan, Indentation scale dependence of tip-in creep be-
can be applied to the loading region. havior in Ni thin films, Scr. Mater. 59 (2008) 195–198.
[23] X. Liu, Q. Zhang, X. Zhao, X. Yang, L. Luo, Ambient-temperature nanoindentation
creep in ultrafine-grained titanium processed by ECAP, Mater. Sci. Eng. A 676
Acknowledgments (2016) 73–79.
[24] Y. Ma, G.J. Peng, D.H. Wen, T.H. Zhang, Nanoindentation creep behavior in a
This work was financially supported by the National Natural Science CoCrFeCuNi high-entropy alloy film with two different structure states, Mater. Sci.
Eng. A 621 (2015) 111–117.
Foundation of China (Grant Nos. 51371089 and 51271152). [25] J. Hu, G. Sun, X. Zhang, G. Wang, Z. Jiang, S. Han, J. Zhang, J. Lian, Effects of
loading strain rate and stacking fault energy on nanoindentation creep behaviors of
References nanocrystalline Cu, Ni-20 wt% Fe and Ni, J. Alloy. Compd. 647 (2015) 670–680.
[26] B. Lucas, W. Oliver, Indentation power-law creep of high-purity indium, Metall.
Mater. Trans. A 30 (1999) 601–610.
[1] H. Van Swygenhoven, Grain boundaries and dislocations, Science 296 (2002) [27] W.C. Oliver, G.M. Pharr, An improved technique for determining hardness and
66–67. elastic modulus using load and displacement sensing indentation experiments, J.
[2] J. Schiøtz, K.W. Jacobsen, A maximum in the strength of nanocrystalline copper, Mater. Res. 7 (1992) 1564–1583.
Science 301 (2003) 1357–1359. [28] H. Conrad, Grain size dependence of the plastic deformation kinetics in Cu, Mater.
[3] R.J. Asaro, S. Suresh, Mechanistic models for the activation volume and rate sen- Sci. Eng. A 341 (2003) 216–228.
sitivity in metals with nanocrystalline grains and nano-scale twins, Acta Mater. 53 [29] S. Mishra, V.K. Beura, A. Singh, M. Yadava, N. Nayan, Rate sensitive behavior of
(2005) 3369–3382. obstacles in age hardenable aluminum alloys, Mater. Sci. Eng. A 729 (2018)
[4] J. Rajagopalan, C. Rentenberger, H.P. Karnthaler, G. Dehm, M.T.A. Saif, In situ TEM 102–105.
study of microplasticity and Bauschinger effect in nanocrystalline metals, Acta [30] J. Hu, W. Zhang, T. Zhang, Y. Zhang, Nanoindentation deformation of refine-
Mater. 58 (2010) 4772–4782. grained AZ31 magnesium alloy: indentation size effect, pop-in effect and creep
[5] M.S. Colla, B. Aminahmadi, H. Idrissi, L. Malet, S. Godet, J.P. Raskin, D. Schryvers, behavior, Mater. Sci. Eng. A 725 (2018) 522–529.
T. Pardoen, Dislocation-mediated relaxation in nanograined columnar palladium [31] L. Lu, R. Schwaiger, Z. Shan, M. Dao, K. Lu, S. Suresh, Nano-sized twins induce high
films revealed by on-chip time-resolved HRTEM testing, Nat. Commun. 6 (2015) rate sensitivity of flow stress in pure copper, Acta Mater. 53 (2005) 2169–2179.
5922. [32] J. Mu, Z. Jiang, W. Zheng, H. Tian, J. Lian, Q. Jiang, High-speed creep process
[6] Z. Shan, E. Stach, J. Wiezorek, J. Knapp, D. Follstaedt, S. Mao, Grain boundary- mediated by rapid dislocation absorption in nanocrystalline Cu, J. Appl. Phys. 111
mediated plasticity in nanocrystalline nickel, Science 305 (2004) 654–657. (2012) 063506.
[7] Y. Wei, A.F. Bower, H. Gao, Enhanced strain-rate sensitivity in fcc nanocrystals due [33] Y. Ma, J. Ye, G. Peng, D. Wen, T. Zhang, Nanoindentation study of size effect on
to grain-boundary diffusion and sliding, Acta Mater. 56 (2008) 1741–1752. shear transformation zone size in a Ni–Nb metallic glass, Mater. Sci. Eng. A 627
[8] Z. Jiang, X. Liu, G. Li, Q. Jiang, J. Lian, Strain rate sensitivity of a nanocrystalline (2015) 153–160.
Cu synthesized by electric brush plating, Appl. Phys. Lett. 88 (2006) 143115. [34] T. Zhang, J. Ye, Y. Feng, Y. Ma, On the spherical nanoindentation creep of metallic
[9] Z. Jiang, H. Zhang, C. Gu, Q. Jiang, J. Lian, Deformation mechanism transition glassy thin films at room temperature, Mater. Sci. Eng. A 685 (2017) 294–299.
caused by strain rate in a pulse electric brush-plated nanocrystalline Cu, J. Appl. [35] Q. Wei, S. Cheng, K. Ramesh, E. Ma, Effect of nanocrystalline and ultrafine grain
Phys. 104 (2008) 053505. sizes on the strain rate sensitivity and activation volume: fcc versus bcc metals,
[10] H. Zhang, Z. Jiang, Y. Qiang, Microstructure and tensile deformation of nanocrys- Mater. Sci. Eng. A 381 (2004) 71–79.
talline Cu produced by pulse electrodeposition, Mater. Sci. Eng. A 517 (2009) [36] H. Conrad, Thermally activated deformation of metals, JOM 16 (1964) 582–588.
316–320. [37] J. Cahn, F. Nabarro, Thermal activation under shear, Philos. Mag. A 81 (2001)
[11] J. Hu, S. Han, G. Sun, S. Sun, Z. Jiang, G. Wang, J. Lian, Effect of strain rate on 1409–1426.
tensile properties of electric brush-plated nanocrystalline copper, Mater. Sci. Eng. A [38] M.J. Kobrinsky, C.V. Thompson, Activation volume for inelastic deformation in
618 (2014) 621–628. polycrystalline Ag thin films, Acta Mater. 48 (2000) 625–633.
[12] J. Hu, J. Zhang, Z. Jiang, X. Ding, Y. Zhang, S. Han, J. Sun, J. Lian, Plastic de- [39] W. Li, K. Shin, C. Lee, B. Wei, T. Zhang, Y. He, The characterization of creep and
formation behavior during unloading in compressive cyclic test of nanocrystalline time-dependent properties of bulk metallic glasses using nanoindentation, Mater.
copper, Mater. Sci. Eng. A 651 (2016) 999–1009. Sci. Eng. A 478 (2008) 371–375.
[13] Y. Jiang, J. Hu, Z. Jiang, J. Lian, C. Wen, Strain rate dependence of tensile strength [40] X. Li, B. Bhushan, Continuous stiffness measurement and creep behavior of com-
and ductility of nano and ultrafine grained coppers, Mater. Sci. Eng. A 712 (2018) posite magnetic tapes, Thin Solid Films 377 (2000) 401–406.
341–349. [41] S. Vachhani, R. Doherty, S. Kalidindi, Effect of the continuous stiffness measure-
[14] P. Spätig, J. Bonneville, J.-L. Martin, A new method for activation volume mea- ment on the mechanical properties extracted using spherical nanoindentation, Acta
surements: application to Ni3 (Al, Hf), Mater. Sci. Eng. A 167 (1993) 73–79. Mater. 61 (2013) 3744–3751.
[15] M. Dao, L. Lu, R.J. Asaro, J.T.M.D. Hosson, E. Ma, Toward a quantitative

41

You might also like