You are on page 1of 11

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/327547278

Depth-sensing time-dependent response of additively manufactured Ti-6Al-


4V alloy

Article · September 2018


DOI: 10.1016/j.addma.2018.09.008

CITATIONS READS

4 196

5 authors, including:

Muztahid Muhammad Mohammad Masoomi


Auburn University Auburn University
10 PUBLICATIONS   29 CITATIONS    15 PUBLICATIONS   60 CITATIONS   

SEE PROFILE SEE PROFILE

Brian Torries Nima Shamsaei


Mississippi State University Auburn University
12 PUBLICATIONS   131 CITATIONS    141 PUBLICATIONS   2,265 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Autonomous Quadcopter for Product Home Delivery View project

Fatigue and Cyclic Deformation of Superelastic and Shape Memory Alloys View project

All content following this page was uploaded by Muztahid Muhammad on 13 September 2018.

The user has requested enhancement of the downloaded file.


Additive Manufacturing 24 (2018) 37–46

Contents lists available at ScienceDirect

Additive Manufacturing
journal homepage: www.elsevier.com/locate/addma

Depth-sensing time-dependent response of additively manufactured Ti-6Al- T


4V alloy
Muztahid Muhammada, Mohammad Masoomib,c, Brian Torriesb,c, Nima Shamsaeib,c,

Meysam Haghshenasa,
a
Department of Mechanical Engineering, University of North Dakota, ND, 58202, USA
b
Department of Mechanical Engineering, Auburn University, Auburn, AL, 36849, USA
c
National Center for Additive Manufacturing Excellence (NCAME), Auburn University, Auburn, AL, 36849, USA

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

Keywords: Depth-sensing (instrumented) indentation testing technique is a robust, reliable, convenient and non-destructive
Nanoindentation characterization method to study small-scale mechanical properties and rate-dependent plastic deformation in
Additive manufacturing metals and alloys at ambient and elevated temperatures. In the present paper, depth-sensing indentation creep
Ti-6Al-4V behavior of an additively manufactured, via laser powder bed fusion (L-PBF) method, Ti-6Al-4V alloy is studied
Creep rate
at ambient temperature. Indentation creep tests were performed through a dual-stage scheme (loading followed
Stress exponent
by a constant load-holding and unloading) at different peak loads of 250 mN, 350 mN, and 450 mN with holding
time of 400 s. Creep parameters including creep rate, creep stress exponent, and indentation size effect were
analyzed, according to the Oliver and Pharr method, at different additive manufacturing scan directions and scan
sizes. To assess processing parameter/ microstructure/ creep property correlations in the additively manufacture
Ti-6Al-4V alloy, microstructural quantitative analyses (i.e. optical microscopy and scanning electron micro-
scopy) were performed as well. The findings of this study, according to stress exponent values, showed that the
controlling mechanism of the creep at ambient temperature for the examined L-PBF Ti-6Al-4V is mainly glide-
controlled dislocation creep. These findings were compared against traditionally processed Ti-6Al-4V as well.

1. Introduction method, seems to be a suitable replacement for the conventional


manufacturing processes of the Ti-6Al-4V alloy. The benefits of additive
Due to its excellent physical and mechanical properties such as light manufacturing of the Ti-6Al-4V alloy include, but not limited to, cost
weight (almost half of steel), high strength (almost similar to steel), and reduction, decreased labor cost and material wastage, smaller foot-
biocompatibility, titanium (Ti) has been a metal of interest for trans- prints, less process, no mold and pattern, shorten fabrication time and,
portation applications in the automotive and aerospace industries, as most importantly, enhanced material properties. Among different AM
well as in biomedical industries [1–4]. The ability to form a stable and techniques, laser powder bed fusion (L-PBF) is a well-developed process
adherent surface oxide upon exposure to water and air has made Ti and to print the Ti-6Al-4V alloy. In the L-PBF process, also known as direct
its alloys excellent candidates for structural efficiency in extreme cor- metal laser sintering (DMLS), high power density laser is used to melt
rosive environments and sea water [1]. In addition, its great strength- and fuse metallic powder together [6]. Cooling rate during the additive
to-weight ratio at elevated temperatures have made Ti a suitable ma- manufacturing process and thermal processing history are quite dif-
terial for high temperature applications [1]. Among different Ti alloys, ferent compared with conventional manufacturing processes. There-
Ti-6Al-4V alloy is known as the ‘workhorse’ alloy [5] which accounts for fore, it is necessary to assess the relationship between the micro-
more than half of all commercial Ti applications [1]. However, con- structure, processing parameters, and properties of the additively
ventional manufacturing processes, like welding, casting, forming, and manufactured Ti-6Al-4V alloys to determine its ability to meet demands
machining of the Ti-6Al-4V alloy, are quite challenging due to high- of engineering design considerations.
temperature oxidation, high chemical reactivity, and poor thermal Consistency of mechanical properties and microstructures in ad-
conductivity. ditive manufacturing is still a critical issue. As Ti-6Al-4V alloy is a two-
Additive manufacturing (AM), a layer-wise material manufacturing phase alloy, mechanical properties can be significantly varied by


Corresponding author.
E-mail address: meysam.haghshenas@engr.und.edu (M. Haghshenas).

https://doi.org/10.1016/j.addma.2018.09.008
Received 29 June 2018; Received in revised form 5 September 2018; Accepted 7 September 2018
Available online 09 September 2018
2214-8604/ © 2018 Elsevier B.V. All rights reserved.
M. Muhammad et al. Additive Manufacturing 24 (2018) 37–46

tailored microstructure [7]. Due to various cooling rates at different levels that impose the mentioned strain accumulations, according to the
positions and distances from the built plate in different additive man- design criteria, are normally in the acceptable (safe) range. To this end,
ufacturing methods, the microstructure, in particular formation of α, understanding the nature of creep sensitivity could be a challenging
can vary a lot [8]. Formation of layers with partial re-melting and so- task but provides a foundation for our understanding on crack devel-
lidification of previous layer and cooling in various ways for various opment in α + β Ti alloys. For instance, ambient-temperature creep in
geometries adds complications to the processes. In spite of many studies the Ti-6Al-4V alloy fuel tanks and fasteners has been reported in the
in the literature to date, rate dependent plastic deformation of the Ti- literature [13].
6Al-4V alloy manufactured by the L-PBF method has not been studied Considering the above-mentioned literature, ambient-temperature
at small scales with microstructural gradients. creep response of an additively manufactured Ti-6Al-4V alloy and
The depth-sensing indentation testing technique (i.e. micro/ nano- correlations between creep parameters (i.e. creep rate, creep stress ex-
indentation) is considered a reliable, convenient, and non-destructive ponent, indentation strain rate sensitivity) and additive manufacturing
testing technique to examine the microstructure/ mechanical property variables (i.e. scan directions and scan sizes) has not yet been docu-
correlation in metals at ambient (room) and elevated temperatures in- mented. In the present study we employ an instrumented (depth-sen-
cluding rate-dependent plastic deformation. Nanoindentation is the sing) indentation testing technique to assess ambient-temperature creep
continuous measurement of force (in the range of nN to mN) and dis- and the corresponding mechanisms of an additively manufactured (AM)
placement (in the range of nm to μm based on the depth-sensitivity or Ti-6Al-4V alloy. Besides, indentation size effect and microstructural
average property required from the substrate). Nano-mechanical assessments are studied in the present paper. To this end, micro-
properties (hardness, indentation stress, indentation strain rate sensi- structural quantitative analyses (i.e. optical microscopy (OM) and
tivity, reduced Young’s modulus, etc.) could be measured from the load scanning electron microscopy (SEM)) are performed to assess micro-
(P) and displacement (h) data recorded throughout the process in the structure of the AM Ti-6Al-4V alloy and the creep property correlations
form of load vs displacement curve. Dislocation activity, phase trans- during the holding time as a function of indenter load.
formation, and shear localization can also be studied with this powerful The findings of this paper provide a baseline to study elevated-
technique [9,10]. temperature creep of AM Ti-6Al-4V alloy and to compare the results
A number of researchers have used this technique and conventional with the conventionally made Ti-6Al-4V alloy. Most of the reported
methods (uniaxial tension) to assess rate-dependent plastic deformation creep results in the literature are based upon conventional (tensile)
in Ti and Ti alloys. Ma et al. [11] studied room temperature creep approach which could be destructive, time-consuming and tough-to-
behavior of Ti-10V-2Fe-3Al based on dislocation mechanism by micro- control (specially at the elevated temperatures) tests. In particular, on
indentation. They found that power-law creep deformation was the the AM aspect, a number of creep test coupons must be printed which
controlling mechanism of creep in Ti-10V-2Fe-3Al alloy. Kumar et al. could be a costly task as well. Having said this, the present paper
investigated creep-fatigue interactions in the Ti-6Al-4V alloy at ambient promotes the application of the depth-sensing indentation testing
temperature [12]. Matsunaga et al. performed creep tests on poly- technique, as a reliable, convenient and non-destructive approach that
crystalline Ti along with pure Mg and pure Zn at ambient temperature can be performed on a small volume of material and can be used toward
and found steady state creep rate around 10−9 s-1 and creep stress assessing time-dependent plastic deformation (creep) in AM materials
exponents around 3.0 at ambient temperature [13]. Hasija et al. de- at ambient and elevated temperatures.
veloped a computational model for Ti-6Al alloys and analyzed time-
dependent plasticity for analyzing creep [14]. Barboza et al. in- 2. Experimental procedures
vestigated creep behavior of conventionally made Ti-6Al-4V alloy
consisting Widmanst ä tten microstructure using conventional uniaxial The material studied in this research is an additively manufactured
test and found different creep stress exponents as 4.4 and 4.1 at 500 °C Ti-6Al-4V alloy fabricated via the L-PBF process in horizontal and
and 600 °C [15]. Kral et al. [16] analyzed rate-dependent plastic de- vertical scan direction. Fig. 1 shows the scan direction and scan size of
formation in conventionally manufactured ultrafine-grained Ti-6Al-4V the samples used in this study. Two rectangular samples with dimen-
using both a conventional uniaxial approach at elevated temperatures sions of 9.8 × 9.8 × 5.60 mm3 (coded as Big sample and shown in
and ambient-temperature indentation creep. Badea et al. [17] analyzed Fig. 1(a) and (b)) and 4.8 × 4.8 × 5.5 mm3 (coded as Small sample and
creep behavior of hot-forged Ti-6Al-4V alloy and compared the creep shown in Fig. 1(c) and (d)) were prepared for both X and Y scan di-
stress exponent and activation energy at different temperature ranges. rection. Tracks for sample “Core X” were parallel to flow (argon) di-
Moreover, at higher stresses, Ti alloys can sometimes exhibit unusually rection and samples were named as “Core No Post X-big” (CNPX-big)
high creep stress exponent values, such as n > 15 [17,18]. and “Core No Post X-small” (CNPX-small). Tracks for sample “Core Y”
It has been reported that the ambient-temperature creep1 occurs in were perpendicular to flow direction and for this research samples were
α+β Ti–6Al–4V [2,6]. In general, the hexagonal close packed (hcp) named as “Core No Post Y-big” (CNPY-big) and “Core No Post Y-small”
crystalline structure of the Ti-6Al-4V is responsible for the ambient- (CNPY-small). “No post” indicates that no post contour exposure was
temperature creep [19,20]. That is, in the hcp structure, during room done on these specimens i.e. after fabrication of each layer, lasers re-
temperature creep, only one slip system is activated which is due to the melted the boundary of parts.
low symmetric structure, generating low work hardening. Considering Masoomi et. al. [25] have shown that by decreasing build area of
this, deformation proceeds at ambient temperature and under stresses additively manufactured parts or using shorter tracks, the temperature
below the yield stress [19,20]. gradient will decrease, and it will cause lower residual stress in the
Harrison et al. [21] developed a model to study time dependent parts. In this study parts with different volume are chosen to demon-
creep of Ti-6Al-4V by strain accumulation generated by dwell time strate this effect. The process parameters chosen to fabricate parts is
fatigue effects at ambient “cold” temperature creep. The occurrence of summarized in Table 1:
ambient temperature (or logarithmic) creep is not unusual in many Rotation was turned off during printing to study the effects of vo-
materials including pure Ti and Ti alloys [22–24]. lume effects without any interference from scan patterns. Porosity level
The key factor in low-temperature (i.e. room temperature) creep in the parts was not experimentally studied in this investigation.
sensitivity of Ti alloys (i.e. Ti-6Al-4V) is strong tendency of Ti in time- However, extensive mechanical and fatigue testing was done on the
dependent strain accumulation at low temperatures [22]. The stress parts [25]. The results suggest that the porosity level should be very low
due to the good behavior of the parts under loading.
Prior to the indentation, the surfaces of the specimens were care-
1
Ambient-temperature creep appears below 0.3–0.4Tm for hcp metals. fully ground with a series of progressively finer sand papers followed by

38
M. Muhammad et al. Additive Manufacturing 24 (2018) 37–46

fine polishing which resulted in scratch free mirror-like surface finish. A


modified Kroll’s reagent of 5 ml HF, 15 ml HNO3, and 80 ml distilled
water was used to reveal the microstructure. Microstructure and grain
sizes of four materials were then examined by Scanning Electron
Microscope (SEM) and Optical Microscopy (OM). Before indentation
testing, the microstructures of as-printed materials were studied using
Scanning Electron Microscopy (SEM, QUANTA FEG 650) and Optical
Microscopy (OM, MM500T).
Instrumented indentation creep tests with dual stage scheme
(loading followed by constant load holding) were performed at ambient
temperature (298 K) using a U9820 A Keysight Nano-Indenter G200. A
self-similar pyramidal Berkovich diamond indenter [10,26] with a face
angle of 65.3° was in this study. Loading was applied to test at a con-
stant indentation loading rate of 10 mN/s, with various peak loads of
250 mN, 350 mN, and 450 mN. The dwell (holding) time of 400 s was
chosen as the creep time. After 400 s, the sample was unloaded to 10%
of the peak load for the purpose of thermal drift corrections. Thermal
drift calibration was done prior to indentation to keep it under
0.05 nm/s, and each test was repeated five times to confirm reprodu-
cibility. The SEM and OM were used to analyze the indentation mor-
phology in terms of any possible sink-in and/ or pile-up.
Fig. 1. Built direction and built size of samples of the additively manufactured
Ti-6Al-4V alloy used in this study.
3. Calculation method

Table 1
To assess rate-dependent deformation behavior of a material by a
Parameters used for fabrication of Ti-6Al-4V parts.
depth-sensing indentation testing technique, it is required to mathe-
System EOS M290 matically represent the variation of indentation displacement as a
function of time. With a Berkovich self-similar pyramidal indenter,
Substrate material Ti-6Al-4V
Powder description Gas-atomized, air-dried experimental data of the holding stage was used to ascertain the creep
Mean particle diameter 35 μm behavior. In a depth-sensing self-similar nanoindentation testing, the
Powder layer thickness 50 μm indentation strain rate can be written as [27,28]:
Powder bed porosity 0.4
Laser spot diameter 100 μm ∂h
∂t 1 P˙
Laser power 170 W ε˙ind = =
Scan speed 1250 mm/s h 2P (1)
Building volume 25 × 25 × 35 cm3
Here, ε̇ind is indentation strain rate, h is the instantaneous indentation
Shielding gas type Argon
Shielding gas inlet temperature 20 °C depth, indentation displacement at a given time, ∂h is rate of indenta-
∂t
Shielding gas inlet flow rate 0.25 m3/s tion depth, Ṗ is loading rate, and P is indentation load. According to this
Chamber wall temperature 20 °C
equation, under a constant indentation load rate test, indentation strain
rate decreases with indentation depth.
Rotation was turned off during printing to study the effects of volume effects
without any interference from scan patterns. Porosity level in the parts was not Indentation stress is calculated as [27,28]:
experimentally studied in this investigation. However, extensive mechanical P
and fatigue testing was done on the parts [25]. The results suggest that the σind =
24.56 × (h + 0.06R)2 (2)
porosity level should be very low due to the good behavior of the parts under
loading. Here, P is indentation load, h is instantaneous contact depth, and R is

Fig. 2. SEM microstructure of (a) CNPX-big (b) CNPY-small.

39
M. Muhammad et al. Additive Manufacturing 24 (2018) 37–46

Fig. 3. Optical microscopy of microstructure of (a) CNPY-small (b) indents of CNPY-big.

following equation:

∂ ln ε˙ind
n=
∂ ln σind (5)

Conventional uniaxial creep testing involves homogeneous loading


of the complete gauge length of the sample, usually lasts between a few
hundred and several thousand hours which results in homogenized
value of stress [32]. However, during an instrumented indentation
creep testing, substantial and complex tri-axial stresses in the GPa
ranges are generated beneath the indenter. This could eventually in-
duce significant creep deformation in the materials even at room tem-
perature, unlike traditional uniaxial creep which is only observable at
elevated temperatures. It is worth mentioning that non-homogeneous
tri-axial stress states and a constantly growing deformation volume
under the indenter during an indentation creep test could result in a
deformation response rather different from that observed during con-
Fig. 4. Indentation morphology of Ti-6Al-4V for CNPY-small. ventional uniaxial creep testing. The dynamics of the deformation in
the indentation test are quite different in comparison to uniaxial creep.
Under the indenter, the deformed volume encompasses the previously
the indenter tip radius due to blunting at tip which was 200 nm for the
undeformed material by continually expanding. Material under the
Berkovich indenter used in this study.
indenter being strained is very likely to expand the cavity with a hy-
It is worth mentioning that local stress ahead of the indenter is large
drostatic core where no deformation was happening along with ex-
tri-axial stress and the test temperature in the current study (ambient
panding plastic/ elastic boundary. It is believed that time dependent
temperature) is well below the recrystallization temperature (i.e.
deformation process depends on rate of proceeding elastic/ boundary
0.5Tm) of the additively manufactured Ti-6Al-4V alloy. Therefore, the
into the material [28].
indentation creep would occur through dislocation creep in which
dislocation glide and the rate of the glide is limited by the presence and
strength of the obstacles within the microstructure. Considering this 4. Results and discussions
mechanism, the modified Dorn equation can be employed, as a con-
stitutive equation, to study ambient-temperature creep of the AM Ti- 4.1. Microstructure
6Al-4V alloy [13,29]:
Miller et al. [33] discovered that in the Ti-Al alloys, low tempera-
AD0 Gb ⎛ b ⎞ p σind n ΔQ ⎞ ture creep is largely dependent on the microstructure and therefore it is
ε˙ind = ⎛ ⎞ exp ⎛−
kT ⎝ d ⎠ ⎝ E ⎠ ⎝ kT ⎠ (3) required to have a detailed understanding on the microstructure of the
test material studied in the present study.
here, σind is the indentation stress, A is a coefficient related to tem-
Additively manufactured Ti-6Al-4V is typically an α-β alloy char-
perature and microstructure, D0 is the diffusion coefficient, G is the
acterized with prior β grains that generates in an epitaxial way through
shear modulus, ΔQ is the activation energy for thermal-activated pro-
several layers, grain boundary α, and α lath size [34–36]. Neikter et al.
cess, n is the stress exponent, E is the Young’s modulus, p is the grain
[37] characterized the microstructure of Ti-6Al-4V resulting from dif-
size exponent, k is the Boltzmann constant, T is the temperature, b is the
ferent additive manufacturing processes and found that powder bed
Burgers vector and d is the grain size.
fusion processes generate smaller prior β grain size than directed en-
At a constant temperature, the steady-state creep rate relation is
ergy deposition additive manufacturing processes. Due to the low to
further simplified per [30,31]:
intermediate cooling rates experienced by Ti-6Al-4V alloy manu-
n
ε˙ind = Bσind (4) factured by L-PBF processes, an α-β lamellar structure associated with
α-phase lamellae in a β-phase matrix is created. The α-lamellae are
here, B is a constant. generated by diffusion controlled nucleation and growth of α platelets
At isothermal conditions, n can be deduced by determining the slope into β-grains [38]. Size of these α platelets is controlled by the cooling
of a ln(ε̇ ) versus ln(σ) plot in the steady state creep condition using rate; an increased cooling rate results in a decreased diffusion rate,

40
M. Muhammad et al. Additive Manufacturing 24 (2018) 37–46

Fig. 5. Indentation load versus depth at different maximum indenter load at (a) CNPX_big (b) CNPX_small (c) CNPY_big (d) CNPY_small.

microscopy of indents along both the horizontal and vertical directions.


Fig. 4 shows a SEM micrograph of an indent made on the specimen.
Some sink-in effects are observed on the indented area. In the sink-in
phenomenon, the flat sides of the impression deform inward around the
indentation, which is a characteristic of elastic conical indentations
[40]. According to Bolshakov and Pharr [41], maximum indentation
depth (hm) and the final plastic depth upon unloading (hf) are two ex-
pedient indicators when assessing pile-up or sink-in characteristics
around the indenter. Sink-in behavior is a dominant factor when the
h
ratio f is less than a critical value (0.7), and material shows strain-
hm
hardening deformation behavior (i.e. ambient-temperature plastic de-
formation).

4.2. Indentation creep behavior

The load/displacement (P-h) curves with a loading rate of 10 mN/s


Fig. 6. Relationship between creep displacement with holding time and creep
and holding time of 400 s for four samples at different peak loads of
rate under maximum indenter load of 250 mN for CNPX-big with a holding time 250 mN, 350 mN, and 450 mN are shown in Fig. 5. Load plateaus are
of 400 s. observed at constant load holding stage and its width increases with
increasing indenter load. Stress distribution in the depth-sensing in-
dentation technique is much more complex than uniaxial tensile/
which subsequently leads to decreased length and thickness of the α-
compressive creep tests, and at low displacement the maximum shear
lamellae associated with higher yield strength [39]. Fig. 2 shows the
stress beneath the indenter exceeds the yield stress of the specimen
SEM micrograph of the microstructures of CNPX-big and CNPY-small
(large tri-axial stresses in the range of some giga-Pascal) [42]. This
samples where the size of α colony is determined by the cooling rate
phenomenon causes the occurrence of creep in the materials, including
from β phase and β grain size, while Fig. 3(a) shows the optical mi-
high melting-point materials, at ambient (room) temperatures, con-
croscopy images of the microstructure of a CNPY-small sample which is
trasting traditional creep tests which show creep mainly at elevated
showing scan tracks of additive manufacturing. Fig. 3(b) shows optical
temperatures [43]. In particular, in the Ti-6Al-4V alloy, ambient-

41
M. Muhammad et al. Additive Manufacturing 24 (2018) 37–46

Fig. 7. Creep displacement vs Constant load holding time at different maximum indenter loads for (a) CNPX-big, (b) CNPX-small, (c) CNPY-small, and (d) CNPY-big.

temperature creep induced by prism and basal slip is reported [14]. displacement is larger at higher indenter loads (450 mN), which in-
Besides, stress states and volume of deformed material are essen- dicates that creep displacement depends on the indentation peak load
tially different in the uniaxial and nanoindentation creep techniques; a (the larger the peak load, the higher the creep displacement will be). In
tri-axial non-homogeneous stress state and a continually growing de- the CNPX-big sample, the maximum displacements at the end of the
formation volume underneath the indenter are two important features dwell time were ∼ 1890 nm, ∼ 2200 nm, ∼ 2500 nm for the in-
of the nanoindentation testing approach. dentation peak loads of 250 mN, 350 mN, and 450 mN, respectively.
From the variation of creep rate, a creep curve originating from the Samples with perpendicular build direction seem to have lower creep
depth-sensing indentation is divided into two separate zones: transient displacement compared to their counterparts. From Fig. 7 it is observed
creep and steady-state creep [44]. Unlike traditional creep curves ob- that, for the 450 mN peak load, creep displacement reaches ∼ 2500 nm
tained from tensile/ compression testing, depth-sensing indentation for both CNPX-big and CNPX-small and 2200 nm and 2400 nm for the
testing does not have a third stage, or accelerated creep, as materials do CNPY-small and CNPY-big samples, respectively, at the end of the
not drastically fail under the indenter during indentation-based creep holding stage. No definite pattern was observed for creep displacement
testing [45]. Nucleated dislocations densified due to high stress at nano- regarding different scan size at maximum indenter loads.
indentation are highly unstable during loading stage and tend to relax Creep rates obtained using Eq. (1) were plotted against holding
during the constant load-holding stage. time, as shown in Fig. 8. With increase in the dwell time, there is a
Fig. 6 shows variation of creep displacement and creep rate over sharp decrease (transient creep) then plateau trend (steady state creep)
time for a CNPX-big sample. At first a sharp decrease in creep rate is in the creep rate. For all four samples, an initial sharp decrease in creep
observed, which then gradually reaches an almost constant stage. For rate of up to ∼ 0.00025 s−1 is observed at the loading stage, followed
indentation depth, a sharp initial rise is observed, which transitions to a by a nearly steady creep rate at holding time.
gradual increase over the holding time. There are numerous instances of ambient temperature creep in-
The relationship between creep displacement and constant load vestigations of Ti/ Ti alloys through modelling and/or traditional creep
holding time for Ti-6Al-4V alloy under different maximum loads at testing technique; however, it is quite challenging to compare the data
different samples is plotted in Fig. 7. All the curves experience an initial obtained through traditional creep tests and nano/ micro-indentation
sharp rise at the primary or transient creep stage and then as the in- creep tests due to their different loading and stress conditions. Harrison
denter penetration continues within the specimen, the creep increases et al. [21] developed a model for predicting ambient temperature time
almost linearly at steady-state stage, which is similar to the uniaxial dependent strain accumulation for Ti-6Al-4V at 20 °C over a range of
tensile/ compression creep. For all 4 samples, creep induced stresses. They recorded large n-values, attributed to the dislocation

42
M. Muhammad et al. Additive Manufacturing 24 (2018) 37–46

Fig. 8. Creep rate vs Holding time at different maximum indenter load for (a) CNPX-big, (b) CNPX-small, (c) CNPY-big, (d) CNPY-small.

creep, and observed that room-temperature stress rupture plot has a followed by an almost consatant creep rate in holding stage. Vertically
much shallower slope than at high temperature [21]. Neeraj et al. scanned samples i.e. samples having vertical (Y-axis) laser track
performed ambient temperature creep for various Ti alloys of different movement experienced slightly more creep rate than their horizontally
microstructures and compositions and creep strain was found in the scanned counter parts. The reason of this variation creep rate between
ranges of ∼ 0.0006 to ∼ 0.03 [24]. Evans [22] performed cold creep of horizontally and vertically scanned samples might be attributed to the
Ti-6Al-4V at different peak loads in the range of 900 MPa to 950 MPa thermal response during manufacturing of the samples [25,48]. Sam-
and found strain in the range of ∼ 0.04 to ∼ 0.14. ples with scanning track parallel to flow of argon gas experienced more
For crystalline materials under uniaxial tension/ compression, work heat due to temperature convection which results in lower thermal
hardening and increased dislocation density are the reasons behind gradient. Samples with perpendicular scanning track experienced less
gradually decreasing creep rates. It is worth mentioning that limited heat which results in lower temperature gradient.
number of slip systems in the hcp Ti-6Al-4V along with relatively low
volumes of dislocation-dislocation interaction at ambient temperature
4.3. Creep stress exponent
lead to the low rate of strain-hardening observed in this alloy.
Secondary, or steady state, creep zones experience dynamic recovery of
The creep stress exponent (n) is used to predict the creep behavior in
work hardening by a “softening” effect [46,47]. From Fig. 8, it is ob-
order to illustrate creep stability and the dominant creep mechanism
served that as the indentation peak load increases, more time is re-
during depth-sensing indentation tests [49]. Creep stress exponents
quired to start this dynamic recovery, or softening, process. As the peak
were calculated from the slope of ln ε̇ind versus ln σind curves using Eqs.
load increases, the time required to reach the peak load at a constant
(2) and (5) in the steady-state creep region. The creep stress exponent
loading rate of 10 mN/s increases. That is, sufficient time would be
depends on the indention peak load. Since the creep stress exponent (n)
available for the plastic deformation to happen (consume) in the
is greater than 3 for all specimens, the indentation creep mechanism is
loading stage (the plastic zone underneath the indenter gradually in-
attributed to dislocation creep [50].
creases as indentation depth increases). Therefore, less creep is ob-
That is, dislocation movement dominates the secondary stage of
served in the holding stage, which postpones the dynamic recovery
creep in the printed Ti-6Al-4V materials at ambient (room) tempera-
process. Considering creep rate vs indentation depth (h) curve in Fig. 9,
ture. Considering the fact that diffusion-controlled creep mechanisms
it is observed that during the loading stage of all four sample types,
are not dominant at room-temperature for the Ti-6Al-4V alloy, creep
creep rate decreases sharply with increasing indentation depth; the
occurs mainly by slip in grains favorably oriented for prism and basal
shallower the indentation depth, the higher the creep rate. This is
slip. By continuing the creep displacement, the applied load through the

43
M. Muhammad et al. Additive Manufacturing 24 (2018) 37–46

Fig. 9. Creep rate vs indentation depth for (a) CNPX-big (b) CNPX-small (c) CNPY-big (d) CNPY-small.

From Fig. 10 showing creep stress exponent (n) across different


samples for different maximum indenter loads, it is observed that, in
both small and large horizontally scanned samples, the creep stress
exponent increases as the indentation peak load increases. There is a
small increase in creep stress exponent from the peak load of 250 mN to
350 mN. When the indentation peak load is 450 mN a sudden increase
in creep stress exponent is observed in both CNPY-small and CNPY-big
samples. From the values of creep stress exponents from Fig. 10, it is
clear that CNPX (horizontally scanned samples) are more ductile than
vertically scanned samples (CNPY).
The smallest n-value of the studied Ti-6Al-4V alloy was recorded as
4.13 for the CNPY-big sample at 250 mN indentation peak load and
highest n-value was recorded as 22.94 for the CNPY-small sample at
450 mN indentation peak load. For the 250 mN and 350 mN peak loads,
n values were found to be around ∼ 4.5 to ∼ 6.5. The measured n-
values in the current study are in fairly good agreement with results
Fig. 10. Creep stress exponent (n) across different samples for different max- reported by Kral et al. [16], who examined the effect of plastic de-
imum indenter loads. formation in coarse-grained (CG) and ultrafine-grained (UFG) Ti-6Al-
4V manufactured by multiaxial forging at 648 K–698 K at stresses ran-
indenter is distributed to the adjacent grains that might be less favor- ging from 300 MPa to 600 MPa. In their study, creep stress exponent for
ably oriented for the slip. This is in agreement with Ma et al. [11] and CG alloy was reported as 17, and that of UFG found was measured to be
Liu et al. [44], who mentioned dislocation climb as the controlling approximately 4. Kral et al. [16], using a nanoindentation approach in
steady-state creep mechanism for Ti-10V-2Fe-3Al alloy and coarse- the UFG and annealed Ti-6Al-4V alloys, reported creep stress exponents
grained Ti, respectively. Although the principal mechanism causing of approximately 4 to 5.
indentation creep is dislocation slip, it has been reported that various
other dislocation behaviors, such as dislocation climb and dislocation
4.4. Indentation size effect (ISE)
nodes, dominate the secondary stage of creep under different maximum
indenter loads, leading to scattered n values [42].
In a depth-sensing indentation test, the representative strain

44
M. Muhammad et al. Additive Manufacturing 24 (2018) 37–46

Fig. 11. Indentation size effect (ISE) under different maximum indenter loads for (a) CNPX-big, (b) CNPX-small, (c) CNPY-big, (d) CNPY-small.

increases with increasing indentation size. Indentation size effect (ISE) larger the observed indentation displacement.
is defined as a phenomenon where indentation hardness or indentation
stress varies as a function of indentation depth or impression size
5. Conclusions
[51–53]. The ISE is usually attributed to geometrically necessary dis-
locations (GNDs) by “mechanism-based gradient plasticity theory” in
Ambient (room) temperature creep behavior of additively manu-
the plastic zone under indents due to strain [5,54–56]. Polishing the
factured Ti-6Al-4V materials, scanned with different print parameters
deformation layer at sub-surface levels may also contribute to in-
(different scan size and scan direction), was investigated by a depth-
dentation size effect. The GNDs are generated in the material to ac-
sensing indentation testing technique based on dislocation mechanisms.
commodate the lattice rotation due to indenter penetration. This cre-
The following conclusions were drawn from the results of this study:
ates extra dislocation in comparison to uniformly strained material in a

• This work has established that the “safe” Ti-6Al-4V alloys are sus-
very small region just below the indenter. Thus, GNDs may be oriented
along non-easy slip crystals, resulting in fundamentally different burger
ceptible to creep phenomenon even at ambient temperature.
• Creep parameters (creep rate and creep stress exponent) are de-
vectors and mobility acting as barriers to ordinary dislocations. Large
amounts of GNDs collectively results in strain gradient or work hard-
pendent upon indentation load and time during constant-load
ening underneath the indenter. Higher indentation loading rate and
holding stage.

density of GNDs result in large strain gradients and work hardening
Dislocation movement (power-law creep) dominates the secondary
effect. Voyiadjis et al. [57] performed indentation experiments on
stage of creep in the printed Ti-6Al-4V materials at ambient (room)
various single and polycrystalline materials and came to the conclusion
temperature.

that indentation depth, temperature, and deformation rate all plays
A clear indentation size effect (ISE response) primarily attributed to
vital roles in the strain gradient. Babu et al. assessed nanomechanical
the GNDs was found in the indentation stress versus indentation
behavior for traditionally built Ti-6Al-4V alloy and found strong in-
depth graphs.

dentation size effects with gradual decreases of nanohardness and
Vertically scanned samples experienced slightly more creep rate
Young’s modulus with the increase of indentation depth [58]. Fig. 11,
than their horizontally scanned counter parts.
which shows the indentation stress vs indentation displacement curve.
Strong indentation size effects are observed for the additively manu-
factured Ti-6Al-4V alloy. In all four sample types with different scan Acknowledgements
sizes and scan directions, the higher the indentation peak load, the
Authors greatly acknowledge fundingprovided by NDEPSCoR

45
M. Muhammad et al. Additive Manufacturing 24 (2018) 37–46

(Grant # 21727). This work was also partially supported by the (2006) 5489–5499.
National Science Foundation under Grant # 1657195. [31] L. Shen, W.C.D. Cheong, Y.L. Foo, Z. Chen, Nanoindentation creep of tin and alu-
minium: A comparative study between constant load and constant strain rate
methods, Mater. Sci. Eng. A 532 (2012) 505–510.
References [32] H.R. Voorhees, M. Prager, Assessment And Use Of Creep-Rupture Properties, ASM
Handb. Mech. Test. Eval. 8 (2000) 383–397.
[33] W.H. Miller, R.T. Chen, E.A. Starke, Microstructure, creep, and tensile deformation
[1] C. Leyens, M. Peters, Titanium and Titanium Alloys, (2003).
in Ti-6Al-2Nb-1Ta-0.8Mo, Metall. Trans. A 18 (8) (1987) 1451–1468.
[2] B.V. Krishna, W. Xue, S. Bose, A. Bandyopadhyay, Engineered porous metals for
[34] B. Baufeld, O. Van Der Biest, R. Gault, Microstructure of Ti-6Al-4V specimens
implants, JOM 60 (5) (2008) 45–48.
produced by shaped metal deposition, Int. J. Mater. Res. 100 (11) (2009)
[3] I.H. Oh, N. Nomura, N. Masahashi, S. Hanada, Mechanical properties of porous
1536–1542.
titanium compacts prepared by powder sintering, Scr. Mater. 49 (12) (2003)
[35] C. de Formanoir, S. Michotte, O. Rigo, L. Germain, S. Godet, Electron beam melted
1197–1202.
Ti-6Al-4V: microstructure, texture and mechanical behavior of the as-built and
[4] M. Niinomi, Mechanical biocompatibilities of titanium alloys for biomedical ap-
heat-treated material, Mater. Sci. Eng. A 652 (2016) 105–119.
plications, J. Mech. Behav. Biomed. Mater. 1 (1) (2008) 30–42.
[36] F. Wang, S. Williams, P. Colegrove, A.A. Antonysamy, Microstructure and me-
[5] R.R. Boyer, An overview on the use of titanium in the aerospace industry, Mater.
chanical properties of wire and arc additive manufactured Ti-6Al-4V, Metall. Mater.
Sci. Eng. A 213 (1996) 103–114.
Trans. A Phys. Metall. Mater. Sci. 44 (2) (2013) 968–977.
[6] S. Bremen, W. Meiners, A. Diatlov, Selective laser melting. A manufacturing tech-
[37] M. Neikter, P. Åkerfeldt, R. Pederson, M.L. Antti, Microstructure characterisation of
nology for the future? Laser Tech. J. 9 (2012) 33–38.
Ti-6Al-4V from different additive manufacturing processes, IOP Conference Series:
[7] R. Filip, K. Kubiak, W. Ziaja, J. Sieniawski, The effect of microstructure on the
Materials Science and Engineering 258 (1) (2017).
mechanical properties of two-phase titanium alloys, J. Mater. Process. Technol. 133
[38] M. Haghshenas, O. Totuk, M. Masoomi, S.M. Thompson, N. Shamsaei, Small-scale
(1–2) (2003) 84–89.
mechanical properties of additively manufactured Ti-6Al-4V, Solid Freeform
[8] L. Ladani, Local and global mechanical behavior and microstructure of Ti6Al4V
Fabrication Symposium, (2017).
parts built using Electron beam melting technology, Metall. Mater. Trans. A Phys.
[39] A. Łukaszek-Sołek, et al., Microstructure and phase transformation of Ti-6Al-4V,
Metall. Mater. Sci. 46 (9) (2015) 3835–3841.
Mater. Sci. Eng. A 2 (1–2) (2013) 27–30.
[9] C.A. Schuh, Nanoindentation studies of materials, Mater. Today 9 (5) (2006) 32–40.
[40] M. Hardiman, T.J. Vaughan, C.T. McCarthy, The effects of pile-up, viscoelasticity
[10] A.C. Fischer-Cripps, Nanoindentation, Nanoindentation (2011) 21–38.
and hydrostatic stress on polymer matrix nanoindentation, Polym. Test. 52 (2016)
[11] X. Ma, et al., Indenter load effects on creep deformation behavior for Ti-10V-2Fe-
157–166.
3Al alloy at room temperature, J. Alloys. Compd. 709 (March) (2017) 322–328.
[41] A. Bolshakov, G.M. Pharr, Influences of pileup on the measurement of mechanical
[12] J. Kumar, S.G.S. Raman, V. Kumar, Creep–Fatigue interactions in Ti-6Al-4V alloy at
properties by load and depth sensing indentation techniques, J. Mater. Res. 13 (4)
ambient temperature, Trans. Indian Inst. Met. 69 (2) (2016) 349–352.
(1998) 1049–1058.
[13] T. Matsunaga, T. Kameyama, K. Takahashi, E. Sato, Intragranular deformation
[42] W.B. Li, J.L. Henshall, R.M. Hooper, K.E. Easterling, The mechanisms of indentation
mechanisms during ambient-temperature creep in hexagonal close-packed metals,
creep, Acta Metall. Mater. 39 (12) (1991) 3099–3110.
Mater. Trans. 50 (12) (2009) 2865–2872.
[43] B.G. Yoo, K.S. Kim, J.H. Oh, U. Ramamurty, J. Il Jang, Room temperature creep in
[14] V. Hasija, S. Ghosh, M.J. Mills, D.S. Joseph, Deformation and creep modeling in
amorphous alloys: Influence of initial strain and free volume, Scr. Mater. 63 (12)
polycrystalline Ti-6Al alloys, Acta Mater. 51 (15) (2003) 4533–4549.
(2010) 1205–1208.
[15] M.J.R. Barboza, et al., Creep behavior of Ti-6Al-4V and a comparison with titanium
[44] X. Liu, Q. Zhang, X. Zhao, X. Yang, L. Luo, Ambient-temperature nanoindentation
matrix composites, Mater. Sci. Eng. A 428 (1–2) (2006) 319–326.
creep in ultrafine-grained titanium processed by ECAP, Mater. Sci. Eng. A 676
[16] P. Kral, J. Dvorak, S. Zherebtsov, G. Salishchev, M. Kvapilova, V. Sklenicka, Effect
(2016) 73–79.
of severe plastic deformation on creep behaviour of a Ti-6Al-4V alloy, J. Mater. Sci.
[45] Y. Ma, G.J. Peng, D.H. Wen, T.H. Zhang, Nanoindentation creep behavior in a
48 (13) (2013) 4789–4795.
CoCrFeCuNi high-entropy alloy film with two different structure states, Mater. Sci.
[17] L. Badea, M. Surand, J. Ruau, B. Viguier, Creep behavior of Ti-6Al-4V FROM 450 ° C
Eng. A 621 (2015) 111–117.
TO 600 ° C, U.P.B. Sci. Bull. 76 (2014) 185–196.
[46] P. De Hey, J. Sietsma, A. Van Den Beukel, Structural disordering in amorphous Pd
[18] W.J. Evans, G.F. Harrison, Power law steady state creep in α/β titanium alloys, J.
40 Ni 40 P 20 induced by high temperature deformation, Acta Mater. 46 (16)
Mater. Sci. 18 (1983) 3449.
(1998) 5873–5882.
[19] E. Sato, T. Yamada, H. Tanaka, I. Jimbo, Categorization of ambient temperature
[47] C. Wang, Q.P. Cao, X.D. Wang, D.X. Zhang, S.X. Qu, J.Z. Jiang, Time-dependent
creep behavior of metals and alloys on their crystallographic structures, J. Japan
shear transformation zone in thin film metallic glasses revealed by nanoindentation
Inst. Light Met. 55 (11) (2005) 604–609.
creep, J. Alloys. Compd. 696 (2017) 239–245.
[20] T. Kameyama, T. Matsunaga, E. Sato, K. Kuribayashi, Suppression of ambient-
[48] T. Bose, M. Rattan, Effect of thermal gradation on steady state creep of functionally
temperature creep in CP-Ti by cold-rolling, Mater. Sci. Eng. A 510–511 (C) (2009)
graded rotating disc, Eur. J. Mech. A/Solids 67 (2018) 169–176.
364–367.
[49] R. Schwaiger, B. Moser, M. Dao, N. Chollacoop, S. Suresh, Some critical experiments
[21] W.J. Harrison, M.T. Whittaker, R.J. Lancaster, A model for time dependent strain
on the strain-rate sensitivity of nanocrystalline nickel, Acta Mater. 51 (17) (2003)
accumulation and damage at low temperatures in Ti-6Al-4V, Mater. Sci. Eng. A 574
5159–5172.
(2013) 130–136.
[50] M.W. Barsoum, Fundamentals of ceramics, Vasa (2003) 622.
[22] W.J. Evans, Time dependent effects in fatigue of titanium and nickel alloys, Fatigue
[51] S. Ando, K. Nakamura, K. Takashima, H. Tonda, {1122} < 1123 > slip in magne-
Fract. Eng. Mater. Struct. 27 (7) (2004) 543–557.
sium single crystal, Japan Inst. Light Met. 42 (12) (1992) 765–771.
[23] H. Tanaka, T. Yamada, E. Sato, I. Jimbo, Distinguishing the ambient-temperature
[52] M. Haghshenas, Y. Wang, Y.T. Cheng, M. Gupta, Indentation-based rate-dependent
creep region in a deformation mechanism map of annealed CP-Ti, Scr. Mater. 54 (1)
plastic deformation of polycrystalline pure magnesium, Mater. Sci. Eng. A 716
(2006) 121–124.
(November (2017)) (2018) 63–71.
[24] T. Neeraj, D.H. Hou, G.S. Daehn, M.J. Mills, Phenomenological and microstructural
[53] N.A. Stelmashenko, M.G. Walls, L.M. Brown, Y.V. Milman, Microindentations on W
analysis of room temperature creep in titanium alloys, Acta Mater. 48 (6) (2000)
and Mo oriented single crystals: an STM study, Acta Metall. Mater. 41 (10) (1993)
1225–1238.
2855–2865.
[25] M. Masoomi, J.W. Pegues, S.M. Thompson, N. Shamsaei, A numerical and experi-
[54] Q. Ma, D.R. Clarke, Size dependent hardness of silver single crystals, J. Mater. Res.
mental investigation of convective heat transfer during laser-powder bed fusion,
10 (4) (1995) 853–863.
Addit. Manuf. 22 (2018).
[55] M.F. Ashby, The deformation of plastically non-homogeneous materials, Philos.
[26] E. Broitman, Indentation hardness measurements at macro-, micro-, and nanoscale:
Mag. 21 (170) (1970) 399–424.
a critical overview, Tribol. Lett. 65 (1) (2017).
[56] Na. Fleck, G.M. Muller, M.F. Ashby, J.W. Hutchinson, Strain gradient plasticity:
[27] W.H. Poisl, W.C. Oliver, B.D. Fabes, The relationship between indentation and
theory and experiment, Acta Metall. Mater. 42 (2) (1994) 475–487.
uniaxial creep in amorphous selenium, J. Mater. Res. 10 (8) (1995) 2024–2032.
[57] G.Z. Voyiadjis, D. Faghihi, Microstructure to macro-scale using gradient plasticity
[28] B.N. Lucas, W.C. Oliver, Indentation power-law creep of high-purity indium,”
with temperature and rate dependent length scale, Procedia IUTAM vol. 3, (2012)
Metall, Mater. Trans. A 30 (March) (1999) 601–610.
205–227.
[29] T.G. Langdon, F.A. Mohamed, A simple method of constructing an Ashby-type de-
[58] B. SridharBabu, A. Kumaraswamy, B. AnjaneyaPrasad, Effect of indentation size
formation mechanism map, J. Mater. Sci. 13 (6) (1978) 1282–1290.
and strain rate on nanomechanical behavior of Ti-6Al-4V alloy, Trans. Indian Inst.
[30] R. Goodall, T.W. Clyne, A critical appraisal of the extraction of creep parameters
Met. 68 (1) (2014) 143–150.
from nanoindentation data obtained at room temperature, Acta Mater. 54 (20)

46

View publication stats

You might also like