mdp2 138

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Received: 30 November 2019 Accepted: 30 December 2019

DOI: 10.1002/mdp2.138

SPECIAL ISSUE ARTICLE

Low temperature stress relief and martensitic


decomposition in selective laser melting produced Ti6Al4V

Gerrit Matthys Ter Haar | Thorsten Hermann Becker

Materials Engineering group, Department


of Mechanical and Mechatronic
Abstract
Engineering, Stellenbosch University, This study contributes to the limited understanding of the initial stages of mar-
South Africa
tensite decomposition and stress relief of SLM-produced Ti6Al4V. Samples are
Correspondence annealed at a temperature range of 427–610 C for hold times ranging from
Gerrit Matthys Ter Haar, Materials 5 minutes and 30 hours. Stress relief is measured through x-ray diffraction,
Engineering group, Department of
while phase transformation is observed using microscopy and x-ray diffraction
Mechanical and Mechatronic
Engineering, Stellenbosch University, techniques. Tensile tests and hardness measurements are used to obtain
South Africa. mechanical properties. Key findings of this study are that ultra-fine vanadium-
Email: gterhaar@sun.ac.za
rich precipitates form at α’ dislocations and grain boundaries during prelimi-
Funding information nary phase transformation. While initial formation of vanadium-rich precipi-
Department of Science and Innovation tates results in the embrittlement of the material, further growth of these
through the Collaborative Program for
Additive Manufacturing (CPAM) precipitates into measurable β-phase is extremely slow but allows for an
improvement of material ductility. The study reveals that phase transformation
is incomplete after 30 hours at 610 C. While an increase in annealing tempera-
ture accelerates phase transformation and improves ductility, this comes at a
cost of material strength.

KEYWORDS
Metal additive manufacturing, Ti6Al4V, heat treatments, tensile properties

1 | INTRODUCTION

SLM has been successfully applied to a range of metal alloys, including Ti6Al4V. This two-phase (α + β) Ti-alloy is pop-
ular in the Ti-industry due to its high specific strength and high-temperature corrosion resistance. Ti6Al4V is also a
material of choice for medical implants due to its biocompatibility. Traditionally processed (i.e. thermo-mechanically
process) Ti6Al4V can achieve up to ~1095 MPa ultimate tensile strength (UTS) with a ~16% fracture elongation.1
SLM-produced Ti6Al4V (without post-processing), on the other hand, achieves ~1200 MPa UTS with only ~4–8%
fracture elongation2 which is below the required ASTM F2924–14 limit of 10%. The poor ductility achieved by SLM var-
iant is argued to be due to the meta-stable martensitic microstructure (denoted α’) which forms as a result of the fast-
cooling process inherent in the SLM process. This high rate of thermal fluctuation of local material in SLM also causes
high residual stress which cause in-situ delamination, part warpage and potential build-failure.3,4 Together, these two
unfavourable properties have shown to decrease part toughness and fatigue life5-7 thereby stimulating on-going
research towards improving material quality. Altering process parameters such as the scan strategy and hatch distance
has a limited effect on reducing residual stress.4,8 Recent work into “intrinsic heat treatments” has shown the possibility
of decomposing martensite in-situ although this comes at a cost of a much longer build time.9,10 In-situ build plate
heating has also shown potential in reducing residual stress.11,12 Currently, the more practical, cost effective and

Mat Design Process Comm. 2021;3:e138. wileyonlinelibrary.com/journal/mdp2 © 2020 John Wiley & Sons, Ltd. 1 of 6
https://doi.org/10.1002/mdp2.138
2 of 6 TER HAAR AND BECKER

conventional approach to improving part quality is through ex-situ post process heat treatments. Ter Haar and Becker13
demonstrated the diffusion controlled α’ martensite decomposition process in the sub-β-transus range of 750–960 C
resulting in a α/β lamellar microstructure. While an increase in ductility is achieved, this comes at a cost of material
strength due to coarsening of α plates following a Hall–Petch relationship.14 Lower temperature annealing limits α
grain growth and thereby a higher material strength.
Little is understood about the nature of phase transformation and material property change at low temperature
heat-treatments (aging) of martensite.1 Popular heat-treatments aimed at stress relief are set between 480–650 C for
1–4 hours.1,5 It is, however, likely that these heat-treatments alter the microstructure in addition to relieving resid-
ual stress. One concern when annealing at low temperatures is the possibility of forming a brittle intermetallic
Ti3Al. This paper aims to improve the general understanding of stress relief and microstructure change at low tem-
peratures (480–610 C) in order to apply this knowledge to achieve a higher quality part (i.e. toward achieving an
optimal strength and ductility). This understanding will further contribute to an improved approach to in-situ base-
plate heating.

2 | EXPERIMENTAL M ETHODOLOGY

Grade 23 Ti6Al4V powder was acquired from TLS Technik GmbH & Co with a particle size of ~20 ~50 μm in diameter.
Powder chemical composition was determined to be (in weight): 90% Ti, 6.08% Al, 3.88% V 0.17% Fe, 0.023% N and
0.09% O. Samples were manufactured using standard process parameters on an EOS M280 machine in an argon
environment.
Heat-treatments were performed in a 5 kW Gallenkamp box furnace. Samples were sectioned with a slow-cutting
diamond-blade machine. Samples intended for scanning electron microscopy (SEM), hardness measurements and x-ray
diffraction were prepared using standard recommended Buehler metallography procedures and equipment for titanium.
Samples intended for Scanning Transmission Electron Microscopy (STEM) were cut using fine ion beam (FIB) milling
in a FEI-HELIOS-Nanolab-650.
Samples were imaged with both electron backscatter diffraction (EBS) in SEM and high-angle annular dark-field
(HAADF) imaging in STEM. Energy-dispersive x-ray spectroscopy (EDS) was used for elemental composition measure-
ments. Electron energy loss spectroscopy (EELS) analysis was used in conjunction with STEM to map elemental com-
position at a higher resolution.
For residual stress measurements, the sin2ψsin2ψ method was followed in accordance to the standard BS
EN15305:2008. Samples were analyses directly after heat treating to avoid erroneous stress states caused by mechanical
surface polishing. A Bruker D8 Advanced XRD machine with a CuKα (1.541838 Å) radiation source set at 40 kV and
40 mA was used. Eleven ψ angles were scanned from −45 to +45 . A 2θ Peak at ~142 (HKL plane 213) was chosen
for stress analysis. The 2θ step size was set at 0.02 and the count time per frame was 2.1 seconds. Data process soft-
ware Diffrac. Leptos 7 was used together with a biaxial stress model.
For XRD phases analysis, wide angle 2θ scans were done with a Bruker D2 PHASER X-ray diffraction machine. The
x-ray generator was set at 30 kV and 10 mA. Samples was scanned with a CuKα (1.541838 Å) radiation source. A hold
time of 0.75 seconds at 0.02 step size. Rietveld refinement was done on certain samples using and software MAUD.15
Tensile tests specimens where machined to a sample geometry recommended by ASTM E8. Sample gauge diameter
of 5 mm and a gauge length of 25 mm. Tensile tests were carried out using an MTS Criterion Model 44 machine at a
cross-head displacement rate of 0.13 mm/min. Micro-indentations were carried out using an Emcotest DuraScan vickers
indentation machine. A diamond-shaped indenter was used, and measurements were carried out according to ASTM
standard E348–10. Load application was 2 kilograms force for 10 seconds.

3 | R E SUL T S

Results of stress relief and microhardness profiles are plotted in Figure 1 (a) and (b) respectively. While a slow rate of
stress relief is observed for temperatures 427 and 480 C, samples annealed at 560 and 610 C showed a high rate of
stress relief in the first 5 minutes. Longer anneal times caused surface oxidation which inhibited clear XRD measure-
ments. All microhardness profiles show an increase in microhardness followed by a slow decrease. The highest peak in
hardness is observed after 1 hours for samples 480 C with a lower peak for 560 C followed by a very gradual degrease
TER HAAR AND BECKER 3 of 6

FIGURE 1 (a) Magnitude of residual stress and (b) microhardness for the annealed samples as a function of time (AB – as built)

for both samples. The sample annealed at 610 C peaked after only 5 minutes and then stabilised at ~350 HV2 after
8 hours.
The STEM analysis depicts the relative microstructural change with time. Figure 2 (a-c) depicts the gradual pre-
cipitation and growth of a heavier phase (light-contrast regions indicated by arrows) at α/α’ grain boundaries and
dislocations. This phase becomes large enough after 30 hours to map with EELS. The result thereof is depicted in
Figure 2 (d & e). Since the light-contrast phase is vanadium-rich, it can be argued to be body-centred cubic (bcc)
β-phase. A reference sample consisting of a stable α + β phase was used as comparison. This reference sample was
annealed at 1000 C for 2 hours followed by slow furnace cooling. EDS showed the reference sample's α-phase con-
sisting of (in weight): 91.2% Ti, 6.8% Al and 2.2% V and the β-phase: 80.1% Ti, 2.5% Al and 17.4% V. EELS mapping
of sample “480 C_30hours” showed similar levels of vanadium concentration in the β-phase; however, a much
higher aluminium concentration in β-phase and a much lower aluminium concentration in the α-phase.

FIGURE 2 HAADF - STEM images of sample 480 C anneal different hold times: arrow point to vanadium enriched regions at
(a) 5 minutes and, (b) 1 hour and (c) 30 hours. (d) EELS mapped region of vanadium (warm-coloured regions), (e) spot and area mappings
of elemental weight analysis
4 of 6 TER HAAR AND BECKER

FIGURE 3 SEM (BSD) images of samples annealed for 30 hours at (a) 480 C (b) 560 C (c) 610 C

Furthermore, the vanadium concentration of this sample was also much higher in the α-phase than that of the
same phase in the reference sample. It should further be noted that no intermetallic Ti3Al precipitation was
observed in any sample through STEM.
In order to track the β-phase growth with temperature, Figure 3 (a-c) illustrates SEM-BSD micrographs of samples
annealed for 30 hours at 480, 560, and 610 C respectively. A clear growth of β phase along α/α’ grain boundaries is
observed. Change in lattice parameters with annealing temperature (30 hours hold time) was obtained from Rietveld
refinement in MAUD of XRD broad-angle data.
Figure 4 (a) depicts a decrease in (0002) peak width with temperature. This indicates a decrease in lattice distortions
caused by dislocations, grain size and residuals stress. Figure 4 (b) depicts the change in hcp c/a lattice parameter ratio
as well as bcc “a” lattice parameter with temperature (no bcc peaks were observed for samples annealed at tempera-
tures 427 and 480 C). These changes are argued to be caused by the change in chemical composition of the phases due
to atom-diffusive driven phase transformation. The” ?” in Figure 4 represents the measured hcp lattice ratio for the hcp
parameters of the reference sample while the “▲” represents the bcc “a” parameter for the reference sample. Volume
fraction of β phase in the reference sample was obtained through image processing to be approximately 10% while the
β-phase volume fraction of sample “610 C_30hours” measured approximately 5–6%.
Tensile test results of selected heat treatments are plotted in Figure 5. As anticipated from the hardness results, the
samples annealed at 480 and 560 C for 1 hours showed an increase in strength at a cost of ductility. Sample
“610 C_1hour” measured the same hardness at the as-built group, however it showed a decrease in strength (relative to
as-built) and a decrease in ductility. Further increase in hold time at 610 C of 8 hours promoted an increase in ductility
with a slight loss of strength.

4 | DISCUSSION

Results of this study document the initial stages of martensite decomposition through the nucleation and growth of
ultra-fine, vanadium-rich precipitates. The fine precipitation process has a clear strengthening and embrittlement
effect due to the increase amount of slip deflection/impediment without a significant contributing to the material's
ability to plastically deform. Whether the precipitates are bcc phase or a transitional “pre-bcc” phase (due to a possi-
ble spinodal rather than classical nucleation-and-growth phase transformation) is not clear and require further
investigation. Once vanadium-rich precipitates have formed, a gradual martensite decomposition phase transforma-
tion (α’àα + β) proceeds through a slow diffusion of vanadium to the β-phase and aluminium to the α-phase. By

F I G U R E 4 XRD results
(a) change in hcp (0002) peak
width and (b) hcp c/a ratio
(black) and bcc”a” lattice
parameter (red). (Star and
triangle indicates the reference
sample's results)
TER HAAR AND BECKER 5 of 6

F I G U R E 5 Stress vs elongation to fracture (circles are plotted


at 2 standard deviations)

comparing the elemental composition, lattice parameters and β-phase volume fraction of the annealed samples with
that of the stable α + β reference sample, it can be deduced that the phase transformation is incomplete. Whereas
the initial stage of the phase transformation embrittles the material, a progression in the transformation allows for a
growth in bcc phase fraction which has the effect of increasing material ductility. While stress relief and martensite
decomposition do occur at temperatures as low as 427 C, this is at an extremely slow rate. Increasing annealing
temperature results in an increased rate of phase transformation (and therefore a ductile β-phase) but promotes
α-grain growth thereby reducing strength. While clear β phase formation and growth was observed, the martensite
decomposition is incomplete after 30 hours at 610 C and further growth in β-phase volume fraction is needed to
improve material ductility. Annealing at a slightly higher temperature such as 650 C might achieve full phase trans-
formation in less than 30 hours thereby providing a more practical cost saving option to optimal mechanical proper-
ties in α’ phase decomposition.

5 | C ON C L U S I ON S

The following conclusions are drawn:

• The initial stages of martensite decomposition were shown to occur at very low temperatures through fine β-particles
which caused an embrittlement of the material.
• The relief of residual stress increases with temperature with a temperature at 560 C and above relieving ~90% of the
residual stress.
• The martensite decomposition is incomplete after 30 hours at 610 C when comparing the sample's lattice parameters
and alloy composition to that of the reference sample.
• While annealing at 610 C for 1 hour achieves adequate stress relief, ductility was shown to only improve after
8 hours. It is therefore advised that samples should be annealed for longer than 8 hours at 610 C or possible at a
slightly higher temperature of 650 C. This would allow for an adequate material ductility with minimal loss in
strength.

ACK NO WLE DGE MEN TS


The authors acknowledge the powder analysis data from Central University of Technology and gratefully the funding
from the Department of Science and Innovation through the Collaborative Program for Additive Manufacturing
(CPAM).

ORCID
Gerrit Matthys Ter Haar https://orcid.org/0000-0002-9520-828X
Thorsten Hermann Becker https://orcid.org/0000-0002-4089-3574
6 of 6 TER HAAR AND BECKER

R EF E RE N C E S
1. Donachie M. Titanium A Technical Guide. 2nd ed. Ohio: ASM; 2000.
2. Yan M, Yu P. Chapter 5 - An Overview of Densification, Microstructure and Mechanical Property of Additively Manufactured Ti-6Al-
4 V — Comparison among Selective Laser Melting, Electron Beam Melting, Laser Metal Deposition and Selective Laser Sintering, and
with Convent. In: Lakshmanan A, ed. Sintering Techniques of Materials. London: InTech; 2015:76-106 doi:10.5772/59275.
3. DebRoy T, Wei HL, Zuback JS, et al. Additive manufacturing of metallic components – Process, structure and properties. Prog Mater Sci.
2018;92:112-224. https://doi.org/10.1016/j.pmatsci.2017.10.001
4. Anderson LS. Evaluating measurement techniques: establishing a testing framework for residual stress in selective laser melted Ti-6Al-
4 V. 2017.
5. Vrancken B, Cain V, Knutsen R, Van Humbeeck J. Residual stress via the contour method in compact tension specimens produced via
selective laser melting. Scr Mater. 2014;87:29-32. https://doi.org/10.1016/j.scriptamat.2014.05.016
6. Leuders S, Lieneke T, Lammers S, Tröster T, Niendorf T. On the fatigue properties of metals manufactured by selective laser melting –
The role of ductility. J Mater Res. 2014;29(17):1911-1919. https://doi.org/10.1557/jmr.2014.157
7. Leuders S, Thöne M, Riemer A, et al. On the mechanical behaviour of titanium alloy TiAl6V4 manufactured by selective laser melting:
Fatigue resistance and crack growth performance. Int J Fatigue. 2013;48:300-307. https://doi.org/10.1016/j.ijfatigue.2012.11.011
8. Kruth J, Badrossamay M, Yasa E, et al. Part and material properties in selective laser melting of metals. Proc SFF Symp. 2010. https://
lirias.kuleuven.be/handle/123456789/265815.
9. Barriobero-Vila P, Gussone J, Haubrich J, et al. Inducing Stable α + β Microstructures during Selective Laser Melting of Ti-6Al-4 V
Using Intensified Intrinsic Heat Treatments. Materials (Basel). 2017;10(3):268-281. https://doi.org/10.3390/ma10030268
10. Haubrich J, Gussone J, Barriobero-Vila P, et al. The role of lattice defects, element partitioning and intrinsic heat effects on the micro-
structure in selective laser melted Ti-6Al-4 V. Acta Mater. 2019;167(April):136-148. https://doi.org/10.1016/j.actamat.2019.01.039
11. Kempen K, Thijs L, Vrancken B, Buls S, Van Humbeeck J, Kruth J-P. Lowering Thermal Gradients in Selective Laser Melting by Pre-
Heating the Baseplate; 2013. https://core.ac.uk/download/pdf/34572651.pdf.
12. Vrancken B. Study of Residual Stresses in Selective Laser Melting. 2016
13. Ter Haar GM, Becker TH. Selective Laser Melting Produced Ti-6Al-4 V: Post-Process Heat Treatments to Achieve Superior Tensile Prop-
erties. Materials (Basel). 2018;11(01):146-160. https://doi.org/10.3390/ma11010146
14. Cao S, Chu R, Zhou X, et al. Role of martensite decomposition in tensile properties of selective laser melted Ti-6Al-4 V. J Alloys Compd.
2018;744:357-363. https://doi.org/10.1016/j.jallcom.2018.02.111
15. Lutterotti L, Wenk H, Matthies S. MAUD (Material Analysis Using Diffraction): a user friendly Java program for Rietveld Texture Analy-
sis and more. In: Proceeding of the Twelfth International Conference on Textures of Materials (ICOTOM-12). Vol.2 Montreal, Canada:
NRC Research Press; 1999:1599-1604.

How to cite this article: Ter Haar GM, Becker TH. Low temperature stress relief and martensitic
decomposition in selective laser melting produced Ti6Al4V. Mat Design Process Comm. 2021;3:e138. https://doi.
org/10.1002/mdp2.138

You might also like