Sunteți pe pagina 1din 15

letter doi:10.

1038/nature23894

3D printing of high-strength aluminium alloys


John H. Martin1,2, Brennan D. Yahata1, Jacob M. Hundley1, Justin A. Mayer1, Tobias A. Schaedler1 & Tresa M. Pollock2

Metal-based additive manufacturing, or three-dimensional (3D) During solidification of these unweldable alloys, the primary equi-
printing, is a potentially disruptive technology across multiple librium phase solidifies first at a different composition from the bulk
industries, including the aerospace, biomedical and automotive liquid. This results in solute enrichment in the liquid near the solidi-
industries. Building up metal components layer by layer increases fying interface, locally changing the equilibrium liquidus temperature
design freedom and manufacturing flexibility, thereby enabling and producing an unstable, undercooled condition6. As a result, there
complex geometries, increased product customization and shorter is a breakdown of the solid–liquid interface leading to cellular or den-
time to market, while eliminating traditional economy-of-scale dritic grain growth with long channels of interdendritic liquid trapped
constraints. However, currently only a few alloys, the most relevant between solidified regions. As temperature and liquid volume fraction
being AlSi10Mg, TiAl6V4, CoCr and Inconel 718, can be reliably decrease, volumetric solidification shrinkage and thermal contraction
printed1,2; the vast majority of the more than 5,500 alloys in use in these channels produces cavities and hot tearing cracks which may
today cannot be additively manufactured because the melting span the entire length of the columnar grain and can propagate through
and solidification dynamics during the printing process lead to additional intergranular regions7,8 (Fig. 1e).
intolerable microstructures with large columnar grains and periodic In contrast, fine equiaxed microstructures more easily accommodate
cracks3–5. Here we demonstrate that these issues can be resolved by strain in the semi-solid state by suppressing coherency that locks the
introducing nanoparticles of nucleants that control solidification orientation of these solid dendrites and promotes tearing9. Producing
during additive manufacturing. We selected the nucleants on the these ideal equiaxed structures requires large amounts of undercool-
basis of crystallographic information and assembled them onto 7075 ing, which has thus far proven difficult in additive processes where
and 6061 series aluminium alloy powders. After functionalization high thermal gradients arise from rastering of a direct energy source in
with the nucleants, we found that these high-strength aluminium an arbitrary geometric pattern10. Here we present a general approach
alloys, which were previously incompatible with additive to control solidification microstructure by promoting nucleation of
manufacturing, could be processed successfully using selective new grains with nanoparticle grain refiners (Fig. 1d). Alloy powder
laser melting. Crack-free, equiaxed (that is, with grains roughly feedstock particles are decorated with lattice-matched nanoparticles
equal in length, width and height), fine-grained microstructures (Fig. 1b) that heterogeneously nucleate the primary equilibrium phases
were achieved, resulting in material strengths comparable to during cooling of the melt pool. By providing a high density of low-
that of wrought material. Our approach to metal-based additive energy-barrier heterogeneous nucleation sites ahead of the solidifi-
manufacturing is applicable to a wide range of alloys and can be cation front, the critical amount of undercooling needed to induce
implemented using a range of additive machines. It thus provides equiaxed growth is decreased11. This allows for a fine equiaxed grain
a foundation for broad industrial applicability, including where structure that accommodates strain and prevents cracking under oth-
electron-beam melting or directed-energy-deposition techniques erwise identical solidification conditions. Using this technology enables
are used instead of selective laser melting, and will enable additive additive manufacturing of previously unattainable high-performance
manufacturing of other alloy systems, such as non-weldable nickel alloys, such as 7075 or 6061 aluminium, with improved properties over
superalloys and intermetallics. Furthermore, this technology could currently available systems.
be used in conventional processing such as in joining, casting and Aluminium alloys are a good demonstration platform for our
injection moulding, in which solidification cracking and hot tearing approach, because the only printable aluminium alloys are based on the
are also common issues. binary Al–Si system and have a wide range of reported properties, but
In metal-based additive manufacturing, application of a direct tend to converge around a yield strength of approximately 200 MPa with
energy source, such as a laser or electron beam, to melt alloy powders a low ductility of 4% (refs 1, 12). The exception is Scalmalloy13,14, which
locally results in solidification rates between 0.1 m s−1 and 5 m s−1, relies on alloying additions of scandium, a rare high-cost metal. In con-
an order of magnitude increase over conventional casting processes. trast, most aluminium alloys used in automotive, aerospace and con-
Given that rastering of this direct energy source (such that it follows sumer applications are wrought alloys of the 2000, 5000, 6000 or 7000
a pattern of slightly overlapping lines in a back and forth pattern) to series, which can exhibit strengths exceeding 400 MPa and ductility of
continuously fuse successive layers of powder is analogous to welding more than 10% but cannot currently be additively manufactured15–17.
processes, it is not surprising that the suite of printable metal alloys These systems have low-cost alloying elements (Cu, Mg, Zn, Si)
are limited to those known to be easily weldable. Application of con- carefully selected to produce complex strengthening phases during
ventional 3D printing methods to ‘unweldable’ high-performance subsequent ageing. These same elements promote large solidification
engineering alloys that cannot accommodate these solidification ranges, leading to hot tearing during solidification, an issue that has
conditions, such as 6000 and 7000 series aluminium alloys and been difficult to surmount for the more than 100 years since the first
high-strengthening-phase (high-γ′) nickel superalloys, results in age-hardenable alloy, duralumin, was developed18,19. The most com-
microstructures with columnar grains and cracks spanning dozens plete study of elemental effects dates back to the late 1940s; however,
of successive print layers3–5. The limitations of the currently printable the mechanistic effect was not fully described until 1999 by Rappaz,
alloys, especially with respect to specific strength, fatigue life and frac- Drezet and Gremaud (RDG)8,20. The RDG model incorporated both
ture toughness, have hindered metal-based additive manufacturing deformation of the semi-solid network and fluid backfill to capture the
from maturing to its full potential. composition and microstructure effects on cavitation-assisted tearing.

1
HRL Laboratories LLC, Sensors and Materials Laboratory, Architected Materials Department, Malibu, California, USA. 2Materials Department, University of California, Santa Barbara, California, USA.

2 1 s e p t e m b e r 2 0 1 7 | V O L 5 4 9 | N A T U RE | 3 6 5
© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
RESEARCH letter

a Standard
a alloy powder b Nanoparticle-enhanced g
powder

Laser or
electron beam

50 μm
15 μm
c Columnar growth Melt pool
d Equiaxed growth
Crack

Powder bed

Shrinkage 2 cm
Strain-tolerant
e f h

z 111
z
y 15 μm
35 μm y
x 100 110
x
Large columnar grains with cracks 1 cm
Crack-free, small equiaxed grains

Figure 1 | Additive manufacturing of metal alloys via selective laser d, Suitable nanoparticles can induce heterogeneous nucleation and facilitate
melting. The central schematic represents an overview of the additive equiaxed grain growth, thereby reducing the effect of solidification
manufacturing process, whereby a direct energy source (laser or electron strain. e, Many alloys exhibit intolerable microstructure with large grains
beam) melts a layer of metal powder (yellow), which solidifies (red to blue), and periodic cracks when 3D-printed using conventional approaches,
fusing it to the previous (underlying) layer of metal (grey). a, Conventional as illustrated by the inverse pole figure. f, Functionalizing the powder
Al7075 powder feedstock. b, Al7075 powder functionalized with feedstock with nanoparticles produces fine equiaxed grain growth and
nanoparticles. c, Many alloys including Al7075 tend to solidify by columnar eliminates hot cracking. g, A 3D-printed, topologically optimized Al6061
growth of dendrites, resulting in cracks due to solidification shrinkage. piston on the build plate. h, 3D-printed Al7075 HRL logo.

Additionally, Gourlay and Dahle21 demonstrated experimentally that the instant of melting provides a high level of mixing and a high density
strain can be accommodated more readily in a fine equiaxed material of nucleation sites.
owing to an increase in the solid fraction at which dendrite coherency Pre-alloyed gas-atomized 7075 and 6061 spherical powders with
occurs and the suppression of large dilatant shear bands which require an average particle size of 45 μ​m were coated with 1 vol% hydrogen-
additional backfilling. Combining the mechanistic effects addressed by stabilized zirconium nucleants using an electrostatic assembly tech-
Gourlay and Dahle and the predictions of the RDG model to minimize nique to ensure uniform distribution in the powder bed and avoid
crack susceptibility has not been effective for highly crack-susceptible settling. Assembled powders were additively manufactured via selec-
alloys such as Al7075 and Al6061, owing to a lack of processing paths tive laser melting using a Concept Laser M2 400W system with an
to produce fine equiaxed grains. We have developed a scalable and 80 mm ×​ 80 mm build volume. Standard machine parameters provided
alloy-agnostic approach to incorporate grain-refining particles into by the manufacturer for a conventional AlSi10Mg alloy were used for
conventional hot-tear-susceptible alloy powders directly to additively all nanoparticle-functionalized 7075 and 6061 powders. After com-
manufacture high-strength crack-free alloys with a fine equiaxed pletion of the build, components were homogenized on the build plate
microstructure (Fig. 2). Conventional alloy powders and nanopar- and aged to a T6 condition in accordance with conventional wrought
ticles are electrostatically assembled, producing a powder feedstock materials. For direct property comparison, parts were also manufac-
with uniformly distributed nanoparticles. Nanoparticle compositions tured from stock 7075, 6061 and AlSi10Mg powders under the same
targeted to each alloy were selected using a software tool that identi- conditions.
fies matching crystallographic lattice spacing and density to provide a Microstructure analysis reveals a substantial difference between com-
low-energy nucleation barrier on the basis of classical nucleation theory ponents additively manufactured from stock powders and those pro-
(Fig. 2f). The software analysed more than 4,500 different powder and duced with nanoparticle-functionalized powder (Figs 1 and 3). Stock
nanoparticle combinations corresponding to more than 11.5 million 7075 and 6061 (Extended Data Figs 6 and 7) exhibit a series of large
matching pairs. Potential matches were sorted by a combined set of columnar grains oriented parallel to the build direction, with cracks
constraints: minimized lattice misfit, similar atomic packing along present in the intercolumnar region and extending through multiple
matched crystallographic planes, thermodynamic stability in the build layers. This is consistent with previously documented attempts
desired alloy, and availability. For the aluminium alloys tested, hydro- at printing wrought aluminium alloys and is driven by the high, direc-
gen-stabilized zirconium particles were selected for their stability in tional heat flux in the additive process, which provides high thermal
air and ability to decompose at the melting temperature, resulting in gradients and minimal undercooling during solidification4. Previous
formation of the favourable Al3Zr nucleant phase22. This phase has additive routes to producing equiaxed grains have focused on manip-
previously been described as a ‘mild’ grain refiner, but can be difficult to ulating the thermal gradient and solidification velocity to induce sub-
incorporate in many aluminium alloys owing to rapid coarsening and stantial undercooling for nucleation of equiaxed microstructures. This
a high liquidus temperature, making gas atomization of additive feed- requires extensive manipulation of parameters including scan strategy
stock difficult22,23. In our approach, incorporation of this particulate at and build temperature and is not extensible to multiple alloy systems,

3 6 6 | N A T U RE | V O L 5 4 9 | 2 1 s e p t e m b e r 2 0 1 7
© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
letter RESEARCH

a b c

5 μm 25 μm 5 μm

d e f
Primary
solidification phase

Lattice-matched
nucleant phase
15 μm 1 μm

Figure 2 | Nanoparticle assembly on additive metal feedstock. The e, Iron powder with TiC nanoparticles. f, Schematic representation of
assembly approach enables the production of various feedstocks with how lattice-matched nanoparticles (bottom phase in blue and yellow)
different nanoparticle assemblies, which can be targeted to induce can induce low-energy-barrier epitaxial growth of solidifying metals
equiaxed grain growth a, Al7075 powder with TiB2 nanoparticles. (top phase in purple), with lattice-matched planes in the unit cells
b, TiAl6V4 powder with ZrH2 nanoparticles. c, Al7075 powder with indicated in green on the right.
WC nanoparticles. d, AlSi10Mg powder with WC nanoparticles.

additive hardware or build geometries10,24. Although the solidification Hunt criterion assumes a steady-state solidification front, whereas the
velocity is relatively high, it alone is not sufficient to induce equiaxed additive process deviates substantially from steady state owing to the
growth per the conventional Hunt criterion for a columnar-to-equiaxed raster pattern and accumulation of residual heat26. As such, solidifica-
transition. In particular, the high thermal conductivity of aluminium tion preferentially occurs through nucleation on existing grains, leading
and the large liquid diffusivities of alloying elements make substantial to the observed grain growth vertical to the build direction with grains
undercooling extremely difficult to achieve with the accessible ranges extending across multiple build layers, as in the inverse pole figure
of solidification velocities and thermal gradients25. In addition, the map in Fig. 1e.

a 635 b fs = 0.2 fs = 0.6 fs = 0.9


AlSi10Mg
610 Al7075

585
Temperature (°C)

560

535

510

485

460
0.0 0.2 0.4 0.6 0.8 1.0
Fraction of solid, fs
c 580 d
Al7075
Al7075 + Zr Al7075 Al7075 + Zr
560
Temperature (°C)

540

520

500

480
25 μm 25 μm
460
0.80 0.85 0.90 0.95 1.00 Cracks Residual porosity
Fraction of solid, fs

Figure 3 | Solidification behaviour of additive aluminium alloys. three panels in each row correspond to the solid fractions indicated by the
a, Solidification curves for Al7075 (orange) and conventional 3D-printed vertical green dashed lined in a. c, Adding zirconium to Al7075 (purple)
aluminium (AlSi10Mg; blue). b, Schematic representation of solidification, has little effect on the solidification behaviour at high solid fractions, at
indicating how solidification over a large temperature range leads to long which alloys are the most tear- and crack-susceptible. d, Polished and
channels of interdendritic fluid that result in cracking (top; such as in etched scanning electron microscopy (SEM) images depicting the resulting
Al7075), whereas a small solidification range leads to short interdendritic microstructures with (right) and without (left) the addition of zirconium.
regions that can easily be backfilled (bottom; such as in AlSi10Mg). The

2 1 s e p t e m b e r 2 0 1 7 | V O L 5 4 9 | N A T U RE | 3 6 7
© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
RESEARCH letter

a 500 Table 1 | Material properties of relevant alloy systems


Yield Ultimate Elastic Elongation
strength strength modulus to failure
400
Material (MPa) (MPa) (GPa) (%)
AM Al7075 (T6) NA 25.5 NA 0.4
Stress (MPa)

300 Wrought Al7075 372–469 462–538 71.7 3–9


(T6; plate)31
AM Al7075 +​  Zr (T6) 325–373 383–417 63–66 3.8–5.4
200
AM AlSi10Mg 209 315 69.4 7.3
Al7075 + Zr AM AlSi10Mg (CL31)32 170–220 310–325 75 2–3
100 Al7075 Measured and literature values for comparative alloys; ranges listed when available. ‘T6’ refers to
AlSi10Mg the heat-treatment condition; ‘plate’ refers to the specific material; AM, additively manufactured;
NA, not available.
0
0 2 4 6 8
Elongation (%)
elimination of cracks that we observed is attributed to the change in
b 5% microstructure. Hot-tearing models, including the RDG model, are
dominated by the final stage of solidification when the fraction of solid
is greater than 0.8 (ref. 27). Many hot-tear-susceptible materials can be
identified from their solidification curves (Fig. 3a). The shapes of these

Principal strain
curves are dictated by the compositions of the constituent alloys and
can be described using a Scheil–Gulliver solidification model based on
the equilibrium phase diagram28. Thermo-Calc software was used to
simulate sequential steps from the liquidus temperature to an approx-
imate solidus temperature, calculating the fraction of solid and com-
position of the new liquid at each point. Susceptible alloys have large
0%
1% 2% 3% 4.5% solidification ranges between the liquidus and solidus temperatures and
Figure 4 | Mechanical testing of 3D-printed aluminium alloys. sharp turnover in the solidification curves at high fractions of solid.
a, Representative tensile curves of the 3D-printed materials in this study. The sharp turnover is typically associated with the increased levels of
b, Representative deformation behaviour of Al7075 +​ Zr, indicating Luders strengthening solute that partitions in the liquid to a high degree during
band propagation due to the refined grain size. The colour scale shows the solidification. Associated thermal shrinkage leads to tearing and cav-
local principal strain, with the total elongation listed under each panel. itation in that thin films of interdendritic liquid that are present at the
high solid fraction. Decreasing the solid fraction at which the turnover
In contrast, the 7075 and 6061 alloys manufactured with grain- occurs or reducing the difference between the solidus and liquidus tem-
refining nanoparticles show uniform equiaxed growth with no cracking. peratures will improve resistance to tearing. A conventional additive
Upon melting, zirconium particulates are pulled into the melt pool and aluminium alloy such as AlSi10Mg has both an early turnover and a
react to form Al3Zr. Al3Zr has more than 20 matching interfaces with small difference in liquidus and solidus temperature, leading to a low
the primary face-centred-cubic aluminium phase, exhibiting less than tendency for cracking during solidification. This is in stark contrast to
0.52% lattice mismatch and 1% variation in atomic density, providing 7075 and other hot-crack-susceptible alloys (Extended Data Fig. 5 with
an ideal low-energy heterogeneous nucleation site. Nucleation of new Al6061 as well).
grains ahead of the solidification front requires both an energetically The shapes of the solidification curves can be shifted with increasing
favourable condition and a large number of nucleation sites to ensure solidification velocity owing to the non-equilibrium partition coeffi-
new grains can form before the main solidification front overtakes new cients; however, we found no evidence of substantial departure from
grains. The columnar growth demonstrated in the unmodified mate- equilibrium29. The addition of zirconium might be expected to shift the
rial indicates that undercooling is present, providing an energetically solidification curve into a more favourable shape; however, as shown in
favourable condition; however, without additional nucleation sites, Fig. 3c, this is not the case. The Al–Zr binary phase diagram indicates
homogenous nucleation requires a substantially higher energy barrier. a peritectic reaction at high mass fractions of aluminium22. As such,
The large number of low-energy-barrier heterogeneous nucleation sites any Al–Zr reactions occur at the beginning of solidification where tear
ahead of the solidification front induces a fine equiaxed structure under resistance is not critical owing to the low volume fraction of solid. More
the same processing conditions as for the unmodified powder. This importantly, the addition of zirconium does not substantially alter the
results in crack-free microstructure with grain sizes of about 5 μ​m, 100 shape of the solidification curve at high fractions of solid, where hot
times smaller than the grains in the unmodified material (Fig. 1e, f). tearing is initiated. As discussed above, the early inclusion of zirconium
The nucleant particles are uniformly incorporated into the micro- induces equiaxed growth, which can more easily accommodate the
structure, which can provide additional strengthening and resistance thermal contraction strains associated with solidification, ultimately
to grain growth owing to pinning effects. resulting in an alloy system that is highly tear resistant, despite con-
The cracking that we observed in the stock material appears consist- ventional wisdom.
ent with the mechanisms of the RDG model8. Columnar grains grow in We performed tensile testing and compared the results against
the direction of the heat flux, leaving a thin layer of interdendritic fluid equivalent specimens produced from unmodified powder to verify
and leading to cavity formation (Fig. 3b). Further thermal shrinkage the crack-free nature of the additively manufactured material. Figure 4
allows this initial cavity to ‘unzip’ and to propagate through interden- displays typical stress–strain curves for each material; the associated
dritic colonies, resulting in large cracks oriented parallel to the colum- yield strength, modulus, ultimate tensile strength and elongation to
nar grains8. Although the RDG model does not explicitly describe the failure are summarized in Table 1. As shown, stock 7075 retains almost
effects of equiaxed microstructure on crack susceptibility, the shift to no strength owing to the large volume of cracks caused by hot tearing.
equiaxed growth drastically reduces the effect of entrapped liquid as Conventional AlSi10Mg shows about 7% ductility but less than half
the grains begin to behave as a low-resistance granular solid21. Fine the strength of the wrought 7075 system, consistent with data pro-
equiaxed semi-solid structures allow easier grain rotation and defor- vided by multiple selective laser melting companies. In comparison,
mation, providing a means to accommodate strain in the semisolid additively manufactured 7075 with the incorporation of Al3Zr nucle-
state and thus preventing crack initiation and growth. The complete ant particles shows an 80% increase in strength over AlSi10Mg, and is

3 6 8 | N A T U RE | V O L 5 4 9 | 2 1 s e p t e m b e r 2 0 1 7
© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
letter RESEARCH

within the expected bounds for its wrought counterpart. The modified 11. Easton, M. A. & Stjohn, D. Grain refinement of aluminum alloys : part I. The
nucleant and solute paradigms — a review of the literature. Metall. Mater. Trans.
Al7075 demonstrates Luders banding during deformation, which is A 30, 1613–1623 (1999).
indicative of an aluminium alloy with grain sizes of less than 10 μ​m 12. Ding, Y. et al. Microstructure and mechanical property considerations in
(Fig. 4)30. The yield strength and elongation of functionalized Al7075 additive manufacturing of aluminum alloys. MRS Bull. 41, 745–751 (2016).
13. Schmidtke, K., Palm, F., Hawkins, A. & Emmelmann, C. Process and mechanical
was produced within reported ranges of wrought Al7075 (Table 1); properties: applicability of a scandium modified Al-alloy for laser additive
however, the ultimate strength difference and lower limit of yield manufacturing. Phys. Procedia 12, 369–374 (2011).
strength can be explained by strain softening from the reduced grain 14. Material Data Sheet — Scalmalloy (APWorks, 2016).
15. Starke, E. A. & Staley, J. T. Application of modern aluminum alloys to aircraft.
size and evaporation of zinc, a major strengthening element, during the Prog. Aerosp. Sci. 32, 131–172 (1996).
laser melting process. Differences between additively manufactured 16. Immarigeon, J. P. et al. Lightweight materials for aircraft applications. Mater.
7075 and conventional wrought material can be remedied by increasing Charact. 35, 41–67 (1995).
the zinc concentration in the feedstock powder to improve strength and 17. Miller, W. et al. Recent development in aluminium alloys for the automotive
industry. Mater. Sci. Eng. A 280, 37–49 (2000).
by optimizing the heat treatment to target an optimum final grain size 18. Society of Automotive Engineers & American Society for Testing and
to eliminate strain softening. Likewise, ductility and elastic modulus Materials. Metals and Alloys in the Unified Numbering System (ASTM
can be increased by improving processing parameters to reduce the International, 2012).
19. Nelson, W. Duralumin Welding. Report No. 399 (National Advisory Committee
porosity caused by excessive laser energy density and trapped gas, a for Aeronautics, 1927).
feature that was not optimized in this study. Additively manufactured 20. Pumphrey, W. I. & Lyons, J. V. Cracking during the casting and welding of
metal parts are often hot-isostatic-pressed to reduce porosity and to the more common binary aluminum alloys. J. Inst. Met. 74, 439–455
(1948).
improve properties, but this was not carried out for this study to pre- 21. Gourlay, C. M. & Dahle, A. K. Dilatant shear bands in solidifying metals. Nature
serve equivalent processing conditions between functionalized and 445, 70–73 (2007).
stock aluminium powders. 22. Murray, J., Peruzzi, A. & Abriata, J. P. The Al-Zr (aluminum-zirconium) system.
J. Phase Equilibria 13, 277–291 (1992).
We have used secondary particulates in additive manufacturing to 23. Murty, B. S., Kori, S. A. & Chakraborty, M. Grain refinement of aluminium and
induce grain refinement of high-strength aluminium alloys of wrought its alloys by heterogeneous nucleation and alloying. Int. Mater. Rev. 47, 3–29
compositions, producing crack-free materials with strengths double (2002).
24. Dehoff, R. R. et al. Site specific control of crystallographic grain orientation
that of the most common additively manufactured aluminium alloy. through electron beam additive manufacturing. Mater. Sci. Technol. 31,
This metallurgical approach is applicable to other industrially relevant 931–938 (2015).
crack-susceptible alloys and can be extended to new families of additive 25. Martorano, M. A., Beckermann, C. & Gandin, C.-A. A solutal interaction
manufacturing materials such as non-weldable nickel alloys, superal- mechanism for the columnar-to-equiaxed transition in alloy solidification.
Metall. Mater. Trans. A 34, 1657–1674 (2003).
loys and intermetallics. Furthermore, our approach provides a metal- 26. Hunt, J. D. Steady state columnar and equiaxed growth of dendrites and
lurgical tool for metals processing, affording a diverse range of alloys eutectic. Mater. Sci. Eng. 65, 75–83 (1984).
for additive manufacturing, accelerating broad adoption of additive 27. Eskin, D. G., Suyitno & Katgerman, L. Mechanical properties in the semi-solid
state and hot tearing of aluminium alloys. Prog. Mater. Sci. 49, 629–711
processes and enabling the design of new alloy systems specifically for (2004).
additive processing. 28. TCAL4: TCS Al-based Alloy Database http://www.thermocalc.com/
media/19881/tcal4.pdf (Thermo-Calc, 2016).
Online Content Methods, along with any additional Extended Data display items and 29. Aziz, M. J. Model for solute redistribution during rapid solidification.
Source Data, are available in the online version of the paper; references unique to J. Appl. Phys. 53, 1158–1168 (1982).
these sections appear only in the online paper. 30. Yu, C. Y., Kao, P. W. & Chang, C. P. Transition of tensile deformation behaviors in
ultrafine-grained aluminum. Acta Mater. 53, 4019–4028 (2005).
received 14 February; accepted 26 July 2017. 31. CL 30AL/CL 31AL Aluminium Alloy (Concept Laser, Hofmann Innovation Group,
2012).
32. Boyer, H. E. & Gail, T. L. Materials Handbook Desk Edition (American Society for
1. Lewandowski, J. J. & Seifi, M. Metal additive manufacturing: a review of Metals, 1985).
mechanical properties. Annu. Rev. Mater. Res. 46, 151–186 (2016).
2. Frazier, W. E. Metal additive manufacturing: a review. J. Mater. Eng. Perform. 23, Acknowledgements We acknowledge financial support by HRL Laboratories,
1917–1928 (2014). LLC, and thank D. Martin for her artistic contribution to the figures, as well as B.
3. Collins, F. R. & Dudas, J. H. Preventing weld cracks in high-strength aluminum Carter of HRL Laboratories, LLC, X. Li of the University of California, Los Angeles,
alloys. Weld. J. 45, 241 (1966). and K. Hemker of John Hopkins University for discussions.
4. Kaufmann, N. et al. Influence of process parameters on the quality of
aluminium alloy EN AW 7075 using selective laser melting (SLM). Phys. Author Contributions J.H.M., B.D.Y., J.M.H., T.A.S. and T.M.P. analysed the results
Procedia 83, 918–926 (2016). and wrote the manuscript. J.H.M., B.D.Y. and T.A.S. designed the experiments.
5. Zhang, H., Zhu, H., Qi, T., Hu, Z. & Zeng, X. Selective laser melting of high B.D.Y. and J.A.M. functionalized the feedstock material and operated the
strength Al-Cu-Mg alloys: processing, microstructure and mechanical Concept Laser M2. J.H.M., B.D.Y. and J.A.M. prepared the metallurgical
properties. Mater. Sci. Eng. A 656, 47–54 (2016). specimens and performed the optical and electron microscopy and the
6. Kurz, W. & Fisher, D. J. Fundamentals of Solidification (Trans Tech Publications, mechanical testing.
1998).
7. Coniglio, N. & Cross, C. E. Initiation and growth mechanisms for weld Author Information Reprints and permissions information is available at
solidification cracking. Int. Mater. Rev. 58, 375–397 (2013). www.nature.com/reprints. The authors declare no competing financial
8. Rappaz, M., Drezet, J. & Gremaud, M. A new hot-tearing criterion. Metall. Mater. interests. Readers are welcome to comment on the online version of the paper.
Trans. A 30, 449–455 (1999). Publisher’s note: Springer Nature remains neutral with regard to jurisdictional
9. Yuan, L., O’Sullivan, C. & Gourlay, C. M. Exploring dendrite coherency with the claims in published maps and institutional affiliations. Correspondence and
discrete element method. Acta Mater. 60, 1334–1345 (2012). requests for materials should be addressed to J.H.M. (jhmartin@hrl.com).
10. Collins, P. C., Brice, D. A., Samimi, P., Ghamarian, I. & Fraser, H. L.
Microstructural control of additively manufactured metallic materials. Annu. Reviewer Information Nature thanks P. Collins, I. Todd and the other
Rev. Mater. Res. 46, 63–91 (2016). anonymous reviewer(s) for their contribution to the peer review of this work.

2 1 s e p t e m b e r 2 0 1 7 | V O L 5 4 9 | N A T U RE | 3 6 9
© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
RESEARCH letter

Methods ageing time for the samples was determined by performing four ageing treatments
Materials. Aluminium alloy 7075 micropowder. Aluminium alloy 7075 micro­ with hold times of 6 h, 12 h, 18 h and 24 h at 120 °C. Vicker’s hardness measure-
powder was purchased from Valimet Inc. The powder consisted of Al (balance), Zn ments were used to determine a range of appropriate ageing time around 18 h (see
(5.40%), Mg (2.25%), Cu (1.54%), Cr (0.19%), Fe (0.17%), Si (0.13%), Mn (0.02%) Extended Data Fig. 3).
and Ti (<​0.01%), in weight per cent. The particle size distribution was bimodal Sectioning and sample preparation. All samples were removed from the build
with peak values at 45 μ​m and 15 μ​m (Extended Data Fig. 1). plates via wire electro discharge machining (EDM). Tensile specimens were sec-
Aluminium alloy 6061 micropowder. Aluminium alloy 6061 micropowder was tioned with wire EDM to a thickness of 2 mm. Tensile specimens were prepared
purchased from Valimet Inc. The powder consisted of Al (balance), Mg (0.83%), for mechanical testing by polishing the surfaces of the gauge section with 240, 360,
Si (0.62%), Fe (0.25%), Cu (0.23%), Cr (0.08%), Mn (0.04%), Zn (0.04%) and Ti 400, 800 and 1,200 grit sand paper by hand. One side of the mechanical test samples
(0.02%), in weight per cent. The average particle size was 45 μ​m. was painted with white and spackled with black paint with an airbrush for digital
CL31 aluminium–silicon–magnesium alloy micropowder. Aluminium–silicon– image correlation using a GOM ARAMIS-3D Motion and Deformation Sensor.
magnesium alloy micropowder was purchased from Concept Laser Inc. The pow- Microstructure blocks were sectioned with a water-cooled saw and mounted in
der consisted of Al (balance), Si (9.0%–10.0%), Mg (0.2%–0.45%), Fe(<​0.55%, epoxy resin for polishing. Grinding was done with 240, 360, 400, 800 and 1,200 grit
trace), Mn(<​0.45%, trace) and Ti(<​0.15%, trace), in weight per cent. Particle size sand paper. Final polishing of the samples was accomplished with 1-μ​m diamond
optimized for selective laser melting and proprietary to the manufacturer. and 50-nm Al2O3 polishing compounds from PACE Technologies. Some polished
Hydrogen-stabilized zirconium. ZrH2 powder was purchased from US Research samples were etched with Keller’s Etch for 10 s to reveal microstructure. Additional
Nanomaterials Inc. imaging was conducted using SEM and electron backscatter diffraction (EBSD).
Lattice matching. To determine the particulate compositions and crystallographic Materials characterization. To observe microstructural differences, mounted sam-
faces with the highest probability of inducing epitaxial, heterogeneous nucleation ples were observed with an optical microscope under polarized light and with SEM.
in the given alloy system, we developed software with Citrine Informatics that Vicker’s micro hardness was performed on mounted samples. The load applied
uses lattice matching algorithms to search through crystallographic databases for was 200 g and indentation sizes were measured and compared (see Extended Data
the highest matching crystal structures. By minimizing the lattice strain between Fig. 3). These hardness measurements indicated that the appropriate ageing time
an inoculating nanoparticulate and the nucleating phase of a molten material, is about 18 h for additively manufactured 7075.
the free-energy barrier to nucleation is reduced, enabling higher nucleation rates Inductively coupled plasma optical emission spectroscopy (ICP-OES) was com-
for any given undercooling. Lattice matching is calculated through an area strain pleted on raw Al7075 +​ Zr powder and on a Al7075 +​ Zr component after being
measurement that accounts for the change in surface area of one indexed lattice additively manufactured. Analysis showed about 25% loss of zinc and about 32%
plane that is needed to match to the area of another indexed lattice plane of another loss of magnesium, both of which are strengthening elements for the 7075 alloy.
material. Mechanical testing. Tensile tests were performed on a servo-electric INSTRON
Selective laser melting. Additive manufacturing of the stock aluminium alloy 5960 frame equipped with a 50-kN load cell (INSTRON). Samples were clamped
and functionalized aluminium alloy powders were performed on a Concept Laser by the ends of the dog-bone-shaped samples. The extension rate was 0.2 mm min−1
M2 selective laser melting machine (specifications are listed in Extended Data and samples were loaded until fracture. Testing was conducted following ASTM E8.
Table 1). Samples consisted of 60 mm ×​ 20 mm ×​ 40 mm tensile block specimens A U-joint was used to account for any misalignment in the sample. Extended Data
and 10 mm ×​ 10 mm ×​ 40 mm blocks for examining microstructure. Images of Fig. 4 displays the stress–strain curves for the Al7075 material tested in this study
the as-printed samples on the build plates can be seen in Extended Data Fig. 2. and typical stress–strain curves for the CL31 stock powder produced.
Samples were processed with the Concept Laser ‘islanding’ scan strategy, which Because cracking tended to orient parallel to the build direction, all tensile
was specifically developed for the CL31 AlSi10Mg alloy material to minimize ther- testing was conducted perpendicular to the expected crack orientation. This
mal and residual stress build-up in the part. Islands that compose the core of the ensured that any residual cracks would have the maximum effect on the tensile
build geometry were 2 mm ×​ 2 mm in size. Standard machine parameters provided properties. Observed ductility in the nanoparticle-functionalized material indicates
by the Concept Laser for conventional AlSi10Mg alloy were used for all builds. a complete elimination of deleterious cracking.
The parameter values are considered proprietary by Concept Laser and cannot Scheil simulation. Scheil simulations were conducted as described in the main
be accessed by the user. The 70 mm ×​ 70 mm build plates were machined out of text. Extended Data Fig. 5 shows the solidification curve for Al6061 in comparison
aluminium alloy 6061 and sandblasted on the surface. Layers of the build were to the solidification curves of Al7075 and AlSi10Mg (which are also depicted in
incremented by a range from 25 μ​m to 80 μ​m depending on part geometry and Fig. 3a). The shape of the Al6061 Scheil curve indicates a high susceptibility to
location in the build. Processing was done under a flowing, inert argon atmosphere hot cracking.
with oxygen monitoring. All processing was completed at room temperature with Microstructure. Extended Data Fig. 6 shows micrographs of additively manufac-
no applied heat to the build plate. Samples were removed from the machine and tured 6061 with and without Zr additions, indicating identical behaviour to 7075.
cleaned of extra powder by sonicating in water. Parts were then dried with clean Additional micrographs of 7075 with and without Zr can also be seen in Extended
compressed dry air. Data Fig. 7. Cracking is seen in the unmodified powder, whereas the addition of
Heat treatment. Samples were then heat treated to a T6 condition. Samples the Zr nucleant induces fine equiaxed grains and eliminates hot cracking. Further
were solutionized at 480 °C in air with a ramp rate of 5 °C min−1 for 2 h and then evidence of cracking between columnar grains in additive manufacturing of 7075
quenched with water at 25 °C. Samples were subsequently aged at 120 °C with a without Zr is shown in Extended Data Fig. 8. The EBSD map is used to clearly show
ramp rate of 4 °C min−1 in air for 18 h and allowed to cool to room temperature. that the cracking is restricted to intergranular regions.
Although additional insight is needed to determine the appropriate heat treat- Data availability. The data that support the findings of this study are available
ment methods for additively manufactured high-strength aluminium alloys, from the corresponding author on reasonable request.

© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
letter RESEARCH

Extended Data Figure 1 | SEM image of aluminium alloy 7075 powder.

© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
RESEARCH letter

Extended Data Figure 2 | As-printed Al7075 parts for tensile testing and microstructure evaluation. a, Stock Al7075. b, Al7075 +​ Zr.

© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
letter RESEARCH

Extended Data Figure 3 | Ageing behaviour of additive Al7075.

© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
RESEARCH letter

Extended Data Figure 4 | Stress–strain curves for the materials tested in this study, indicating high repeatability in the Al7075 + Zr material.
The colour indicates the material type, with curves of the same colour indicating replicate samples.

© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
letter RESEARCH

Extended Data Figure 5 | Scheil solidification curves for aluminium alloys AlSi10Mg, Al7075 and Al6061. The curve for Al6061 is from ref. 28.

© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
RESEARCH letter

Extended Data Figure 6 | Micrographs of etched Al6061, processed as received. Large cracks are observed in the absence of Zr (left). With the addition
of Zr nanoparticles, no cracking is observed, but there is some residual porosity (right). Rows indicate increasing magnification.

© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
letter RESEARCH

Extended Data Figure 7 | Micrographs of etched Al7075, processed as received. Large networks of cracks are observed in the absence of Zr (left). With
the addition of Zr nanoparticles, no cracking is observed, but there is some residual porosity (right). Rows indicate increasing magnification.

© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
RESEARCH letter

Extended Data Figure 8 | EBSD inverse pole figure of 3D-printed stock 7075 indicating large networks of columnar cracking. Build direction is
vertical to the page.

© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.
letter RESEARCH

Extended Data Table 1 | Specifications of the Concept Laser M2 selective laser melting system

© 2017 Macmillan Publishers Limited, part of Springer Nature. All rights reserved.

S-ar putea să vă placă și