Lebrun 2015
Lebrun 2015
PII: S0261-3069(15)00272-1
DOI: http://dx.doi.org/10.1016/j.matdes.2015.05.026
Reference: JMAD 7257
Please cite this article as: LeBrun, T., Nakamoto, T., Horikawa, K., Kobayashi, H., Effect of Retained Austenite on
Subsequent Thermal Processing and Resultant Mechanical Properties of Selective Laser Melted 17-4 PH Stainless
Steel, Materials and Design (2015), doi: http://dx.doi.org/10.1016/j.matdes.2015.05.026
This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Effect of Retained Austenite on Subsequent Thermal Processing and Resultant Mechanical
‡: Corresponding Author
Abstract
Thermal processing and mechanical characterization of tensile specimens fabricated from bulk 17-4
PH material produced by selective laser melting (SLM) is presented. Prior to specimen evaluation, various
industry standard and non-standard heat treatments were performed. The focus of this thermal processing
was to evaluate the effect of aging on mechanical performance with or without a prior solution heat
treatment. Volumetric fraction of metastable austenite was measured to vary with heat treatments
performed and the initial conditions from which aging was initiated. Material aged following a solution
heat treatment was found to have an austenite reversion from a fully martensitic structure to volumetric
fractions as high as 20.4%. Whereas, initial concentrations of retained austenite in SLM as-fabricated
material increased following a peak-age heat treatment, but decreased with higher temperature, overaged
heat treatments. Specimens with large amounts of austenite demonstrated stress-induced transformation to
martensite during tensile testing. This behavior was reflected in substantially reduced yield strengths,
increased work hardening rates across greater ranges of strain, and delayed onset of localized plastic
deformation. Ultimate tensile strengths for varieties of specimens aged to the peak-aged condition were
similar, but differed greatly at strain levels where this strength was achieved.
Keywords: Additive Manufacturing; Heat Treatment; Stainless Steel; Selective Laser Melting; Retained
Austenite
1. Introduction
1
Hardware production via additive manufacturing (AM) technologies has proven versatile across a
variety of industries as a means to bypass many aspects of existing, traditional manufacturing processes and
relatively new material processing technology may not be fully understood or appreciated. Of the alloys
currently deployed in commercial AM applications, a small number may be further thermally processed to
achieve desired strength and toughness characteristics by way of precipitation hardening (PH) kinetics. The
effectiveness of these subsequent thermal processing steps when performed on bulk material produced by
AM remains largely unstudied. Similarly, some investigators have identified enhanced mechanical strength
as a result of AM [1,2], but consideration for how these properties are altered as a result of additional
thermal processing is lacking. Of the materials available that may take advantage of further PH through
aging, this study has focused on 17-4 PH stainless steel (AISI 630) as processed by selective laser melting
(SLM).
17-4 PH is primarily a lath martensitic (bct, α’) stainless steel at room temperature following the
allotropic transformation from an austenitic structure (fcc, γ-iron) upon cooling. Strength may be enhanced
through subsequent aging as a consequence of the nucleation and growth of Cu-rich precipitates. These
precipitates initially form as coherent spherical bcc structures that may transform into enlarged fcc non-
coherent structures at greatly extended aging times [3]. With aging, a reversion of martensite back to the
parent austenite structure may be noted. This behavior is believed to be a result of localized chemical
stabilization following the diffusion of Ni and Cu coinciding with the formation of precipitates, increasing
localized concentration of austenite stabilizing elements and locally lowering the martensitic transformation
AM processing of 17-4 PH has been studied by some regarding initial mechanical and material
properties. Previous work by the authors of this paper [6] outlined the relative increases in mechanical
strength and the strain rate sensitivity associated with the highly refined microstructure as a result of
selective laser melting (SLM). The small, prior-austenite grain sizes observed for the as-fabricated 17-4 PH
material is attributable to the rapid melt and solidification of the working powder during processing.
Several studies by others have investigated the retention of metastable austenite following the formation of
bulk material by SLM and other similar processing techniques. In one case, a concentration ratio as high as
72% austenite to the 28% martensite was reported for 17-4 PH at room temperature [7]. The retention of
2
this metastable phase is largely seen as a consequence of mechanical stabilization due to the residual
thermal stresses within the bulk material. Additionally, the influence on the residual phase fractions and
texturing by processing parameters, initial working powder fabrication methods, and SLM processing
environment have also been identified [8]. How this retained austenite and its evolution with additional
standard and non-standard thermal processing affect mechanical properties for SLM 17-4 PH is the focus of
this study.
2. Experimental Procedures
Water-atomized pre-alloyed 17-4 PH powder was used as the working material to fabricate high-
density bulk primitives. Resultant density was measured using the Archimedes method on polished cubic
specimens. The averaged density, where surface defects were removed, was 7.7 grams per cubic
centimeter. This amounted to a relative density of 98.7% when compared to fully densified material. Figure
1 highlights the particle shape and regularity of the powdered material used. The two dissimilar particle
shapes, irregular and spherical, are a result of the water atomization process used in its fabrication. Table 1
outlines the chemical composition of the powder as compared to the ASTM A564/A564M [9] material
standard maximums and ranges. Table 2 details the powdered material size distribution as percent weight
Figure 1
Table 1
Table 2
Bulk material was fabricated using an EOS M280 SLM machine (EOS, GmbH, Germany). The total
quantity of primitives and subsequent tensile specimens were built across two different operations, but the
same powder source and operational settings with the machine system were used. In an effort to minimize
oxidation, a protective nitrogen atmosphere was maintained by an attached nitrogen generator throughout
the entire SLM process. The M280 system uses a 400 W capable Nd:YAG fiber laser system to scan, melt,
and consolidate the powdered material into bulk form. Laser operational settings such as scanning speed,
power, focus (laser spot diameter), layer thickness and scanning pattern may be adjusted based on the
material and purpose of the build. Settings deployed in this study included a laser power of 190 W,
scanning speed of 787 mm/s, laser spot diameter of 100 μm, and a scan centerline spacing of 110 μm (10
3
μm gap between individual scanning paths). The laser was operated in continuous wave mode. Layer
thickness was controlled by the vertical displacement of the build chamber’s baseplate between laser
scanning. In this study a layer thickness of 40 μm was used. While scanning patterns are a focus of some
studies [10], this investigation used a single pattern of parallel scan lines with each layer’s orientation
The focus of this experimental investigation was to characterize the mechanical properties of 17-4
PH produced via SLM. However, these evaluations were intended to be irrespective of the geometries
fabricated by this process (i.e. independent of the shape of the volume submitted to the machine for
building). It is well recognized that resultant surface finishes associated with this form of AM are currently
inferior to what is capable by traditional machining [11,12]. Tensile specimens were therefore machined
from within the bulk material primitives so as to eliminate surface-related effects and artifacts of the SLM
process. Tensile specimens evaluated in this study were all fabricated with their loading axes parallel to the
SLM building plane (i.e. layers are oriented in-line with the load applied during tensile testing). Figure 2
depicts the geometry of the compact-size tensile specimens produced. Tensile specimens of this geometry
were selected because of the requirements associated with additional, dynamic tensile testing performed
outside of this study and the need for congruency in mechanical behavior irrespective of specimen
geometry.
Figure 2
Following initial machining, a subset of the tensile specimens did not undergo additional thermal
processing. These were evaluated with material in a condition hereafter referred to as “as-fabricated”. A
variety of material standards govern heat treatments of 17-4 PH, but for the purpose of this study the
ASTM A564/A564M-10 standard is referenced [9]. The remaining specimens underwent the standardize
H900 (480 °C for 1 hr), H1025 (550 °C for 4 hrs), and H1150 (620 °C for 4 hrs) heat treatments. Half of
those treated to each temper were initially treated with a solution heat treatment to the Condition A state
(1040 °C for 30 min) and quenched prior to the subsequent aging. The reported treatment temperatures and
times for both varieties of specimens are in accordance with the aforementioned material standard;
however, the minimum yield and ultimate tensile strengths, elongations, and hardness values are for
materials aged after an initial solution heat treatment. Some researchers working with this material
processed by SLM have evaluated properties without this initial step and treated the as-fabricated condition
4
as a solution treated state [8]. A solution heat treatment was included as part of the preparatory processing
for half of the specimens evaluated to measure its effect. For shorthand notation and disambiguation, aged
specimens that have had a solution heat treatment prior to further aging will be referred to in this paper with
the shorthand "CA" prefix attached to the name of the aged condition (e.g., CA-H900). All heat treatments
were performed in a quartz chambered electric vacuum furnace. At the end of the designated heat
treatment, all specimens were then exposed to laboratory air and quenched in water.
Material samples not intended for mechanical evaluation underwent the same thermal processing.
These bulk samples were evaluated by X-ray diffraction (XRD) (spectral) analysis. XRD analysis was
performed using a Rigaku RINT Ultima III X-Ray Diffractometer with a Co (Kα) X-ray source (λ = 1.79
Å). Rupture surfaces of tensile specimens were evaluated by standard scanning electron microscopy using a
JEOL JSM-5310LV Scanning Microscope. Transmission electron microscopy (TEM) images were
Tensile testing was performed using a Shimadzu AG-Xplus, a screw-driven uniaxial tension testing
machine. Cross-head speed was set to 0.5 mm/min for an equivalent 1.04x10-3 s-1 strain rate. Because of the
specimens’ relatively small gage length, specimen strain was evaluated by measuring cross-head
displacement. Elastic compliance of the testing apparatus was corrected in all recorded tests to arrive at the
correct specimen displacement and corresponding elongation with respect to the applied load. All testing
Engineering stress-strain curves typical of samples evaluated for each condition are presented in Fig.
3 (a) and (b). Figure 3 (a) shows curves for specimens following aging heat treatments to the CA-H900,
CA-H1025, and CA-H1150 states. These curves are also plotted alongside a reference curve for a typical
SLM specimen following only a solution heat treatment. Figure 3 (b) shows a similar comparison, but with
specimens treated to the same aging conditions without the intermediate, post-SLM solution heat treatment.
Figure 3 (b) also shows a stress-strain curve for a typical as-fabricated specimen without any additional
thermal processing following machining. Averaged values and ranges for micro-Vickers hardness, proof
stress (0.2% elastic offset), ultimate tensile strength (UTS), uniform elongation (UEL), and hardness for the
5
various tested conditions are summarized in Table 3. Reference minimums per the material standard are
included in parenthesis. In the cases where an upper and lower yield point are observed, the upper yield
point is presented in the table. All averages and ranges were of 3 measurements for each condition.
Table 3
Specimens shown in Fig. 3 (a) demonstrate an evolution of mechanical behavior across the aging
conditions evaluated consistent with standard thermal processing for 17-4 PH [9,13,14]. Aging to the CA-
H900 results in the peak-age condition characterized by the high yield and ultimate tensile strengths. The
two other aged conditions, where tempering was performed at higher temperatures and for longer durations,
demonstrated an expected reduction in mechanical strength and accompanied increase in elongation. Total
elongation following uniform plastic strain for all solution heat treated specimens was impaired due to local
stress concentrations at embedded porosity sites. For samples of the peak-aged condition, shortly after the
onset of localized plastic deformation (necking) failure was observed to occur. And for progressively
longer and higher temperature heat treated specimens, both UEL and total elongation were observed to
increase. The increase in UEL also indicates the material’s increasing ability to strain harden after initial
A largely different stress-strain response is observed for specimens following thermal processing
without an initial solution heat treatment. Both the as-fabricated and peak-age specimens yield at reduced
stress levels, but rapidly work harden across early plastic strain. The general shape of the stress-strain curve
observed of the as-fabricated condition is preserved following the peak-age H900 heat treatment, but at
elevated stress levels with noted increases in both YS and UTS. However, the shoulder associated with
initial yielding of the as-fabricated condition shows both an upper and lower yield point. This demonstrates
a gradual relaxation in stress with continued elongation for approximately 1% of the nominal strain before
once again strengthening. Unlike the CA-H1025 and CA-H1150 heat treatments, the H1025 and H1150
show an increase from the initial as-fabricated condition in both yield and ultimate strength with reductions
in total and uniform plastic elongation. This is in contrast with the behavior observed for those two similar
Figure 4 (a) and (b) show the work-hardening rates as a function of strain across all conditions
tested. These curves are generated from the individual tests depicted in both Fig. 3 (a) and (b) and are also
6
representative of the behavior for all tests conducted from each condition evaluated. The work hardening
rate was calculated by performing a finite difference of the true stress, divided by the corresponding
increment of true strain. Without in situ measurements of specimen diameter, the approximate work
hardening rates calculated by this method are only reliable under the assumption that uniform plastic
deformation is maintained until the onset of necking. Values calculated here are with units of MPa per unit
strain. The trend for each family of thermal processing conditions, those with a prior solution heat
treatment and those without, illustrate different behavior across the aged conditions. The two curves for the
as-fabricated and H900 conditions in particular depict strain-hardening rates that are not monotonically
decreasing with increasing strain. The reversal in strain-hardening rates explains the sigmoidal shape of the
stress-strain curve. Additionally, the H900 curve is observed to be negative for a short region of strain. This
corresponds to the softening observed post initial yielding outlined in Fig. 3 (b). When compared to the as-
fabricated condition, the H900 specimens show reduced work hardening rates over early stages of plastic
strain following the recovery from initial softening, but remain elevated by comparison after 5% strain and
persist across greater ranges of uniform elongation. Despite the increase in both yield and ultimate tensile
strengths observed for the H900 specimens over the as-fabricated condition, the relative increase in strength
as a result of this work hardening is slightly diminished. Relative differences between averages of yield
strength and UTS for the as-fabricated and H900 specimens are 594 and 472 MPa, respectively. This
illustrates a reduced capacity to work harden for this particular sequence of thermal processing and these
two conditions.
For those tested at the CA-H900 and CA-H1025 treatments, little to no work hardening capacity
over extended regions of strain is demonstrated by the rapid transition from nearly vertical to nearly flat
curves. The CA-H1150 shows a slightly elongated range of strain over which elevated work hardening
occurs. For specimens without an initial solution heat treatment and aged for an extended period (H1025
and H1150) an opposite trend in behavior is observed. Strain-hardening rates more rapidly transition
towards a constant value earlier in strain. The general shape for these two conditions progressively
Other researchers have noted by hardness testing that the absolute and relative strengths attained
post-H900 heat treatment, without a preliminary solution heat treatment, are consistent with those observed
7
for traditionally processed 17-4 PH [8]. However, micro Vickers hardness measurements made and
reported here for the H900 condition were found to approach, but not exceed the minimum required per the
material standard for this heat treatment (average 375 HV, 38.5 HRC equivalent, as compared to the
material standard minimum 40 HRC). Whereas, performing a solution heat treatment prior to the peak age
resulted in hardness measurements that well exceeded the minimum required (CA-H900) (average 417 HV,
42.5 HRC equivalent). These observations are also supported by the dissimilarity in the observed yield
strengths between the two conditions. Whereas, the ultimate tensile strengths were found to be similar
(1417 and 1444 MPa for H900 and CA-H900, respectively). Unlike hardness measurements, tensile testing
has highlighted the different yielding and plastic deformation behavior associated with the two thermal
Across all heat treatments performed it is useful to compare the relative increases in strengthening
associated with aging of the material. The relative increase in strength following an aging treatment may be
Where St may describe the relative increase in strength due to aging over an untreated specimen (either
from the as-fabricated condition or solution heat treated condition, as the respective baseline condition), So,
and the post-treated aged strength, S. For example, the relative increase in strength of a specimen in the
CA-H1025 condition is based on the relative difference to a sample in the Condition A state. And a H1025
specimen is directly compared to the strength measured from specimens of the as-fabricated condition. As
is common with investigations into aging kinetics, previous researchers relied upon measured hardness
values to establish this relationship as a relative change in strength. For the purpose of this study, true stress
at UEL is instead used. Similarly, tempering charts commonly provide hardness as a function of
temperature at which aging is performed. These charts only outline the relationship across a single
temperature with varying exposure times. However, the various standardized heat treatments evaluated
were performed across a variety of different temperatures and durations. Relating the effects of aging
across these two variables is necessary. The Larson–Miller or Hollomon–Jaffe tempering parameter (P)
[14,15] may be used to interrelate these as it is a time-compensated temperature term. The Hollomon-Jaffe
8
Where T is the absolute temperature in degrees Kelvin, t is the tempering time in hours, and C is a material
Figure 5 shows the relationship between these two expressions for the tensile specimens evaluated.
Both exhibit a downward trend off of the peak-age condition commonly characterized by a tempering
parameter of 15 for 17-4 PH. However, samples without an initial solution heat treatment preceding aging
did not exhibit the expected negative relative strengthening for larger values of P above values of 17.9.
These trends in behavior may be partially explained by a combination of interrelated mechanisms. To help
elucidate those differences, it is useful to examine the phase presence and volumetric fractions of
Figure 5
Figures 6 shows the XRD spectrums for the various heat treatment conditions as measured across
the vertical reference plane (SLM layers oriented normal to the surface examined). Mechanical processing
and preparation for XRD analysis may disturb the austenite and initiate a strain induced transformation to
martensite due to the localized stresses associated with grinding and polishing. Therefore, the observed
concentrations of austenite may be artificially low from what is initially produced via SLM. Observations
of these measurements, however, still point to reasons for the observed mechanical behavior. Table 4
details the calculated volume fractions of martensite and austenite observed for all specimen conditions
from the XRD line spectra. Quantitative X-ray diffraction phase analysis and calculated spectra were
performed using the Rietveld method provided by the MAUD software package by a full pattern fitting
From Fig. 6, specimens with an initial solution heat treatment show increases in the major and minor
γ-iron peaks, e.g. (111)γ or (200)γ, that progress with increased exposure temperature and time. The initial
state prior to these aging treatments, Condition A, contains no discernable fraction of the metastable
austenite phase post SLM-processing that is otherwise found in the as-fabricates state. Aging from this
initial condition shows a gradual reintroduction of austenite by reverting martensite to its parent austenitic
structure. This observation is consistent for standardized thermal processing of conventional 17-4 PH
[14,18]. This process of austenite reversion has been generally attributed to the chemical stabilization by
9
the diffusion of alloying element constituents, commonly toward lath boundaries. More specifically, as Cu
diffuses to form precipitates, the local M temperature is effectively lowered at the sites of their formation
f
and thereby permits austenite chemical stabilization at room temperature [19]. Viswanathan et al. identified
the presence of austenite grains to be closely accompanied by coarsened Cu-precipitates at lath boundaries
Figure 6
Table 4
The evolution from the as-fabricated condition is not as straightforward. Following the initial H900
treatment, an increase in austenite is measured. Treatments to H1025 and H1150 conversely see reductions
in the intensity of all γ-iron indices. At their peak, the H900 condition demonstrated an austenite volume
fraction of 40.5%, an increase from the as-fabricated fractional percent of 36.0%. The reduced strain-
hardening capacity observed for both the H1025 and H1150 specimens correlates with the similar reduction
in retained austenite following heat treatment. Work-hardening rates for both of these conditions begin to
approach those observed for the family of heat treatments that included a preliminary solution heat
treatment. The resultant concentrations at these two processing conditions is believed to be a combination
of two competing mechanisms for its retention. The reduction in volume fraction between H900 and H1025
conditions is likely a result of the release of thermal stresses associated with the SLM process when
samples are exposed to increasingly elevated temperature and longer duration heat treatments. It has been
suggested that the retention of austenite within SLM fabricated martensitic steels is a result of mechanical
stabilization by hydrostatic pressures generated during initial formation [20]. With a reduction in
mechanical stabilization, austenite is free to complete its transformation to martensite when cooled below
the expected Mf temperature. And by extension, it is outright eliminated in the case of a solution heat
treatment at 1040 °C for 30 minutes. While, as previously mentioned, the reversion or local stabilization of
austenite by the diffusion of Cu and Ni during aging maintains the observed austenite concentrations above
zero.
Comparative differences in relative strengthening between the two CA-H900 and H900 conditions
highlight the significance (see Fig. 5) of austenite concentrations and the material’s ability to age-harden.
Specimens with large starting volume fractions of austenite and heat treated to the peak-age condition do
10
not demonstrate the same relative strengthening capacity as that demonstrated by those initially solution
heat treated, such as CA-H900. In these cases, the austenite does not age-harden and consequently inhibits
strengthening by the intended growth of Cu-precipitates [5,21]. Precipitate nucleation requires a saturated
state. And in this case, austenite maintains a higher solubility of copper, thereby reducing age hardening
throughout. This is most evident when YS is considered for the two variations of the peak-age. Samples
aged without the solution heat treatment, H900, exhibited a 30.1% lower yield on average to those with
such preliminary treatment, CA-H900. Despite the differences between the relative strengthening for the
these two peak-aged conditions, the H900 specimens showed a capacity to strain harden throughout the
comparatively elongated uniform plastic deformation region. The plastic deformation behavior for both the
as-fabricated and H900 conditions with their high volume fractions of austenite demonstrate the capacity
for transformation induced plasticity (TRIP). This is further supported by two XRD measurements made
Figures 7, 8, and 9 show TEM imagery collected of the two overaged states, CA-H1150 and H1150.
Figure 7 (a) shows a bright field image of a CA-H1150 specimen. The dark field reflection highlights the
reverted austenite structure interspersed amongst the martensite lath structure adjacent to lath boundaries.
Figure 8 shows a magnified view of overaged precipitates commonly visible throughout this sample's
microstructure. Circular or oval-shaped precipitates were observed uniformly dispersed outside of austenite
grains, occasionally coinciding with dislocations. The long axis of the oval-shaped precipitates was found
to be limited to approximately 20 nm in length. In contrast to Fig. 7, Fig. 9 shows the same aging treatment
applied without an initial solution treatment. The dark field reflection shows austenite without the same
distribution or frequency as the CA-H1150. The precipitate structure observed in Figure 8 for the CA-
H1150 specimen was not observed with the same regularity. Twinning is observed in the selected area
diffraction patterns for both instances of austenite in the CA-H1150 and H1150 specimens.
Figure 8
Figure 10 shows an additional two XRD spectra measured from within a sectioned, as-fabricated
tensile specimen post-test. The upper curve demonstrates the combination of both martensite and austenite
as measured from within the threaded grip section of the specimen. No plastic deformation occurred as part
11
of the tensile test in this region of the sample. The lower curve shows the reduced volume fraction of
austenite as measured from within the gauge section adjacent to the fracture surface. The region of the
sample where localized plastic deformation occured (necking region) is a smaller target to successfully
measure for tensile samples of this shape and size. Therefore, these measurements made by XRD are
limited to the remainder of the gauge section. The measured phase presence and volumetric fractions
highlight the austenite to martensite transformation within the region largely deformed by uniform plastic
deformation. The calculated volume fraction of the grips section of the sample was found to be unchanged
from the bulk material also examined by XRD. The gauge section was measured to have fallen from 36.0%
martensite of the as-fabricated material. Similar behavior is noted in the peak-aged condition without an
Figure 10
Figures 11 (a) and (b) depict the fracture surfaces at low magnification for specimens in the CA-
H900 and H900 conditions. These images were taken at angle to highlight the macro-scale texturing in their
respective rupture surfaces. Despite the differences in total elongation and work hardening, these two
figures show the similarity in fracture behavior between the peak age conditions. Of note is the planarity
observed of both specimens. Figure 11 (c) shows a similar view, but of the as-fabricated condition. In
contrast to Figs. 11 (a) and (b), the material without further thermal processing is found to have a highly
irregular fracture surface. A greater degree of shearing is observed throughout and surrounding a less clear
fracture plane. Some isolated pores are observed to be atop local mounds of material or in the base of a
corresponding divots in the fracture surface. Figure 12 highlights such an example of a pore isolation with
Figure 12
Figures 13 and 14 show at high magnification the texturing across the surface observed for the CA-
H900 and as-fabricated specimens. The tensile loading direction was normal to the these images. When
viewed at this scale and normal to the fracture surface, the micro-scale dimple structure shown in both
12
figures suggests localized ductility. However, the formation of micro-voids and their coalescence into
larger dimple structures commonly observed in traditionally processed 17-4 PH is not observed here. When
the fracture plane is normal to the applied load, the observed texturing at the scale shown in Figs. 13 and 14
is consistent across all SLM specimens despite heat treatment condition. The shared fracture surface texture
at the smallest size scale suggests that mechanics of deformation achieve commonality between all
specimens despite their differing initial microstructural conditions. For specimens with high volumetric
fractions of austenite, this follows the transformation of that phase to martensite as a result of plastic strain
Figure 13
Figure 14
3.4 Discussion
High volumetric fractions of austenite are found to persist after the formation of bulk material via
SLM. Application of a solution heat treatment at 1040 °C exceeds the transus temperature of 727 °C where
all martensite reverts to the parent austenite structure (α' ⟶ γ) and residual stresses have been eliminated.
Cooling to room temperature initiates the reverse transformation, γ ⟶ α', starting at ≈ 100 °C (MS) with a
transformation finish temperature of ≈ 32 °C (Mf). Subsequent aging treatments nucleate and grow Cu-rich
precipitates that further coarsen with extended time and temperature. Reversion of martensite to austenite
aging treatment. These austenite grains are observed to nucleate on lath and parent austenite grain
boundaries. TEM observations show such formations interspersed within the lath structure of the overaged
CA-H1150 specimens.
condition with a rapid transition in work hardening and limited ranges of plastic strain before the onset of
incipient necking. Ductility of these specimens is limited due to early fracture from embedded porosity. For
CA-H1025 and CA-H1150 specimens, progressive reversion of the primary martensite to austenite, along
with the coarsening of the precipitate structure, both reduce yield and ultimate tensile strengths while also
extending the range of UEL and total elongation. Softening of the material via increased fractions of the
weaker phase and a likely transition to Orowan bowing around coarsened precipitates are reasoned for the
13
transition in behavior for overaged specimens.
As-fabricated specimens and those further aged without an initial solution heat treatment
demonstrate different behavior. Temperatures applied for all examined post-processing heat treatments do
exceed the temperature range over which the γ ⟶ α' martensitic transformation is to occur and therefore
maintain austenite stability throughout treatment. The solubility of Cu within these austenite grains is
higher than the surrounding martensite and therefore does not precipitate. Their absence is noted in TEM
observations of H1150 specimens. However, the temperatures associated with the overaged conditions,
H1025 and H1150, do begin to sufficiently relieve the expected residual thermal stresses [19].
Consequently, austenite that remains stable during the heat treatment is then permitted to fully transform
into martensite upon cooling in the absence of mechanical stabilization. This is observed as reductions in
austenite volume fraction, as noted for H1025 and H1150 specimens via XRD spectra analysis.
Additionally, any metastable martensite that may reside in the as-fabricated material is permitted to
austenite stabilizing elements, these grains will revert to a stable lath martensite upon return to room
temperature. The lower temperature, peak-age treatment of H900 does not demonstrate the dramatic
reductions in retained austenite likely because the aging temperature is insufficient to relieve the
the austenite volume fractions for H900 are believed to originate from the stabilization that accompanies
the initial diffusion and precipitation of Cu in the fraction of the structure that is initially stable martensite.
With high volume fractions of austenite the H900 tensile specimens closely reproduce the stress-strain
response of the as-fabricated material, but at elevated stress levels due to the limited precipitation
hardening.
Enhanced ductility with rapid transitions in work hardening rates of the as-fabricated and H900
martensite. Due to the lower yielding stresses needed to initiate the martensitic transformation, stresses
within tensile specimens do not attain levels necessary to initiate premature failure by crack formation at
embedded porosity sites early in plastic strain. This behavior is in contrast to what is observed in CA-H900
specimens. For those with higher concentrations of austenite, plastic deformation is permitted to continue
until sufficient transformation to martensite raises localized stress to levels necessary to initiate necking
14
and ultimately failure.
With greater relief of thermal stresses accompanying over-aged treatments and accompanying
reductions in austenite concentration, the now largely martensite structures for H1025 and H1150
specimens do not carry the same ability to transform with applied stress. Work hardening rates for these
specimens rapidly transition across reduced ranges of strain. This is similarly reflected in the reduced UEL
and total elongation observed from the initial condition of the as-fabricated state. Reductions in volumetric
fractions of austenite for the H1025 and H1150 treatments, where martensite remains the primary structure,
do not show the same transformation induced strengthening associated with the as-fabricated and H900
conditions.
Quasi-static tensile testing was performed on specimens machined from bulk 17-4 PH produced via
selective laser melting. Mechanical performance varied greatly as result of the evolution of phases resulting
from the various thermal processing performed. From these evaluations several conclusions regarding the
1. 17-4 PH produced by SLM and in the as-fabricated state was measured to yield at reduced stress levels.
Onset and continued plastic deformation was dominated by the stress-induced transformation of the
retained austenite to martensite. Plastic deformation and specimen elongation via phase transformation
resulted in significantly increased work hardening rates over lengthened regions of uniform plastic
deformation. The delay of incipient necking was a result of the increased work hardening capacity
driven by martensite formation. Continued plastic deformation beyond necking remained relatively
short due to the presence of embedded porosity, but total elongation was enhanced despite the presence
of porosity.
2. Solution heat treatment of as-fabricated material was sufficient to eliminate the metastable austenite
phase. Subsequent aging of the material from this condition resulted in yield and tensile strengths
exceeding standardized minimums. Austenite reversion to levels as high as 20.4% by volume were
measured at increased aging temperatures and times. An increase in both uniform and total elongation
resulted.
3. Initially high volumetric fractions of austenite in the as-fabricated material inhibit strengthening by
15
aging. However, increased temperature heat treatments progressively transform SLM retained austenite
to martensite. The likely relief of residual stresses permits the transformation of austenite to martensite
upon cooling post-treatment. The over aging of the H1025 and H1150 heat treatments does not result in
the expected negative relative strengthening observed for material initially treated with a solution heat
4. Aging efficacy from the as-fabricated condition is limited due to the higher solubility of precipitating
elements within the austenite phase. Relative strengthening post treatment is reduced when compared
Acknowledgements
This research received financial support from Grants-in-Aid for Scientific Research in an
Innovative Area (24120008) from the Ministry of Education, Culture, Sports, Science, and Technology
(MEXT). The authors would also like to thank Dr. Takahiro Kimura of the Technology Research Institute
of Osaka Prefecture who provided assistance in the fabrication of our bulk material.
References
[1] Zhang, K., Wang, S., Liu, W., & Shang, X. Characterization of stainless steel parts by Laser Metal
[2] Amato, K. N., Gaytan, S. M., Murr, L. E., Martinez, E., Shindo, P. W., Hernandez, J., … Medina, F.
Microstructures and mechanical behavior of Inconel 718 fabricated by selective laser melting. Acta Mater
[3] Murayama, M., Hono, K., & Katayama, Y. Microstructural evolution in a 17-4 PH stainless steel
after aging at 400 °C. Metall Mater Trans A 1999;30(2):345–353. DOI: 10.1007/s11661-999-0323-2
[4] Wang, J., Zou, H., Li, C., Qiu, S., & Shen, B. The effect of microstructural evolution on hardening
behavior of type 17-4PH stainless steel in long-term aging at 350 °C. Mater Charact 2006;57:274–280.
DOI: 10.1016/j.matchar.2006.02.004
[5] Hsiao, C. N., Chiou, C. S., & Yang, J. R. (2002). Aging reactions in a 17-4 PH stainless steel. Mater
[6] LeBrun, T., Tanigaki, K., Keitaro, H., & Kobayashi, H. Strain rate sensitivity and mechanical
anisotropy of selective laser melted 17-4 PH stainless steel. Mechanical Engineering Journal, JSME,
16
2014;1(5):1–13. DOI: 10.1299/mej.2014smm0049
[7] Facchini, L., Vicente, N., Lonardelli, I., Magalini, E., Robotti, P. and Molinari, A. Metastable
Austenite in 17–4 Precipitation-Hardening Stainless Steel Produced by Selective Laser Melting. Adv Eng
Mater 2010;12:184–188.
[8] Murr, L. E., Martinez, E., Hernandez, J., Collins, S., Amato, K. N., Gaytan, S. M., & Shindo, P. W.
Microstructures and Properties of 17-4 PH Stainless Steel Fabricated by Selective Laser Melting. J Mater
[9] ASTM Standard A564/A564M-13 (2013). Standard Specification for Hot-Rolled and Cold-Finished
Age-Hardening Stainless Steel Bars and Shapes, ASTM International, DOI: 10.1520/A0564_A0564M,
www.astm.org
[10] Li, R., Shi, Y., Wang, Z., Wang, L., Liu, J., & Jiang, W. Densification behavior of gas and water
atomized 316L stainless steel powder during selective laser melting. Appl Surf Sci 2010;256(13):4350–
[11] Syed, W. U. H., Pinkerton, A. J., & Li, L. A comparative study of wire feeding and powder feeding
in direct diode laser deposition for rapid prototyping. Appl Surf Sci 2005;247(1-4): 268–276. DOI:
10.1016/j.apsusc.2005.01.138
[12] Santos, E. C., Shiomi, M., Osakada, K., & Laoui, T. (2006). Rapid manufacturing of metal
10.1016/j.ijmachtools.2005.09.005
[13] ASM International. Handbook Committee. ASM Handbook. Vol. 1., “Properties and Selection:
[14] Mirzadeh, H., & Najafizadeh, A. Aging kinetics of 17-4 PH stainless steel. Mater Chem Phys,
[15] Hollomon, J. H., & Jaffe, L. D.. Time-temperature relations in tempering steel. Trans. AIME
1945;162:223-249.
[16] Lutterotti, L., Matthies, S., Wenk, H. R., Schultz, A. S., & Richardson Jr, J. W. Combined texture
and structure analysis of deformed limestone from time-of-flight neutron diffraction spectra. J Appl Phys
[17] Lonardelli, Ivan, et al. Texture analysis from synchrotron diffraction images with the Rietveld
17
method: dinosaur tendon and salmon scale. J Synchrotron Radiat 2005;12(3):354-360. DOI:
10.1107/S090904950500138
[18] Viswanathan, U. Kinetics of precipitation in 17–4 PH stainless steel. Mater Sci Tech
[19] Starr, T., Rafi, K., Stucker, B., & Scherzer, C. Controlling phase composition in selective laser
melted stainless steels. In Proceedings of the Solid Freeform Fabrication Symposium 2012;439–446
[20] Colaço, R., & Vilar, R. Stabilisation of retained austenite in laser surface melted tool steels. Mat Sci
[21] Salje, G., & Feller-Kniepmeier, M. The diffusion and solubility of copper in iron. J Appl Phys
18
Figure 1: SEM imagery of the pre-alloyed 17-4 PH powder used to fabricate the bulk primitives.
Note the dual shape morphology attributable to the water atomization process.
Figure 3 : Nominal stress-strain curves for (a) samples initially solution treated prior to aging, (b)
samples aged without an initial solution heat treatment. The indicated aged condition is shown
adjacent to the terminus of the stress-strain trace. For (a) the initial condition, Condition A, is
shown as a dashed trace and labeled ‘CA’. For (b) the as-fabricated condition is shown as a
Figure 4 (a): Variation of work-hardening rate across UEL for heat treated specimens with a
preceding solution heat treatment. (b) Variation of work-hardening rate across UEL for heat
treated specimens without a preceding solution heat treatment. The as-fabricated condition is also
included.
Figure 5: Effect of tempering parameter, P, on aging behavior for the two different sets of tensile
term, P=T(C + log(t)) × 10 , where T is the absolute temperature in Kelvin, t is the tempering
-3
time in hours, and C is the materials constant (ranging from 10 to 20, 20 is commonly used in
many applications).
Figure 6: XRD spectra for material samples aged concurrently with tensile specimens.
Figure 7: (a) TEM Bright field image of a material sample aged to CA-H1150. (b) Dark field
image showing fcc reflection of reverted austenite. (c) Corresponding SAED pattern with
19
twinning indicated by the secondary spots in the austenite structure. [1 2] diffraction plane
shown.
Figure 8: TEM bright field image of the copper precipitate structure of an overaged CA-H1150
specimen. Precipitates were of circular or oblong shaped with long axes approximately 20 nm in
size. Some, but not all, precipitates are associated with dislocations. Arrows indicate example
precipitates.
Figure 9: (a) TEM bright field image of a material sample aged to H1150 (without a preceding
solution heat treatment). (b) Dark field image showing fcc reflection of retained austenite
structure. (c) Corresponding SAED pattern with twinning. fcc [011] diffraction plane shown.
Figure 10: XRD spectra as measured from within the both the gauge section and threaded grips
Figure 11: (a) Angled view of a CA-H900 specimen. (b) Angled view of a H900 specimen. (c)
Figure 12: Angled view of a planar pore geometry penetrating into the tensile loading direction of
an as-fabricated specimen. The fracture surface is not limited to the pore’s periphery, but found to
locally remove surrounding material to form a divot in the fracture surface. Tensile loading
Figure 13: High magnification view of the rupture surface as imaged from the CA-H900
specimen. The orientation of the applied load was normal to this image.
20
Figure 14: High magnification view of the rupture surface as imaged from the as-fabricated
specimen. The orientation of the applied load was normal to this image.
21
Table 1: Composition of 17-4 PH Powder
Composition (w%)
C Mn P S Si Cr Ni Nb Cu Fe
Powdered Working Material 0.054 0.18 0.011 0.005 0.41 16.51 4.16 0.42 3.97 bal
ASTM A564/A564M Limits 0.07 1.00 0.040 0.030 1.00 15.00-17.50 3.00-5.00 NS 3.00-5.00 bal
Limits are in percent maximum unless shown as a range
NS – Not Specified
ASTM A564/A564M-10 minimum values are shown in parentheses for aging treatments that follow a solution heat treatment.
Condition A 100 0
22
Figure 1
Figure 2
1600
CA-H900
1400
1200
600
400
200
0
0 0.05 0.10 0.15 0.20
Nominal Strain
(a)
1600
1400
H900
H1025
1200
Nominal Stress (MPa)
H1150 As-Fabricated
1000
800
600
400
200
0
0.00 0.05 0.10 0.15 0.20
Nominal Strain