You are on page 1of 13

Materials Science & Engineering A 738 (2018) 44–56

Contents lists available at ScienceDirect

Materials Science & Engineering A


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

Impact of composition on the heat treatment response of additively T


manufactured 17–4 PH grade stainless steel
S.D. Mereditha,b, J.S. Zubackb, J.S. Keista, T.A. Palmerb,c,

a
Applied Research Laboratory, The Pennsylvania State University, University Park, PA 16802, USA
b
Department of Materials Science and Engineering, The Pennsylvania State University, University Park, PA 16802, USA
c
Department of Engineering Science and Mechanics, The Pennsylvania State University, University Park, PA 16802, USA

ARTICLE INFO ABSTRACT

Keywords: Precipitation hardened (PH) martensitic grade stainless steels are commonly used in additive manufacturing
Stainless steel (AM) processes, but the heat treatment response can vary depending upon the powder feedstock composition.
Precipitation hardening (PH) As-built AM 17–4 PH grade stainless steel fabricated using argon and nitrogen atomized feedstocks in separate
Additive manufacturing (AM) powder bed fusion systems responded differently to standard heat treatment cycles. Materials fabricated from
Nitrogen
argon atomized feedstocks, containing low levels of nitrogen (0.01 wt%) and retained austenite (< 1%), re-
Gas atomization
Aging
sponded as expected to standard solutionizing and aging heat treatment cycles. In contrast, materials fabricated
from nitrogen atomized feedstocks, containing between 0.06 and 0.12 wt% nitrogen and up to 81% retained
austenite, did not display peak aging with standard heat treatments and deviated from the expected overaging
response with increasing aging temperatures. At the highest nitrogen and retained austenite levels, peak aging is
found to occur at temperatures in excess of 680 °C, even though the material still contains retained austenite
levels in the 20% range. These unexpected changes in the heat treat response are closely linked to differences in
the nitrogen composition of the powder feedstock. Changes in the Ni and Cr equivalent values determined by
other primary alloying elements also impact the heat treat response, even though the alloy compositions still fall
within standard alloy element composition ranges.

1. Introduction processing, and it is seeing widespread use in powder bed fusion (PBF)
processes [10–19]. In the wrought condition, the nitrogen level is not
Precipitation-hardened (PH) martensitic grade stainless steels are specified, but AM processed 17–4 PH grade stainless steel can display
widely used in the aerospace, marine, and power generation industries wide variations in nitrogen concentration originating primarily from
[1–4]. Strengthening in these alloys is derived from a combination of the atomization conditions for the powder feedstock or to a lesser ex-
the martensitic microstructure and the formation of sub-micron copper- tent through the selection of processing gases. For example, previous
rich precipitates. These precipitates form during post-process heat studies [2,20] have shown that Ar-atomized feedstock was consistently
treatments, which can be tailored to change the precipitate size and martensitic, whereas nitrogen atomized feedstock, containing five times
shape and the hardness and strength of the post-processed component the nitrogen content [20], was primarily austenite. Increases in the
[2–6]. In the peak-aged (H900) condition [7], elliptical Cu-rich pre- nitrogen concentration can influence mechanical properties by sup-
cipitates approximately 25 nm in length are formed [5]. As the material pressing the martensite start temperature (Ms) and thereby promote
is over-aged, Ostwald ripening mechanisms become dominant, leading greater retention of austenite at room temperature [2,21–24], which
to a coarsening of the precipitates, a decrease in their number, a loss of could explain the large variations of measured properties [25]. In ad-
coherency with the matrix, and tempering of the martensitic matrix dition to being softer than martensite [2], higher levels of austenite
[4–6,8,9]. The combination of these changes in the precipitate and slow the formation of precipitates within the 17–4 PH grade due to the
matrix morphologies results in a decrease in mechanical strength but an higher solubility and lower diffusivity of copper within the austenite
improvement in ductility and impact toughness [1,7]. phase [3,20,26].
The weldability of 17–4 PH grade (UNS S17400) stainless steel Previous investigations of AM 17–4 PH grade stainless steels have
makes it an attractive alloy system for additive manufacturing (AM) been primarily directed at evaluating the heat treatment response of as-


Corresponding author.
E-mail addresses: ScottM@psu.edu (S.D. Meredith), jsz125@psu.edu (J.S. Zuback), jsk25@arl.psu.edu (J.S. Keist), tap103@psu.edu (T.A. Palmer).

https://doi.org/10.1016/j.msea.2018.09.066
Received 19 July 2018; Received in revised form 17 September 2018; Accepted 18 September 2018
Available online 19 September 2018
0921-5093/ © 2018 Elsevier B.V. All rights reserved.
S.D. Meredith et al. Materials Science & Engineering A 738 (2018) 44–56

built material through hardness measurements in the solutionized concentrations than the argon atomized powder feedstocks.
[6,24,27], solutionized and aged [6,12,13], and aged only conditions Due to differences in the recoating and process parameters for the
[2,3,6,11]. For example, Murr et al. [2] applied conventional peak- two systems, there are different powder size requirements for each AM
aging heat treatments directly to as-deposited laser PBF AM 17–4 PH PBF system used. While the powder suppliers specified the particle size
grade stainless steel and found that as-built materials which exhibited a distributions, other powder characteristics required additional analysis.
predominantly martensitic structure showed an increase in hardness. Each as-received powder feedstock was characterized using a range of
On the other hand, other as-built materials, which were primarily conventional powder metallurgy methods for evaluating the flowability
austenitic, did not respond to the aging heat treatment. Similarly, Rafi [30], apparent and tap density [31,32], as well as the particle size
et al. [3] observed a higher proportion of retained austenite in the as- distribution [33]. A standard Hall funnel [30] was used to measure
built material when nitrogen atomized powder was processed under a flowability by the Hall flow test and combined with a 25 cm3 brass cup
nitrogen atmosphere (resulting in 0.15 wt% N) rather than an argon [31] to measure apparent density. Average values were calculated from
atmosphere. An initial heat treatment (788 °C, 2 h), which falls below at least three measurements for each powder. Since no powder flowed
the standard solutionizing temperature (1038 ± 14 °C) [28], raised the freely, flow was assisted by continuous agitation of the powder. An
yield and ultimate tensile strengths, but no additional hardening was average tap density was also determined from at least three measure-
observed following a subsequent standard peak-aging heat treatment. ments, after settling the powder with a QuantaChrome Dual Autotap
Additional work directed at exploring the heat treatment response of model DA-1. The particle size distributions of the powders were as-
water atomized powder feedstocks, which are not as susceptible to wide sessed by both a light scattering technique (Malvern Masersizer 2000)
changes in nitrogen concentrations, also has shown varying degrees of as well as image analysis (Malvern Morphologi® G3), with the latter
retained and reverted austenite. Aging of these AM materials directly providing additional information on particle shape, including mea-
from the as-built condition resulted in mechanical properties that de- surements of the convexity and circularity of the particles. Convexity
viated from those obtained in wrought materials subjected to peak- represents the ratio of the convex hull perimeter of two-dimensional
aging and overaging heat treatments [6]. profiles of the particles to the actual perimeters, and therefore ap-
Although standard wrought heat treatments are insufficient for proaches unity as surface roughness decreases. Circularity accounts for
consistently achieving peak aging in AM 17–4 PH grade stainless steel deviations in either surface roughness or shape by comparing the actual
[6,12], the role of composition on the retained austenite levels and the particle perimeters to the circumference of a circle with an equivalent
resulting heat treat response has not been systematically addressed. In area to the two-dimensional profile.
order to identify the role of chemistry on the resulting heat treatment The measurement of retained austenite fractions in the powders and
response, as-built and post-process heat treated AM 17–4 PH grade AM fabricated materials was performed using X-Ray Diffraction (XRD)
stainless steels produced using several nitrogen and argon atomized techniques on a PANalytical X′Pert Pro MPD with an Empyrean Cu
powder feedstocks were characterized and compared to the expected anode (λ = 1.54059 Å) operated at 45 kV and 40 mA. XRD data were
heat treat response for wrought components. The heat treat response of obtained using a PIXcel 1D detector with a 0.020 mm Ni β filter,
AM material fabricated from argon atomized powder feedstocks with 0.04 mm Soller slit, 10 mm beam mask, and a 2° antiscatter slit. A 2θ
low nitrogen concentrations, on the order of 0.01 wt% N, is closest to range of 40°–100° was used with an automatic spot size of 5 or 10 mm,
that expected in wrought materials. On the other hand, nitrogen ato- depending upon the size of the specimen. Individual peaks, which are
mized feedstocks, which can contain nitrogen concentrations in excess schematically shown in Fig. 1, were indexed and the integrated in-
of 0.12 wt% N, display extremely large retained austenite levels on the tensities for each peak measured using MDI Jade 2010 (version 3.6.5).
order of 81%, which shift peak aging temperatures to levels in excess of The amount of retained austenite was then calculated using the direct
680 °C. While nitrogen concentrations play a prominent role in de- comparison method [34,35].
termining the heat treatment response, other alloying elements such as The most prominent peaks for each phase were generally FCC (111)
Cr, Ni, C, Mo, and Mn can also impact the amount of retained austenite and BCC (110) at approximate two-theta values of 43.6° and 44.5°,
[29] and heat treatment response of this alloy system. The role of these respectively. These two peaks were used to determine the lattice spa-
alloying elements, even though they fall within the specified composi- cing by Bragg's Law, from which the lattice constant and unit cell vo-
tion ranges, are also shown to impact the resulting heat treat response. lume could be calculated. The proportional parameter Rihkl [34] was
then determined for each hkl peak using tabulated values [36] for the
multiplicity factor, atomic scattering factor, temperature factor, and the
2. Experimental
structure factor appropriate for each phase, as listed in Table 2. The
austenite volume fraction, Vγ, was then calculated using the following
The specified alloying element compositions for wrought 17–4 PH
relationship [34]:
grade stainless steel are listed in Table 1 [7]. Like other commercial
alloy systems, individual alloying element compositions can vary over q P q
1 I j 1 Ii 1 I j
rather wide ranges and still meet chemistry requirements. Since the V = / +
q R P R i q R
standard specification for this alloy does not include nitrogen, its j=1 j i=1 j=1 j
(1)
composition can vary rather widely and impact the properties of AM
fabricated material in unknown ways. In order to investigate the impact where the ratio of integrated intensity, I, to the corresponding pro-
of nitrogen composition on the heat treat response, four powder feed- portional parameter, R, is averaged over the number of peaks con-
stocks satisfying the 17–4 PH martensitic grade stainless steel chemical sidered for each phase over the selected two-theta range (40–100°). For
requirements [7] were sourced from selected powder vendors to meet this study, the value of P corresponds to four peaks for the α phase and
the particle size distribution suitable for use on a 3D Systems ProX 200 q corresponds to five peaks attributed to γ phase. Due to the low carbon
and an EOS M280 PBF system. Nitrogen atomized and argon atomized content of this ferrous alloy system (≤ 700 ppm), the lattice parameters
powder feedstocks were obtained for each AM PBF system, and a of BCT martensite would not significantly differ from BCC ferrite [37],
summary of the powder feedstock chemistries, measured at a certified so the two phases were considered comparable in XRD analysis.
testing laboratory,1 is provided in Table 1. It is clear that the choice of The appropriate powder feedstocks were then used on 3D Systems
atomization gas impacts the nitrogen concentration, with the nitrogen ProX 200 and EOS M280 laser PBF systems using the recommended
atomized powder feedstock displaying much higher nitrogen operating parameters to fabricate a series of simple build geometries.
Each build consisted of a series of bars, each having an approximate
cross-section of 400 mm2, with build heights between 4.31 and
1
Westmoreland Testing & Research, Inc., Youngstown, PA. 9.52 mm. During processing, either an argon or nitrogen atmosphere

45
S.D. Meredith et al. Materials Science & Engineering A 738 (2018) 44–56

Table 1
Summary of chemical compositions for wrought [7], nitrogen and argon atomized powder feedstocks, and as-built materials for 17–4 PH grade stainless steels.
Additional calculations of the Nickel and Chromium Equivalents as calculated per the DeLong constitution diagram [42] as well as austenite volume fraction as
measured by XRD.
3D Systems ProX 200 EOS M280

Ar atomized feedstock N2 atomized feedstock Ar atomized feedstock N2 atomized feedstock

Specified ranges (wt%) Praxair lot As-built As-built Phenix lot As-built As-built LPW lot As-built EOS lot As-built
FE-276–3 (Ar) (N2) 14D1147 (Ar) (N2) UK5032 (N2) F471301 (N2)

15.0–17.5 Cr 16.6 15.6 15.7 15.8 15.7 15.7 16.4 16.5 15.2 15.2
3.0–5.0 Ni 4.3 4.2 4.2 4.1 4.2 4.2 4.1 4.3 4.3 4.6
3.0–5.0 Cu 3.2 3.3 3.3 3.7 3.9 4.0 4.0 4.2 4.3 4.3
≤ 1.00 Si 0.81 0.70 0.68 0.80 0.66 0.66 0.43 0.38 0.72 0.67
≤ 1.00 Mn 0.19 0.14 0.17 0.61 0.54 0.51 0.19 0.15 0.68 0.57
0.15–0.45 Nb(+Ta) 0.20 0.22 0.23 0.24 0.24 0.23 0.29 0.25 0.26 0.25
≤ 0.07 C 0.01 0.005 0.010 0.03 0.04 0.03 0.02 0.017 0.06 0.058
≤ 0.040 P 0.013 0.010 0.010 0.017 0.019 0.019 0.016 0.019 0.008 0.018
≤ 0.030 S 0.006 0.004 0.004 0.004 0.002 0.003 0.002 0.002 0.004 0.005
— Mo 0.13 0.012 0.020 0.92 0.098 0.100 0.06 0.009 0.89 0.098
— N 0.01 0.010 0.017 0.06 0.088 0.091 0.01 0.027 0.12 0.142
— O 0.06 0.052 0.085 0.09 0.084 0.098 0.04 0.036 0.02 0.031
Delong equivalent Nieq (wt%) 5.0 4.8 5.1 7.1 8.3 8.1 5.1 5.7 10.0 10.8
Creq (wt%) 18.0 16.8 16.8 18.0 17.0 16.9 17.3 17.2 17.3 16.4
Creq/Nieq ratio 3.6 3.5 3.3 2.5 2.1 2.1 3.4 3.0 1.7 1.5
%Austenite < 1% 0% 0% 20% 9% 14% 3% < 1% 97% 81%

alloying element compositions in the as-built condition is shown in


Table 1.
Selected samples were then subjected to a series of heat treatments
to measure the heat treatment response in the as-built AM material.
Samples were placed in a preheated, argon-purged CM, Inc Rapid Temp
Furnace for a selected time and then allowed to air cool to room tem-
perature. Three sets of standard wrought heat treatment cycles were
performed on the as-built AM material, including direct aging heat
treatments, solution only heat treatments, and a two-stage solution and
aging heat treatment. For the standard solution heat treatment, a re-
commended temperature of 1040 °C [28] was used, but since a specific
duration is not specified, times of 15, 30, 60, and 90 min were in-
vestigated. An initial solutionizing step of 30 min was used for the
combined heat treatment, and aging treatments of four hours in dura-
Fig. 1. Representative XRD pattern showing observed peaks in the 40–100°
two-theta range for AM 17–4 PH material.
tion were carried out at temperatures of 495 °C, 580 °C, and 620 °C to

Table 2
Summary of coefficients for analyzing X-ray diffraction data for measurement of retained austenite levels.
h k l Atomic scattering Structure factor, Multiplicity factor, p Lorentz polarization Temperature factor, Calculated proportional
factor, fFe F factor, LP e−2M parameter, Ri

BCC phase 1 1 0 17.5 2f 12 11.4 0.95 283.2


2 0 0 14.8 2f 6 4.9 0.92 41.6
2 1 1 13.0 2f 24 3.1 0.88 80.1
2 2 0 11.8 2f 12 2.7 0.84 27.3
FCC phase 1 1 1 17.6 4f 8 11.9 0.96 211.2
2 0 0 16.5 4f 6 8.4 0.94 97.7
2 2 0 13.7 4f 12 3.7 0.89 55.2
3 1 1 12.3 4f 24 2.8 0.86 65.9
2 2 2 11.9 4f 8 2.7 0.85 19.8

was used on the 3D Systems ProX 200 system, while only a nitrogen replicate the H925, H1075, and H1150 standard heat treatments, re-
atmosphere was used on the EOS M280 system. Fabricated specimens spectively [7]. Additional aging temperatures of 550 °C, 650 °C, 680 °C,
were removed from the base plate by electrical discharge machining. 720 °C, and 760 °C, after solutionizing for either 30 or 60 min, were also
The chemistry of the as-built material from each build was then ana- performed on selected specimens.
lyzed at a certified testing laboratory2 using a combination of com- Prior to XRD characterization and microhardness measurements,
bustion infrared detection (C, S), inert gas fusion (N, O), and direct specimens were mounted in an epoxy thermosetting powder, and then
current plasma emission spectroscopy. A summary of the measured ground through a series of silicon carbide abrasive grinding media to a
FEPA P4000 rating microgrit. Specimens were then polished with a
3 µm diamond suspension followed by a 1 µm diamond suspension for
2 2 min each on a Struers Pedemax-2. A final polishing step using a
Luvak Laboratories, Inc; Boylstown, MA.

46
S.D. Meredith et al. Materials Science & Engineering A 738 (2018) 44–56

0.06 µm colloidal silica suspension was performed for 8 min to produce addition, the higher surface area associated with fine particle sizes also
a mirror finish. A Nikon Epithot microscope equipped with a Digital promotes cohesion through Van der Waals forces and high surface en-
Sight DS-Fi2 camera was used to capture optical micrographs of the ergy [40]. These factors inhibit flowability and resulted in the higher
fabricated material following 6–11 s immersion in a solution of Marble's flow times and angle of repose for the finer powder feedstock. The other
etchant (50 ml HCl + 50 ml H2O + 10 g CuSO4) [38] diluted with an powder feedstocks display similar particle size distributions and fewer
equal weight of glycerol. To assess the hardening response throughout differences in other measurements that might impact the resulting
the heat treatments, a minimum of 30 Vickers microhardness mea- fabrication of test specimens. Image analysis results of the powders
surements were recorded on each specimen at a load of 300 gF with a shown in Table 3 revealed that flowability generally worsened as cir-
Leco M-400-G1 Hardness Tester. Since hardness specifications for cularity and convexity, which is a measure of surface roughness, de-
wrought 17–4 PH grade are provided on the Rockwell C scale [7], a viated from the ideal value of unity.
hardness conversion table for non-austenitic steels was used to identify In addition to these traditional particle characterization measure-
the corresponding Vickers hardness numbers [39]. A Philips XL-30 ments, the chemical compositions of the argon and nitrogen atomized
Environmental Scanning Electron Microscope was used to capture powder feedstocks were measured and summarized in Table 1. Even
secondary electron (SE) images of the powders. Energy-dispersive X-ray though the alloying elements in each powder feedstock fall within the
spectroscopy (EDS) maps were acquired using Oxford Instruments specified composition limits, there are notable differences in composi-
Aztec 3.1 SP1 software in conjunction with an X-MaxN detector (Model tions for the Cr, Cu, and Mn alloying elements. The switch from argon
51-XMX1005) affixed to an FEI Helios NanoLab 660 scanning electron to nitrogen atomization also produces rather significant differences in
microscope. the nitrogen composition. In the nitrogen atomized feedstocks, the ni-
Since 17–4PH stainless steel is a multicomponent alloy with over 10 trogen concentrations fell in a range between 0.06 wt% and 0.12 wt%,
alloying elements comprising roughly one fourth of the alloy mass, a while the argon atomized feedstocks displayed nitrogen concentrations
computational thermodynamic approach was used to investigate phase at the 0.01 wt% level. These differences are important because nitrogen
equilibria and transformation temperatures. All calculations were per- is a well-known austenite stabilizer and has an impact on the micro-
formed using JMatPro® (Sente Software Ltd., United Kingdom), which structure formed during AM processing [2,3,24].
uses the CALPHAD approach to determine phase equilibria as a function Differences in the other alloying element compositions, such as Cr,
of temperature and alloy composition. To compare phases of builds Ni, and Cu, among others, have additional impacts on both solidifica-
fabricated with different feedstocks, equilibrium calculations were tion and solid state phase transformations during cooling and post
performed spanning temperature ranges to include the solid and liquid process heat treatments. The impact of changes in these alloying ele-
regions. Additionally, Scheil simulations were used to compare the re- ments can be captured through the calculation of Cr and Ni equivalent
lative amounts of ferrite and austenite formed upon solidification. The values (Creq and Nieq) and the use of the Schaeffler diagram [41]. The
Scheil calculations take elemental segregation into account and use the Creq and Nieq values are calculated based upon the following relation-
assumptions that there is no diffusion in the solid phases, diffusion in ships [42] and summarized in Table 1:
the liquid is infinitely fast, and local equilibrium exists at the solid-
Ni eq (wt%) = [Ni] + 30 × [C] + 30 × [N] + 0.5 × [Mn] (2)
liquid interface. To highlight the effect of nitrogen content on mar-
tensite formation, the chemical compositions for each fabricated com- Creq (wt%) = [Cr] + [Mo] + 1.5 × [Si] + 0.5 × [Nb] (3)
ponent were used to estimate the martensite start (Ms) temperature as a
function of nitrogen composition while assuming the other elements where all elements are given in wt%. While the Creq values do not
remained as measured. The calculations were performed exclusively for display significant differences with changes in the atomization gas,
the austenite composition to account for some solute being consumed there is a large increase in the Nieq values for the nitrogen atomized
by secondary phases coexisting in equilibrium with austenite. feedstocks. These changes in Nieq values have a significant impact on
the Creq/Nieq ratio, with the nitrogen atomized powder feedstocks
3. Results and discussion displaying much lower values (1.7–2.5) than the argon atomized
powder feedstocks (3.4–3.6). In general, the higher Nieq values and
3.1. Role of atomization condition on powder feedstock lower Creq/Nieq ratios indicate that the austenite phase becomes more
stable.
Differences in the two AM PBF systems employed here require These Creq/Nieq ratios can also help predict the solidification mode
powder feedstocks with different particle size distributions. These dif- for a given stainless steel composition when plotted on the Schaeffler
ferent particle size requirements are based, in part, on the recoating diagram [41], as shown in Fig. 3. Values corresponding to the powder
systems and powder spreading used for each machine. For example, the feedstock compositions in this work are compared to the highlighted
3D Systems ProX 200 uses a blade and roller system to place the in- region representing the elemental ranges specified for wrought 17–4 PH
dividual powder layers, whereas the EOS M280 system uses a stiff metal grade material [7], as well as reported chemistries in the as-built con-
recoating blade. Images of the as-received powder feedstocks are shown dition from other AM studies. The argon atomized powder feedstocks
in Fig. 2(a) and (b) for those used in the 3D Systems ProX 200 and for both systems and the nitrogen atomized feedstock for the 3D Sys-
Fig. 2(c) and (d) for those used in the EOS M280 system. In all cases, the tems ProX 200 system fall in the specified 17–4 PH grade range on the
gas atomization process produces relatively spherical particles, with the plot. For these Creq/Nieq ratios, which fall in a range between 2.5 and
powders used for the EOS M280 appearing to be coarser than those 3.6, the solidified material is expected to contain a combination of
used for the 3D Systems ProX 200 system. ferrite and martensite, with some austenite also present in the nitrogen
Results of additional characterization of the powder size, mor- atomized feedstock for the 3D Systems ProX 200 system. However, the
phology, and associated flowability and density measurements are nitrogen atomized feedstock for the EOS M280 system has a Creq/Nieq
summarized in Table 3. Differences in the particle size distribution for ratio of 1.7 and falls in a region of the Schaeffler diagram that is
the powder feedstocks used in each system are apparent, even though dominated by austenite during solidification. Clausen et al. [23] simi-
the powders generally have a D50 value around 28 µm. The nitrogen larly observed a fully austenitic product in this region while Stoudt
atomized powder feedstock for the 3D Systems ProX 200 machine, et al. [24,27] reported a retained austenite level of 50%. Retained
which is the size recommended by the machine manufacturer, dis- austenite fractions of 36% [6], 72% [43], 62% and 94% [44] have also
played the lowest particle sizes, which fall between 4 and 30 µm. Since been observed at lower Nieq values, but the nitrogen compositions were
finer particles entrap more air during packing, the apparent density was not reported and not taken into account when calculating these values.
considerably lower than that displayed by the other powders. In Measurements of the retained austenite using XRD techniques

47
S.D. Meredith et al. Materials Science & Engineering A 738 (2018) 44–56

Fig. 2. Micrographs of feedstock powders for (a) argon atomized for ProX 200, (b) nitrogen atomized for ProX 200, (c) argon atomized for EOS M280, and (d)
nitrogen atomized for EOS M280.

confirm the presence or absence of austenite with the changes in


composition of the powder feedstocks. The argon atomized powder
feedstocks, which displayed nitrogen concentrations in the range of
0.01 wt% and Creq/Nieq ratios between 3.4 and 3.6, exhibited low levels
of retained austenite, ranging from < 1–3% volume fractions. In con-
trast, the nitrogen atomized powders exhibited much higher levels of
retained austenite, which can be correlated with the much higher ni-
trogen concentrations. For example, the nitrogen atomized powder
feedstock for the 3D Systems ProX 200 displayed a retained austenite
level of 21% with a nitrogen concentration of 0.06 wt% and a Creq/Nieq
ratio of 2.5. However, the highest retained austenite level of 97% was
observed in the nitrogen atomized powder feedstock for the EOS M280,
which had a nitrogen concentration of 0.12 wt% and a Creq/Nieq ratio of
only 1.7, largely driven by its high Nieq value of 10.
Fig. 3. Modified Schaeffler diagram [41] showing the range of Nieq and Creq
3.2. Impact of powder feedstock composition on as-built material values expected for the wrought specification as well as the four powder
feedstocks and as-built compositions from literature as calculated from the
In order to evaluate the impact of the different powder feedstock DeLong diagram equations [42].
compositions on the resulting as-built microstructures, builds were
fabricated from each of the powder feedstocks sourced for the re- After processing, the compositions of the primary alloying elements
spective PBF machines. However, the selection of the shielding gas used and retained austenite levels were measured and compared with those
to establish the processing atmosphere can also impact the as-built in the respective powder feedstocks and are summarized in Table 1.
chemistry. Separate builds processed with either argon or nitrogen Overall, there was little change in the compositions of the primary al-
shielding gases was performed on the 3D Systems ProX 200 machine, loying elements, such as Cr, Ni, and Cu, for all the builds except the
while only nitrogen shielding gas was used on the EOS M280. argon atomized feedstocks used in the 3D Systems ProX 200. In this
Representative microstructures orientated perpendicular to the layer case, a decrease in Cr content for both the argon and nitrogen proces-
planes are shown in Fig. 4. EDS maps did not exhibit observable ele- sing conditions is observed, but the change in Cr content still falls
mental segregation for components fabricated on the EOS M280 from within the specified composition range. This loss in Cr content, though,
either the argon atomized or nitrogen atomized feedstock.

Table 3
Summary of powder characterization results for the 17–4 PH grade stainless steel powder feedstocks.
Feedstock PSD (volume basis) Image analysis App. density (g/cm3) Tap density (g/cm3) Hall flow ratea (s/50 g) Angle of repose ( ± 0.5°)

D10 D50 D90 Convexity Circularity


(µm) (µm) (µm)

ProX Ar atom. 15 27 47 0.936 0.909 3.95 5.0 24.8 25.5°


N2 atom. 4 13 30 0.992 0.969 2.72 4.7 290.2 51.5°
EOS Ar atom. 17 28 44 0.983 0.944 3.45 4.5 12.3 27.5°
N2 atom. 26 29 58 0.950 0.982 3.83 5.1 24.5 35.0°

a
Standard Hall flow method resulted in “no flow” condition for all powders; the listed flow times correspond to continuous agitation of the powder and should be
considered for relative comparisons only.

48
S.D. Meredith et al. Materials Science & Engineering A 738 (2018) 44–56

Fig. 4. Micrographs of as-built condition for (a) argon atomized and (b) nitrogen atomized for EOS M280 under N2 atmosphere.

had a minimal impact on the Creq/Nieq ratios for the as-built material, nitrogen does not seem to substantially increase the austenite fraction
indicating that this change in chemistry will have minimal impact on in the build, it has been shown in previous work to impact the micro-
the resulting solidification and microstructural development during AM structure and hardness of the martensite phase [45–47].
processing.
Even though changes in processing gases had no discernible impact
on the compositions of the other major alloying elements, the nitrogen 3.3. Heat treatment response of as-built material
and oxygen concentrations are more likely to be impacted by changes in
the processing conditions. When an argon shielding gas was used in the Given the differences in the as-built hardness and retained austenite
3D Systems ProX 200 system, there was little change in the nitrogen levels for the four powder feedstocks, changes in the heat treat response
levels in the builds when compared with the as-received powder com- of these materials are expected. Since several post-processing heat
position shown in Table 1. With the use of a nitrogen shielding gas, treatment routes have been previously explored
slight increases were observed in the nitrogen concentration of the as- [2,3,6,12,13,20,24,27], the impact of similar heat treatments on these
built material, but the overall magnitude of this increase was rather different as-built microstructures is investigated. The standard heat
minimal. The oxygen content varied within 0.03 wt% of the levels ob- treatments for wrought material includes an optional homogenizing
served in the feedstock, indicating that the process did not suffer from step at 1149 °C for at least 90 min [48] followed by a solutionizing step
significant atmospheric contamination. at 1040 °C (“Condition A”) and a subsequent precipitation heat treat-
A further examination of the as-built microstructures shows a ment, such as 1 h at 482 °C for the peak-aged condition [1,7,28]. The
change in the amount of retained austenite for the different powder first heat treatment routes to be investigated here involve overaging
feedstocks. Based on the retained austenite measurements summarized directly from the as-built condition. Measurements of the retained
in Table 1, there is a decrease in the measured retained austenite when austenite levels after aging for four hours at temperatures of 495 °C,
moving from the powder feedstock to the as-built microstructure for all 580 °C, and 620 °C are summarized in Table 4. The hardness results are
conditions. The retained austenite levels for both argon atomized graphically compared to the expected heat treatment responses for
powder feedstocks approach zero when processed on either system, standard solutionizing and aging heat treatments for wrought materials
regardless of shielding gas. In the case of the nitrogen atomized powder [7] in Fig. 5(a) and (b) for the 3D Systems ProX 200 and EOS M280
feedstocks, the retained austenite levels also decreased in the as-built processing systems, respectively.
conditions from that observed in the powder feedstock. For example, In the as-built condition, the hardness values for the different
the nitrogen atomized powder feedstock used in the 3D Systems ProX components consistently fall below the maximum hardness require-
200 system decreased from a level of 21% to between 9% and 14% ments for the solutionizing heat treatment. At these hardness levels, PH
when processed in argon and nitrogen shielding gases, respectively. The grades in the as-built condition have previously been considered
nitrogen atomized powder with the highest Nieq value (10) also showed equivalent to the solutionized condition specified for wrought materials
a decrease in retained austenite, albeit only from 97% to 81%. [3,49], and were subjected to follow on aging treatments without so-
These differences in the amount of retained austenite in the builds lutionizing [2,3,6,49]. When going from the as-built condition directly
fabricated from the different powder feedstocks correspond to differ- into an aging heat treatment, the hardness values change depending
ences in the hardness measurements. As shown in Table 4, the as-built upon the amount of retained austenite, which is dictated primarily by
material fabricated using the nitrogen atomized feedstock in the EOS the original atomization condition for the powders. For example, the
M280 system displayed the highest level of retained austenite (ap- argon atomized materials maintained low levels of retained austenite
proximately 81%) and the lowest hardness levels (260 ± 7 HVN). For regardless of system or shielding gas, and exhibit near-peak aging fol-
the builds fabricated from the argon atomized feedstock in both AM lowing treatment at 495 °C. While slightly higher hardness levels were
systems, little or no retained austenite is present, and the hardness le- observed when the 3D Systems feedstock was processed under nitrogen,
vels increase to approximately 300 HVN. However, the builds fabri- all of the argon atomized feedstocks satisfy the standard wrought
cated from the nitrogen atomized powder feedstock on the 3D Systems hardness requirements [7] for this condition. When the same feedstock
ProX 200 system, which display austenite levels of approximately 14% was processed under the different atmospheres, slightly higher hardness
when processed in nitrogen and 9% when processed in argon, display levels were observed for the builds processed under nitrogen.
much higher hardness values, on the order of 350 HVN. Even though As expected with overaging, the measured hardness levels in the
the as-built material fabricated under a nitrogen shielding gas displays argon atomized powder feedstock builds decrease when aged at higher
a higher retained austenite level, there is not a substantial increase in temperatures. At both the 580 °C and 620 °C aging heat treatment
hardness between the nitrogen and argon shielding gases. This increase temperatures, the argon atomized materials consistently fell below the
in hardness, though, appears to be related to the rather high nitrogen minimum hardness values for wrought materials [7]. Although the
concentration, which is in the range of 0.09 wt%. While this level of austenite fraction in the as-built materials was lower than the feedstock,
the austenite levels increased after aging, as shown in Table 4,

49
S.D. Meredith et al. Materials Science & Engineering A 738 (2018) 44–56

Table 4
Measured percentages of retained austenite for the powder feedstock, as-built material, and after various
post-processing steps.

indicating that reverted austenite formed. processed in this system displays a similar aging response as the cor-
The nitrogen atomized feedstock processed in the 3D Systems ProX responding builds processed in the 3D Systems machine, the highly
200, also shown in Fig. 5(a), generally met the hardness requirements austenitic material builds display the opposite behavior. As shown in
for the three aging temperatures [7]. Following the near-peak aging Fig. 5(b), the nitrogen atomized powder feedstock builds showed no
heat treatment at 495 °C, similar hardness levels to those obtained with change in the as-built hardness value following the 495 °C aging
the argon atomized feedstocks are observed, and the use of a nitrogen treatment. However, the hardness values began to increase as the aging
shielding gas also results in slightly higher hardness values. For the temperatures increased, with the highest hardness values observed after
overaging treatments at temperatures of 580 °C and 620 °C, the builds aging at a temperature of 620 °C. This behavior can be connected, in
fabricated using the nitrogen atomized feedstocks also easily met the part, to the high levels of retained austenite in these builds, which
minimum hardness values defined by the wrought standard [7]. An decrease as the aging temperature is increased. In the as-built condi-
increase in the hardness values is observed in the builds aged at 620 °C, tion, the retained austenite level is measured to be 81%, which in-
representing an unexpected deviation from the typical over-aging re- creases to 87% after aging at 495 °C and decreases to 40% after aging at
sponse but generally still satisfied the acceptable range of hardness 620 °C.
values. Since these builds were subjected to aging heat treatments without a
Deviations from the expected heat treat response of the as-built prior solutionizing heat treatment, the as-built microstructure, which
material are most pronounced in the nitrogen atomized feedstock builds was dominated by the rapid solidification conditions prominent in the
fabricated on the EOS M280. While the argon atomized feedstock PBF process, governed the aging response. The solutionizing heat

Fig. 5. Comparison of Vickers microhardness measurements made on materials aged directly from the as-built condition processed in the (a) 3D Systems ProX 200
and (b) EOS M280 system. The minimum and maximum specified hardness values for wrought materials shown as bars at each aging temperature [7].

50
S.D. Meredith et al. Materials Science & Engineering A 738 (2018) 44–56

Fig. 6. Summary of Vickers microhardness measurements for variations in solutionizing times on materials fabricated on the (a) 3D Systems ProX 200 and (b) EOS
M280 processing systems. The maximum specified hardness for solutionized wrought material is shown by the bar [7].

treatment, however, is designed to minimize this variation in the nitrogen atomized feedstock, however, displayed an inverse trend, with
starting microstructure prior to the aging heat treatment. While 30 min the hardness values increasing as the aging temperature increased.
at 1040 °C is the minimum recommendation [48], it is not clear whe- After the significant decline in retained austenite following the solution
ther this time frame is sufficient to adequately solutionize the as-built heat treatment, the retained austenite fractions measured after the
microstructure prior to aging. In order to evaluate the impact of the different aging heat treatments remained relatively unchanged.
duration of the solutionizing heat treatment, times between 15 and Since the hardness values were continuing to increase for the ni-
90 min at a temperature of 1040 °C were then investigated to determine trogen atomized feedstock, it could be deduced that the material is still
if the difference between a thermo-mechanically processed micro- in an underaged condition. In order to identify the peak aging condi-
structure and an AM as-built microstructure is significant. tions, a range of higher temperatures were investigated for the mate-
The impact of changing solutionizing times on the hardness of these rials fabricated from the nitrogen atomized feedstocks on the EOS M280
AM specimens is shown in Fig. 6(a) and (b) for builds fabricated in the system. Non-standard aging heat treatments up to 760 °C were used to
3D Systems ProX 200 and EOS M280 systems, respectively. Measure- monitor the continued aging response by the hardness and retained
ments of the retained austenite fractions were also performed and are austenite levels. Solutionizing times of both 30 min and 60 min were
summarized in Table 4. Similar to the as-built condition, the hardness tested to determine whether the longer duration would impact the
values generally fall below the specified limit for solutionized wrought aging response. As shown in Fig. 9, increasing aging temperatures up to
materials [7]. In the case of the argon atomized feedstocks, little to no approximately 680 °C produces an increase in hardness levels, with the
retained austenite was measured in the solutionized material, and the hardness at 680 °C producing a peak aged condition which meets the
retained austenite levels in the as-built nitrogen atomized material also general wrought specification for the standard H900 heat treatment
decreased. Among the fully martensitic components (< 5% FCC), in- (1 h at 482 °C)[7]. Aside from an initial increase in hardness at the
creased nitrogen and carbon content corresponded to higher hardness lower temperatures, the longer solutionizing time had minimal impact
values. For these materials, duration of the solutionizing heat treatment on the resulting aging behavior. Even with this shift to higher tem-
had no meaningful impact on the hardness measurements, and there- peratures, the retained austenite levels nevertheless remained around
fore a 30 min solutionizing time was maintained. However, the nitrogen 20% and did not significantly change.
atomized feedstock processed on the EOS M280 system, which had the
highest nitrogen concentration and Nieq value, still retained more than 3.4. Role of composition on phase equilibrium
20% austenite and exhibited a shift at longer solutionizing times.
The impact of a solutionizing heat treatment on the aging response Differences in the heat treat responses for the different materials can
of builds fabricated from these four powder feedstocks was then in- be correlated with the composition of the powder feedstocks. Each of
vestigated. Microstructures obtained after this combined solutionizing the alloying elements influence the evolution of the microstructure;
and aging heat treatments for the four powder feedstocks fabricated however, some elements, particularly nitrogen, have a much larger
under nitrogen are shown in Fig. 7. The corresponding hardness values effect on phase equilibrium. The selection of argon and nitrogen ato-
compared to the wrought specifications for these different heat treat- mization gases has the greatest impact on the nitrogen concentration in
ments are illustrated in Fig. 8(a) and (b) for the respective processing the powder feedstock, with nitrogen levels an order of magnitude
systems. Overall, the materials fabricated on the 3D Systems ProX 200 higher in the nitrogen atomized feedstock than in the argon atomized
from the argon and nitrogen atomized feedstocks, meet the wrought feedstock. Since the nitrogen concentration is not specified in the
specifications for each heat treatment [7], as shown in Fig. 8(a). While wrought condition, it can vary widely, as demonstrated in the measured
the argon and nitrogen atomized feedstocks have minimal amounts of nitrogen compositions of these four powder feedstocks summarized in
retained austenite when processed under argon, the same material Table 1.
shows increasing levels of retained austenite when processed with a These changes in chemistry can then lead to variations in the phase
nitrogen shielding gas. These higher nitrogen levels also consistently regions and solidification behavior for each of the powder feedstocks.
produce higher hardness levels across all of the heat treatment tem- As shown in the Schaeffler diagram in Fig. 3, the Nieq and Creq values
peratures. for the highly austenitic nitrogen atomized feedstock used in the EOS
On the other hand, the materials fabricated from the argon and M280 system place it outside the delineated region representing the
nitrogen atomized feedstocks in the EOS M280 system displayed op- wrought specification [7,41,42]. However, the powder feedstocks that
posite behaviors, as observed when aged directly from the as-built fall within this range display heat treatment responses much closer to
condition. As shown in Fig. 8(b), the material fabricated from the argon that expected for the wrought material. While the impact of these
atomized feedstock met the wrought specifications for hardness across compositions on the solidification can be captured here, the role that
the range of aging heat treatments [7]. The material fabricated from the changes in composition play on the phase equilibrium and resulting

51
S.D. Meredith et al. Materials Science & Engineering A 738 (2018) 44–56

Fig. 7. Microstructures following a two-stage


heat treatment of builds fabricated from each
of the feedstocks under a nitrogen atmosphere,
consisting of a 30 min solutionization plus
aged 4 h at the specified temperature, (a) argon
atomized for ProX 200, aged @495 °C; (b) ni-
trogen atomized for ProX 200, aged @495 °C;
(c) argon atomized for EOS M280, aged @
495 °C; and (d) nitrogen atomized for EOS
M280, aged @680 °C.

heat treat response can be further explored. Fig. 10(a) and (b) show the equilibrium phases for the argon and
Point calculations were used to determine the mass fraction of each nitrogen atomized feedstock chemistries fabricated on the EOS M280
phase and then plotted as a function of temperature for the chemistries system. Table 5 provides a summary of the temperature boundaries for
for each system and atomization condition. Liquid, austenite, ferrite, the phase regions calculated for each of the powder feedstock compo-
carbides/nitrides, and Cu phases were all considered in the calcula- sitions. Overall, the impact of the differences in chemistry between the
tions. For both systems, the amount of carbon and nitrogen in the nitrogen and argon atomized powder feedstocks is significant. While
powder feedstock is higher for the nitrogen atomized conditions. As a the argon atomized feedstock solidifies first as ferrite and continues to
result, the mass fractions of carbides and nitrides in the nitrogen ato- austenite, the first phase to solidify in the nitrogen atomized feedstock
mized feedstock is higher than in the argon atomized feedstock, and is austenite, with no ferrite appearing until a temperature of approxi-
their phase regions also extend to higher temperatures. However, the mately 700 °C. When elemental segregation is taken into account during
combined mass fraction of all carbides and nitrides is very small, in- solidification, the nitrogen atomized feedstock is fully austenitic com-
dicating that these phases have little to no effects on austenite and pared to just 35% austenite in the argon atomized material as the last
ferrite equilibrium. Additionally, the changes in powder chemistries for liquid solidifies under Scheil conditions. The change in chemistry be-
all conditions have little effect on the Cu-phase equilibrium as the mass tween the nitrogen and argon atomized powder feedstocks also im-
fractions and temperature ranges show negligible changes. pacted the temperature ranges for the liquidus and solidus regions as

Fig. 8. Summary of Vickers microhardness measurements for as-built materials subjected to a two-stage solutionizing and aging heat treatment fabricated on the (a)
3D Systems ProX 200 and (b) EOS M280 processing systems. The minimum and maximum specified hardness values for wrought materials shown as bars at each
aging temperature [7].

52
S.D. Meredith et al. Materials Science & Engineering A 738 (2018) 44–56

EOS M280 feedstock chemistries. In addition, the changes in the tem-


peratures, as summarized in Table 5, are much lower than those ob-
served for the powder feedstocks used in the EOS M280 system.
In addition to the impact of these composition changes on the phase
regions for the primary phases, there is also a rather significant impact
on the martensite start (Ms) temperatures. In general, lower values of Ms
that approach room temperature signify that the martensite transfor-
mation is less likely to complete, resulting in higher fractions of re-
tained austenite. The Ms temperature for the different powder feed-
stocks is often experimentally determined using a linear combination of
alloying elements in the following form:

Ms = k 0 + ki (wt %i) (4)


i

where k0 is a positive constant and ki are coefficients that highlight the


effect of each alloying element on martensite formation. The de-
Fig. 9. Summary of Vickers microhardness measurements on as-built materials termined ki coefficients for interstitial elements like C and N are often
subjected to a combined solutionizing and aging heat treatment for nitrogen an order of magnitude larger than those for common substitutional
atomized feedstocks processed in the EOS M280 system. The minimum and elements such as Cr, Ni, Mo, and Mn [50]. The accuracy of these em-
maximum specified hardness values for the peak aged H900 heat treatment for
pirical approaches often suffers when extending the relations to wider
wrought materials shown as dashed bars [7].
ranges of chemical compositions, and some important elements are
often not investigated. Thermodynamic approaches are more favorable
since they allow for more versatility over larger ranges of chemical
composition and do not treat each element exclusively as an austenite
or ferrite stabilizer [51]. In these calculations, the Ms temperature is
defined as the temperature at which the difference between the Gibbs
free energy of the austenite and ferrite phases reaches a critical value
for martensite nucleation [52]. The calculations are performed for the
austenite composition, removing the possibility of secondary phases
impacting the martensite transformation from the austenite parent
phase. Although previous work has shown larger prior austenite grain
size to favor higher Ms temperatures [50], the calculations do not take
this effect into account.
Fig. 12 illustrates how the Ms temperature changes with both the
composition of the as-built material and the nitrogen concentration.
Each curve on the plot represents the composition measured for each of
the six builds, with the actual nitrogen concentration marked, and
clearly reveals the role of chemistry on the Ms temperature. Even dis-
regarding differences in nitrogen concentration, the variation in com-
positions within the different components results in a change of nearly
60 °C in the Ms temperature. When changes in nitrogen concentration
are taken into account, the disparities between the predicted Ms tem-
perature become amplified. The highest projected Ms temperatures
corresponds to the builds with the lowest nitrogen concentrations and
retained austenite measurements, while the Ms temperature for the
components with the highest nitrogen concentrations and retained
austenite levels is estimated to be nearly 100 °C lower.

4. Summary and conclusions

The heat treatment response of AM PBF 17–4 PH grade stainless


steels has been shown to deviate from standard wrought heat treatment
practices. These variations in the heat treat response have been corre-
lated with changes in the composition of the powder feedstock, which
Fig. 10. Plots showing the phase stability over the solid and liquid regions for can be further connected with the gas atomization conditions. As-built
the (a) argon atomized feedstock and (b) nitrogen atomized feedstock materials
material fabricated using argon and nitrogen atomized feedstocks in
fabricated in the EOS M280 system.
different PBF systems has been subjected to a series of post process heat
treatment cycles, including aging directly from the as-built condition
well as the solid state austenite and ferrite phase fields. and following an initial solutionizing heat treatment. AM material
Changes in the phase regions for the argon and nitrogen atomized fabricated from argon atomized feedstocks, which contained 0.01 wt%
powder feedstocks fabricated on the 3D Systems ProX 200 system, nitrogen and retained austenite levels < 1%, displayed heat treatment
however, are much less pronounced. As shown in Fig. 11(a) and (b), the responses most closely resembling those expected in wrought materials.
change in atomization conditions and nitrogen concentrations do not In contrast, nitrogen atomized feedstocks, containing 0.12 wt% ni-
significantly impact the primary phase regions for the austenite and trogen and up to 81% retained austenite significantly deviate from
ferrite phases. Upon Scheil solidification, the solid fraction of austenite standard wrought heat treatment responses. For these highly austenitic
for both atomization conditions is much more comparable than for the components, the observed peak aging temperature increases from

53
S.D. Meredith et al. Materials Science & Engineering A 738 (2018) 44–56

Table 5
Summary of calculated equilibrium phase regions and volume fractions during Scheil solidification for the four powder feedstock chemistries.
3D Systems ProX 200 EOS M280

Ar atomized N2 atomized Ar atomized N2 atomized

Equilibrium calculations
Liquidus temperature (°C) 1453 1441 1444 1429
Solidus temperature (°C) 1384 1362 1367 1367
Austenite region(s) (°C) 1399–501, Below 431 Below 1398 1409–519, Below 405 Below 1429
Ferrite region(s) (°C) 1453–1104, Below 800 1441–1184, Below 765 1444–1202, Below 781 Below 695
Cu phase formation temperature (°C) 944 986 994 1027
Scheil solidification calculations
Solidification range (°C) 1453–1225 1441–1225 1444–1230 1429–1245
Volume fractions (Ferrite/Austenite) 0.80/0.19 0.67/0.32 0.64/0.35 0.004/0.99

Fig. 12. Calculated martensite start temperature as a function of weight percent


nitrogen for each build composition.

equivalence, increasing from 5 wt% for the argon atomized feed-


stocks to levels above 10 wt%, thus promoting higher levels of re-
tained austenite in the powder feedstocks.
• A decrease in the measured retained austenite when moving from
the powder feedstock to the as-built microstructure was observed for
all conditions. The retained austenite levels for both argon atomized
powder feedstocks approached zero, regardless of production
system or shielding gas. However, higher austenite levels, reaching
levels of 81% for the nitrogen atomized feedstocks, were observed in
materials with higher concentrations of austenite-stabilizing ele-
ments, which was primarily dictated by the feedstock composition
and only marginally influenced by the choice of shielding gas.

Fig. 11. Plots showing the phase stability over the solid and liquid regions for
• Increased levels of retained austenite impede the normal overaged
heat treat response. The AM components fabricated from low aus-
the (a) argon atomized feedstock and (b) nitrogen atomized feedstock materials tenite feedstocks followed the expected trend of declining hardness
fabricated in the 3D Systems ProX 200 system.
values as the material was overaged at increasing temperatures.
Although partially austenitic AM specimens responded normally at
482 °C to temperatures nearing 680 °C. An analysis of the role of com- low overaging temperatures, hardness deviated from the trend and
position has shown that increasing levels of nitrogen leads to higher increased at higher temperatures.
levels of retained austenite, causing the phase regions in the alloys to be
shifted and a decrease in the Ms temperature. The combination of these
• Highly austenitic material with retained austenite levels ap-
proaching 22% actually exhibited its lowest hardness values for heat
effects causes the heat treatment response in these AM components to treatments closest to the conventional peak age condition for 17–4
deviate from the expected wrought conditions. The primary conclusions PH grade stainless steel. For the material fabricated from the ni-
in this work are provided below: trogen atomized powder processed under nitrogen, the peak aged
condition was not achieved until 4 h at 680 °C.
• The atomization condition plays a prominent role in altering the
nitrogen concentration in 17–4 PH grade stainless steel powder
feedstocks. Argon atomized feedstocks contained approximately
• Changes in composition, particularly of nitrogen originating from
the atomization gas, influence the phase equilibrium, even though
0.01 wt% N, while the nitrogen concentration in the nitrogen ato- the compositions remain within specified ranges. As more nitrogen
mized powders increased to levels between 0.06 and 0.12 wt%. is introduced, the temperature boundaries of the austenite phase
These higher nitrogen concentrations have a strong impact on nickel region significantly widens. A shift of nearly 100 °C in the

54
S.D. Meredith et al. Materials Science & Engineering A 738 (2018) 44–56

martensite start temperature is predicted between the builds with Powder Bed Fusion additive manufactured 17–4 stainless steel, Mater. Des. 133
the highest and lowest nickel equivalency values, which is heavily (2017) 205–215, https://doi.org/10.1016/j.matdes.2017.07.047.
[15] M. Masoomi, N. Shamsaei, R.A. Winholtz, J.L. Milner, T. Gnäupel-Herold,
influenced by nitrogen concentration. A. Elwany, M. Mahmoudi, S.M. Thompson, Residual stress measurements via neu-
tron diffraction of additive manufactured stainless steel 17–4 PH, Data Br. 13
Acknowledgements (2017) 408–414, https://doi.org/10.1016/j.dib.2017.06.027.
[16] B. AlMangour, J.M. Yang, Integration of heat treatment with shot peening of 17–4
stainless steel fabricated by direct metal laser sintering, JOM 69 (2017) 2309–2313,
The authors wish to thank the Office of Naval Research https://doi.org/10.1007/s11837-017-2538-9.
Manufacturing Technology program and the Applied Research [17] H. Irrinki, T. Harper, S. Badwe, J. Stitzel, O. Gulsoy, G. Gupta, S.V. Atre, Effects of
powder characteristics and processing conditions on the corrosion performance of
Laboratory’s Institute for Manufacturing and Sustainment Technologies 17–4 PH stainless steel fabricated by laser-powder bed fusion, Prog. Addit. Manuf. 3
which is funded under the Naval Sea Systems Command (NAVSEA), (2018) 39–49, https://doi.org/10.1007/s40964-018-0048-0.
United States, contract #N00024-12-D-6404 for funding this work, as [18] Z. Hu, H. Zhu, H. Zhang, X. Zeng, Experimental investigation on selective laser
melting of 17–4 PH stainless steel, Opt. Laser Technol. 87 (2017) 17–25, https://
well as the resources of Penn State's Center for Innovative Materials
doi.org/10.1016/j.optlastec.2016.07.012.
Processing through Direct Digital Deposition (CIMP-3D), where sample [19] A.A. Adeyemi, E. Akinlabi, R.M. Mahamood, K.O. Sanusi, S. Pityana, M. Tlotleng,
fabrication was performed by Mr. J.J. Blecher of 3D Systems and Mr. Influence of laser power on microstructure of laser metal deposited 17–4 PH
C.J. Dickman and Dr. K.C. Meinert. The authors also extend their gra- stainless steel, IOP Conf. Ser. Mater. Sci. Eng. 225 (2017) 012028, https://doi.org/
10.1088/1757-899X/225/1/012028.
titude to several individuals who provided significant contributions, [20] T.L. Starr, K. Rafi, B. Stucker, C.M. Scherzer, Controlling phase composition in se-
such as E.A. Good in metallographic specimen preparation, E.T. McHale lective laser melted stainless steels, in: Proceedings Solid Freeform Fabrication
for contributing XRD scans, and N.J. Carrier for providing SEM of the Symposium, 2012, pp. 439–456.
[21] D.J. Kotecki, T.A. Siewert, WRC-1992 constitution diagram for stainless steel weld
powders. metals: a modification of the WRC-1988 diagram, Weld. J. 71.5 (1992) 171–178.
[22] J.C. Lippold, D.J. Kotecki, Welding Metallurgy and Weldability of Stainless Steels,
Data availability John Wiley & Sons, 2005.
[23] B. Clausen, D.W. Brown, J.S. Carpenter, K.D. Clarke, A.J. Clarke, S.C. Vogel,
J.D. Bernardin, D. Spernjak, J.M. Thompson, Deformation behavior of additively
The raw/processed data required to reproduce these findings cannot manufactured GP1 stainless steel, Mater. Sci. Eng. A. 696 (2017) 331–340, https://
be shared at this time as the data also forms part of an ongoing study. doi.org/10.1016/j.msea.2017.04.081.
[24] S. Cheruvathur, E.A. Lass, C.E. Campbell, Additive manufacturing of 17–4 PH
stainless steel: post-processing heat treatment to achieve uniform reproducible
Appendix A. Supporting information microstructure, JOM 68 (2016) 930–942, https://doi.org/10.1007/s11837-015-
1754-4.
[25] T. Debroy, H.L. Wei, J.S. Zuback, T. Mukherjee, J.W. Elmer, J.O. Milewski,
Supplementary data associated with this article can be found in the
A.M. Beese, A. Wilson-heid, A. De, W. Zhang, Progress in materials science additive
online version at doi:10.1016/j.msea.2018.09.066. manufacturing of metallic components – process, structure and properties, Prog.
Mater. Sci. 92 (2018) 112–224, https://doi.org/10.1016/j.pmatsci.2017.10.001.
References [26] M. Murayama, K. Hono, Y. Katayama, Microstructural evolution in a 17–4 PH
stainless steel after aging at 400 °C, Metall. Mater. Trans. A (1999), https://doi.org/
10.1007/s11661-999-0323-2.
[1] AK Steel product data bulletin, 17–4 PH Stainless Steel, (n.d.). 〈http://www. [27] M.R. Stoudt, R.E. Ricker, E.A. Lass, L.E. Levine, Influence of postbuild micro-
aksteel.com/pdf/markets_products/stainless/precipitation/17-4_PH_Stainless_ structure on the electrochemical behavior of additively manufactured 17–4 PH
Steel_PDB_201512.pdf〉. (Accessed 1 August 2017). stainless steel, JOM (2017), https://doi.org/10.1007/s11837-016-2237-y.
[2] L.E. Murr, E. Martinez, J. Hernandez, S. Collins, K.N. Amato, S.M. Gaytan, [28] ASTM Standard A564/A564M, Standard Specification for Hot-Rolled and Cold-
P.W. Shindo, Microstructures and properties of 17-4 PH stainless steel fabricated by Finished Age-Hardening Stainless Steel Bars and Shapes, 2013. 〈http://dx.doi.org/
selective laser melting, J. Mater. Res. Technol. (2012), https://doi.org/10.1016/ 10.1520/A0564_A0564M-13〉.
S2238-7854(12)70029-7. [29] G. Krauss, Steels: Processing, Structure, and Performance, ASM International, 2015.
[3] H.K. Rafi, D. Pal, N. Patil, T.L. Starr, B.E. Stucker, Microstructure and mechanical [30] ASTM Standard B213, Standard Test Methods for Flow Rate of Metal Powders Using
behavior of 17-4 precipitation hardenable steel processed by selective laser melting, the Hall Flowmeter Funnel, (n.d.). 〈http://dx.doi.org/10.1520/B0527-15〉.
J. Mater. Eng. Perform. (2014), https://doi.org/10.1007/s11665-014-1226-y. [31] ASTM Standard B212, Standard Test Method for Apparent Density of Free-Flowing
[4] R. Bhambroo, S. Roychowdhury, V. Kain, V.S. Raja, Effect of reverted austenite on Metal Powders Using the Hall Flowmeter Funnel, 2013. 〈http://dx.doi.org/10.
mechanical properties of precipitation hardenable 17-4 stainless steel, Mater. Sci. 1520/B0212-13〉.
Eng. A 568 (2013) 127–133, https://doi.org/10.1016/j.msea.2013.01.011. [32] ASTM Standard B527, Standard Test Method for Tap Density of Metal Powders and
[5] C.N. Hsiao, C.S. Chiou, J.R. Yang, Aging reactions in a 17–4 PH stainless steel, Compounds, 2015. 〈http://dx.doi.org/10.1520/B0527-15〉.
Mater. Chem. Phys. (2002), https://doi.org/10.1016/S0254-0584(01)00460-6. [33] ASTM Standard B822, Standard Test Method for Particle Size Distribution of Metal
[6] T. LeBrun, T. Nakamoto, K. Horikawa, H. Kobayashi, Effect of retained austenite on Powders and Related Compounds by Light Scattering, B822. 97, 2001, pp. 1–5.
subsequent thermal processing and resultant mechanical properties of selective 〈http://dx.doi.org/10.1520/B0822-10.2〉.
laser melted 17–4 PH stainless steel, Mater. Des. 81 (2015) 44–53, https://doi.org/ [34] ASTM Standard E975, Standard Practice for X-Ray Determination of Retained
10.1016/j.matdes.2015.05.026. Austenite in Steel with Near Random Crystallographic Orientation, 2013. 〈http://
[7] ASTM Standard A693, Standard Specification for Precipitation-Hardening Stainless dx.doi.org/10.1520/E0975-13〉.
and Heat-Resisting Steel, Plate, Sheet, and Strip, 2013. 〈http://dx.doi.org/10.1520/ [35] Y. Waseda, E. Matsubara, K. Shinoda, X-Ray Diffraction Crystallography, Springer,
A0564_A0693-13〉. Berlin, 2011, https://doi.org/10.1007/978-3-642-16635-8.
[8] F.C. Campbell (Ed.), Elements of Metallurgy and Engineering Alloys, ASM [36] B.D. Cullity, S.R. Stock, Elements of X-Ray Diffraction, 3rd ed., Prentice Hall, Upper
International, 2008. Saddle River, NJ, 2001.
[9] C.R. Brooks, Principles of the Austenitization of Steels, Springer Science & Business [37] L. Cheng, A. Böttger, T.H. de Keijser, E.J. Mittemeijer, Lattice parameters of iron-
Media, 1992. carbon and iron-nitrogen martensites and austenites, Scr. Metall. Mater. 24.3
[10] A. Yadollahi, N. Shamsaei, S.M. Thompson, A. Elwany, L. Bian, Effects of building (1990) 509–514.
orientation and heat treatment on fatigue behavior of selective laser melted 17–4 [38] ASTM Standard E407, Standard Practice for Microetching Metals and Alloys, 2015.
PH stainless steel, Int. J. Fatigue 94 (2017) 218–235, https://doi.org/10.1016/j. 〈http://dx.doi.org/10.1520/E0407-07R15E01〉.
ijfatigue.2016.03.014. [39] ASTM Standard E140, Standard Hardness Conversion Tables for Metals
[11] B. AlMangour, J.M. Yang, Understanding the deformation behavior of 17–4 pre- Relationship Among Brinell Hardness, Vickers Hardness, Rockwell Hardness,
cipitate hardenable stainless steel produced by direct metal laser sintering using Superficial Hardness, Knoop Hardness, Scleroscope Hardness, and Leeb Hardness,
micropillar compression and TEM, Int. J. Adv. Manuf. Technol. 90 (2017) 119–126, 2013. 〈http://dx.doi.org/10.1520/E0140-12B〉.
https://doi.org/10.1007/s00170-016-9367-9. [40] L.F. Pease, W.G. West, Fundamentals of powder metallurgy, Metal. Powder Ind.
[12] M. Mahmoudi, A. Elwany, A. Yadollahi, S.M. Thompson, L. Bian, N. Shamsaei, Fed. (2002).
Mechanical properties and microstructural characterization of selective laser [41] A.L. Schaeffler, Constitution diagram for stainless steel weld metal, Met. Prog. 56
melted 17–4 PH stainless steel, Rapid Prototyp. J. (2017), https://doi.org/10.1108/ (1949) 680 (680B).
RPJ-12-2015-0192. [42] C.J. Long, W.T. Delong, The Ferrite Content of Austenitic Stainless Steel Weld
[13] S. Pasebani, M. Ghayoor, S. Badwe, H. Irrinki, S.V. Atre, Effects of atomizing media Metal, 1973.
and post processing on mechanical properties of 17–4 PH stainless steel manu- [43] L. Facchini, N. Vicente, I. Lonardelli, E. Magalini, P. Robotti, M. Alberto, Metastable
factured via selective laser melting, Addit. Manuf. 22 (2018) 127–137, https://doi. austenite in 17-4 precipitation-hardening stainless steel produced by selective laser
org/10.1016/j.addma.2018.05.011. melting, Adv. Eng. Mater. (2010), https://doi.org/10.1002/adem.200900259.
[14] A. Kudzal, B. McWilliams, C. Hofmeister, F. Kellogg, J. Yu, J. Taggart-Scarff, [44] M. Averyanova, P. Bertrand, B. Verquin, Effect of initial powder properties on final
J. Liang, Effect of scan pattern on the microstructure and mechanical properties of microstructure and mechanical properties of parts manufactured by selective laser

55
S.D. Meredith et al. Materials Science & Engineering A 738 (2018) 44–56

melting, in: B. Katalinic (Ed.), Proceedings 21st International DAAAM Symposium, [48] SAE International Aerospace Material Specification, Steel, Corrosion Resistant,
DAAAM International, Vienna, Austria, EU, 2010. Investment Castings 16Cr–4.1Ni–0.28Cb–3.2Cu Homogenization and Solution Heat
[45] F. Bahrami, A. Hendry, Microstructure and mechanical behaviour of nitrogen al- Treated or Homogenization, Solution, and Precipitation Heat Treated, 2015. 〈www.
loyed martensitic stainless steel, Mater. Sci. Technol. 11.5 (1995) 488–497. sae.org〉.
[46] S. Saeedipour, A. Kermanpur, A. Najafizadeh, Effect of N on phase transformations [49] K.M. Coffy, Microstructure and Chemistry Evaluation of Direct Metal Laser Sintered
during martensite thermomechanical processing of the nano/ultrafine-grained 201L 15-5 Stainless Steel, University of Central Florida, 2014.
steel, J. Mater. Eng. Perform. 25 (2016) 5502–5512, https://doi.org/10.1007/ [50] S.-J. Lee, K.-S. Park, Prediction of martensite start temperature in alloy steels with
s11665-016-2387-7. different grain sizes, Metall. Mater. Trans. A (2013) 3423–3427 (44.8).
[47] S.C. Krishna, N.K. Karthick, A.K. Jha, B. Pant, P.V. Venkitakrishnan, Microstructure [51] N. Saunders, Z. Guo, a.P. Miodownik, J.-P. Schillé, The calculation of TTT and CCT
and properties of nitrogen-alloyed martensitic stainless steel, Metallogr. diagrams for general steels, JMatPro Softw. Lit. (2004) 1–12 (10.1.1.565.9926).
Microstruct. Anal. 6 (2017) 425–432, https://doi.org/10.1007/s13632-017- [52] G. Ghosh, G.B. Olson, Computational thermodynamics and the kinetics of marten-
0381-6. sitic transformation, J. Phase Equilibria. 22.3 (2001) 199–207.

56

You might also like