You are on page 1of 31

Review SCIENCE PROGRESS

Science Progress

A review on the production 2023, Vol. 106(1) 1–31


© The Author(s) 2023

of 17-4PH parts using press Article reuse guidelines:


sagepub.com/journals-permissions

and sinter technology DOI: 10.1177/00368504221146060


journals.sagepub.com/home/sci

Mohammed Qasim Kareem1,2 ,


Tamás Mikó1, Gréta Gergely1
and Zoltán Gácsi1
1
Institute of Physical Metallurgy, Metalforming and Nanotechnology,
University of Miskolc, Hungary
2
Faculty of Engineering, Al-Qasim Green University, Al-Qasim/Babylon,
Iraq

Abstract
The press and sinter method remains the standard among powder metallurgy processes for pow-
dered stainless-steel materials. It delivers low cost, low oxidation rate, and adequate corrosion
resistance. Furthermore, 17-4PH is a martensitic stainless-steel that is commonly used for high-
strength and medium-ductility stainless steel parts. However, a few studies have investigated the
press and sinter method for producing 17-4PH parts. This shortage is due to the high hardness
(low compressibility) of 17-4PH powder. Thus, the main objective of this study is to evaluate
the press and sinter method in terms of the manufacturing process, the influencing factors, and
the theoretical basis of press and sinter methods in conjunction with metal injection molding tech-
nology for the production of 17-4PH parts. First, the literature and monographs are examined and
summarized to cover the previous results, research progress, development trends, and applica-
tions of press and sinter method 17-4PH parts. Following the theoretical analysis, the practical
investigation was conducted by producing parts with cold pressing from 800 to 1600 MPa, fol-
lowed by sintering: the sintering temperature was 1200 °C for one hour under a protective vac-
uum atmosphere. ImageJ analysis was performed to measure the sinter density. The results
showed an increase in relative sinter density from 84.43% to 96.43% for 800 and 1600 MPa,
respectively, while the earlier results reached 93.47%. Overall, the press and sinter method
enables the production of high-hardness 17-4PH parts with a high density, without using additives
like lubricants, wax, or alloying elements.

Keywords
Press and sinter, metal injection molding, stainless steel, 17-4PH, heat treatment, corrosion
performance

Corresponding author:
Mohammed Qasim Kareem, Institute of Physical Metallurgy, Metalforming and Nanotechnology, University
of Miskolc, Miskolc, Hungary.
Emails: mohammed.qasim@uni-miskolc.hu; mohammed.qasim@wrec.uoqasim.edu.iq

Creative Commons Non Commercial CC BY-NC: This article is distributed under the terms of the Creative
Commons Attribution-NonCommercial 4.0 License (https://creativecommons.org/licenses/by-nc/4.0/)
which permits non-commercial use, reproduction and distribution of the work without further permission provided the original
work is attributed as specified on the SAGE and Open Access page (https://us.sagepub.com/en-us/nam/open-access-at-sage).
2 Science Progress 106(1)

Introduction
There are many different powder metallurgical manufacturing processes. For example,
metal powder injection molding (MIM), additive manufacturing (AM), and hot isostatic
pressing. In addition, the press and sinter method (PSM) remains the leader in manufac-
turing ferrous parts among all other powder metallurgical operations worldwide.
According to Figure 1(a), the ferrous products obtained through PSM accounted for
63% of the total. A report on the metal powder market size in the United States from
2016 to 2027, which is shown in Figure 1(b), explains the high demand for PSM products
compared to others.
Ferrous powder, such as 17-4PH (precipitation-hardened martensitic stainless steel),
has high hardness, 3–4 times stronger than austenitic stainless steel such as 316.
Precipitation-hardened stainless steel is usually chosen to produce stainless materials
with high strength and medium ductility.3,4 17-4 PH parts are typically used in the aero-
space, chemical, petrochemical, food, and general metal processing industries.5–8
Designers and engineers prefer 17-4PH stainless steel because of its beneficial properties
such as high strength, hardness, easy modification by heat treatment, and corrosion resist-
ance at temperatures below 300 °C.6,9,10 This is attributed to the alloy containing more
chromium and nickel.4,11
However, PSM could not previously be used to produce high-density parts from
17-4PH powder due to the poor compressibility of the powder. Therefore, the second-best
sintering method has been MIM, according to Figure 1(b). MIM includes the preparation
of the starting material (polymer blend with metal-like properties), injection molding,
debinding, and sintering.12–15 Similarly, PSM involves the compaction of the powder
(metal or metal alloy, with or without polymers), usually in molds, followed by sintering
of the molded bodies.16,17 The thermal debinding in a high-temperature burnout furnace
should be done before sintering in the MIM process, which takes up to 50 h.7 In contrast,
there is no debinding step in PSM. Another advantage of PSM is that the particles bond
together directly (without organic materials such as wax), resulting in less oxidation com-
pared with MIM.18 Therefore, PSM does not cause high pollution.19 MIM is similar to

Figure 1. The use of press and sinter technology compared to other processes for producing
powdered materials in (a) Europe in 2019 and (b) the US market over a range of years
(reproduced from Refs.,1,2 respectively).
Kareem et al. 3

PSM because both produce near-to-net shapes and complicated geometries, but MIM is
more expensive than PSM.7 Figure 2 shows the previous studies that performed the densi-
fication of sintered 17-4PH parts by PSM and MIM. Fewer PSM studies were published
than MIM studies. The low compressibility of the 17-4PH powder, which leads to low
densification, was the main reason for fewer PSM studies.20–22 PSM can produce stainless-
steel parts with a sintered density of 6.8 to 7.3 g/cm3.23 In contrast, the theoretical density of
17-4PH materials reported in the literature ranges from 7.75 to 7.85 g/cm3.24
PSM stainless steel parts are cheaper than MIM stainless-steel parts, especially for 103
to 106 pieces with weights ranging from 0.1 to 9 kg per piece.19,34,35 PSM has several
advantages for the mass production of small, intricately shaped components.36 For
example, money can be saved for gear pressing and sintering due to the reduced
number of production operations.25 PSM stainless steels exhibit higher quality than
other types of stainless steel in terms of cost, control of size, and resistance to wear
and corrosion.23
However, the high porosity of PSM parts made of 17-4PH is still the main issue. Steels
with higher density typically have greater mechanical properties.37 Thus, earlier studies
attempted to increase the sinter density of 17-4PH parts produced by PSM. In 1994,
Reinshagen and Witsberger demonstrated that aged, sintered PSM 17-4PH parts could

Figure 2. Earlier attempts for proving densification of 17-4PH parts by the press and sinter
compared to the injected and sintered parts (constructed from Refs.3,9,7,11,14,20,26,27,28,29,30,31,32).
4 Science Progress 106(1)

achieve a relative sinter density of 93.47%.11 Moreover, Schade et al. discovered that
cutting the carbon content of conventional water-atomized 17-4PH powder
(0.023 wt.% carbon) in half with a high-performance atomization (HPAT) technique
increased the relative sinter density to 94.75%.20 Additionally, in 2008, water-atomized
PSM 17-4PH parts that were sintered and heat-treated reached the maximum relative
sinter density of 92.45%.11 In 2013, it was found that increasing the sintering temperature
improved the sinter density by 92.19%.3 Moreover, adding 0.6 wt.% of boron increased
the relative sinter density of PSM 17-4PH parts from 84.54% to 96.4%.9 Figure 3 shows
the results of previous studies regarding how to increase the density of sintered PSM
17-4PH parts by increasing the cold pressing or sintering temperature (without alloying
additions, postprocessing, or lowering the carbon content).
As a result of the above literature, PSM is still the most demanded technology world-
wide in the powder metallurgy market for producing stainless-steel parts. According to
Figure 1(b), PSM will also be demanded in the future. However, PSM is still unable to
provide high-density 17-4PH parts. The densification is considerably affected by many
forming factors such as cold pressing, powder compressibility, powder particle shape,
particle size distribution, lubricant addition, sintering temperature and time, and sintering
atmosphere. Therefore, this review article focuses on the 17-4PH materials produced by
PSM and compares them with those produced by MIM, summarizes the latest research on
the forming factors, investigates the causes of low densification results, and finds the

Figure 3. The densification results from previous press and sinter 17-4PH parts as a function of
cold pressing and sintering temperature (obtained by Refs.3,11,20,27,33).
Kareem et al. 5

theoretical solutions to approve the densification property. This report also includes prac-
tical experiments that validate the academic gaps in a parallel manner.

Cold pressing
The densification of green compacts (i.e. powders before sintering) affects the strength
and shrinkage ratio of the sintered parts. Lower green density means lower sinter
density20 and increased sinter shrinkage.38 Consequently, low sinter density leads to
poor mechanical properties.39
Cold pressing is the process of repackaging metal powders. Pressing begins with the
reduction of the pore space and increasing the powder’s efficiency, eliminating the
bridges created during filling, and forming new surfaces free of oxide.40 However, the
metal powder’s compressibility (hardness) is the main factor that directly affects green
densification. The high compressibility of metal powder (low hardness) leads to an ele-
vated green density. The effect of cold pressing on metal powders with different hardness
(pure iron 104 HV, titanium 144 HV, and 304L 205 HV) was investigated.41 The findings
indicated that the powder hardness and densification behavior had an inverse relationship.
Moreover, different powders of stainless steel with different compressibility (434L,
304L, and 17-4PH) were cold pressed up to 690 MPa.20 The results showed that the
parts made from 434L and 304L (high ferrite content and high compressibility) had a
higher density than those made from 17-4PH (containing martensite and low compress-
ibility). Since ferrite is less hard than martensite, it is possible to achieve higher green and
sinter densities.42
Previously, the effect of cold pressing on the green densification of PSM 17-4PH parts
was investigated by different studies, which is shown in Figure 4(a). The highest cold
press applied by previous studies was 760 MPa. As a result, the maximum relative
green density was approximately 80.67%. Additionally, the effect of particle size

Figure 4. The effect of cold pressing and particle size on green densities of 17-4PH pressed parts
according to (a) earlier studies (constructed from Refs.3,11,20,27) and by (b) the practical work in
this study.
6 Science Progress 106(1)

distribution has been investigated, which is also presented in Figure 4(a). The lowest
average particle size was around 45 µm. However, the increase in cold pressing has
had a more significant impact than reducing particle size on improving green
densification.26
In the present work, Oerlikon water-atomized powder consisting of martensitic
stainless steel (17-4PH) with a particle size of − 15/ + 45 µm was sieved using a
32-µm mesh and cold-pressed under vacuum up to 1600 MPa. The results shown in
Figure 4(b) indicate that cold pressing increased the green density, and at 1600 MPa,
the relative green density was around 90%, which was much higher than in previous
studies.

Sintering
The energy stored in the cold-pressed metal part is released during sintering. Sintering is a
complicated process in which the development of the microstructure occurs through the
action of several different transport mechanisms.43 In the solid state, the sintering of crys-
talline solids can occur by vapor transport, and surface, lattice, or grain boundary diffu-
sion, in addition to plastic deformation by dislocation migration.44 In the sintering of
powdered steel, the green parts are solidified to the desired composition using tempera-
tures below the melting point (but above half the melting temperature) for the appropriate
amount of time.40 During the sintering process, there are three kinetic stages: bonding
after the reduction of oxides from the surfaces of powder particles (bonding between adja-
cent particles with the formation necks); densification (adhesion between the solid and
porous phases); and grain coarsening (significant grain growth and interaction between
pores and grain boundaries).45 Thus, a higher green density, which means there are
more points of contact between the particles, speeds up adhesion and densification
during sintering.40
Most metal (M) oxides in steel powdered materials are reduced by reduction reactions
to start the necking stage during sintering. The reduction media can be (a) hydrogen, (b)
carbon, or (c) carbon monoxide.46 Moreover, the reduction process can take place at tem-
peratures of less than 500 °C. Hydrogen comes from moisture in the powder and an
atmosphere that contains hydrogen. Carbon comes from graphite lubricants or steel
powder mixtures.
Mx Oy + yH2  xM + yH2 O (1)

Mx Oy + yC  xM + yCO (2)

Mx Oy + yCO  xM + yCO2 (3)


The signals for the beginning of the sintering process of 17-4PH parts, manufactured by
PSM or MIM, can be constructed by dilatometric measurements of sintering peaks. The
majority of studies discovered that, regardless of lubricant, debinding conditions, or pro-
tective atmosphere, the sintering behavior of PSM and MIM 17-4PH parts is
similar.3,13,27,47 According to those studies, several dilatometric peaks can describe the
sintering process and phase transformations of 17-4PH sintered parts. Consequently,
Kareem et al. 7

there are three primary stages of the sintering process: heating, isothermal holding, and
cooling. As shown in Figure 5, the sintering peaks behaved the same way during
heating and isothermal sintering with a hydrogen protection atmosphere. They showed
the following peaks:

1. Between 725 and 750 °C, the peak of thermal expansion shows that martensite
BCC is transforming into austenite FCC.
2. The initial sintering shrinkage peak appeared between 964 and 1050 °C.
3. The isothermal sintering peak can be seen at temperatures ranging from 112 to
1143 °C. This peak started with heating shrinkage and ended with delayed
shrinkage.
4. The hydrogen peak shows that high sinterability can be reached by reducing the
silicon oxides that cover the 17-4PH powder between 1250 and 1308 °C. This
peak is reached when hydrogen is used as a protective atmosphere.

To confirm the first peak (thermal expansion), water-atomized 17-4PH powder was
examined before sintering, and the microstructure showed a solidified martensite, as
shown in Figure 6(a). After sintering at 780 °C and subsequent quenching, the

Figure 5. Sintering shrinkage peaks of dilatometric measurements of 17-4PH parts, manufactured


by metal injection molding regardless the heating conditions. (a) Reproduced from Ref.13 and (b)
from Ref.47
8 Science Progress 106(1)

Figure 6. The microstructure of 17-4PH (a) as water-atomized powder, (b) powder quenched
from 780 °C (reproduced from Ref.47), (c) powder after cold pressing and sintering at 1360 °C in
pure H2 for one hour (reproduced from Ref.3). M and δ refer to martensite and delta-ferrite,
respectively.

microstructure showed lathe martensite, emphasizing austenitization. Another example of


the microstructure of a PSM 17-4PH part is shown in Figure 6(b). At the end of one hour
of sintering at 1360 °C in pure hydrogen, the microstructure showed martensite, delta-
ferrite (δ-ferrite), and pores. A detailed explanation will be given in the following sections
of this review.
In addition to increasing cold pressing, PSM stainless-steel parts can be improved by
changing the sintering conditions (temperature, time, and atmosphere), and thus control-
ling the grain growth, oxidation, and the ratio and shape of pores. High sintering tempera-
tures can eliminate tiny pores, make the remaining pores spherical, and homogenize the
stainless-steel structure. Sometimes, higher temperatures can distort the PSM steel’s final
dimensions, but spherical pores increase the mechanical properties of a sample compared
to irregular pores. The pores are the most critical factors when it comes to predicting and
modifying the mechanical properties.

Factors related to the properties of sintered


stainless-steel materials
Sintering time and temperature
Densification. Increasing the sintering temperature and time affects the ratio and shape of
the pores. The effect of sintering temperature and time on the densification of 17-4PH
parts produced by PSM has been studied.27 The results are shown in Figure 7. They
show higher relative sinter densities with increasing sintering temperature and time.
The highest and lowest relative densities were 92.19% (using the highest temperature
Kareem et al. 9

Figure 7. Sintered density of 17-4PH materials as a function of the sintering time and
temperature and (reproduced from Ref.27).

of 1300 °C and a time of 120 min) and 85% (using a low sintering temperature of 1200 °
C), respectively.
The previous experiments performed to improve the sintering densification of PSM
17-4PH parts are summarized in Figure 8(a). Increasing the sintering temperature from
1300 to 1340 °C for the same sintering time and cold pressing (60 min, 600 MPa) resulted
in a 2.3% increase in the relative sinter density. In addition, increasing the sintering time
from 20 to 60 min at a fixed cold pressing and sintering temperature (690 MPa, 1260 °C)
was investigated.20 The results showed an increase in the relative sintering density from
89.6% to 93.47% for 20 and 60 min, respectively.
For the practical experiments conducted herein, the sintering conditions were estab-
lished as follows: one hour and up to seven hours as holding times; 1200 °C as a sintering
temperature; and vacuum as a protective atmosphere. The relative sinter density was mea-
sured by ImageJ analysis as the average value of 12 readings for 12 optical photos taken
from different positions for each cross-sectionally cut and polished sample of PSM
17-4PH. These densification results are shown in Figure 8(b). The sintered densities
increase with the green densities, Figure 4(b) for comparison. The relative sinter
density of the green samples cold-pressed at 800 and 1600 MPa increased from
84.43% to 96.43%, respectively. In earlier research, the maximum relative sinter
density was 93.47%. In addition, an extension of the sintering time, up to seven hours,
improved the densification properties. The optical photos and the results of ImageJ ana-
lysis of the practical part of this review are shown in Figure 9. The fraction of pores
decreases with increasing cold pressing, and the shapes of pores change from irregular
to more spherical shapes. According to a previous study, increasing the cold pressing
from 200 to 600 MPa transformed the shapes of the pores of 316L into regular shapes,
and the volume fraction of regular pores increased from 10% to 22%.50
In a previous study, the relative density of 17-4PH samples prepared by MIM was con-
trolled from 61% to 99% by increasing the sintering temperature from 1200 to 1350 °C.28
The results showed that not only did the pore fraction decrease but also the shapes of the
pores became more regular as the sintering temperature increased. Additionally, the
10 Science Progress 106(1)

Figure 8. The sinter densification of press and sinter 17-4PH products at (a) different conditions
(produced from Refs.3,11,20,27); (b) condition of 120° C for 1 h using vacuum protective
atmosphere (measured by ImageJ, present work).

relative sinter density of MIM 17-4PH parts increased up to 98% when the sintering tem-
perature was increased from 1150 to 1300 °C under the same production conditions.48 In
addition, the effect of sintering time and temperature on the densification of a MIM
17-4PH part containing 1% FeB was investigated.51 The results showed that changes
in sintering temperature, from 1265 to 1285 °C, and time, from 10 to 45 min, led to a
nearly complete density (around 99%) due to the formation of the eutectic liquid
phase, which exhibited high diffusivity and rapid densification. However, Figure 10
depicts the appearance of the δ-ferrite phase during sintering at 1300 °C, which may
Kareem et al. 11

Figure 9. The optical microscopic photos of press and sinter 17-4PH samples with indicated
different cold pressings and calculated porosity (%) by ImageJ; sintered at 1200 °C for one hour.

be responsible for the rise in the density of 17-4PH sintered parts. The light gray is the
martensitic phase, and the δ-ferrite is the dark gray phase. The microstructure of the
parts sintered at 1150 °C included both austenitic grain boundaries and martensitic layers.
However, increasing the sintering time at elevated temperatures can lead to overheat-
ing of the steel structure. For example, the effect of using different sintering times on the
properties of HK30, an austenitic stainless steel, was studied.49 Sintering at 1270, 1280,
and 1300 °C revealed that, for the same holding time, density increases with increasing
temperature. At 1280 °C, the highest density of 7.61 g/cm3 was reached after seven hours
of holding time, and further holding led to a slightly lower density, indicating the
12 Science Progress 106(1)

Figure 10. Microstructures of 17-4PH produced by metal injection molding after sintering at (a)
1150 °C and (b) 1300 °C (reproduced from Ref.48). M, γ, and δ refer to martensite, austenite, and
delta-ferrite, respectively.

possibility of overheating. A higher sintering temperature of 1300 °C led to the overheat-


ing of the microstructure, reducing the bond strength between grain boundaries. As a
result, the density and mechanical properties decreased further.

Mechanical properties. The porosity ratio affects the mechanical and physical properties
of powdered steel parts. The balanced effects of reduced porosity and grain coarsening
enhanced tensile strength and elongation, while yield strength remained unchanged.52
Usually, a decrease in porosity raises the yield strength, but it was terminated by the
grain-coarsening phenomenon. Moreover, by increasing the sintering time from 20 to
60 min, the sintered density of conventional water-atomized 17-4PH green compacts
has also increased from 6.8 to 7.3 g/cm3.26 The high density resulted in high mechanical
properties. Conversely, pores reduce mechanical properties by acting as stress concentra-
tions and promoting fracture growth.53 In another study, when the sintering time was
increased from 20 to 40 min, the sinter density increased, and subsequently, the hardness
increased from 59 to 65 HRA.20 In addition, the effect of different sintering temperatures
on MIM 17-4PH parts was investigated.14 The results showed that the 17-4PH parts had a
higher relative density at the high sintering temperature of 1360 °C than at 1200 °C, but
there was no significant change in hardness. Table 1 shows the mechanical properties of
previously studied PSM 17-4PH parts made using different conditions.

Pores and grain size. Multiple factors influence the shape of the pores, including the
powder particle morphology, sintering temperature and time, and the atmosphere.
Using gas-atomized and water-atomized powders for producing MIM 17-4PH parts
has been investigated.31 The sintering results at 1250 °C showed that the pores
between particles were quite different for the MIM 17-4PH made of gas-atomized and
water-atomized powders (see Figure 11). Spherical powder particles produce more
rounded pores compared to irregular powder particles. The shape of the particles in
gas-atomized samples is spherical, while it is irregular in water-atomized samples.31 In
another study, the MIM 17-4PH parts sintered at 1300 °C had round pores compared
Table 1. The mechanical properties of 17-4PH parts manufactured by press and sinter and compared to wrought processing.

Rel. Sint. Hardness, UTS, YS, Elong.,


Conditions Heat treatment density % HV MPa MPa % Ref.
Kareem et al.

20
HPA 17-4PH, P at 690, S at 1260 °C, - 94.75 - - -
60 min., H2 Aging, 538 °C, 1 h - ̴ 237 819 688 1.7
11
WA 17-4PH, P at 760, S at 1275 °C, - 92.45 188 869 768 2.5
45 min.
3
17-4PH, P at 600, 1340 °C for 60 min., - 90.91 800 - -
in H2 Annealing1040 °C +Ag.480 °C 92.19 332 (10) 1100 - -
27
17-4PH, P at 600, Sintered1300 °C, - 92.19 - - - -
120 min., in H2
11
17-4PH Aged at 482 °C 93.47 - 1080 - 4.2
17-4PH, low C and N2 >93.47 - 773 - 1.5
33
17-4PH, by PLT, P at 0.40, S 1240 °C, - 84.2 200 - - -
90 min. in Ar+20% H2
17-4PH, by PLT, P at 0.40, S 1290 °C, - 93.0 450
90 min. in Ar+20% H2
Casting 17-4PH - - - 1000– - 8–15
1300
9
17-4PH, P at 600, S 1260 °C, 45 min., - 84.51 190 (10) - - -
in H2
17-4PH +0.6 wt.% B, P at 600, - 96.00 350 - - -
Sintered1260 °C for 45 Min., in H2
56
Cast 17-4PH H925 (496 °C, 4 h, air cooling) - - ̴ 1380 ̴ 1206 6– 9
H1100 (594 °C, 4 h, air cooling) - - ̴ 1103 ̴ 965 9–13
54
Wrought 17-4PH rod H1150 (620 °C, 4 h) - 340 (1) 1017 ± 914 -
2 ±7
55
17-4PH cast - - 366 (10) - - -
1040 °C, 20 min., quenched and then - 396 (10) - - -
heated to 450 °C
56
17-4PH wrought sheet (1 mm thick) H900 (482 °C for 1 h) - - ̴ 1300 - 5–23
13

(Continued)
14

Table 1. (continued)

Rel. Sint. Hardness, UTS, YS, Elong.,


Conditions Heat treatment density % HV MPa MPa % Ref.
57
17-4PH warm forming sheet - - - 1105 986 6.8
(1.6 mm thick)
22
17-4PH wrought - - - 1050 - 5–14
H900 - - 1500 - 10–18
58
17-4PH rod (8 mm dia.) * HQ - 340 (30) 1030 940 -
HQ + H900 - 460 (30) 1402 1384 -
59
17-4PH rod Solutioned at 1040 °C aging 430 °C, 4 h - 455 - - -
for 60 min. RD + aging 410 °C, 4 h. - 480 - - -
RD + aging 500 °C, 4 h. - 400 1295 1280 -
aging 500°C, 4 h - - 1300 1200 -
60
Cast 17-4PH aging at 482 °C - - 1255 1130 -
26
WA 17-4PH, P at 1600, S at 1200 °C, - 94.8 287 - - -
60 min.

332(10): 332 per Vickers with 10 kgf as test load; HPA 17-4PH: high performance atomized 17-4PH mixed with 0.75 wt.% Acrawax C; HQ: 1040 °C for 1 h and oil
quenching; P: cold-pressing in MPa; PLT: powder layering technique; RD: 60% reduction at 500 °C; S: sintered; UTS: ultimate tensile strength; WA17-4PH: water-atomized
17-4 PH with 1.0 wt.% Li stearate; YS: yield strength.
* UTS and YS according to correlation model of Yrieix and Guttman; - means not included.
Science Progress 106(1)
Kareem et al. 15

Figure 11. Pores in (a) gas and (b) water atomized 17-4PH parts sintered at 1250 °C,
manufactured by metal injection molding (reproduced from Ref.31).

to those sintered at 1150 °C.62 Moreover, adding alloying elements can affect the pore
morphology. Increasing the boron content in MIM 17-4PH parts reduced the volume
of the pores and changed their shapes to be more spherical.65 Consequently, the mechan-
ical properties of the sintered 17-4PH materials may be related to the pore shapes, which
are related to the form of the powder particles, as mentioned earlier. The tensile properties
of MIM 17-4PH parts made from gas-atomized powder over a range of sintering tempera-
tures (1250−1350 °C) were around 10% higher than those of these parts made from
water-atomized powder.66 More rounded pores also led to higher ductility.11 Regular
pore geometries and low porosity at elevated sintering temperatures contributed to the
increase in the tensile strength of parts made of chromium alloy steel that were produced
by PSM.38
Furthermore, high porosity causes the hardness values to vary within the same man-
ufactured part. For wrought steel, the porosity reduces heat transfer, favoring bainite for-
mation rather than martensite, which may explain the tremendous difference in hardness
with decreasing density.64 Additionally, the martensite phase is harder than the bainite
phase.
The effect of sintering temperature or time on the shapes of the pores has also been
observed in PSM 316L parts. Both the 17-4PH and 316L parts showed similar sintering
behavior.27 Increasing the sintering temperature from 1200 to 1300 °C for PSM 316L
parts decreased the pore fraction from 20% to 12%, and the shapes of the pores
changed from irregular to more regular.61 However, sintering at a temperature of 1200
°C was considered insufficient and resulted in low mechanical properties.63
Increasing the sintering time and temperature increases grain growth and decreases the
porosity ratio and the diameter of the pores.52 For example, the average grain sizes of the
samples were 45, 64, and 76 μm after sintering at 1280 °C for 5, 7, and 9 h, respectively.67
The higher the sintering temperature, the larger the average grain size.68 However, tensile
strength and elongation depend more on porosity than on grain size, because the fracture
is mostly caused by the growth of micropores and coalescence.52
In the case of PSM 17-4PH parts, there is no study directly showing the influence of
sintering temperature or time on the size and form of porosity. There is also a lack of
16 Science Progress 106(1)

suitable sintering conditions to solidify the green compacts of 17-4PH with enhanced
mechanical properties. As mentioned above, a sintering temperature of 1200 °C leads
to insufficient sintering, even for 316L, which has greater compressibility than
17-4PH. Therefore, it is necessary to investigate the optimal sintering conditions for
PSM 17-4PH parts.

Particle size and carbon content


Densification. Fine powder particles and low-carbon content improve the densification
properties of 17-4PH parts. The effect of gas-atomized 17-4PH powders in the range
of 10 to 32 µm was investigated by MIM.66 The result showed that the particle size dis-
tribution at higher sintering temperatures (above 1300 °C) had little effect on the tensile
strength. On the contrary, the highest mechanical properties (tensile strength, yield
strength, and elongation) were obtained at the lowest sintering temperature (1149 °C)
with the lowest particle size (10 µm).
There is an inverse connection between the carbon content and the densification prop-
erty of powdered steel materials. The effect of using water-atomized 17-4PH powder was
studied, but it was prepared using two different methods: conventional atomization
(CAT) and HPAT.20 HPAT 17-4PH powder has lower carbon and nitrogen content
and finer particles than the CAT powder. The results showed that the sintered parts
made of HPAT powder had a higher density (7.4 g/cm3) than those made of CAT
powder (7.3 g/cm3), regardless of the sintering conditions. Moreover, parts with low-
carbon content in the range of 0.02 to 0.04 wt.% gave optimum combinations of strength
and ductility, although the wide specification range for carbon is 0.00 to 0.070 wt.% in
the MIM 17-4PH parts.11 The percentage of carbon in low-carbon stainless steel must
be kept low to prevent carbide precipitation from chromium carbides. For example, in
austenitic stainless steels that are cooled slowly, the carbon content must be in the
range of 0.02–0.03 wt.%. The results showed that, as the amount of carbon in the material
increased, the tensile strength, yield strength, and hardness all increased, but elongation
decreased.40
Nitrogen has a comparable effect to carbon on martensite formation. Control of nitro-
gen content is equally essential for maximizing ductility.11 Nitrogen acts as an austenite
stabilizer in 17-4PH parts. The denser austenite slows the rate of shrinkage and densifi-
cation. It also prevents the formation of the favorable ferrite phase, which hinders densi-
fication.47 During sintering, nitrogen penetrates the steel and forms nitrides. This impedes
the diffusion of iron or chromium and restricts the densification rate.67
The phase diagram of 17-4PH with a carbon content of less than 0.04 wt.% indicates
that the liquid phase occurs slightly above 1360 °C (see Figure 12).7 The formation of
δ-ferrite during high sintering temperatures and the development of the microstructure
during sintering are summarized in the following structures: ferrite (α); austenite (γ);
and γ + δ-ferrite. The phase diagram shows that the formation of δ-ferrite strongly
depends on the carbon content, and δ-ferrite drops with increasing carbon content.32
However, an investigation of the microstructural evolution of MIM 17-4 PH stainless
steel at room temperature revealed that a martensitic phase occurs.69 The martensite trans-
forms into γ at approximately 700 °C, which is consistent with the corresponding phase
Kareem et al. 17

Figure 12. Phase diagram of 17-4PH alloy (reproduced from Ref.7).

diagram in Figure 13. The second-phase transformation is the transformation of γ to


δ-ferrite, which starts at nearly 1200 °C and is related to the dilatometric peaks explained
in the Sintering section.
The above studies discussed how carbon content affects density, mechanical proper-
ties, and ferrite volume fraction. However, the effect of carbon content on carbide forma-
tion and oxidation (reduction) remains unclear in PSM 17-4PH parts. In addition, the
conducive relationship between carbon content and sintering conditions (temperature
and time) for the formation of carbides and oxides has not been clarified.

Residual carbon effect. Furthermore, the formation of δ-ferrite decreased the pores and
improved the densification of sintered 17-4PH materials.47 This phenomenon is observed
because the diffusion volume of the ferrite phase structure (BCC) is faster than in the aus-
tenite phase structure (FCC) of low-carbon steels.7 According to Baba and Kyogoku, an
abrupt decrease in tensile strength occurred at high residual carbon content, which they
attributed to the effect of retained austenite.6 The consequence of retained austenite on the
mechanical behavior of 17-4PH parts produced by the selective laser melting method (an
AM process) was investigated. The results showed that the parts with large amounts of
retained austenite exhibited stress-induced transformation to martensite during tensile
18 Science Progress 106(1)

Figure 13. Phage diagram of 0.1 C%-steel, 18% chromium, and 4% nickel (reproduced from
Ref.69).

testing. This behavior was associated with much lower yield strengths, faster strain hard-
ening over a wider range of strains, and a later start to local plastic deformation.70
A study was conducted to investigate the relationship between carbon content and
densification property and retained austenite formation in the area of press and sinter
of 17-4PH parts, as mentioned in the literature. However, the influence of carbon
content on the volume fraction of δ-ferrite, the retained austenite, and phase transform-
ation during sintering requires further research. Additionally, more research is necessary
regarding the effects of carbon content on the mechanical and physical properties of PSM
17-4PH parts.

Sintering atmosphere and oxidation


Atmosphere. By controlling the phases in the microstructures of 17-4PH materials, it is
possible to determine a connection between the sintering atmosphere and the densifica-
tion property. δ-ferrite formation in PSM 17-4PH parts produced under different sintering
atmospheres has been studied.47 The results illustrated that the appearance of δ-ferrite
began in a pure hydrogen (H2) environment with a sintering temperature above 1220 °
C, which can be seen in Figure 14(a) along the grain boundaries (whitest color). Other
results obtained using H2 and hydrogen/nitrogen (H2/N2) atmospheres at an elevated sin-
tering temperature (1360 °C) revealed that no δ-ferrite phase formed and that the micro-
structure contained significant amounts of porosity, as shown in Figure 14(b).
Unfortunately, these results cannot be compared with other studies (e.g. Figure 12)
Kareem et al. 19

Figure 14. Microstructures of press and sinter 17-4PH parts (a) quenched from indicated
temperatures in pure H2, and (b) quenched after sintering for 1 h at 1365 °C under the specified
atmospheres (reproduced from Ref.47).

due to the absence of carbon content. Additionally, it was mentioned that δ-ferrite forma-
tion can occur at a temperature lower than 1220 °C, as shown in Figure 15, according to
the two-phase master sintering curve model.69 Moreover, Ar + 20% H2 was used as an
atmosphere to sinter PSM 17-4PH green compacts.33 A range of sintering temperatures
from 1240 to 1290 °C for 90 min was used. A chromium–nickel–martensite matrix with a
few small islands of δ-ferrite at the grain boundaries comprised the final microstructure.
Thus, the formation temperature of the δ-ferrite depends on the chemical composition of
the powder, the sintering atmosphere, and the residual carbon.31
The exact role of the N2 atmosphere was found in another type of steel. The influence
of the sintering atmosphere on the densification properties of HK30 materials was
studied.67 The results showed that the density was lower when sintered with N2 than
with Ar. However, HK30 parts sintered in N2 had a higher hardness and tensile strength
than those sintered in Ar. This was because N2 was dissolved in the austenitic microstruc-
ture and formed precipitates at the grain boundaries.
Overall, sintering in a vacuum atmosphere lowers the oxygen content. For example,
the results of sintering water-atomized iron powder pre-alloyed with 1.8 wt.% Cr
showed that the oxygen content in the green compacts before sintering was 0.15 wt.%,
while the highest oxygen content after sintering in a vacuum was 0.02 wt.%.38
To achieve a higher final density, oxidation during sintering should be avoided
entirely. Sintering in an Ar atmosphere or vacuum is the preferred choice for 17-4PH
20 Science Progress 106(1)

Figure 15. Delta-ferrite percentage with sintering temperature of 7-4PH parts manufactured by
metal injection molding (reproduced from Ref.69).

materials. Depending on the sintering atmosphere, Cr2O3 may be formed. The effect of
using different sintering atmospheres (vacuum, Ar, and N2/H2) to produce MIM parts
from gas-atomized 17-4PH stainless-steel powder was investigated by X-ray analysis.18
The results are illustrated in Figure 16. No oxygen was found in the case of vacuum or Ar,
whereas oxygen was observed when N2 + H2 was used as a protective atmosphere. Cr2O3
was created after sintering in the N2 + H2 atmosphere, but no Cr2O3 was formed when
sintering in the Ar and vacuum atmospheres. In addition, it was found that the increase
in Cr2O3 was related to the higher sintering temperature Cr2O3 formed on the surface
of the sintered particles, which stopped diffusion and made the material less dense.

Oxides source. Steel powder atomized with water contains more oxide than powder ato-
mized with gas. Therefore, water-atomized sintered powdered materials will always
contain small amounts of fine oxides.11 In addition, these oxides might exist either on
the surface or within the particles. As mentioned in Section 3, surface oxides need to
be reduced for sinter necks to form between powder particles.
The transformation of the surface oxides and their entrapment inside the sinter necks
can occur during the heating phase.38 Oxidized particles enriched with Si and Mn were
Kareem et al. 21

Figure 16. Spectra of line EDS analysis of metal injected molding 17-4PH parts sintered at 1050 °
C under (a) vacuum or Ar, (b) mixture of N2 and H2 atmosphere (reproduced from Ref.18).

found on the surface of H13 steel powder atomized with water.68 After sintering, these
oxide particles of about 1–3 μm were randomly distributed in the grain boundaries and
matrix. Additionally, the gas-atomized powder has a more regular shape than the water-
atomized 17-4PH powder. Therefore, the gas-atomized powder results in more regular
pores (see Figure 11). These regular pores reduce the oxidation rate during the sintering
process because they reduce the amount of exposed surface area.36
Furthermore, different oxides can form in water-atomized steel powders, such as Cr2O3,
NiO, and SiO2. Si is considered critical because most of the oxidation that occurs during
water atomization is the oxidation of Si to SiO2, causing reduction. This reduction
depends on the sintering conditions and results in residual oxides in the sintered component.
Mn, however, increases oxidation during water atomization.40 Several critical oxides in aus-
tenitic stainless steels were residues from the water atomization process: Cr2O3, MoO3, and
SiO2.16 Cr2O3 has high stability and is challenging to reduce, while Mo and Si must be
removed before or during the sintering process. Moreover, the mechanical properties of
stainless-steel parts can be improved by reducing oxides. For example, it was found that
the mechanical properties of MIM 17-4PH parts made from gas-atomized powder were
improved compared with MIM 17-4PH parts made from water-atomized powder, mainly
due to the lower Si, O, and SiO2 contents in the case of gas-atomized 17-4PH powder.24
Even though certain oxides, such as Cr2O3, MoO3, and SiO2, were found in the pressed
and sintered austenitic stainless-steel parts, those analyses did not adequately investigate the
oxides that may be present in the PSM 17-4PH parts. The phases with different sintering
temperatures were well explained,47 but the study did not address the oxides.

Heat treatment
Heat treatment is the postprocessing of a manufactured metal part with the intention to
modify properties, such as strength, ductility, and hardness, by controlling the
22 Science Progress 106(1)

microstructure and achieving the desired phases, grain growth, and precipitation. For
instance, after annealing a CoCrFeMnNi alloy, the strain increased by 5%.71 In addition,
ferrite dissolution reduced the hardness of 316 stainless steel by 25% when solution-
treated at 1050 °C.72 Thus, many studies investigated different conditions for sintered
parts produced from water- and gas-atomized 17-4PH using PSM, MIM, and fused
deposition modeling.11,24,73 However, each experiment used a different atmosphere,
hold time, heating rate, and cooling rate. Solution annealing, H900, and H1100 made
up most of these conditions, as presented in Figure 17. The volume fraction of
δ-ferrite increased after the solution heat treatment. Therefore, this condition is appropri-
ate for a reasonable ductility requirement. An increase in strength, hardness, and a
decrease in ductility is noted after applying H900 (generally 482 °C for 1 h), which
formed Cu precipitates (observed by optical microscope) and less δ-ferrite compared
to the as-sintered parts. Thus, δ-ferrite impacts tensile strength and ductility.7
Increasing the volume fraction of ferrite in low-carbon steel leads to a further rise in duc-
tility and a loss in tensile strength due to the elasticity.74 However, an accumulation of
δ-ferrite and larger Cu precipitates occurred during H1100 (552 °C for 1 to 2 h).
Consequently, an elemental map was used to examine the black spots (not the big
pores), which were distributed in the microstructure of the water-atomized MIM
17-4PH parts.24 The results revealed the presence of SiO2 due to the presence of Si
near oxygen sites, in addition to Cu precipitates (see Figure 18).
After annealing, H900, and H1100 heat treatments, it can be deduced that the
gas-atomized MIM 17-4PH parts had more δ-ferrite and less SiO2 than the water-
atomized parts. Table 2 summarizes the mechanical properties of PSM 17-4PH parts
as a result of the heat treatment conditions.

Figure 17. Optical microstructures from indicated heat-treatment conditions of 17-4PH parts
manufactured by (a) metal injection molding,24 (b) press and sinter.11 M and δ refer to martensite
and delta-ferrite, respectively.
Kareem et al. 23

Figure 18. The elemental map of Si, O, and Cu after heat treatment for 1-h under vacuum
followed by quenching in N2 (reproduced from Ref.24).

Table 2. Represents the changing in mechanical properties according to the heating treatment
conditions of 17-4PH sintered parts.

UTS, Elong.,
Conditions Heat-treatment Hardness MPa % Ref.
11
17-4PH, by PSM As sintered 89 HRB 869 2.5
Solutioned at 1065 °C, 1.5 h, in 90 HRB 816 4.5
hydrogen, fast cooling 2 °C/ min
Aged at 482 °C, 2 h, in nitrogen. 98 HRB 1136 1.4
Aged at 552 °C, 4 h, in nitrogen 95 HRB 1029 3.3
73
17-4PH, by FDM As sintered - 800 -
3D printing Annealing + aging at 480 °C, 1 h - 1140 -
17-4PH with 0.01 Aged at 538 °C, 1 h. ̴ 237 HV 819 1.7 20

C%, by PSM
17-4PH, by PSM Annealing + aging at 480 °C 332 1100 - 3

HV10
65
17-4PH, by MIM Solutioned at 1050 °C, 1 h, quenched in 34 RC 976 1.7
water, and aged in argon at 480 °C
for 4 h
The following references for comparison purposes
Cast 17-4PH Aged at 496 °C, 4 h - ̴ 1380 6–9 56

Aged at 594 °C, 4 h - ̴ 1103 9–13


54
Wrought 17– Aged at 620 °C, 4 h 340 HV1 1017 -
4PH rod

-: not included; 332 HV10: 332 per Vickers with 10 kgf as a test load; Elong.: elongation percent; HRB: Rockwell
hardness; PSM: press and sinter; UTS: ultimate tensile strength.

Corrosion performance
The corrosion of stainless-steel materials in aqueous media is primarily caused by an elec-
trochemical mechanism. This mechanism involves two types of reactions: anodic (oxida-
tion of the metal) and cathodic (reduction in the aqueous medium).75,76 These processes
produce a protective layer of Fe2O3.76 In the case of 17-4PH materials, Fe2O3 and/or
Cr2O3 might form, which function as a barrier between the metal and the service
24 Science Progress 106(1)

environment.77 In addition, the following equations show an electrochemical reaction in a


solution of sulfuric acid in water.78

Oxidation reaction:
Iron oxides to form ferrous ions Fe  Fe2+ + 2e− (1) and ferric ions
Fe  Fe3+ + e− (2). Consequently, the formation of the protective film of iron
2+

oxide occurs:
3H2 O + 2Fe3+  Fe2 O3 + 6H + (4)
The reduction reactions are in sulfuric acid by the evaluated hydrogen:
2H + +2e−  H2 (5)

H2 SO4  H + +HSO−
4 (6)

HSO− +
4  H +SO4
2−
(7)
Thus, H+, Fe2+, Fe3+, HSO− 2−
4 , and SO4 ions form during the corrosion reaction of iron
alloys in a sulfuric acid aqueous medium.
Although few studies have reported the corrosion properties of 17-4PH powdered sin-
tered materials, the porosity ratio and pore geometry are the main defects responsible for
the lower corrosion performance of powdered stainless steels. Isolda Costa et. al. inves-
tigated the effect of pores of 17-4PH MIM parts with a porosity of 3% in naturally aerated
phosphate buffer solution at 25 ± 2 °C.79 The results showed that these parts were sub-
jected to slightly more pitting corrosion than the parts fabricated by conventional metal-
lurgy due to the built-in porosity. The pores increase the geometrical area of the alloy
exposed to the corrosive medium and induce localized corrosion.78,79 Furthermore, the
effect of an aerated medium of 3% NaCl electrolyte on the pitting resistance of
17-4PH parts manufactured by MIM and conventional technology was compared.80
Pitting attack by the chloride ions toward the protective layer of both conventional and
MIM 17-4PH parts was observed. This attack was slightly higher for MIM parts com-
pared to conventional parts due to the higher porosity in MIM parts. Additionally, this
undesired influence of pores was found in additive-manufactured 17-4PH parts. Two dif-
ferent powder particle size distributions were used to investigate the effect of pores size
on the corrosion performance of 17-4PH parts manufactured by laser powder bed fusion
(LPBF).77 The smaller pores from finer powder with D50 = 17 µm resulted in improved
corrosion performance compared to the larger pores from wide-range powder with D50 =
17 to 43 µm, at an energy density of 64 J/mm3. Moreover, the pores, especially the large
pores, are responsible for the worse corrosion behavior of the LPBF parts compared with
forged parts.81
Surface roughness, nonmetallic inclusions, and phase microstructure all play a role in
corrosion performance. Improving surface roughness by the shot peening process
resulted in a decreased the corrosion rate of 17-4PH parts manufactured by direct
metal laser sintering.76 These surfaces work as a concentration of hydrogen, the same
as porosity in an acid medium.78 Additionally, the inclusions, such as MnS, stressed mar-
tensite, incoherent Cu precipitates, and M23C6 carbides, lead to an acceleration of the
Kareem et al. 25

corrosion rate in the 17-4PH parts manufactured by conventional metallurgy.75


Therefore, heat treatment is necessary to control the microstructure of 17-4PH parts.
The effect of aging temperature on the corrosion performance of PSM 17-4PH parts sin-
tered at 1340 °C and solutioned at 1040 °C in a diluted sulfuric acid medium was inves-
tigated.78 The results revealed decreasing corrosion rates with decreasing aging
temperatures from 500 to 480 °C. This can happen as a result of coherent Cu precipitates
and stress relief in the martensite matrix.75
More research into corrosion performance is required, particularly for 17-4PH parts
manufactured by PSM. Since the porosity is the main factor that influences the passivity
of the stainless-steel powdered materials, the corrosion performance should be investi-
gated in parallel with the densification of the PSM 17-4PH parts and in different corrosive
aqueous media, for example, NaCl and H2So4. Additionally, different heat treatment con-
ditions need to be performed on these parts to investigate the effects of other important
microstructural factors on the corrosion rate. These factors can be the coherent distribu-
tion of Cu precipitates, the structure of the martensitic matrix, and the volume fraction of
δ-ferrite and austenite.

Conclusions and prospects


PSM remains the leader among the powder metallurgy processes used for manufacturing
stainless-steel materials because it enables low cost, low oxidation rate, and adequate cor-
rosion resistance. According to the data provided in the metal market powder size report
from 2016 to 2027, the ferrous parts produced by PSM have the highest demand among
other technologies. However, using PSM in the production of 17-4PH parts is still limited
due to the low compressibility of this martensitic stainless-steel powder, which leads to
low densification. Therefore, this review analyzed the previous investigations to examine
the factors influencing the limitations of 17-4PH parts. These factors included the 17-4PH
powder characteristics (particle size, atomization method, chemical compositions, and
particle morphology) and PSM parameters (cold pressing, sintering temperature, sinter-
ing time, and sintering atmosphere). Additionally, this paper found a series of scientific
gaps in the literature. Moreover, the previous work on PSM 17-4PH parts helped to sum-
marize the optimal parameters that were applied and suggest theoretical solutions to
improve the densification properties. The conclusions can be summarized as follows:

1. Previous PSM studies used a range of cold pressing up to 760 MPa, and the
maximum relative green density was around 80%. Therefore, establishing a
clear relationship between cold pressing (higher than 760 MPa) and green
density is necessary. Additionally, most PSM studies used a particle size distribu-
tion of 30 to 100 µm. Thus, further studies are needed to investigate 17-4PH with
a particle size smaller than 30 µm. Moreover, a sintering time of 60 min at a tem-
perature of 1260 °C resulted in sufficient sintering of the PSM 17-4PH parts. The
increase in density did not change much when the sintering temperature was
raised above 1300 °C.
2. The high diffusion rate of δ-ferrite demonstrated the densification property of
PSM 17-4PH parts. δ-ferrite can be formed at a sintering temperature lower
26 Science Progress 106(1)

than 1220 °C. Carbon, nitrogen, and the sintering atmosphere are the main factors
affecting the volume fraction of δ-ferrite. However, the influence of carbon
content and residual carbon on the carbide formation, the volume fraction of
retained austenite, oxidation, microstructural phases, and mechanical properties
of parts made from PSM 17-4PH has yet to be carefully examined. Moreover,
vacuum and argon sintering atmospheres were effective for 17-4PH parts to
remove oxides and reduce porosity due to the high affinity of carbon to
oxygen, but the nitrogen atmosphere caused a decrease in density.
3. Cr2O3 can form in sintered 17-4PH parts at elevated sintering temperatures and is
also related to the sintering atmosphere (hydrogen and nitrogen). Cr2O3 was not
found when a vacuum was used as the protective atmosphere. Further studies need
to investigate the dependence of oxide formation on the sintering temperatures of
the PSM 17-4PH parts and the possibility of elimination of oxides by reducing the
pore ratio.
4. Gas-atomized 17-4PH powder has a more uniform shape than water-atomized
powder, which improved the resulting strength of sintered parts. Gas-atomized
17-4PH powder also led to less oxidation during the sintering process. This
was because the pores were more regular.
5. For high strength and hardness with acceptable ductility, H900 was a preferred
heat treatment condition. The solution and annealing heat treatments were suitable
for increasing the ductility. Using the H1100 condition resulted in lower strength
and hardness and higher ductility than H900. Additionally, as the amount of
δ-ferrite in the microstructure of 17-4PH parts increased, their tensile properties
decreased.

Correspondingly, applying a cold pressing of more than 760 MPa was one of the theor-
etical suggestions that was determined in this review. Herein, the effect of applying a
range of cold pressing (800–1600 MPa) using water-atomized 17-4PH powder with a
low particle size distribution ( + 15/−32 m) was investigated. The experimental results
showed an improvement in green densification with increasing cold pressing from
78% at 800 MPa to 90% at 1600 MPa. In addition, the relative sinter density determined
by ImageJ increased from 84.43% to 96.43% for the samples cold pressed at 800 and
1600 MPa, respectively.
Finally, the literature did not mention how an extended sintering time affects the prop-
erties of PSM 17-4PH parts. Most studies used a sintering time of 20 to 90 min. However,
it was mentioned in the literature that increasing the sintering time to seven hours at a
sintering temperature higher than 1280 °C led to overheating and negatively affected
the densification of PSM HK30 steel parts. Therefore, the effects of increasing the sinter-
ing time (over two hours) on the final PSM 17-4PH properties are unknown. However, at
a sintering temperature of 1200 °C, the experimental results showed that increasing the
sintering time up to seven hours had a better impact on densification.
In summary, it was found that green densification was the most critical factor, which
directly influenced the sintering density, and consequently, the mechanical properties.
Increasing the sintering temperature was also beneficial, but it may cause oxidation
and carbide precipitation. Moreover, increasing sintering time can result in improved
Kareem et al. 27

density, but it must be controlled to keep grains from overgrowing. Furthermore, the
initial 17-4PH powder characteristics affected the final product density. Finally, PSM
can be used to achieve near full density, high mechanical properties, and economic
17-4PH products.

Acknowledgments
This work was financed by the UMA3 project, which has received funding from the European
Union’s Horizon 2020 research and innovation program under grant agreement no. 952463.

Declaration of conflicting interests


The author(s) declared no potential conflicts of interest with respect to the research, authorship, and/
or publication of this article.

Funding
The author(s) disclosed receipt of the following financial support for the research, authorship, and/
or publication of this article: This work was supported by the UMA3 project (grant no. 952463).

ORCID iD
Mohammed Qasim Kareem https://orcid.org/0000-0001-9140-5940

References
1. Frank P. An overview on powder metallurgy (PM). In: UMA3 Summer School on Powder
metallurgy and additive manufacturing by Fraunhofer IFAM in Bremen, Germany (31
May-2 June 2022)
2. Grand View Research Company. Metal Powder Market Size Analysis Report, 2020–2027.
Report ID: GVR-3-68038-948-7, Research Reports in Advanced Materials, 2022.
3. Kazior J, Szewczyk-Nykiel A, Pieczonka T, et al. Properties of precipitation hardening 17-4
PH stainless steel manufactured by powder metallurgy technology. Adv Mater Res 2013;
811: 87–92.
4. Patibandla AR. Effect of process parameters on surface roughness and porosity of direct metal
laser sintered metals. Master thesis, University of Cincinnati, Cincinnati, USA, 2018.
5. Mohammad Moradi J, Emadoddin E and Omidvar H. Transient liquid phase bonding of
17-4-ph stainless steel using conventional and two-step heating process. Met Mater Int
2020; 27: 5268–5277.
6. Zhang H. Powder injection molding of 17-4ph stainless steel. In: Metal Powder Industries
Federation (ed) Powder injection molding symposium 1992, pp.219–227.
7. Schroeder R, Hammes G, Binder C, et al. Plasma debinding and sintering of metal injection
moulded 17-4PH stainless steel. Materials Research 2011; 14: 564–568.
8. Inovar Communications Ltd. The processing of 17-4 PH MIM in China: company visits HIP
for MIM: cost calculations. Inovar Communications Ltd. 2018; 12: 49–76.
9. Szewczyk-Nykiel A. The effect of the addition of boron on the densification, microstructure
and properties of sintered 17-4 PH stainless steel. Technical Transactions Mechanics. In:
Czasopismo Techniczne, 2014: 86–96.
28 Science Progress 106(1)

10. Wang FZ, Wang QZ, Yu BH, et al. Interface structure and mechanical properties of Ti(C,
N)-based cermet and 17–4PH stainless steel joint brazed with nickel-base filler metal BNi-2.
J Mater Process Technol 2011; 211: 1804–1809.
11. Samal PK, Nandivada N, Hauer I, et al. Properties of 17-4ph stainless steel produced via press
and sinter route. Adv Powder Metallurg Particulat Mater 2008; 7: 109–120.
12. Altıparmak SC, Yardley VA, Shi Z, et al. Challenges in additive manufacturing of high-
strength aluminium alloys and current developments in hybrid additive manufacturing. Int J
Lightweight Mater Manuf 2021; 4: 246–261.
13. Wu Y, German RM, Blaine D, et al. Effects of residual carbon content on sintering shrinkage,
microstructure and mechanical properties of injection molded 17-4 PH stainless steel. J Mater
Sci 2002; 37: 3573–3583.
14. Singh G, Missiaen JM, Bouvard D, et al. Additive manufacturing of 17–4 PH steel using metal
injection molding feedstock: analysis of 3D extrusion printing, debinding and sintering.
Additive Manufac 2021; 47: 102287.
15. Zaky MT, Soliman FS and Farag AS. Influence of paraffin wax characteristics on the formu-
lation of wax-based binders and their debinding from green molded parts using two compara-
tive techniques. J Mater Process Technol 2009; 209: 5981–5989.
16. Kozub B, Kazior J and Szewczyk-Nykiel A. Sintering kinetics of austenitic stainless steel aisi
316l modified with nanographite particles with highly developed bet specific surface area.
Materials (Basel) 2020; 13: 4569.
17. Cristofolini I, Rao A, Menapace C, et al. Influence of sintering temperature on the shrinkage
and geometrical characteristics of steel parts produced by powder metallurgy. J Mater Process
Technol 2010; 210: 1716–1725.
18. Ye H, Liu XY and Hong H. Sintering of 17-4PH stainless steel feedstock for metal injection
molding. Mater Lett 2008; 62: 3334–3336.
19. Bai Y, Li L, Fu L, et al. A review on high velocity compaction mechanism of powder metal-
lurgy. Sci Prog 2021; 104: 1–20.
20. Schade CT, Stears PD, Lawley A, et al. Precipitation-hardening PM stainless steels. Int J
Powder Metall 2007; 43(4).
21. Irrinki H. Mechanical properties and microstructure evolution of 17-4 PH stainless steel pro-
cessed by laser-powered bed fusion. Master thesis, University of Louisville, Louisville, USA,
2016.
22. Dobson S, Vunnam S, Frankel D, et al. Powder variation and mechanical properties for SLM
17-4 PH stainless steel. In: Texas university/ Mechanical Engineering (ed) International solid
freeform fabrication symposium, University of Texas at Austin, Austin, USA, 2019, pp. 478–
493.
23. Muratal O and Yamanoglu R. Production of 316l stainless steel used in biomedical
applications by powder metallurgy. In: Scientific meeting on electrical-electronics &
biomedical engineering and computer science, IEEE, Istanbul, Turkey, 24–26 April
2019, pp.1–4.
24. Wu MW, Huang ZK, Tseng CF, et al. Microstructures, mechanical properties, and fracture
behaviors of metal-injection molded 17-4PH stainless steel. Met Mater Int 2015; 21: 531–537.
25. Dizdar S, Skoglund P and Bengtsson S. Process, quality and properties of high-density P/M
gears. Adv Powder Metall Part Mater 2003; 9: 9–36.
26. Qasim Kareem M. Theoretical and practical solutions for producing 17-4ph sintered com-
pacted parts. Seminar Report, University of Miskolc Antal Kerpely Doctoral School of
Materials Science and Technology, Miskolc, Hungary, January 2022.
27. Szewczyk-Nykiel A and Bogucki R. Sinter-bonding of aisi 316l and 17-4 PH stainless steels. J
Mater Eng Perform 2018; 27: 5271–5279.
Kareem et al. 29

28. Sung HJ, Ha TK, Ahn S, et al. Powder injection molding of a 17-4 PH stainless steel and the
effect of sintering temperature on its microstructure and mechanical properties. J Mater
Process Technol 2002; 130: 321–327.
29. Khalil AK and Kim SW. Relationship between binder contents and mechanical properties of 17-4
PH stainless steel fabricated by PIM process and sintering. Met Mater Int 2006; 12: 101–106.
30. Shi J, Cheng Z, Gelin JC, et al. Sintering of 17-4PH stainless steel powder assisted by micro-
wave and the gradient of mechanical properties in the sintered body. Int J Adv Manufac
Technol 2017; 91: 2895–2906.
31. Gulsoy HO, Ozbek S and Baykara T. Microstructural and mechanical properties of injection
moulded gas and water atomised 17-4 PH stainless steel powder. Powder Metall 2007; 50:
120–126.
32. Chang CW, Chen PH and Hwang KS. Enhanced mechanical properties of injection molded
17-4PH stainless steel through reduction of silica particles by graphite additions. Mater
Trans 2010; 51: 2243–2250.
33. Firouzdor V and Simchi A. Co-sintering of M2/17-4PH powders for fabrication of functional
graded composite layers. J Compos Mater 2010; 44: 417–435.
34. Sundaram MV, Khodaee A, Andersson M, et al. Experimental and finite element simulation
study of capsule-free hot isostatic pressing of sintered gears. Int J Adv Manuf Technol 2018;
99: 1725–1733.
35. Sundaram MV. Novel approaches for achieving full density powder metallurgy steels.
Doctoral dissertation, Chalmers Tekniska Hogskola, Sweden, 2019.
36. Bautista A, Moral C, Velasco F, et al. Density-improved powder metallurgical ferritic stainless
steels for high-temperature applications. J Mater Process Technol 2007; 189: 344–335.
37. Kareem MQ and Dorofeyev V. Effect of fullerenes additions on physical–mechanical proper-
ties of hot-forged iron-based powder materials. IOP Conf Series Earth Environ Sci 2021; 877:
012009.
38. Vattur Sundaram M, Karamchedu S, Gouhier C, et al. Vacuum sintering of chromium alloyed
powder metallurgy steels. Met Powder Rep 2019; 74: 244–250.
39. Coovattanachai O, Lasutta P, Tosangthum N, et al. Analysis of compaction and sintering of
stainless steel powders. Chiang Mai J Sci 2006; 33: 293–300.
40. Al-Mangour B. Powder metallurgy of stainless steel: state-of-the art, challenges, and develop-
ment. Nova Sci Publ 2015: 37–80.
41. Kondo S, Kume Y, Kobashi M, et al. Densification behavior of different metal powders by
compression and shear combined loading. Procedia Eng 2014; 81: 1175–1179.
42. Schade C. Processing, microstructures and properties of a dual phase precipitation-hardening
PM stainless steel. Doctoral dissertation, Drexel University, Philadelphia, USA, 2010.
43. Li D, Chen S, Shao WQ, et al. Application of master sintering curve theory to predict and
control the sintering of nanocrystalline tio2 ceramic. Key Eng Mater 2008; 368: 1588–1590.
44. Johnson DL. Solid-state sintering. In Brook RJ (ed) Concise encyclopedia of advanced
ceramic materials. Pergamon: Elsevier, 1991, pp.454–458.
45. Bordia RK, Kang SJL and Olevsky EA. Current understanding and future research directions at
the onset of the next century of sintering science and technology. J Am Ceram Soc 2017; 100:
2314–2352.
46. Bergman O. Studies of oxide reduction and nitrogen uptake in sintering of chromium-alloyed
steel powder. Doctoral dissertation, KTH Industrial Engineering and Management, Stockholm,
Sweden, 2008.
47. Blaine DC, Wu Y, Schlaefer CE, et al. Sintering shrinkage and microstructure evolution during
densification of a martensitic stainless steel. In: Proceedings Sintering conference, 2003.
Materials Research Institute, Pennsylvania State University, Pennsylvania, USA.
30 Science Progress 106(1)

48. Imgrund P, Rota A and Simchi A. Microinjection moulding of 316L/17-4PH and 316L/Fe
powders for fabrication of magnetic–nonmagnetic bimetals. J Mater Process Technol 2008;
200: 259–264.
49. Hu Y, Li Y, Lou J, et al. Effects of sintering temperature and holding time on densification
and mechanical properties of MIM HK30 stainless steel. Int J Metallurg Metal Physics
2018; 3: 22.
50. Kandala SR, Balani K and Upadhyaya A. Mechanical and electrochemical characterization of
supersolidus sintered austenitic stainless steel (316 L). High Temp Mater Processes 2019; 38:
792–805.
51. Gülsoy HÖ, Salman S and Özbek S. Effect of FeB additions on sintering characteristics of
injection moulded 17-4PH stainless steel powder. J Mater Sci 2004; 39: 4835–4840.
52. Yoon TS, Lee YH, Ahn SH, et al. Effects of sintering conditions on the mechanical properties
of metal injection molded 316L stainless steel. ISIJ Int 2003; 43: 119–126.
53. Erden MA, Yaş ar N, Korkmaz ME, et al. Investigation of microstructure, mechanical and
machinability properties of Mo-added steel produced by powder metallurgy method. Int J
Adv Manuf Technol 2021; 114: 2811–2827.
54. Eskandari H, Lashgari HR, Ye L, et al. Microstructural characterization and mechanical prop-
erties of additively manufactured 17–4PH stainless steel. Mater Today Commun 2022; 30:
103075.
55. Bala P, Krawczyk J, Pawlowski B, et al. The kinetics of phase transformations during continu-
ous heating from as-quenched state of 17-4PH steel. Metal 2015.
56. Susan DF, Crenshaw TB and Gearhart JS. The Effects of Casting Porosity on the Tensile
Behavior of Investment Cast 17-4PH Stainless Steel. Journal of Materials Engineering and
Performance 2015; 24 (8): 2917–2924.
57. Trzepieciń ski T, Pieja T, Malinowski T, et al. Investigation of 17-4PH steel microstructure and
conditions of elevated temperature forming of turbine engine strut. J Mater Process Technol
2018; 252: 191–200.
58. Wilcox H, Lewis B and Styman P. Evaluation of the mechanical properties of precipitation-
hardened martensitic steel 17-4ph using small and shear punch testing. J Mater Eng
Perform 2021; 30: 4206–4216.
59. Isogawa S, Yoshida H, Hosoi Y, et al. Improvement of the forgability of 17-4 precipitation
hardening stainless steel by ausforming. J Mater Process Technol 1998; 74: 298–306.
60. Murthy AS. Role of alloy additions on strengthening in 17-4 PH stainless steel. Doctoral dis-
sertation, Missouri University of Science and Technology, Rolla, Missouri, USA, 2012.
61. Kurgan N, Sun Y, Cicek B, et al. Production of 316L stainless steel implant materials by powder
metallurgy and investigation of their wear properties. Chin Sci Bull 2012; 57: 1873–1878.
62. Furlan KP, Binder C, Klein AN, et al. Thermal stability of the MoS2 phase in injection
moulded 17-4 PH stainless steel. J Mater Res Technol 2012; 1: 134–140.
63. Kurgan N. Effect of porosity and density on the mechanical and microstructural properties of
sintered 316L stainless steel implant materials. Mater Des 2014; 55: 235–241.
64. Chagnon F and Trudel Y. Effect of density on mechanical properties of sinter hardened P/M
materials. Advances in P/M & Particulate Materials, 1999; 3: 112–119.
65. Gülsoy HÖ. Influence of nickel boride additions on sintering behaviors of injection moulded
17-4 PH stainless steel powder. Scr Mater 2005; 52: 187–192.
66. Murray K, Coleman AJ, Tingskog TA, et al. Effect of particle-size distribution on processing
and properties of MIM 17-4 PH. Int J Powder Metall 2011; 47: 21–28.
67. Yu Y, Li Y, Lou J, et al. Microstructure and mechanical properties of metal injection
molding HK30 stainless steel sintered in N2 and Ar atmosphere. Int J Metallurg Metal
Physics 2019; 4: 30.
Kareem et al. 31

68. Zhang B, Yang F, Qin Q, et al. Characterisation of powder metallurgy H13 steels prepared
from water atomised powders. Powder Metall 2020; 63: 9–18.
69. Jung ID, Ha S, Park SJ, et al. Two-phase master sintering curve for 17-4 PH stainless steel.
Metall Mater Trans A 2016; 47: 5548–5556.
70. LeBrun T, Nakamoto T, Horikawa K, et al. Effect of retained austenite on subsequent thermal
processing and resultant mechanical properties of selective laser melted 17–4 PH stainless
steel. Mater Des 2015; 81: 44–53.
71. Oliveira JP, Shamsolhodaei A, Shen J, et al. Improving the ductility in laser welded joints of
CoCrFeMnNi high entropy alloy to 316 stainless steel. Mater Des 2022; 219: 110717.
72. Rodrigues TA, Escobar JD, Shen J, et al. Effect of heat treatments on 316 stainless steel parts
fabricated by wire and arc additive manufacturing: microstructure and synchrotron X-ray dif-
fraction analysis. Additive Manuf 2021; 48: 102428.
73. Abe Y, Kurose T, Santos M, et al. Effect of layer directions on internal structures and tensile
properties of 17-4ph stainless steel parts fabricated by fused deposition of metals. Materials
(Basel) 2021; 14: 243.
74. Sunil B and Rajanna S. Evaluation of mechanical properties of ferrite-martensite DP steels pro-
duced through intermediate quenching technique. SN Appl Sci 2020; 2: 1–8.
75. Alnajjar M. Corrosion properties of 17-4 PH martensitic stainless steel obtained by additive
manufacturing. Doctoral dissertation, Lyon University, Lyon and Saint-Étienne, France, 2019.
76. Michla JR, Nagarajan R, Krishnasamy S, et al. Conventional and additively manufactured
stainless steels: a review. Trans Indian Inst Met 2021; 74: 1261–1278.
77. Irrinki H, Harper T, Badwe S, et al. Effects of powder characteristics and processing conditions
on the corrosion performance of 17-4 PH stainless steel fabricated by laser-powder bed fusion.
Prog Additive Manuf 2018; 3: 39–49.
78. Szewczyk-Nykiel A and Kazior J. Effect of aging temperature on corrosion behavior of sin-
tered 17-4 PH stainless steel in dilute sulfuric acid solution. J Mater Eng Perform 2017; 26:
3450–3456.
79. Costa I, Rogero SO, Saiki M, et al. Corrosion resistance and cytotoxicity study of 17-4PH
steels produced by conventional metallurgy and powder injection molding. Mater Sci
Forum 2008; 591: 18–23.
80. Costa I, Franco CV, Kunioshi CT, et al. Corrosion resistance of injection-molded 17-4PH steel
in sodium chloride solution. Corrosion 2006; 62: 357–365.
81. Garcia-Cabezon C, Castro-Sastre MA, Fernandez-Abia AI, et al. Microstructure–hardness–
corrosion performance of 17–4 precipitation hardening stainless steels processed by selective
laser melting in comparison with commercial alloy. Met Mater Int 2022; 10: 1–6.

Author biographies
Mohammed Qasim Kareem is a PhD student in materials engineering. His area of research is
casting, powder metallurgy, AM of metal, heat treatment and microstructure analysis.

Tamás Mikó holds a PhD in Material Sciences and Technologies. His area of research is Material
Sciences and Technologies.

Gréta Gergely is an associate professor in materials science. Her area of research is Al alloys, com-
posite materials, bioceramics and image analysis.

Zoltán Gácsi is a professor in Physical Metallurgy. His area of research is computer model for phys-
ical metallurgy process, Solidification, Quantitative microscopy, Stereology, and Image analysis.

You might also like