Microstructure, Mechanical and Tribological Properties of Nickel-Aluminium Bronze Alloys Developed Via Gas-Atomization and Spark Plasma Sintering
Microstructure, Mechanical and Tribological Properties of Nickel-Aluminium Bronze Alloys Developed Via Gas-Atomization and Spark Plasma Sintering
Microstructure, Mechanical and Tribological Properties of Nickel-Aluminium Bronze Alloys Developed Via Gas-Atomization and Spark Plasma Sintering
PII: S0921-5093(17)31203-0
DOI: http://dx.doi.org/10.1016/j.msea.2017.09.047
Reference: MSA35514
To appear in: Materials Science & Engineering A
Received date: 6 July 2017
Revised date: 9 September 2017
Accepted date: 11 September 2017
Cite this article as: Wenzheng Zhai, Wenlong Lu, Po Zhang, Mingzhuo Zhou,
Xiaojun Liu and Liping Zhou, Microstructure, mechanical and tribological
properties of nickel-aluminium bronze alloys developed via gas-atomization and
spark plasma sintering, Materials Science & Engineering A,
http://dx.doi.org/10.1016/j.msea.2017.09.047
This is a PDF file of an unedited manuscript that has been accepted for
publication. As a service to our customers we are providing this early version of
the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting galley proof before it is published in its final citable form.
Please note that during the production process errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
Microstructure, mechanical and tribological properties of nickel-aluminium bronze
Wenzheng Zhai, Wenlong Lu1, Po Zhang, Mingzhuo Zhou, Xiaojun Liu, Liping Zhou
The State Key Laboratory of Digital Manufacturing Equipment and Technology, School of Mechanical Science and
Engineering, Huazhong University of Science and Technology (HUST), Wuhan 430074, PR China
Abstract: This work investigated the effect of sintering temperatures (600-750 °C) on mechanical and
tribological behaviors of nickel-aluminium bronze (NAB) alloys developed by gas atomization and spark
plasma sintering. Results indicated an increase of the volume fraction of B2-type NiAl precipitates with
increasing sintering temperature, leading to an improvement of the yield strength and the wear resistance.
Specifically, detailed microstructural analyses of sintered NAB alloys at 750 °C showed the presence of
ultrafine grains with an average size of 367 nm, nanoscale twins with a volume fraction of 11.7% and
dislocations with a density of 1.3 ± 0.1 × 1014 m-2. A much higher yield strength of 620 MPa was obtained in
NAB, if compared to that of conventional cast counterparts (280-440 MPa). Estimations of strengthening
mechanisms suggested the predominant mechanism of grain boundary strengthening (335MPa) for NAB
alloys with contributions from precipitate strengthening (54 MPa), dislocation strengthening (75 MPa), twin
boundary strengthening (89 MPa) and solid solution strengthening (50 MPa). Moreover, dislocations was
blocked at twin boundaries to form complex dislocation barriers and networks, further contributing to the high
strength. The high wear resistance of NAB could be ascribed to the reduction of the local stress around crack
tips due to the high elastic modulus mismatch (ENiAl/ECu), and to the crack extension toughening near the
interface of the matrix and the precipitate caused by the high plastic mismatch (σNiAl/σCu).
1
Corresponding author.
E-mail addresses: hustwenlong@mail.hust.edu.cn (W.L. Lu)
1
1. Introduction
Coherent twin boundaries (TBs) inside grain interiors show a strengthening effect similar to that of grain
boundaries (GBs), and ultrafine-grained copper containing a high density of nanoscale twins has the ultrahigh
strength [1-3]. Additionally, precipitation hardening is an attractive strengthening mechanism for malleable
materials, including most structural alloys of aluminium, magnesium, nickel, titanium, copper and stainless
steels [4-6]. Among these alloys, the nickel-aluminium bronze (NAB) has been extensively used in marine
engineering applications owing to its valuable mechanical and physical characteristics, as well as the
superior corrosion resistance [7, 8]. The fretting wear often occurs between contact surfaces, resulting in
severe galling failure and fretting fatigue of NAB. In the literature, many reports indicated that the
precipitation during aging could increase the strength, hardness and wear resistance of alloys without a
considerable decrease in the corrosion resistance [9-12]. However, previous studies also showed that the
precipitation strengthening of the NAB alloy by conventional aging might be limited due to the
non-uniformly distributed precipitates and coarse particles in the matrix [13, 14].
The cast NAB alloy presents complex microstructures which mainly consist of the Widmanstätten α
phase with coarse grains, intermetallic κ phases and the retained β phase [15]. The κ-phases containing κI-,
κII-, κIII- and κIV-phase are main precipitates in NAB which are distinguishable by the morphology, location
and distribution in the microstructure [13, 16, 17]. The coarse α grains in the cast NAB alloy significantly
decrease its tensile strength [18]. Also, these precipitates show strong effects on the corrosion, fatigue,
The corrosion resistance of the cast NAB alloy is reported to be initially confined to eutectoid α + κIII
structure with slight attack of the copper rich α-phase. This can be attributed to the presence of a greater
proportion of cathodic phase (κIII) in the eutectoid structure as compared to that in the α-grains and also to
2
the closely adjacent anodic and cathodic areas in the eutectoid structure [16]. The κIII-phase normally
appears in a lamellar or a globular (degraded lamellar) form and is described as the nickel-rich (NiAl) phase.
It often grows normal to the α/β boundary, as well as forms at the boundary of the κI-phase. Recently, the
effect of second phases on the fatigue crack growth behavior in the as-cast and annealed NAB has been
analyzed by Xu et al. [14]. The coarse dendritic κII particles in NAB were demonstrated to have an
accelerative effect on the fatigue crack propagation by cracking through α and κII interfaces. It is noteworthy
that the uniformly distributed κIV precipitates with fine grains improves the growth resistance of fatigue
crack in the NAB alloy without the expense of its corrosion resistance. The κIV-phase is a fine precipitate
within the α-phase and is considered to be iron-rich with the chemical compositions of 3Cu-11Al-7Ni-73Fe
The mechanical and tribological behaviors of alloys depend on the amount, morphology and size of
precipitates [20, 21]. Although thermal treatments including dissolution and ageing usually control
characteristics of precipitates of casting parts, casting produces coarse segregated microstructures which
limit their broad applications. Powder metallurgy has demonstrated to be an effective way to reduce
segregations and promote a finer microstructure of alloys that can exhibit a good combination of mechanical
and tribological properties [22, 23]. Spark plasma sintering (SPS) is a consolidation technique combining a
uni-axial compaction with the current. With regard to alloy powders, densification can be achieved at
significantly lower temperatures and faster thermal cycles, thereby inhibiting grain growth and preserving a
fine microstructure. Meanwhile, gas-atomization with SPS offers many advantages including control of
contamination, grain refinement and avoidance of segregation [22]. Gas-atomization enables the large-scale
production of powders (2 kg/min) with a uniform microstructure due to rapid solidification [24, 25]. It has
been confirmed that precipitates presented in the gas-atomized powders and the fraction of precipitates could
remain rather constant after SPS process, which facilitated the enhancement of mechanical properties of
3
alloys [20]. The SPS technique has been applied recently to fabricate Cu alloys with finer grain sizes and
While the SPS process has been proved to be one of the most promising and widely studied method to
consolidate Cu-based alloys keeping fine grain sizes, a very limited attention has been attracted to control
the precipitation of NAB alloys for enhancing their mechanical and tribological performance. To the best of
our knowledge, there has been no report on NAB alloys through the gas-atomization and SPS processes. In
order to gain a full knowledge of the mechanical and tribological behaviors of NAB alloys with
uniformly-distributed fine precipitates by SPS, a more precise relation among the sintering temperature, the
Thus, the aim of the present work is to estimate quantitatively the impact of the SPS sintering
temperature on microstructures of the NAB alloy in terms of the grain size, the chemical composition, and
the size of precipitates. Additionally, the effect of precipitates and nanoscale twins on the mechanical and
tribological properties of NAB alloys is also systematically analyzed by taking into account the Orowan
strengthening mechanism and the Hall-Petch relationship. Mechanical and tribological properties of
SPS-sintered NAB alloys are compared with those of casting and heat-treated counterparts to highlight the
2. Experimental details
The gas-atomized NAB powders, supplied by Metal Products Trade Co. Ltd., were used as the initial
material for the SPS process. Scanning electron microscopy (SEM, JSM-7600F, Jeol Ltd., Tokyo, Japan)
image is shown that the raw powders mainly contain spherical particles having diameters between 5 and 25
μm (Fig. 1). Table 1 gives the chemical composition of the gas-atomized powders. Cylindrical specimens
with a diameter of 26 mm and a height of 7 mm were sintered from the gas-atomized NAB powders by an
4
SPS apparatus (LABOX-1575, SINTER LAND, Japan). From the analysis of Al-Fe [29] and Al-Ni [30]
binary phase diagrams for binary Fe-(20-25) at% Al and Ni-50 at% Al alloys, the Fe3Al (D03) and NiAl (B2)
domains can be obtained at the temperature above 400 ℃, while 650 ℃ is the temperature of the theoretical
ending of the Fe3Al phase. To track the potential phase transformation of the precipitate during the sintering
process, different final sintering temperatures varying from 600 to 750 ℃ were applied with a holding time
of 5 min and a pressure of 50 MPa in a graphite die. NAB-600, NAB-650, NAB-700 and NAB-750 denote
diffractometer with Cu Kα radiation, a JSM-7600F field emission scanning electron microscope (FE-SEM)
and a JEOL JEM 2100 transmission electron microscope (TEM) with Energy Dispersive X-Ray
Spectroscopy (EDX). To determine the evolution of the precipitation during the SPS process, the samples
were polished followed by etching with a solution of 5 g FeCl3 + 15 mL HCl + 80 mL H2O. The etched
specimens were examined and the volume fraction of the precipitation were analyzed using Image-Pro Plus
software. Morphologies of surfaces after wear tests were investigated by the combination of an OLYMPUS
DSX510 3D optical microscope (OM) and the FE-SEM. SEM imaging with the energy dispersive
spectroscopy (EDS) served for the investigation of wear mechanisms of sintered bulk materials, and
profilometry was conducted to provide a quantitative comparison of the wear volume to determine the wear
Mechanical properties were assessed by rate-controlled: (i) compression tests on cylindrical specimens
with a diameter of 3.0 mm and a height of 4.5 mm; (ii) tensile tests on samples with a width of 2.0 mm, a
thickness of 1.8 mm and a gage length of 6.0 mm. All samples were sectioned by wire cutting from SPS
disks. The tests were conducted at a strain rate of 1.0 × 10-3 s-1 using a Shimadzu AG-100KN device. Five
specimens were tested for each sintering condition. The yield strength, the ultimate strength and the strain at
Friction and wear tests were carried out using a torsional fretting test system with a flat-on-flat
configuration under contact loads of 43-106 N and a frequency of 2 Hz for 20000 cycles. The diagram of the
configuration can be found in our previous study [31]. The angular displacement amplitude (θ) of tests was
1.5°. These experimental parameters were set based on the practical operating conditions [32]. The
counterpart was 42CrMo4 steel with a hardness of 25 HRC and a roughness of Ra = 0.2 μm. All tests were
conducted at least three times to make sure the repeatability of experimental results at same conditions, and
3. Results
Fig. 2 shows the evolution of XRD patterns of NAB samples sintered at 600, 650, 700 and 750 °C,
which confirms that the main peak of all samples is indexed to the α phase (JCPDF no. 50-1295). Fe3Al
precipitates can be also observed in all samples, and the (220) peak related to the Fe3Al phase increases with
6
increasing the sintering temperature from 600 to 650 °C. Besides, the peak splitting presents on the right of
the Fe3Al peak (denoted by the dashed arrow) of the sample sintered at 700 °C (as shown in the inset of Fig.
2). In particular, complete bifurcation occurs for NAB-750 with the reduced Fe3Al content and the increased
NiAl content. As the sintering temperature increases, the (220) peak shifts to a higher angle and the NiAl
content increases, indicating a distortion of lattice parameters in the Fe3Al-NiAl system with an increased
mole fraction ratio of NiAl/Fe3Al. In the D03 structure, three types of ordered domain boundaries including
B2(I)-, B2(II)- and D03-types could be present at lower temperatures [33]; whereas, for the Fe3Al
composition, Ni substitute for the Fe position leads to the lattice distortion induced by different atomic radii
of Ni (rNi = 0.122 nm) [34] and Fe (rFe = 0.124 nm) [35], resulting in a shift of Fe3Al peak to a higher angle.
The partial substitution of Ni for Fe decreases the lattice parameter of Fe3Al, thus reducing the lattice misfit
strain energy for nucleation of nanoparticles. The critical energy for nucleation of NiAl nanoparticles is
reduced, which could promote to a transformation of precipitates from the Fe3Al phase to the NiAl phase.
Consequently, the peak of the Fe3Al phase in XRD patterns decreases with the raise of sintering temperature
(650-750 °C).
According to XRD information, the grain diameter d and microstrain ε of the NAB bulk materials can
be obtained from the true XRD peak broadening B of the samples based on the Williamson-Hall method as
[36]
𝐾𝜆
B cos 𝜃𝐵 = + s 𝜃𝐵 (1)
𝑑
where λ represents the wavelength of Cu Kα X-ray, K is ~0.9, and θB is the Bragg angle. Therefore, the
dislocation density ρ can be calculated from the grain diameter d and microstrain ε by [37]
2√3
𝜌= (2)
𝑑𝑏
√2
where b is the Burgers vector, and equals to a for face-centered cubic (fcc) metals (a is the lattice
2
parameter). Through Eqs. (1) and (2), the grain diameter and the dislocation density for sintered bulk
7
materials are ~281 nm and 1.3 ± 0.1 × 1014 m-2, respectively.
For each sintering temperature condition, the NAB bulk materials have a relative density close to 100%,
indicating the full densification of the alloy at the investigated temperature range. Fig. 3 gives SEM images
at two magnifications showing the microstructures for the bulk materials sintered at 600, 700 and 750 °C.
For each condition, the existence of Fe3Al and NiAl precipitates is observed both at the GBs and in grain
interiors. The amount the Fe3Al precipitate reduces with increasing sintering temperatures; in contrast, the
content of the B2-type NiAl precipitate increases. This confirms the increase of the NiAl peak observed by
XRD. From the SEM and XRD analyses, the total volume fraction of two precipitates is determined as 5%
Table 2. EDX composition analyses of the precipitates in the NAB alloy after the SPS process at 600 °C
Precipitate Al Mn Fe Ni Cu
1 13.9 1.1 40.7 7.3 37.0
2 16.5 1.9 47.1 10.7 23.8
3 13.0 1.6 39.6 10.8 34.9
4 10.2 2.5 38.6 9.2 39.5
Fig. 4 displays TEM characterizations of the sintered bulk sample at 750 °C. The average grain size of
367 nm is determined from Fig. 4a, which is larger than the grain diameter calculated by the XRD result.
This is because XRD measures coherent diffraction domains that include dislocation cells and subgrains [38,
39]. The selected-area electron diffraction (SAED) pattern (inset in Fig. 4a) illustrates that the ultrafine
grains have random crystallographic orientations. The precipitates also can be clearly observed either in
bright field (Fig. 4b) or dark field (Fig. 4c) images. The lamellar structure in Fig. 4c is confirmed to
correspond to the B2 phase, while the elliptical structure corresponds to the ordered Fe3Al phase, with their
diffraction patters given in the inset of Fig. 4c. The distribution of precipitates is quite uniform with an
average grain size of 47 nm (Fig. 4d). Since no needle-shape or disk-shape precipitates can be observed, the
8
precipitates are considered to have a spherical shape. Several larger precipitates appear in a lamellar
microstructure, which would be the NiAl phase. Fig. 4e gives an example of the EDX spectra of the
precipitates. Four spherical precipitates in the NAB specimen are examined, and composition results are
shown in Table 2. The composition of precipitates shows the Fe/Al atomic ratio of ~3, matching the XRD
results of Fe3Al phase. The high Cu concentrations arise from overlapping of the precipitate and the matrix.
TEM bright field images of NAB-750 also reveal twins in grains with the thickness of 45-67 nm, as
displayed in Fig. 5a. Besides, numerous twins are observed in the alloy (Fig. 5b). The occurrence of twins in
the NAB alloys would arise from the formation of precipitates with a local stress during the gas-atomization
process. Meanwhile, twins can be generated during the SPS process due to the compression condition. To
determine the mean value of the twin thickness, more than 60 twins are measured in order to guarantee a
good statistics. The distribution of the twin thickness based on the observation is summarized in Fig. 5c,
demonstrating that 83.4% of the twin thickness varies from 44 nm to 74 nm. The twin volume fraction (Fv)
Fv = 2e / (t + 2e) (3)
where e and t are the average twin thickness (Fig. 5c) and mean twin spacing (Fig. 5d), respectively. For the
mean twin spacing test, more than 30 areas are considered. However, the twins are not observed to be
present in all grains and their total volume fraction needs to be calculated by the consideration of the
concentration (26%) of grains distributed by twins. Therefore, the twin volume fraction is calculated to be
11.7%. This result of the twin volume fraction is consistence with the previous work of Cu-based alloys [38].
It has been proved that at the macro-scale, increasing the fraction of special boundaries could enhance the
Fig. 6 displays further TEM observations of nanostructures around the twins, indicating that dislocations
are captured between TBs. The high magnification image displays the inhibition of dislocation movement at
9
the TB, as denoted by arrows in Fig. 6b. Previous studies also reported that TBs could act as dislocation
barriers and promot to a high fracture toughness and strengths due to the inhibition of the dislocation
movement [22, 41]. Some curved dislocations can be observed, suggesting that they could be introduced by
a dislocation multiplication mechanism. As discussed by Wen et al. [38] the pressure applied during the SPS
process would generate some dislocations. Meanwhile, a high density of precipitates in the sintered bulk
sample and a significant fraction of twins in the grain interiors further reduce the recovery rate.
Fig. 7a displays representative compressive stress-strain curves of sintered NAB alloys at 600 and
750 °C. The yield strengths of NAB-600 and NAB-750 are 540 ± 10 and 620 ± 10 MPa, respectively, and
the ultimate strengths are 1700 ± 30 and 1480 ± 30 MPa at strain of ~44%. Specifically, the values of the
compressive yield strength of NAB-600 and NAB-750 are close to those of the tensile yield strength (from
the tensile stress-strain curves given in Fig. 7b), which are 545 ± 10 and 630 ± 10 MPa, respectively. Table 3
gives the comparison of yield and ultimate strengths between NAB alloys in this work and in previously
reported works where the NAB alloys were fabricated by other methods [14, 42-45] such as casting,
annealing and normalizing. Our SPS-sintered bulk materials exhibit a high ultimate strength value for NAB
alloys, and show one of the highest reported yield strength values.
Table 3. Comparison of strengths of NAB-600 in this work and the previous reported NAB alloys.
Alloy Processing condition Yield strength (MPa) Ultimate strength (MPa) Ref.
NAB-750 Gas-atomization and SPS 620-630 1480 This work
NAB Casting 379 540 [42]
NAB - 295 800 [43]
NAB Casting 259 690 [14]
NAB Annealing 270 600 [14]
NAB Normalizing 432 760 [14]
NAB Casting 312 722 [45]
NAB Annealing 304 695 [45]
NAB Normalizing 439 791 [45]
NAB Quenching and aging 921 949 [45]
10
NAB Quenching and aging 561 769 [45]
NAB Quenching and aging 521 758 [45]
NAB Casting 284 691 [44]
NAB Annealing 280 673 [44]
Since friction and wear have traditionally been linked to mechanical properties and the composition,
tribological behaviors of NAB samples sintered at 600, 700 and 750 °C are compared in this subsection. Fig.
8a-c show evolutions of friction coefficients of NAB specimens as a function of fretting cycles tested at
contact loads of 43, 86 and 106 N. Friction coefficients evolve with the first 103 fretting cycles and finally
arrive at relatively stable values. It is essential to note that for any given contact load, no obvious difference
can be observed in the friction coefficients of NAB alloys, implying that the phase transformation of
precipitates from Fe3Al to Fe3Al + NiAl has a minimal effect on the friction behavior of bulk materials. As
the contact load increases, the friction coefficients of all samples slightly decrease. This can be explained by
the fact that the most of particles lost their cutting edges, accompanied by the particle damage during the
abrasive process [46, 47]. With different mechanical properties, the NAB alloys sintered at different
temperatures are expected to exhibit varying damage resistance. Many works have focused on investigating
the effect of mechanical properties on the abrasion resistance of materials and correlating their tribological
performances with their composition-related mechanical properties [48, 49]. Indeed, compared with NAB
alloys obtained at 650 and 700 °C, a relative higher damage resistance of the NAB alloy obtained at 750 °C
can be observed for all test conditions, as illustrated from the 3D morphologies of the wear tracks (Fig. 9).
Fig. 10a presents quantitative measurements of wear in NAB specimens. Values of the wear volume are
substantially higher for the NAB-600 compared to those of the NAB-750. This suggests that the precipitates
have a strong impact on the wear behavior of the NAB alloys, despite the minimal effect on the friction
performance. At 106 N, wear tracks of the NAB alloys sintered at 600 and 750 °C with different precipitates
are examined by SEM, as shown in Fig. 10b and c. In these images, the plastic deformation and abrasion
11
wear showing surface plowing by asperities, as well as cracked transfer layers can be seen, and some
patches of the compressed transfer layer remain on wear tracks (identified by white arrows). Specifically, the
crack propagation can be observed on the high-magnification image of the wear track of the NAB-600 (the
inset in Fig. 10b). Although such a crack could extend from its original length during the specimen
preparation, inter- or trans-granular cracks are not detected on the magnification image of the wear track of
the NAB-750 (the inset in Fig. 10c). Therefore, it can be assumed that NAB-750 with Fe3Al + NiAl
precipitates has a higher wear resistance than NAB alloys with a single Fe3Al-precipitate phase. This can be
further confirmed by the dissipated energy of NAB alloys during fretting, which can be defined as
Ed = μ‧ P‧ d (Joules) (4)
where μ is the average friction coefficient, P is the applied load and d is the total fretting distance. Fig. 10d
presents the wear volume as a function of the dissipated energy for NAB alloys. The wear volume and the
dissipated energy of all NAB specimens fit with linear relationships. The slope of NAB-600 is relatively
higher than that of NAB-600, indicating a higher wear rate of the NAB-600 (181397 μm3/J) compared with
those of the NAB-700 (157680 μm3/J) and the NAB-750 (150346 μm3/J).
4. Discussion
Despite the application of different sintering temperatures during SPS processes, the final average grain
size of all bulk materials is maintained at ~ 400 nm (Fig. 3 and Fig. 4a) due to high cooling and fast
solidification rates. This work demonstrates that the gas-atomized and spark plasma sintered NAB alloys
with ultra-fine grains can exhibit extraordinary compressive ultimate strengths of 1480-1700 MPa at room
temperature with a plastic strain of > 40% (Fig. 7). NAB-750 shows the excellent yield strength of 620-630
MPa. Although these high ultimate strengths were obtained by the compression mode, the results would
12
boost more research interests and efforts into the methodology for the development of ultrahigh strength of
Cu-based alloys using gas-atomization and spark plasma sintering. This is inferred from results of the
comparison with strengths of many conventional high-strength NAB alloys, in which various heat treatment
methods were used (Table 3). The present bulk materials exhibit better strengths (yield strength of 540 MPa,
ultimate strength of 1480 MPa) even at the sintering temperature of 600 °C compared with the cast
counterparts.
Both the strength and the wear resistance of the sintered bulk materials increase with increasing the
sintering temperature. Besides the fine grain size of sintered alloys, microstructure characterizations indicate
that the volume fraction of the B2-type NiAl precipitate increases with the sintering temperature. With the
Fe3Al precipitate, the NiAl phase uniformly distributes in the NAB alloy to form intragranular and
intergranular precipitates. The increase in the strength and the wear resistance of SPS-consolidated NAB can
be attributed to the grain refinement and the fine dispersion of the precipitation in bulk materials, since
varying degrees of segregation are commonly observed in the cast alloys after heat treatment [50].
Compared with the Fe3Al precipitate, the NiAl phase with higher elastic modulus of 231 GPa and yield
strength of 400 MPa further promotes the wear resistance of NAB bulk materials. Consequently, NAB-750
Furthermore, a high dislocation density in the sintered bulk samples is observed. This could be caused
by the high density of precipitation existing in the grain interiors, which reduces the recovery rate.
Nanoscale twins are retained with average twin thickness of 50 nm, which all facilitate the high strength and
the wear resistance. Thus, based on microstructural observation as well as mechanical and tribological test
results, five contributions to the high strength and the excellent wear resistance of the sintered bulk samples
13
4.2. Strengthening mechanisms for sintered bulk materials
The GB strengthening can be estimated through the Hall-Petch law [51]. A decrease of the grain size
could induce an increase of the total grain boundary length, hindering the motion of dislocations. For
precipitate strengthening, since the precipitates have an average particle size of 47 nm, the Orowan by-pass
mechanism is expected to be operative for the dislocation motion rather than the particle shearing
mechanism. In addition, the microstructural analysis shows a high dislocation density particularly at the TBs.
Therefore, the dislocation strengthening relation should be considered, which relates the critical stress for an
onset of slip to the square root of the dislocation density. Furthermore, coherent twins observed in ultrafine
grains contribute to the high strength similar to that of the GB via Hall-Petch type strengthening. Finally, the
solute atoms in the Cu solid solution contribute strengthening to the metal matrix through the generation of
an elastic strain caused by the atom misfit, and the mismatch of elastic moduli between solute and solvent
atoms.
4.2.1. GB strengthening
It is demonstrated that GB strengthening follows the Hall-Petch relationship, which relates the yield
where KHP is the Hall-Petch coefficient, σ0 constitutes the overall resistance of the crystal lattice to the
dislocation motion, and d is the average grain diameter (325 nm). KHP and σ0 are assumed to be 0.17
MPa‧ m1/2 and 36.2 MPa, respectively, for polycrystalline Cu with d > 2·10-7 m [52]. Therefore, the total
14
4.2.2. Precipitate strengthening
𝑟̅ = √2/3𝑟 (7)
where M is the Taylor factor of the fcc polycrystalline matrix (3.06) [38], G and b are the shear modulus
(42.1 GPa) and the magnitude of the Burger vector (0.256 nm) [53] of the matrix, respectively. υ represents
the Poisson’s ratio (0.34) of the matrix [52]. r is the mean precipitate radius of 23.5 nm, and λ is the effective
where f is the volume fraction of precipitates (5%). Finally, the contribution from the precipitates
where α is a constant (taken as 0.2 for fcc metals), ρ is the dislocation density of 1.3 ± 0.1 × 1014 m-2. M, G
and b have the same significance as above. Hence, the calculated Δσd is 75 MPa.
4.2.4. TB strengthening
where KTB is a constant and can be approximated to the Hall-Petch constant of the corresponding material. Vf
is the total volume fraction of twins. λTB is the average TB thickness, which is 50 nm from the TEM
15
observation (Fig. 5c). Previous works by Wen et al. [38] demonstrated that the volume fraction of twins is
~10% in the SPS-consolidated Cu-based alloy, which is similar with the result in this work (11.7%). With
This contribution solid solution strengthening (Δσss) also has a significant effect on the yield strength.
For conventional binary solid solutions, well-established models are valid for the related strengthening
mechanisms governed by elastic interactions between dislocations and the solute atom-induced stress
[54-56]. As NAB is the Cu-based alloy with a complex composition, Senkov et al. [57] proposed an equation
where A is a dimensionless parameter of 0.1. G and C are the shear modulus and concentration of solute
atoms, respectively. ε is the lattice strain caused by solute atoms. In an AlCoCrCuFeNi system, Ganji et al.
[58] suggested that Al might cause more strain to the copper based lattice with a strain of 12%. For the
current composition (AlCuFeNiMn), it is approximated that the strain of 12% caused by Al atom could be
taken as the effective strain. Therefore, the concentration of Al solute atoms is 0.09 based on the analysis of
the fraction of Al in the matrix and precipitates. Hence, the strength increment of Δσss is 50 MPa.
With the above-mentioned calculation, the contribution proportions from ΔσHP, ΔσOr, Δσd, ΔσTB and Δσss,
are 54%, 8.7%, 12%, 14.4% and 8%. Meanwhile, a proportion from the slightly underestimated yield
strength (~3%) can be obtained, which may result from dislocation-twin interactions observed in Fig. 6. The
interactions between dislocation strengthening and TB strengthening mechanisms could create more
pronounced effects [43, 59, 60]. As noted, a fraction of dislocations is observed to present and concentrate at
TBs. Recent study demonstrates that the dislocation interaction with and accumulation at twin boundaries
significantly improve the strength of fcc metals and alloys [61] because dislocations are blocked by the TBs
16
due to the discontinuity in slip systems of two crystals. Thus, in order to slip from one grain to the twin,
cross slip or dislocation nucleation would occur at the TB. However, the sessile dislocations could serve as
barriers to the motion of other dislocations and prevent dislocation nucleation in the system. This is
consistent with TEM observations in which the NAB-750 sample contains a high-density of dislocations at
the TBs (Fig. 5). Molecular dynamics (MD) simulation works by Wu et al. [62] showed that the sessile and
glissile dislocations together with the TBs could form complex dislocation barriers and networks, as well as
point defects during the plastic deformation, contributing to the high strength.
The XRD, SEM and TEM characterizations indicate that NAB alloys sintered at different temperatures
exhibit significant differences in the precipitation, which consequently has a strong effect on the tribological
properties of alloys as determined from the flat-on-flat fretting wear tests. NAB alloys sintered at 750 °C
consist of ultra-fine grains (the matrix) and a large number of Fe3Al + NiAl precipitates existing at the GBs
(Fig. 3e and f). Compared with NAB-750, the NAB alloys prepared at 600 °C contain the matrix and the
Fe3Al precipitate. Both NAB-600 and NAB-750 show friction coefficients between 0.4 and 0.5 at contact
loads of 43-106 N, which possibly benefit from the intrinsic friction behavior of the copper matrix [63-65].
However, obviously different wear behaviors of the two NAB alloys are evident from the fretting wear tests
and SEM observations, in which cracks can be identified on the wear tracks of NAB-600. This can be
concluded that the precipitation exhibits more significant effect on the wear performance than on the friction
behavior. It is rather intuitive that the relative higher wear rate of NAB-600 (181397 μm3/J) compared with
that of the NAB-750 (150346 μm3/J) is related to the crack initiation and propagation on the wear tracks of
NAB-600 during fretting (Fig. 10b and c). This difference of the wear resistance of NAB alloys sintered at
600 and 750 °C motivates the discussion of crack evolution mechanisms during wear in the following
17
sub-sections.
As mentioned previously, NAB alloys contain different precipitates with Fe3Al in NAB-600 and with
Fe3Al + NiAl in NAB-750. The elastic modulus for Fe3Al and NiAl is 141 GPa [66] and 231 GPa [67],
respectively; while the modulus of copper is 129.8 GPa [68]. We first explore the influence of the elastic
modulus mismatch, which may have a significant effect on crack-tip stress fields. Two cases that are the
moderate (𝐸𝐹𝑒3 𝐴𝑙 /ECu) and the high (ENiAl/ECu) plastic mismatches are addressed. During the wear process,
the high shear stress caused by asperities could locally exceed the yield stress of NAB alloys, leading to the
crack initiation on the surface. As fretting proceeds, the crack-tip stress field would be locally continuous
across a transformation boundary. Therefore, the stress continuity across interfaces can be maintained only
by the locally relieving strain [69] both in the matrix and precipitates. When the elastic modulus mismatch is
considered, Robertson et al. [69] observed strain reversals near the crack tip, effectively providing a strain
accommodation mechanism with the reduction of the local stress. MD simulations demonstrated that traction
distributions near the crack tips were noticeably lower for the high elastic mismatch case than for the
moderate one, whereas the total displacements were about the same; meanwhile, at a given displacement, the
It can be noted that the crack propagation observed in this work is also associated with the moderate
mismatch during the wear process. In addition, SEM observations indicate the complex nature of the
structural evolution including the severe plastic deformation in the NAB system during fretting. This result
strongly suggests that the elastic modulus mismatch could not be considered as the only cause of the crack
propagation at the surface during wear, since deformations of the matrix and precipitates will both occur.
Thus, the plastic mismatch between the matrix and precipitates is also discussed in the next sub-section.
18
4.3.2. Plastic mismatch between the matrix and precipitates
The yield strength of Fe3Al and NiAl are 330 MPa [71] and 400 MPa [14], respectively. Therefore,
compared with the yield strength of copper (250 MPa [14]), the moderate (𝜎𝐹𝑒3 𝐴𝑙 /σCu) and the high (σNiAl/σCu)
plastic mismatch cases will be considered. Since the volume fraction of precipitates is 5%, a microcrack
caused by the fretting wear tends to occur near the interface of the matrix and the precipitate, and the stress
intensity factor could decrease at a given applied load due to the build-up of plastic strains [72]. This
corresponds to an increase in effective toughness, which is strongly dependent on the distance of the crack
from the interface. Moreover, as the crack is very close to the interface, a decrease of crack-tip stresses
would occur with the crack extension at the high plastic mismatch. This was demonstrated by
Charalambides et al. [73]. For the case of the high plastic mismatch, the precipitate phase (NiAl) can be
considered as a more rigid material which would lead to a reduction of crack-tip stresses with the fewer
anti-shielding effect, resulting in higher failure loading. Therefore, more plastic deformations would occur
which increase the effective toughness in subsequent crack extension increments. The higher stress is then
required for the further crack propagation. Conversely, for the moderate plastic mismatch, a less rigid
precipitate (Fe3Al) could amplify crack-tip stresses, leading to a high anti-shielding effect and low failure
loads. It has been indicated that such high anti-shielding effect could cause the less plastic deformation, and
there would be less extension toughening [72]. The failure load in crack extension increments is then limited,
Consequently, it can be concluded that the higher plastic mismatch (σNiAl/σCu), which causes the crack
extension toughening near the interface of the matrix and the precipitate, may also have a strong effect on
19
5. Conclusions
Bulk nickel-aluminium bronze alloys with the excellent wear resistance and superior strengths were
developed via gas-atomization and SPS. Detailed comparative research was carried out to determine the
Microstructural characterizations of NAB-750 showed that its high strength was attributed to the uniformly
distributed precipitates (Fe3Al and NiAl) both in ultrafine-grain interiors and at GBs, the high dislocation
density, the TBs inside grain interiors and the solid solution. The wear resistance could be related to the
elastic modulus mismatch and the plastic mismatch between the matrix and precipitates. The following
Fine intragranular and intergranular precipitates with a volume fraction of 5% are formed after the
sintering process. The NiAl content is increased by increasing the sintering condition to 750 °C, while
the Fe3Al content is decreased. Their total fraction and size remain rather constant for each sintering
temperature.
At 750 °C, the sintered bulk materials show an average grain size of 367 nm with a high density of
twins (11.7 vol%), a dislocation density of 1.3 ± 0.1 × 1014 m-2 and an Al-solute concentration of 0.09.
predominant mechanism for the sintered bulk materials with contributions from precipitate
strengthening (54 MPa), dislocation strengthening (75 MPa), TB strengthening (89 MPa) and solid
solution strengthening (50 MPa). In addition, high-density dislocations are blocked at TBs to form
complex dislocation barriers and networks, further contributing to the high strength of NAB alloys.
NAB alloys exhibit excellent wear properties, with an increase in wear resistance with increasing the
sintering temperature. The excellent wear resistance can be ascribed to the reduction of the local stress
around the crack tip due to the high elastic modulus mismatch (ENiAl/ECu), and the crack extension
20
toughening near the interface of the matrix and the precipitate caused by the high plastic mismatch
(σNiAl/σCu).
Acknowledgments: The article was supported by the National Basic Research Program of China [grant
No. 2014CB046705]; the China Postdoctoral Science Foundation [Grant No. 2016M600586]; and the
References
[1] L. Lu, Y. Shen, X. Chen, L. Qian, K. Lu, Ultrahigh strength and high electrical conductivity in copper,
[2] L. Lu, X. Chen, X. Huang, K. Lu, Revealing the maximum strength in nanotwinned copper, Science 323
(2009) 607-610.
[3] K. Lu, L. Lu, S. Suresh, Strengthening materials by engineering coherent internal boundaries at the
[4] S.T. Chen, W.Y. Tang, Y.F. Kuo, S.Y. Chen, C.H. Tsau, T.T. Shun, J.W. Yeh, Microstructure and properties
of age-hardenable Al x CrFe 1.5 MnNi 0.5 alloys, Materials Science & Engineering A 527 (2010)
5818-5825.
[5] M.H. Chuang, M.H. Tsai, W.R. Wang, S.J. Lin, J.W. Yeh, Microstructure and wear behavior of AlxCo1.5
[6] C.W. Tsai, M.H. Tsai, J.W. Yeh, C.C. Yang, Effect of temperature on mechanical properties of
Al0.5CoCrCuFeNi wrought alloy, Journal of Alloys & Compounds 490 (2010) 160-165.
[7] S.p. Fonlupt, B. Bayle, D. Delafosse, J.L. Heuze, Role of second phases in the stress corrosion cracking of
21
[8] J.A. Wharton, R.C. Barik, G. Kear, R.J.K. Wood, K.R. Stokes, F.C. Walsh, The corrosion of
[9] J.P. Tu, W.X. Qi, Y.Z. Yang, F. Liu, J.T. Zhang, G.Y. Gan, N.Y. Wang, X.B. Zhang, M.S. Liu, Effect of
aging treatment on the electrical sliding wear behavior of Cu-Cr-Zr alloy Wear, 249 (2001) 1021-1027.
[10] S.A. Kouhanjani, A. Zare-Bidaki, M. Abedini, N. Parvin, Influence of prior cold working on the
tribological behavior of Cu-0.65 wt.%Cr alloy, Journal of Alloys & Compounds 480 (2009) 505-509.
[11] G.C. Weatherly, P. Humble, D. Borland, Precipitation in a Cu-0,55 wt.% Cr alloy, Acta Metallurgica 27
(1979) 1815-1828.
[12] G. Purcek, H. Yanar, O. Saray, I. Karaman, H.J. Maier, Effect of precipitation on mechanical and wear
[13] Y. Lv, Y. Ding, Y. Han, L.C. Zhang, L. Wang, W. Lu, Strengthening mechanism of friction stir processed
and post heat treated NiAl bronze alloy: Effect of rotation rates, Materials Science & Engineering A
(2016).
[14] X. Xu, Y. Lv, M. Hu, D. Xiong, L. Zhang, L. Wang, W. Lu, Influence of second phases on fatigue crack
growth behavior of nickel aluminum bronze, International Journal of Fatigue 82 (2015) 579-587.
[15] F. Hasan, A. Jahanafrooz, G.W. Lorimer, N. Ridley, The morphology, crystallography, and chemistry of
1337-1345.
[16] J.A. Wharton, K.R. Stokes, The influence of nickel–aluminium bronze microstructure and crevice
seawater&mdash-Effect of galvanic coupling to UNS S31603 Corrosion Science 121 (2017) 43-56.
[18] D.R. Ni, P. Xue, Z.Y. Ma, Effect of Multiple-Pass Friction Stir Processing Overlapping on Microstructure
22
and Mechanical Properties of As-Cast NiAl Bronze, Metallurgical and Materials Transactions A 42 (2011)
2125-2135.
[19] S. Neodo, D. Carugo, J.A. Wharton, K.R. Stokes, Electrochemical behaviour of nickel-aluminium bronze
38-46.
[20] M. Mondet, E. Barraud, S. Lemonnier, J. Guyon, N. Allain, T. Grosdidier, Microstructure and mechanical
properties of AZ91 magnesium alloy developed by Spark Plasma Sintering, Acta Materialia 119 (2016)
55-67.
[21] H.Y. Zhang, Y.H. Lu, M. Ma, J. Li, Effect of precipitated carbides on the fretting wear behavior of Inconel
[22] H.S. Kim, P. Dharmaiah, B. Madavali, R. Ott, K.H. Lee, S.J. Hong, Large-scale production of (GeTe)x
(AgSbTe 2 )100-x (x=75, 80, 85, 90) with enhanced thermoelectric properties via gas-atomization and
[23] W. Zhai, X. Shi, Z. Xu, A. Zhang, Investigation of the friction layer of Ni3Al matrix composites, Wear
[24] K.C. Park, P. Dharmaiah, H.S. Kim, S.J. Hong, Investigation of microstructure and thermoelectric
properties at different positions of large diameter pellets of Bi0.5Sb1.5Te3 compound, Journal of Alloys &
[25] M. Omori, Sintering, consolidation, reaction and crystal growth by the spark plasma system (SPS),
[26] A.S. Rogachev, K.V. Kuskov, N.F. Shkodich, D.O. Moskovskikh, A.O. Orlov, A.A. Usenko, A.V. Karpov,
I.D. Kovalev, A.S. Mukasyan, Influence of high-energy ball milling on electrical resistance of Cu and
Cu/Cr nanocomposite materials produced by Spark Plasma Sintering, Journal of Alloys & Compounds
23
688 (2016) 468-474.
[27] J. Galy, M. Dolle, T. Hungria, P. Rozier, J.P. Monchoux, A new way to make solid state chemistry: Spark
plasma synthesis of copper or silver vanadium oxide bronzes, Solid State Sciences 10 (2008) 976-981.
[28] K.N. Zhu, A. Godfrey, N. Hansen, X.D. Zhang, Microstructure and mechanical strength of near- and
sub-micrometre grain size copper prepared by spark plasma sintering, Materials & Design 117 (2017)
95-103.
[29] K. Han, I. Ohnuma, R. Kainuma, Experimental determination of phase equilibria of Al-rich portion in the
Al-Fe binary system, Journal of Alloys & Compounds 668 (2016) 97-106.
[30] N.S. Kanhe, A.K. Tak, A.B. Nawale, S.A. Raut, S.V. Bhoraskar, A.K. Das, V.L. Mathe, Understanding
plasma assisted gas phase condensation method, Materials & Design 112 (2016) 495-504.
[31] W. Lu, P. Zhang, X. Liu, W. Zhai, M. Zhou, J. Luo, W. Zeng, X. Jiang, Influence of surface topography on
torsional fretting wear under flat-on-flat contact, Tribology International 109 (2017) 367-372.
[32] M. Godjevac, Wear and Friction in a Controllable Pitch Propeller, Mechanical Maritime & Materials
Engineering (2010).
[33] H.Y. Yasuda, T. Nakajima, K. Nakano, K. Yamaoka, M. Ueda, Y. Umakoshi, Effect of Al concentration
[34] Z.B. Jiao, J.H. Luan, M.K. Miller, C.Y. Yu, C.T. Liu, Effects of Mn partitioning on nanoscale
precipitation and mechanical properties of ferritic steels strengthened by NiAl nanoparticles, Acta
[35] L. Anthony, B. Fultz, Effects of early transition metal solutes on the D03-B2 critical temperature of Fe3Al,
[36] G.K. Williamson, W.H. Hall, X-ray line broadening from filed aluminium and wolfram, Acta
24
Metallurgica 1 (1953) 22-31.
[37] G.K. Williamson, R.E. Smallman, III. Dislocation densities in some annealed and cold-worked metals
from measurements on the X-ray debye-scherrer spectrum, Philosophical Magazine 1 (1956) 34-46.
[38] H. Wen, T.D. Topping, D. Isheim, D.N. Seidman, E.J. Lavernia, Strengthening mechanisms in a
high-strength bulk nanostructured Cu-Zn-Al alloy processed via cryomilling and spark plasma sintering,
[39] L. Balogh, T. Ungár, Y. Zhao, Y.T. Zhu, Z. Horita, C. Xu, T.G. Langdon, Influence of stacking-fault
[40] R.L. Fullman, Measurement of particle sizes in opaque bodies, Trans. AIME 197(1954) 447-452.
[41] P. Zhou, Z.Y. Liang, R.D. Liu, M.X. Huang, Evolution of dislocations and twins in a strong and ductile
[42] C. Wang, C. Jiang, Y. Zhao, M. Chen, V. Ji, Surface mechanical property and residual stress of peened
nickel-aluminum bronze determined by in-situ X-ray diffraction, Applied Surface Science 420 (2017)
28-33.
[43] F.U. Zhongtao, W. Yang, S. Zeng, B. Guo, H. Shubing, Identification of constitutive model parameters for
nickel aluminum bronze in machining, Transactions of Nonferrous Metals Society of China 26 (2016)
1105-1111.
microstructure evolution of as-cast Nickel Aluminum Bronze alloy, Materials & Design 60 (2014)
233-243.
[45] Z. Wu, Y.F. Cheng, L. Liu, W. Lv, W. Hu, Effect of heat treatment on microstructure evolution and
25
(2015) 260-270.
[46] A. Khellouki, J. Rech, H. Zahouani, Energetic analysis of cutting mechanisms in belt finishing of hard
[47] A. Jourani, B.Hagège, S. Bouvier, M. Bigerelle, H. Zahouani, Influence of abrasive grain geometry on
friction coefficient and wear rate in belt finishing, Tribology International 59 (2013) 30-37.
[48] X. Xu, F.H. Ederveen, S.V.D. Zwaag, W. Xu, Correlating the abrasion resistance of low alloy steels to the
standard mechanical properties: A statistical analysis over a larger data set, Wear 368-369 (2016) 92-100.
properties of substrate and coating on wear performance of TiN- or DLC-coated 316LVM stainless steel,
[50] D. Samanta, N. Zabaras, Control of macrosegregation during the solidification of alloys using magnetic
[51] G.E. Dieter, Mechanical Metallurgy (SI Metric Adaptation), McGraw-Hill Higher Education (2011).
[52] Y. Estrin, N.V. Isaev, S.V. Lubenets, S.V. Malykhin, A.T. Pugachov, V.V. Pustovalov, E.N. Reshetnyak,
V.S. Fomenko, L.S. Fomenko, S.E. Shumilin, Effect of microstructure on plastic deformation of Cu at
[53] Y.J. Li, X.H. Zeng, W. Blum, Transition from strengthening to softening by grain boundaries in
[54] R. Labusch, A Statistical Theory of Solid Solution Hardening, Physica Status Solidi, 41 (2010) 659-669.
[56] L. Lu, R. Schwaiger, Z.W. Shan, M. Dao, K. Lu, S. Suresh, Nano-sized twins induce high rate sensitivity
26
[57] O.N. Senkov, J.M. Scott, S.V. Senkova, D.B. Miracle, C.F. Woodward, Microstructure and room
temperature properties of a high-entropy TaNbHfZrTi alloy, Journal of Alloys & Compounds 509 (2011)
6043-6048.
[58] R.S. Ganji, P.S. Karthik, K.B.S. Rao, K.V. Rajulapati, Strengthening mechanisms in equiatomic ultrafine
grained AlCoCrCuFeNi high-entropy alloy studied by micro- and nanoindentation methods, Acta
[60] G. Csiszar, L. Balogh, A. Misra, X. Zhang, T. Ungar, The dislocation density and twin-boundary
frequency determined by X-ray peak profile analysis in cold rolled magnetron-sputter deposited
[61] Y.T. Zhu, X.L. Wu, X.Z. Liao, J. Narayan, L.J.Kecskés, S.N. Mathaudhu, Dislocation-twin interactions in
[62] Z.X. Wu, Y.W. Zhang, D.J. Srolovitz, Dislocation-twin interaction mechanisms for ultrahigh strength and
[63] H.M. Cao, X. Zhou, X.Y. Li, K. Lu, Friction mechanism in the running-in stage of copper: From plastic
[64] P. Zhang, W.L. Lu, X.J. Liu, M.Z. Zhou, W.Z. Zhai, G.P. Zhang, X.Q. Jiang, Torsional fretting wear
behavior of CuNiAl against 42CrMo4 under flat on flat contact, Wear 380 (2017) 6-14.
[65] P. Zhang, W.L. Lu, X.J. Liu, W.Z. Zhai , M.Z. Zhou, X.Q. Jiang, A comparative study on torsional fretting
and torsional sliding wear of CuNiAl under different lubricated conditions, Tribology International 117
(2017) 78-86.
[66] C.T. Liu, K.S. Kumar, Ordered intermetallic alloys, part I: Nickel and iron aluminides, JOM 45 (1993)
27
38-44.
[67] R.D. Noebe, R.R. Bowman, M.V. Nathal, Physical and mechanical properties of the B2 compound NiAl,
[68] R.W. Hertzberg, Deformation and fracture mechanics of engineering materials, J. Wiley 1996.
[69] S.W. Robertson, A. Mehta, A.R. Pelton, R.O. Ritchie, Evolution of crack-tip transformation zones in
superelastic Nitinol subjected to in situ fatigue: A fracture mechanics and synchrotron X-ray
[70] X.W. Zhou, N.R. Moody, R.E. Jones, J.A. Zimmerman, E.D. Reedy, Molecular-dynamics-based cohesive
zone law for brittle interfacial fracture under mixed loading conditions: Effects of elastic constant
[71] G. Hasemann, J.H. Schneibel, E.P. George, Dependence of the yield stress of Fe3Al on heat treatment,
[72] M.T. Tilbrook, I.E. Reimanis, K. Rozenburg, M. Hoffman, Effects of plastic yielding on crack
[73] P.G. Charalambides, P.A. Mataga, R.M. Mcmeeking, A.G. Evans, Steady-State Mechanics of a Growing
Crack Paralleling an Elastically Constrained Thin Ductile Layer, Applied Mechanics Reviews 43 (1990).
28
Figure captions
Fig. 1. SEM image of raw powders, the inset shows the particle size between 5 and 25 μm.
Fig. 2. Evolution of XRD patterns of NAB samples sintered at 600, 650, 700 and 750 °C, the inset shows the
peak splitting of the Fe3Al peak at different sintering temperatures.
Fig. 3. SEM images at two magnifications showing the microstructures for the bulk materials sintered at (a
and b) 600, (c and d) 700 and (e and f) 750 °C
Fig. 4. TEM characterizations of the sintered bulk sample at 750 °C: (a) an average grain size of 367; (b and
c) both bright and dark field images indicate the presence of precipitates; (d) size distribution of precipitates;
(e) an example of the EDX spectra of the precipitates; the inset in the image a indicates the ultrafine grains
with random crystallographic orientations; the inset in b shows the SAED pattern of the matrix; the inset in c
gives the SAED patterns of the matrix, B2 phase NiAl and D03 phase Fe3Al in [1̅11]
Fig. 5. (a) TEM bright field image of NAB-750 reveals twins in grains; (b) numerous twins observed in the
NAB-750; (c) distribution of the twin thickness; (d) distribution of twin spacing.
Fig. 6. (a) TEM image showing dislocations between TBs; (b) a high magnification image displaying the
inhibition of dislocation movement at the TB.
Fig. 7. Representative (a) compressive and (b) tensile stress-strain curves of NAB-600 and NAB-750.
Fig. 8. Evolutions of friction coefficients of NAB-600, NAB-700 and NAB-750 as a function of fretting
cycles tested at contact loads of (a) 43, (b) 86 and (c) 106 N.
Fig. 9. 3D morphologies of the wear tracks of NAB-600, NAB-700 and NAB-750 after wear tests at 43, 86
and 106 N.
Fig. 10. (a) Wear volumes of NAB-600, NAB-700 and NAB-750 tested at 43, 86 and 106 N; (b and c) wear
tracks of NAB-600 and NAB-750 at 106 N; (d) the wear volume as a function of the dissipated energy for
NAB alloys.
Fig. 11. Strengthening mechanisms based on the microstructural observations including GB strengthening,
precipitate strengthening, dislocation strengthening, TB strengthening and solid solution strengthening.
29
Fig. 1. SEM iamge of raw powders, the inset shows the particle size between 5 and 25 μm.
30
Fig. 2. Evolution of XRD patterns of NAB samples sintered at 600, 650, 700 and 750 °C, the inset shows the
peak splitting of the Fe3Al peak at different sintering temperatures.
31
Fig. 3. SEM images at two magnifications showing the microstructures for the bulk materials sintered at (a
and b) 600, (c and d) 700 and (e and f) 750 °C
32
Fig. 4. TEM characterizations of the sintered bulk sample at 750 °C: (a) an average grain size of 367; (b and
c) both bright and dark field images indicate the presence of precipitates; (d) size distribution of precipitates;
(e) an example of the EDX spectra of the precipitates; the inset in the image a indicates the ultrafine grains
with random crystallographic orientations; the inset in b shows the SAED pattern of the matrix; the inset in c
gives the SAED patterns of the matrix, B2 phase NiAl and D03 phase Fe3Al in [1̅11]
33
Fig. 5. (a) TEM bright field image of NAB-750 reveals twins in grains; (b) numerous twins observed in the
NAB-750; (c) distribution of the twin thickness; (d) distribution of twin spacing.
34
Fig. 6. (a) TEM image showing dislocations between TBs; (b) a high magnification image displaying the
inhibition of dislocation movement at the TB.
35
Fig. 7. Representative (a) compressive and (b) tensile stress-strain curves of NAB-600 and NAB-750.
36
Fig. 8. Evolutions of friction coefficients of NAB-600, NAB-700 and NAB-750 as a function of fretting
cycles tested at contact loads of (a) 43, (b) 86 and (c) 106 N.
37
Fig. 9. 3D morphologies of the wear tracks of NAB-600, NAB-700 and NAB-750 after wear tests at 43, 86
and 106 N.
38
Fig. 10. (a) Wear volumes of NAB-600, NAB-700 and NAB-750 tested at 43, 86 and 106 N; (b and c) wear
tracks of NAB-600 and NAB-750 at 106 N; (d) the wear volume as a function of the dissipated energy for
NAB alloys.
39
Fig. 11. Strengthening mechanisms based on the microstructural observations including GB strengthening,
precipitate strengthening, dislocation strengthening, TB strengthening and solid solution strengthening.
40