1. Introduction
Heavy tungsten alloys (HTAs) of the W-Ni-Fe system with the W contents over 85-90% have a number of unique characteristics: simultaneous increased density, strength, and plasticity at room temperature [
1,
2,
3,
4,
5], high characteristics of dynamic strength [
6,
7], high radiation resistance, etc. The industrial tungsten alloys with 90-95%W obtained by conventional liquid-phase sintering (LPS) have the density of 17-19 g/cm
3, the ultimate strength up to 1000 MPa, and elongation to failure up to 25-30% at room temperature. Additional strain strengthening allows improving the ultimate strength up to 1400-1600 MPa preserving satisfactory ductility [
8,
9,
10,
11]. These features make HTAs interesting for various construction applications as well as interesting object for investigations of the effect of the interphase boundaries on the mechanical properties of the alloys [
12,
13,
14,
15].
It is worth noting that the technologies of strain hardening of the HTAs have approached the limits of their capabilities at present. For further improving the properties of the HTA, modern technologies of powder metallurgy are used often [
16,
17] including the additive manufacturing methods [
18,
19,
20] and microwave sintering method [
21]. Modern methods of fabrication of nanopowder compositions W-Ni-Fe are being developed extensively [
5,
18,
22].
The analysis of the literature shows Spark Plasma Sintering (SPS) to be one of promising methods of fabricating the HTAs [
23,
24,
25,
26,
27]. An opportunity to vary main process parameters affecting the microstructure parameters of the HTAs most essentially (the heating rate, the temperature and time of sintering, the magnitude of the pressure applied, etc.) directly in the course of sintering provides a great flexibility to SPS method in controlling the mechanical properties of the HTAs. An opportunity of sintering the materials at very high heating rates (up to 2500
oC/min) is an important feature of SPS technology, which allows reducing the grain growth rate [
23,
27]. The fine-grained HTAs obtained by SPS have simultaneously high strength and hardness [
28,
29,
30,
31,
32,
33,
34,
35,
36,
37,
38,
39,
40,
41,
42,
43,
44] (
Table A1, see
Appendix A).
Joint application of High Energy Ball Milling (HEBM) and SPS is one of promising methods of further improving the mechanical properties of the HTAs [
1,
2,
8,
9,
12,
13,
14,
15,
16,
17]. This allows ensuring an outstanding increase in the strength and hardness of the HTAs (
Table A1, see
Appendix A). It is interesting to note that the results of [
27,
28,
33,
39] evidence the possibility of formation of strongly supersaturated solid solution of tungsten in the γ-phase based on nickel during HEBM that leads to an additional acceleration of sintering of the W-based nanopowders. The formation of the supersaturated solid solution of tungsten atoms in the γ-phase leads to manifestation of some unexpected effects. The effect of non-monotonous dependence of the density on the sintering temperature [
28,
29,
35,
43,
44,
45] as well as the effect of reduction of the activation energy of sintering of the nanopowders [
45] deserve high attention. One should note separately the effect of decreasing of the density of the tungsten alloys sintered from mechanically activated nanopowders [
33,
39,
45], the nature of which remains unclear now. Also, the effect of nonmonotonous (with a maximum) dependencies of the HTA density on the time of preliminary HEBM of the powders observed in [
45] is interesting.
From the practical viewpoint, the result of increase in the hardness, the static strength and the dynamic one of the HTAs obtained by SPS is important.
The analysis of the literature shows that the W-7%Ni-3%Fe alloys obtained by SPS from nanopowders (t
HEBM = 20 min) have a higher dynamic strength as compared to the coarse-grained alloys [
42,
46]. The decreasing of the grain sizes was noted to result in simultaneous increase in the static strength, the dynamic one, and in the penetration depth of a HTA sample into a steel barrier. At the throwing speed of 1194 m/s, the penetration depth reached 15.2 mm whereas for the coarse-grained HTAs this magnitude varied from 9.4 to 12.8 mm at similar testing regimes [
46]. The formation of strongly supersaturated solid solution in the γ-phase may be one of the origins of manifestation of increased dynamic strength of the HTAs. In our opinion, it allows increasing the strength of the interphase (α-γ)-boundaries and, as a consequence, ensuring additional increase in the strength of HTAs.
The investigation of the effect of the HEBM time (tHEBM) of the powders on the density, microstructure parameters, and mechanical properties of heavy tungsten alloy W-7%Ni-3%Fe obtained by SPS was the main goal of the present work.
2. Materials and Methods
The object of investigations was W-7%Ni-3%Fe alloy. The chemical compositions of the initial powders α-W, β-Ni, and α-Fe are presented in
Table 1. The mean particle sizes in the initial powders α-W, β-Ni, and α-Fe according to Fischer were R
0 = 0.8 μm, 20 μm, and 11 μm respectively.
The initial coarse-grained compositions were obtained by mixing the powders α-W, β-Ni, and α-Fe in a FRITSCH® Pulverisette 6 planetary ball mill (Idar-Oberstein, Germany). The mixing time was 15 h, the mixing rate was 100 rpm. HEBM of the compositions W-7%Ni-3%Fe was performed in APF-3 high energy planetary mill (Russia). The acceleration of the milling bodies was 60g, the mixing rate was 1450 rpm. The containers and milling balls of 6-10 mm in diameter were made from industrial alloy W-7%Ni-3%Fe. The mass ratio of the balls and the powders was 10:1. HEBM with the durations of 5, 10, 20, and 40 min was performed in argon ambient with addition of ethanol. To minimize the heating of the powders, HEBM was performed in several stages each 5 min long.
The HTA samples were obtained by conventional sintering in hydrogen (Method I) and by SPS (Method II).
To obtain the HTA samples by Method I, the nanopowders were pressed into the samples of 30 mm in diameter and 5-6 in height in advance. The uniaxial pressing by the pressure of 150 MPa was performed at room temperature in a steel mold. The sintering was performed in a two-stage regime: Stage I – heating with the rate 25 oC/min up to 950 oC and holding for 2 h for the reduction of the powders and removing oxygen; Stage II – heating up to the sintering temperature (1250, 1300, 1350, and 1400 oС) in 100 min and holding for ts = 1 h. The accuracy of measuring the temperature was ±6 oC. After sintering, the samples were subjected to annealing in vacuum (10-5 Torr) at 1000 oС, 2 h. As the reference samples, the coarse-grained alloys W-7%Ni-3%Fe obtained by sintering the coarse-grained powder compositions in hydrogen were used.
Method II was implemented using Dr. Sinter model SPS-625 setup (SPS SYNTEX Inc., Japan). The sintering of the samples of 30 mm in diameter and 4-4.5 mm in height was performed in vacuum (6 Pa) when applying a uniaxial pressure of 70 MPa. A single-stage sintering regime was used – heating with the rate of 100 oC/min up to the sintering temperature (1200 oC). The uniaxial pressure was applied simultaneously with the beginning of heating up. The holding at the sintering temperature was absent (ts = 0 min). The temperature was measured by an optical pyrometer Chino IR-AH (Japan) focused onto the surface of the graphite mold. The uncertainty of the temperature measurement was ±5 oC. As the reference samples, the nanopowders after HEBM subjected to annealing in hydrogen at 900 oC, 1 h in advance (before SPS) were used.
In the course of heating up, the effective shrinkage (Leff, mm) and shrinkage rate (Seff, mm/s) of the nanopowders were measured using a dilatometer. The contribution of the thermal expansion (L0, mm) of the graphite mold and punch was taken into account on the base of the results of experiments with empty mold. The temperature curves of shrinkage of the nanopowders (L, mm) were calculated according to the formula: L(T) = Leff(T) – L0(T).
The residual graphite was removed from the surfaces of the sintered samples of 30 mm in diameter by waterjet cleaning. The mechanical grinding was performed using Struers Secotom 10 setup, polishing – Buehler Automet 250 setup. In the course of mechanical treatment, the layers of ~300-350 μm, were removed from the sample surfaces, which an increased carbon concentration can be observed in due to high temperature interaction of the material with the graphite mold [
47,
48].
The microstructure investigations of the alloys were carried out using Leica IM DRM optical metallographic microscope, Jeol JSM-6490 scanning electron microscope (SEM), and TESCAN Vega 2 SEM with Oxford Instruments INCA 350 EDS microanalyzer. The density of the samples was measured by hydrostatic weighting in distilled water using Sartorius CPA balance. The uncertainty of measuring the density was ± 0.01 g/cm
3. The magnitude of the theoretical density was accepted to be equal to ρ
th = 17.245 g/cm
3. The analysis of the chemical composition was performed using Ultima 2 ICP atomic-emission spectrometer and Leco RHEN-602 analyzer. The X-ray diffraction (XRD) phase analysis was performed using Shimadzu XRD-700 diffractometer (CuK
α emission, λ = 1.54056 Å, scan rate 0.5
o/min, exposure 2 s). The phase analysis of the samples was performed by Rietveld method. The contribution of the apparatus broadening into the XRD peaks was taken into account according to [
45]. The analysis of internal stresses and calculation of the sizes of coherent scattering regions were performed by Williamson-Hall method [
49].
To study the mechanical properties of the sintered samples, the stress-relaxation testing technique was used allowing determining the magnitudes of the macroelasticity stress (lattice friction stress) (σ
0) and of the yield strength (σ
y) in the compression tests [
45]. To measure σ
0 and σ
y, the cylindrical samples of 3 mm in diameter and 6 in height were used. The uncertainties of measuring σ
0 and σ
y were ±20 MPa. The microhardness (Hv) of the alloys was measured using HVS-1000 hardness tester at the load of 1 kg.
The magnitude of compressive dynamic ultimate strength (Ys) was investigated by Kolsky bar method using Hopkinson split rod at the strain rate of ~10
3 s
-1 [
50]. An example of a dynamic strain curve of a tungsten alloy is presented in
Figure 1. The ballistic characteristics of the material were determined by measuring the penetration depth (H) of the sample of 3 mm in diameter and 25 mm long (
Figure 2a) into a steel target of 30 in thickness. The penetration depth of the sample was measured on the metallographic cross-section with Leica IM DRM optical microscope using Good Grains software (
Figure 2b). The uncertainty of determining the magnitude of H was ± 0.01 mm. The steel hardness was 27-30 HRC. The samples were accelerated with a 12 mm two-stage gas gun powered by compressed air. The impact angle of the sample with the steel plate was 90
o. The test procedure was described in [
46]. The samples for testing were cut out from the central parts of the workpieces of 30 mm in diameter by spark cutting in distilled water.
4. Discussion
The nonmonotonous dependence of the density of the samples obtained by SPS on the HEBM time (
Figure 13) is one of the most interesting results obtained in the present work. The decrease in the alloy density from 98.4% down to 90.9% for the samples sintered from the non-annealed nanopowders and from 98.7% down to 94.6% for the ones sintered from the annealed nanopowders was observed when increasing HEBM time from 0 up to 20 min.
Note that large pores were absent in the microstructure of the sintered samples. The comparison of the microstructure of the samples sintered from the coarse-grained powders and from the nanopowders after t
HEBM = 20 min (
Figure 8 and
Figure 9) did not reveal any essential increase in the sizes and volume fraction of the pores. So far, one can conclude that the decrease in the density of the alloy sintered from the nanopowders originates not from the increase in the volume fraction of the pores but from other factors. This conclusion is supported indirectly by the analysis of the curves L(T) and S(T) presented in
Figure 7. As it has been shown above, the maximum values of the shrinkage L
max and of the shrinkage rate S
max were observed in the case of SPS of the powders subjected to HEBM for 20 min.
To answer the question of the origin of the non-monotonous dependence of the density on the HEBM time, let us analyze the SPS mechanisms of the W-Ni-Fe nanopowders. First, it is worth noting that the characteristic sintering temperature T
s = 1200
oC is ~0.4T
m(W) and, according to [
54], the diffusion intensity and the strain rate in tungsten are very low. (Here T
m(W) = 3695 K is the melting point of tungsten [
54]).
As one can see in
Figure 9a,b, the curve L(T) has usual three-stage character: the shrinkage of the nanopowders is finished almost completely within the intensive shrinkage stage (Stage II) in the temperature range from 750-800
oC to 1000-1050
oC. The intensity of shrinkage of the nanopowders within the third stage of sintering (Т
s > 1000-1050
oC), which the grain growth occurs at often is very low.
The analysis of the compaction kinetics of the nanopowders at Stage II can be made using Young-Cutler model [
55] developed for the analysis of the compaction curves of the fine-grained powders in the continuous heating regime. The Young-Cutler model [
55] describes the initial stage of non-isothermic sintering of spherical particles in the conditions of simultaneous volume and grain boundary diffusion as well as plastic deformation processes. According to [
55], the slope of the dependence ln(T∂ε/∂T) – T
m/T corresponds to the effective activation energy of the non-isothermic sintering process
mQs2 where
m is a numerical coefficient depending on the diffusion mechanism, T
m = T
m(Ni) = 1723 K is the melting point of the γ-phase [
1,
2]. The magnitude of
m = 1/3 for the case of the grain boundary diffusion,
m = 1/2 for the case of the volume diffusion in the crystal lattice,
m = 1 for the case of the creep of the material [
55].
The dependencies ln(T·∂ε/∂T) – T
m/T are presented in
Figure 14. As one can see in
Figure 14, the dependencies ln(T·∂ε/∂T) – T
m/T have usual two-stage character that confirms the correctness of application of the Young-Cutler model to analyze the sintering kinetics of the W-Ni-Fe alloy. The values of the effective sintering activation energy at the intensive shrinkage stage (
mQs2) are presented in
Table 4. The mean uncertainty of determining the activation energy
mQs2 was ±0.2 kT
m.
Table 4.
SPS activation energies of nanopowders W-7%Ni-3%Fe.
Table 4.
SPS activation energies of nanopowders W-7%Ni-3%Fe.
|
Nanopowders after HEBM |
Nanopowders after HEBM and annealing in hydrogen |
tHEBM, min |
Stage II |
Stage III |
Stage II |
Stage III |
mQs2, kTm
|
m |
Qs2, kTm / kJ/mol |
Qs3, kTm / kJ/mol |
mQs2, kTm
|
m |
Qs2, kTm / kJ/mol |
Qs3, kTm / kJ/mol |
0 |
3.9 |
1/3 |
11.7 / 167 |
16.1 / 230 |
3.2 |
1/3 |
9.6 / 137 |
17.2 / 246 |
5 |
6.5 |
1 |
6.5 / 93 |
17.8 / 255 |
5.9 |
1 |
5.9 / 84 |
15.4 / 221 |
10 |
6.0 |
6.0 / 86 |
19.1 / 273 |
4.9 |
4.9 / 70 |
14.7 / 210 |
20 |
5.2 |
5.2 / 75 |
18.9 / 271 |
4.6 |
4.6 / 65 |
15.8 / 226 |
40 |
7.0 |
7.0 / 100 |
19.2 / 275 |
6.4 |
6.4 / 92 |
16.8 / 240 |
In the case of the coarse-grained powders, the magnitude of the effective SPS activation energy
mQs2 was close to the activation energy of the grain boundary diffusion in nickel (
Qb(Ni) = 115 kJ/mol [
54]) at
m = 1/3. This allows suggesting the kinetics of high-speed sintering of the coarse-grained powders W-7%Ni-3%Fe at the intensive shrinkage stage to be controlled by the intensity of the grain boundary diffusion in the γ-phase.
A good correspondence of the SPS activation energy of the nanopowders to the data on the activation energy of the diffusion processes reported in the literature was observed for
m = 1 (
Table 4). The value of
m = 1 allows concluding the sintering kinetics of the nanopowders W-7%Ni-3%Fe to be determined by the Coble creep intensity. This conclusion agrees qualitatively with M.F. Ashby deformation-mechanism maps [
54], in which the SPS regimes used in the present work correspond well to the creep area for nickel (stress σ = 70 MPa ~ 9⋅10
-4G where G = 78.9 GPa [
54] is the shear modulus for nickel, temperature 1200
oС). In our opinion, large lengths of the grain boundaries in the sintered materials is the origin of the realization of Coble creep mechanism in SPS of nanopowders since no intensive grain growth takes place at this sintering stage yet. Note also that the decreasing of the grain sizes would lead to an essential increase in the creep rate in the metal materials [
56]. Note also that Coble creep is one of main compaction mechanisms in SPS of WC [
57], WC-Co [
58,
59], and YAG:Nd-W [
60] nanopowders.
Note that sometimes lower values of the activation energy of the grain boundary diffusion in fine-grained nickel are reported in the literature as well. In particular, the creep activation energy of 110-115 kJ/mol and the activation energy of the grain boundary diffusion of ~66 kJ/mol were reported for ultrafine-grained nickel with the grain sizes 0.3-0.5 μm [
61].
The annealing of the nanopowders in hydrogen leads to a decrease in the SPS activation energy (
Table 4). In our opinion, it originates from, first of all, the reduced concentration of oxygen adsorbed on the nanopowder particle surfaces. This leads to an increase in the diffusion creep rate in the nanopowders W-7%Ni-3%Fe at elevated temperatures. Thin tungsten oxide layers, which can form on the W nanoparticle surfaces when heating in vacuum can be another origin of decrease in the creep rate. The oxide nanoparticles located at the tungsten grain boundaries can lead to a decrease in the creep rate at the second SPS stage.
In the high temperature range, the slope of the dependence ln(T·∂ε/∂T) – T
m/T (
Figure 14) becomes negative. It means that one has to use another approach to estimate the sintering activation energy in the higher temperature range. According to [
62], one can estimate the activation energy at this stage using the model of diffusion dissolving of pores located near the grain boundaries in the fine-grained materials. The correctness of application of this procedure was demonstrated earlier in [
45,
60,
63,
64]. The activation energy at the non-isothermic sintering stage
Qs3 can be determined from the slope of the dependence ρ(T)/ρ
th in the double logarithmic axes ln(ln(α⋅ρ/ρ
th/(1-ρ/ρ
th)) – T
m/T where α is the coefficient of compaction when pressing (α = 0.45 for nanopowders α-W) (
Figure 15). The mean uncertainty of determining the activation energy
Qs3 was ±1.5 kT
m. The calculated values of
Qs3 are presented in
Table 4. For the annealed nanopowders, the value of
Qs3 was ~2-3 kT
m less than the one for the mechanically activated nanopowders.
The calculated values of the SPS activation energy correspond well to the activation energy of heterodiffusion of the tungsten atoms in the γ-phase (taking into account the corrections for the decrease in the activation energy of the grain boundary diffusion in the fine-grained materials [
65]). Available data on sintering activation energies of W-Ni-Fe tungsten alloys give dramatically different values;
Qs for W-8.4Ni-3.6Fe alloy is 250 kJ/mol according to [
66], whereas reference [
67] gives
Qs = 367 kJ/mol for W-8.4Ni-3.6Fe alloy and 480 kJ/mol for 95W-3Ni-2Fe. In [
68], the SPS activation energy for the fine-grained alloy W-5.6%Ni-1.4%Fe was shown to depend on the heating rate – at V < 30
oC/min the magnitude of
Qs = 454 kJ/mol, at V > 200
oC/min the magnitude of
Qs = 200 kJ/mol. The diffusion activation energy for
181W in nickel crystal lattice is
Qv ~268-309 kJ/mol [
8,
69]. In the cases of the volume diffusion in the Ni-W system and in the γ-Fe-W one, the values of
Qv are 295-306 kJ/mol and 268 kJ/mol, respectively [
70,
71]. So far, one can conclude the compaction of the W-Ni-Fe nanopowders at high temperatures to be determined by the intensity of the heterodiffusion of the tungsten atoms in the crystal lattice of the γ-phase.
Summarizing the results of analysis, one should stress that the intensive shrinkage of the mechanically activated nanopowders W-7%Ni-3%Fe goes at low temperatures and is characterized by small values of compaction activation energy. In our opinion, low activation energy of sintering in nanopowder materials is the origin of the reduction of the activation energy of sintering of the W-Ni-Fe nanopowders.
According to the theory of the nonequilibrium grain boundaries [
65], the increasing of the free (excess) volume of the grain boundaries Δα is the origin of the decrease in the activation energy of the grain boundary diffusion in the strongly deformed fine-grained metals and alloys. The magnitude of Δα is proportional to the density of orientation mismatch dislocations (OMDs) captured by the grain boundaries during severe plastic deformation. In the course of HEBM, the grinding of the γ-phase particles with the FCC lattice is difficult due to high ductility of these ones. In the course of HEBM, the grinding of the γ-phase particles and the accumulation of the OMDs at the grain boundaries take place [
65]. The diffusion permeability of the γ-phase grain boundaries after HEBM appears to be higher than the one of “conventional” grain boundaries in the γ-phase formed during sintering individual γ-phase particles to each other. The enhanced diffusion permeability of the nonequilibrium grain boundaries to accelerated diffusion of the tungsten atoms in the γ-phase, and to enhanced Coble creep rate. These factors will lead to a decrease in the SPS activation energy of the nanopowders W-Ni-Fe.
The second factor promoting the decrease in the activation energy of the SPS of nanopowders can be the nonequilibrium state of their crystal lattice.
As it has been shown above, the asymmetric distortion of the (110) α-W XRD peak towards higher reflection angles after HEBM was observed. In our opinion, it means that the crystal lattice constant in the surface layers of the W particles appears to be smaller than the one in the central parts of the W particles. So far, the asymmetry of the (110) α-W XRD peak originates from the increase in the concentrations of the Ni and Fe atoms in the surface layers of the α-W particles. the concentrations of the Ni and Fe atoms in the surface layers of the α-W nanoparticles and from the increase in the concentration of the W atoms in the γ-phase particles with the nonequilibrium grain boundaries.
Tungsten is known to have a high solubility in nickel. At elevated temperatures, the concentration of the W atoms in the γ-phase can reach 25-28% [
4]. Partial dissolving of the α-W nanopowders in the γ-phase takes place during sintering that leads to a decrease in the density of the sintered W-Ni-Fe samples. This process can be accelerated by larger specific surface area of the α-W nanoparticles, strong distortion of the crystal lattice in the surface layers of the α-W nanoparticles, and high diffusion permeability of the γ-phase nonequilibrium grain boundaries.
Now, let us consider the origins of the non-monotonous dependence of the density of the W-7%Ni-3%Fe alloy on the HEBM time (
Figure 13). This effect was observed, in particular, for the nanopowders annealed in hydrogen and, consequently, its nature is not related to the oxidation of the nanopowders in the course of storing.
The primary origin of the decrease in the alloy density is the strain-stimulated dissolving of the W atoms in the γ-phase lattice during HEBM. This effect can have a considerable scale for the W-7%Ni-3%Fe alloy since the volume fraction of the γ-phase in this alloy exceeds 20%. The increasing of the HEBM time lead to an increase in the volume fraction of the tungsten particles dissolved in the γ-phase. The decrease in the initial density of the W-Ni-Fe nanopowders would lead to a decrease in the final density of the alloy at given sintering regimes provided all other conditions being equal.
It is worth noting that the acceleration of the grain boundary diffusion processes and the increase in the creep rate should lead to an increase in the rate of solid phase sintering of the W-Ni-Fe nanopowders. This effect would compensate partly the negative effect of HEBM on the density of the W-7%Ni-3%Fe alloy.
The question of the origins of the increase in the SPS activation energy at long HEBM times (tHEBM = 40 min) is more complex. In our opinion, this effect is caused simultaneously by three factors.
First, it is worth noting that the intensity of the strain-induced dissolving of the tungsten particles can decrease at long HEBM times due to the achievement of the solubility limit of W in the γ-phase. In this case, the increasing of the HEBM time leads to an increase in the creep rate due to an increase in the nonequilibrium degree of the grain boundaries in the γ-phase and a decrease in the grain sizes in the alloy.
The analysis of the XRD phase analysis results shows that the micro-strain of the α-W crystal lattice begins to decrease after 40 min of HEBM (
Figure 5b). This effect can originate from the decrease in the concentrations of the Ni and Fe atoms in the surface layers of the α-W nanoparticles as well as from the decrease in the dislocation density in the α-W particles. The heating of the powders during HEBM is the most probable origin of this effect. The increase in the temperature during HEBM will reduce the distortion degree of the α-W crystal lattice and, as a consequence, the tendency of the nanoparticles to dissolving in the strongly deformed γ-phase. It will lead in an increase in the density of the tungsten alloy.
The increase in the density of agglomerates after 40 min of HEBM is the third probable origin. The increase in the density of agglomerates can originate from the intensive plastic deformation of the tungsten nanoparticles during HEBM as well as from the heating of the samples. The increase in the packing density of the nanoparticles inside the agglomerates will lead to an increase in the initial density of the powder composition W-Ni-Fe and, as a consequence, will provide an opportunity to achieve a higher final density of the tungsten alloy when sintering in the same conditions.
Finally, let us discuss briefly the results of investigations of the mechanical properties of the HTAs. As one can see in
Figure 16a, the dependencies of hardness and yield strength on the grain size can be described by Hall-Petch relation with a good accuracy. Traditionally, the dependence of the strength characteristics on the grain size in the coarse-grained HTAs is suggested to have more complex character. In particular, the magnitudes of the ultimate strength and of the yield strength in the tension tests depend strongly on the volume fraction of the γ-phase particles as well as on the relations of the lengths of the intergranular boundaries W-W and the ones of the interphase boundaries W-γ [
5,
11,
31,
32]. The realization of Hall-Petch effect in the ultrafine-grained HTAs can be a consequence of considerably greater lengths of the grain boundaries that leads to the proportional decrease in the sizes (thicknesses) of the interphase boundaries α-W - γ-phase.
Knowing the magnitudes of the yield strength σ
y, of the macroelasticity stress σ
0, and of the grain size d one can calculate Hall-Petch coefficient according to the formula K
HP = (σ
y - σ
0)⋅d
1/2 for each alloy reflected in
Table 2 and
Table 3.
Figure 16b presents the dependence of K
HP on the grain size. As one can see in
Figure 16b, the increasing of the sintering temperature and the grain growth lead to an increase in Hall-Petch coefficient. The maximum magnitudes of K
HP were achieved in the HTAs sintered at 1450 and 1500
oC (
Table 2). In our opinion, this evidences an increase in the adhesion strength of the interphase boundaries with increasing sintering temperature. The increasing of the concentration of the W atoms in the γ-phase can also contribute additionally into the increase in K
HP.
No essential differences in the magnitudes of KHP for the alloys obtained from the non-annealed and annealed nanopowders were observed. This allows suggesting the differences in the strength characteristics of the HTAs observed to originate from, first of all, the differences in the grain sizes.
Figure 1.
Typical stress – strain and strain rate – strain curves for a tungsten alloy sample recorded in a dynamic compression test according to Kolsky bar method.
Figure 1.
Typical stress – strain and strain rate – strain curves for a tungsten alloy sample recorded in a dynamic compression test according to Kolsky bar method.
Figure 2.
Tungsten alloy samples for testing (a) and example of measuring the sample penetration depth into the steel target (b).
Figure 2.
Tungsten alloy samples for testing (a) and example of measuring the sample penetration depth into the steel target (b).
Figure 3.
SEM images of W-7%Ni-3%Fe powder agglomerates for different HEBM: times (a) 0 min; (b) 5 min; (с) 10 min; (d) 20 min; (e, f) 40 min.
Figure 3.
SEM images of W-7%Ni-3%Fe powder agglomerates for different HEBM: times (a) 0 min; (b) 5 min; (с) 10 min; (d) 20 min; (e, f) 40 min.
Figure 4.
SEM images of the nanoparticles in the powder agglomerates 90%W-7%Ni-3%Fe for different HEBM times: (a) tHEBM = 10 min; (b) tHEBM = 40 min.
Figure 4.
SEM images of the nanoparticles in the powder agglomerates 90%W-7%Ni-3%Fe for different HEBM times: (a) tHEBM = 10 min; (b) tHEBM = 40 min.
Figure 5.
Results of XRD phase analysis of the powders 90%W-7%Ni-3%Fe: (
а) overview XRD curves after different HEBM times; (
b) effect of the HEBM time on the broadening of the (110) α-W pears [
45]; (
c) effect of the HEBM time on the CSR size (1) and on the magnitude of the micro strain of the α-W crystal lattice (2) [
45].
Figure 5.
Results of XRD phase analysis of the powders 90%W-7%Ni-3%Fe: (
а) overview XRD curves after different HEBM times; (
b) effect of the HEBM time on the broadening of the (110) α-W pears [
45]; (
c) effect of the HEBM time on the CSR size (1) and on the magnitude of the micro strain of the α-W crystal lattice (2) [
45].
Figure 6.
Appearance of the W-7%Ni-3%Fe alloy samples sintered in hydrogen from the nanopowders obtained by HEBM. Sintering temperature: (a) 1450 oС; (b) 1500 oС.
Figure 6.
Appearance of the W-7%Ni-3%Fe alloy samples sintered in hydrogen from the nanopowders obtained by HEBM. Sintering temperature: (a) 1450 oС; (b) 1500 oС.
Figure 7.
Microstructure of W-7%Ni-3%Fe alloy sintered from coarse-grained (a, c, e) and nanopowder (tHEBM = 20 min) (b, d, f). Sintering temperature Ts = 1250 oC (a, b), 1400 oC (c, d).
Figure 7.
Microstructure of W-7%Ni-3%Fe alloy sintered from coarse-grained (a, c, e) and nanopowder (tHEBM = 20 min) (b, d, f). Sintering temperature Ts = 1250 oC (a, b), 1400 oC (c, d).
Figure 8.
Microstructure of the W-7%Ni-3%Fe alloy sintered from the nanopowder (tHEBM = 20 min) at the temperatures 1150 oC (a), 1200 oC (b), and 1400 oC (с).
Figure 8.
Microstructure of the W-7%Ni-3%Fe alloy sintered from the nanopowder (tHEBM = 20 min) at the temperatures 1150 oC (a), 1200 oC (b), and 1400 oC (с).
Figure 9.
Temperature curves of shrinkage (a, b) and shrinkage rate (c, d) for the W-7%Ni-3%Fe nanopowders after HEBM (a, c), and after HEBM and annealing in hydrogen (b, d).
Figure 9.
Temperature curves of shrinkage (a, b) and shrinkage rate (c, d) for the W-7%Ni-3%Fe nanopowders after HEBM (a, c), and after HEBM and annealing in hydrogen (b, d).
Figure 10.
Microstructure of the W-7%Ni-3%Fe alloy obtained by SPS at 1200 oC from non-annealed (а, c, e) and annealed (b, d, f) nanopowders. HEBM time: (a, b) 0 min; (c, d) 10 min; (e, f) 40 min.
Figure 10.
Microstructure of the W-7%Ni-3%Fe alloy obtained by SPS at 1200 oC from non-annealed (а, c, e) and annealed (b, d, f) nanopowders. HEBM time: (a, b) 0 min; (c, d) 10 min; (e, f) 40 min.
Figure 11.
Microstructure of the samples obtained by SPS from the annealed (a) and non-annealed (b) nanopowders after HEBM (tHEBM = 40 min). The areas with the abnormally large grains in the microstructure of the W-7%Ni-3%Fe alloy obtained by SPS at 1200 oC from non-annealed (с, d, e) and annealed (f) nanopowders: (c) tHEBM = 10 min, (d) tHEBM = 20 min; (e, f) tHEBM = 40 min.
Figure 11.
Microstructure of the samples obtained by SPS from the annealed (a) and non-annealed (b) nanopowders after HEBM (tHEBM = 40 min). The areas with the abnormally large grains in the microstructure of the W-7%Ni-3%Fe alloy obtained by SPS at 1200 oC from non-annealed (с, d, e) and annealed (f) nanopowders: (c) tHEBM = 10 min, (d) tHEBM = 20 min; (e, f) tHEBM = 40 min.
Figure 12.
Results of XRD analysis of the W-7%Ni-3%Fe alloy obtained by SPS at 1200 oC. Effect of the HEBM time on the parameters of the (100) α-W peak in the samples sintered from the non-annealed (a) and annealed nanopowders (b).
Figure 12.
Results of XRD analysis of the W-7%Ni-3%Fe alloy obtained by SPS at 1200 oC. Effect of the HEBM time on the parameters of the (100) α-W peak in the samples sintered from the non-annealed (a) and annealed nanopowders (b).
Figure 13.
Dependencies of relative density of the W-7%Ni-3%Fe alloy sintered at 1200 oC on the HEBM time: (1) – densities of the samples sintered from the non-annealed nanopowders; (2) – densities of the samples sintered from the annealed nanopowders.
Figure 13.
Dependencies of relative density of the W-7%Ni-3%Fe alloy sintered at 1200 oC on the HEBM time: (1) – densities of the samples sintered from the non-annealed nanopowders; (2) – densities of the samples sintered from the annealed nanopowders.
Figure 14.
Dependencies (T·∂ε/∂T) – Tm/T for the W-7%Ni-3%Fe alloy. Analysis of the compaction mechanisms of the non-annealed nanopowders at Stage II.
Figure 14.
Dependencies (T·∂ε/∂T) – Tm/T for the W-7%Ni-3%Fe alloy. Analysis of the compaction mechanisms of the non-annealed nanopowders at Stage II.
Figure 15.
Dependencies ln(ln(α⋅ρ/ρth/(1-ρ/ρth)) – Tm/T for non-annealed nanopowders W-7%Ni-3%Fe.
Figure 15.
Dependencies ln(ln(α⋅ρ/ρth/(1-ρ/ρth)) – Tm/T for non-annealed nanopowders W-7%Ni-3%Fe.
Figure 16.
Analysis of investigations of the mechanical properties of the HTAs 90%W-7%Ni-3%Fe: (
a) dependencies of the hardness and of the yield strength on the grain size; (
b) dependence of Hall-Petch coefficient on the grain size. Analysis of the results presented in
Table 2 and
Table 3.
Figure 16.
Analysis of investigations of the mechanical properties of the HTAs 90%W-7%Ni-3%Fe: (
a) dependencies of the hardness and of the yield strength on the grain size; (
b) dependence of Hall-Petch coefficient on the grain size. Analysis of the results presented in
Table 2 and
Table 3.
Table 1.
Chemical composition of initial powders (wt.%).
Table 1.
Chemical composition of initial powders (wt.%).
Powders |
O |
Fe |
C |
S |
P |
Ni |
Co |
Si |
Cu |
Mo |
Mn |
α-W |
8⋅10-2
|
2⋅10-2
|
1⋅10-2
|
- |
5⋅10-3
|
1⋅10-2
|
- |
5⋅10-3
|
1⋅10-2
|
4.5⋅10-2
|
2⋅10-3
|
β-Ni |
3⋅10-1
|
1.5⋅10-3
|
1⋅10-1
|
6⋅10-4
|
1⋅10-3
|
- |
7⋅10-4
|
1⋅10-3
|
1⋅10-3
|
- |
3⋅10-4
|
α-Fe |
2⋅10-1
|
- |
4.8⋅10-2
|
4⋅10-3
|
- |
- |
- |
1⋅10-2
|
- |
- |
- |
Table 2.
Microstructure characteristics and mechanical properties of the W-7%Ni-3%Fe alloy samples obtained from the nanopowders by sintering in hydrogen.
Table 2.
Microstructure characteristics and mechanical properties of the W-7%Ni-3%Fe alloy samples obtained from the nanopowders by sintering in hydrogen.
|
Characteristics of alloy obtained from coarse-grained powders |
Characteristics of alloy obtained from nanopowders |
Ts, oC |
ρ, g/cm3
|
d, μm |
σ0, MPa |
σy, MPa |
Hv, GPa |
Ys, MPa |
H, mm |
ρ, g/cm3
|
d, μm |
σ0, MPa |
σy, MPa |
Hv, GPa |
Ys, MPa |
H, mm |
1250 |
17.94 |
5-10 |
520 |
1150 |
4.2 |
1830 |
3.8 |
17.79 |
1-3 |
960 |
1540 |
7.9 |
1900 |
5.1 |
1300 |
18.02 |
~5-10 |
460 |
990 |
4.2 |
1700 |
4.15 |
17.93 |
1-3 |
1000 |
1340 |
7.5 |
2050 |
4.68 |
1350 |
18.06 |
~10 |
290 |
790 |
4.1 |
1650 |
4.23 |
17.95 |
1-3 |
860 |
1200 |
7.2 |
1870 |
3.72 |
1400 |
18.14 |
~20 |
230 |
740 |
4.0 |
1640 |
4.05 |
17.49 |
3-5 |
300 |
740 |
6.9 |
1500 |
3.1 |
1450 |
18.11 |
40-45 |
220 |
600 |
3.8 |
1590 |
3.7 |
17.05 1
|
~10 |
- |
- |
6.5 |
- |
- |
1500 |
18.06 |
~50 |
200 |
690 |
3.6 |
1580 |
3.2 |
16.87 1
|
~22 |
- |
- |
6.2 |
- |
- |
Table 3.
Microstructure characteristics and mechanical properties of the W-7%Ni-3%Fe alloy samples obtained by SPS from nanopowders (Ts = 1200 oC).
Table 3.
Microstructure characteristics and mechanical properties of the W-7%Ni-3%Fe alloy samples obtained by SPS from nanopowders (Ts = 1200 oC).
|
Characteristics of alloy obtained from non-annealed nanopowders |
Characteristics of alloy obtained from annealed nanopowders |
tHEBM, min |
ρ, g/cm3
|
d, μm |
σ0, MPa |
σy, MPa |
Hv, GPa |
Ys, MPa |
H, mm |
ρ, g/cm3
|
d, μm |
σ0, MPa |
σy, MPa |
Hv, GPa |
Ys, MPa |
H, mm |
0 |
16.97 |
1.3 |
920 |
1850 |
4.2 |
- |
4.9 |
17.02 |
1.2 |
1050 |
1930 |
4.3 |
- |
5.1 |
5 |
16.64 |
0.9 |
1330 |
2160 |
4.5 |
2280 |
5.8 |
16.79 |
1.0 |
1400 |
2270 |
4.8 |
2350 |
5.1 |
10 |
16.45 |
0.8 |
1450 |
2180 |
4.7 |
- |
5.7 |
16.92 |
0.7 |
1520 |
2310 |
4.9 |
- |
5.4 |
20 |
15.68 |
0.7 |
1500 |
2370 |
4.8 |
- |
5.7 |
16.31 |
0.6 |
1610 |
2350 |
5.3 |
- |
6.6 |
40 |
16.78 |
0.7 |
1480 |
2350 |
4.7 |
2480 |
5.6 |
17.04 |
0.6 |
1530 |
2320 |
4.9 |
2630 |
6.1 |