4.3. Influence of C-np on Microstructure: Metallurgical Phases, RESIDUAL Austenite Content, Texture and Grain Size
By investigating the chemically etched cross sections of the specimens via OM, SEM imaging and EDS area scanning, no obvious differences in the microstructure were detected. As a consequence of the process-inherent high cooling rates, both the pure AISI H11 sample as well as the specimens manufactured with the C-np-enriched AISI H11 powder mixtures have a fine-grained, predominately dendritic microstructure (see
Figure 4b,c,e,f). As exemplarily illustrated in the SEM images presented in
Figure 4b,c,e,f, a fine needle-like inner structure is discernable within the dendrites for all investigated tool steel specimens. It is most likely to be martensite, and includes amounts of ferritic phases (e.g., ferritic bainite) and a retained austenite (RA).
Moreover, as could be verified with EDS area mapping (see
Figure 5,
Figure A1 and
Figure A2), the metallic main alloying elements chromium (Cr), molybdenum (Mo), manganese (Mn) and vanadium (V) occur concentrated at the inter-dendritic spaces or at the grain boundaries, respectively. In contrast, the iron (Fe) content in these areas decreases noticeably (see
Figure 4b,c). This points out the formation of strength- and hardness-increasing carbides, preferably Cr- and Mo-mixed carbides from type M
7C
3 or M
23C
6 as well as VC- and Mo
2C-special carbides. These observations are also in good accordance with the scientific outcomes of others who reported on the formation of a dense carbide network, which mainly consists of fine Cr-mixing carbides, in AISI H11 or similar CrMo-alloyed hot-work tool steel components manufactured with PBF-LB/M [
37].
Furthermore, by investigating the microstructure with respect to the sample height, it is discernable that all specimens have differences in local or at least layer-specific inner structure. Low-magnification OM analysis of etched cross sections revealed two characteristic regions (compare to
Figure 4a,d). Independently, on the processed tool steel powder mixture, the bottom part is slightly over-etched and appears darker in the microscopic images. This seems to be an indicator for the presence of multiple re-austenitized and cyclically tempered martensites. In contrast to the middle and upper section of the tool steel samples, the martensite takes on a brighter coloration, which is characteristic for freshly formed martensite when V2A Beize is used as the etchant.
As already described in [
29] and also reported by others who investigated similar tool steels in DED-LB/M [
17,
35], the layer-specific microstructural differences can be attributed to a kind of in-process heat treatment of the lower weld track layers during the layer-by-layer manufacturing process. As exemplarily illustrated in
Figure 6, the first layer (black curve) and the lower layers are cyclically exposed to high thermal loads and are re-melted and re-austenitized several times at the beginning of the manufacturing process during the placement of further weld tracks next to or on top.
By continuing the deposition process, thermal energy is permanently introduced into the layer structure by the absorbed laser power, leading to a steady increase in the average interlayer temperature, and a reduction of prevailing temperature gradients and local cooling rates at the top layers. Fresh martensite, which has already formed, is cyclically exposed to a high thermal load and, depending on the prevailing temperature level, can transform itself back into a softer austenite. Depending on the exact process parameters, the deposition strategy, the sample geometry and the prevailing thermal boundary conditions, the interlayer temperature at the top of the specimen can achieve a constant level in the order of or above the martensite start temperature M
s of the processed tool steel after the deposition of a certain number of weld track layers. At this point, a conversion of austenite into martensite can no longer take place during the running DED-LB/M process (compared to the black curve in
Figure 6). Rather, the phase transformation into martensite can only occur after the manufacturing process has finished and the complete specimen cools down below M
S. As shown in
Figure 6, taking the grey curve illustrating the temperature development of a middle weld track layer n as the example, it can be concluded that (in addition to martensite) ferrite and bainite are also proportionally contained in the microstructure with a high probability, due to the process-specific in situ heat treatment related to the layer-by-layer manufacturing approach. However, the proportion of ferritic phases should be smaller compared to martensite, particularly in the C-np-reinforced tool steel powder blends as carbon is a strong austenite former and counteracts the formation of ferritic phases in the steel samples. The TTT diagrams calculated for the processed tool steel powder mixtures in JMatPro support this assumption (see
Figure A1).
Nevertheless, to obtain a comprehensive understanding of the impact of C-np on the microstructures of the materials, XRD and EBSD examinations were carried out. The target parameters of these investigations were the RA-c, the texture, the crystallite size of martensite and of the RA and the grain size of martensite.
Figure 7 exemplarily presents the measured diffraction diagrams and the reconstructed curves using the Rietveld approximation for the bottom region of specimens in as-built conditions. Regardless of the samples examined, the diffraction patterns could be satisfactorily modelled and reconstructed using the Rietveld approach.
In the diffractograms, diffraction peaks mainly occurred at the reference positions of the RA and ferrite or martensite. In contrast, no diffraction peaks were observed at the reference positions of alloy-specific carbides. In particular, Cr- and Mo-mixed carbides, which (like martensite and ferrite) have a cubic lattice structure, cannot be detected uniquely in the diffraction diagram, since their peaks overlap with those of the α-Fe. On the other hand, primary carbides, e.g., VC or Mo2C, are only present in low concentrations of less than or equal to 1%, which is why they are below the detection limit/measurement sensitivity of the XRD device used. The reflections at the reference positions of ferrite and martensite could not be clearly assigned to the respective phases, due to the high peak widths in combination with the limitations in instrument resolution. Nevertheless, it was found that the diffractograms of the Rietveld method could be better described with a combination of both phases (ferrite and martensite). Without considering ferrite, the quality of the Rietveld fit obtains poor quality for all samples and the deviations between the reconstructed diffraction diagrams, and the experimental data are noticeably larger. Thus, both phases are probably present in the microstructures of the analyzed specimens. Considering that bainite consists of dislocation-rich ferrite surrounded by cementite (Fe3C, iron carbide), the detected ferrite peaks most likely indicate the presence of bainitic ferrite in the microstructure. This would also match well with the above-described theory of in-process heat treatment. A reliable determination of both phase fractions cannot take place due to the high correlation of the scaling factors of martensite and ferrite. This is also not relevant with regard to the quantification of the Ra content since the diffraction peaks of the RA occur as single peaks.
A total of 10 martensite, 5 ferrite and 5 retained austenite reflections were independently considered for the reconstruction of the experimentally determined diffraction patterns on the examined tool steel.
Table A1 (
Appendix C) summarizes the peaks considered with respect to the metallurgical phase and their crystallite orientation (Miller indices).
Although all specimens have the same martensite, ferrite and RA peaks, the intensities and half widths of individual reflections differ depending on the concentration of the added C-np or the total carbon content, respectively. By comparing intensities of austenite peaks of equal spatial orientations (same Miller indices) with each other (
Figure 5b–d), it becomes obvious that a higher carbon content leads to higher peak intensities and a stabilization and enhanced formation of the high-temperature austenitic phase in the steel samples. In contrast to this, the intensities of the diffraction peaks of martensite and of ferrite seem to be weakened with increasing C-np content, pointing out a decrease in the concentration of these metallurgical phases (see
Figure 7b–d). These observations correlate well with the outcomes of the JMatPro calculations (see
Appendix A) and can be explained by the fact that carbon acts as an austenite former. Consequently, a higher total carbon content in the steel sample leads to a decrease in the martensite transformation temperatures (M
s, M
50%, M
90%). This leads to a stabilization and higher concentrations of the RA in the C-np-enriched tool steel specimens compared to the pure AISI H11 sample.
Figure 8a shows the mean values of the RA content determined using XRD for specimens under as-built conditions and post-heat-treated samples. The mean value of the RA-c of specimens in as-built conditions increases from 9.5 ± 2 vol.% for the unmodified AISI H11 sample to 15.2 ± 2.6 vol.% for the sample manufactured out of powder mixture No. 1 and to 21.4 ± 7.1 vol.% for the specimen containing a maximum C-np content of 0.2 wt.-% (powder mixture No. 2). Consequently, the addition of 0.2 wt.-% C-np more than doubled the RA-c on average. The large standard deviations of the mean values in the order of 20% or more indicate a location dependency of the RA content (see
Figure 8a and
Table 6).
Considering the diffraction diagrams of the local measurements carried out for the sample manufactured using the powder mixture No. 1 (see
Figure A3 in
Appendix D), it becomes obvious that the intensities of the peaks of the same phase and with the same orientation are dependent on the measurement position.
There seems to be a tendency for the intensities of the austenite peaks in the bottom region of the sample to be larger than in the middle or upper sample area. By comparing the individual measuring values of the RA content (listed in
Table 6), with respect to the measuring position, the maximum RA content is independently detected in the bottom region on the investigated tool steel powder mixture. The RA content tends to decrease within higher weld track layers and reaches its minimum in the middle and top section of the specimens. The observed local dependency of the RA content can probably be traced back to the in-process heat treatment of the lower weld track layers using strong thermal cycling and with repeated re-melting and re-austenitizing at the beginning of the manufacturing process.
Figure 8b,c show the mean values of the domain size of martensite and of the RA for the processed tool steel powders and with respect to the heat treatment performed. As already mentioned in
Section 3, the values are derived based on the Rietveld analysis. The domain size of martensite tends to decrease with increasing content of nanoscale carbon additives. In the AISI H11 sample, the measured mean value corresponds to 330 ± 61 Å, and takes on a smaller mean value of 180 ± 117 Å for the specimen produced with the powder mixture No. 1 containing 0.1 wt.-% C-np. The smallest value of 135 ± 20 Å is obtained for the specimen with a maximum C-np concentration of 0.2 wt.-%.
In contrast to this, the domain size of the RA rises up with the increasing amount of C-np added to the AISI H11 host powder. For the pure tool steel sample, the mean value of the RA domain size is equivalent to 175 ± 71 Å, and increases to 265 ± 35 Å for the specimen manufactured with 0.1 wt.-% C-np. For the specimen with a maximum C-np-content of 0.2 wt.-%, the domain size takes on a maximum mean value of 345 ± 17 Å. The increase in the austenite domain size is caused by the concentration of the RA significantly increasing with the amount of C-np in the powder or the total carbon content, respectively. In this way, the probability for the formation of the blocky RA is enhanced.
The smaller measuring values of the domain size of martensite indicate a finer martensitic (ferritic) structure in the C-np-enriched specimens. Consequently, it seems that the C-np promotes the formation of finer martensite grains within the dendrites or primary austenite grains, respectively.
However, the domain size determined with XRD cannot be equated with the grain size of martensite in the classical understanding of this matter. Rather, this quantity describes the area within a grain in which the lattice planes are oriented parallel to one another, and (associated with this) in which the reciprocal lattice vectors have the same orientation. Nevertheless, small domain size values point out a fine-grained structure with high probability, as long as there is no dominant preferential direction of the crystal orientation.
In order to verify and quantify the impact of C-np on grain size and to confirm the results of the XRD investigations, EBSD measurements were carried out with the pure tool steel sample and the sample manufactured out of the powder mixture with maximum C-np content of 0.2 wt.-%. Thereby, the focus was placed on specimens in as-built conditions.
Figure 9 exemplarily shows the determined phase fractions of the RA (FCC) and martensite (BCC) and the inverse pole figures for the sample manufactured with 0.2 wt.-% C-np. The measurements were carried out at three characteristic positions: top, center and bottom.
Based on the EBSD analysis the RA content, the unmodified AISI H11 sample is determined to be 1.3% in the top layer, about 1.6% in the center region and 1.8% in the bottom section. Contrary to the value of the RA content measured, the C-np-reinforced tool steel sample is 4.3% in the top layer, 4.8% in the center and 5.4% in the bottom section, which is noticeably larger. The observed dependency of the RA content on the C-np concentration in the powder agrees well with the results of the XRD measurements; also, if the absolute values are not directly comparable due to the different measuring methods, the tendency is the same. Consequently, the EBSD analysis confirms that the added nanoscale additives act as austenite formers, stabilizing the formation of the RA. Additionally, the EBSD examinations also confirm the local dependency of the RA content.
By analyzing the inverse pole figures, presented in
Figure 9 and
Figure 10, no preferred direction of a C-np-concentration-dependent crystal orientation is detected. As can be seen from the inverse pole figures, the crystallites are in equal orientations, both in the pure tool steel sample and the specimen manufactured out of powder mixture No. 2 (+0.2 wt.-% C-np). This result is also in good accordance with the findings of the XRD investigations, in which a preferred orientation of crystallites could not be clearly identified.
However, C-np seem to have an impact on the resulting grain size (martensite needle size).
Figure 11a–d present the distribution as-area-weighted fraction of the analyzed martensite grain cross sections for the pure sample and the specimen modified by adding 0.2 wt.-% C-np for two different threshold angles (low-angle grain boundaries ≤ 10° and high-angle grain boundaries ≤ 15°). Measurements were carried out at the center position for both specimens. By analyzing the distribution of the measured grain cross sections, it is discernable that the modification of the AISI H11 powder with C-np seem to favor the formations of grains with smaller cross sections (finer martensite), which is indicated by larger values of the area-weighted fraction for small grain cross sections. This relationship was observed for both low-angle as well as high-angle grain boundaries.
Table 7 summarizes the results of the grain size analysis, including the number of grains considered and the mean values of the grain cross sections. The mean values of the grain cross sections (area mean) are smaller in both the small-angle grain boundaries as well as the large-angle boundaries when C-np is added to the host powder, and the total carbon content increases. Consequently, this indicates a finer martensitic/ferritic microstructure for the specimens manufactured using the C-np-enriched tool steel powder mixtures.
One possible reason for the observed grain refinement could be that in the end, due to the addition of the C-np, more carbon is available for the precipitation of carbides from the molten phase, particularly VC and Mo2C primary carbides. These particles have high melting points and can act as additional crystallization nuclei during solidification and the cooling down of laser-induced melt pools, which leads to an increase in the crystallization rates and promotes fine grain formations. Furthermore, the precipitated carbides inhibit the growth of the primarily austenite grains during solidification and the cooling down period and ultimately result in a finer martensitic/ferritic structure within the dendrites. A further reason for the detected grain refinement could be related to a change of the martensite morphology caused by the increased total carbon content in the C-np-enriched tool steel specimens. In addition, it could be that the cooling rate increases during the solidification process due to the modification of the starting material with the carbon nanoparticles and the associated change in the thermal properties (e.g., increase in thermal conductivity). This in turn would favor the formation of finer grains.
4.4. Influence of Heat Treatments (S1-HT, S2-HT) on Microstructural Properties
Independently, on the processed tool steel powder mixture and after one-time tempering at a temperature of 580 °C for 2 h (S1-HT), the microstructure still appears very similar to the inner structure under an as-built condition. Following the EDS mappings and OM of the etched cross sections presented in
Figure 12 and
Figure 13, the process-specific fine-grained dendritic microstructure is still present after carrying out S1-HT.
Furthermore, EDS area mappings confirm that the carbide formers (main alloying elements), Cr, Mo, Mn and V, also accumulate at the inter-dendritic spaces, as it is the case for as-built specimens (see
Figure 13).
Additionally, the fine needle-like structure (martensite, bainitic ferrite) inside the dendrites seems to be more or less unchanged. However, the effect of the tempering process on the microstructure is clearly visible in the RA content determined with XRD, which can be compared to
Figure 8a and the reference peaks for RA shown in
Figure 14a–d.
The RA content significantly decrease for all specimens, wherein the tempering process has the greatest effect on the pre-alloyed tool steel specimen due to the tempering parameters which were chosen according to the material standard of AISI H11 [
30]. For the pure AISI H11 sample, the RA content takes on a mean value of 1.3 ± 0.4% and is lower than the as-built sample by a factor of 7.27. The specimen manufactured with powder mixture No. 1 (+0.1 wt.-% C-np) has an average RA content of 6 ± 1.8%, which is a decrease of about 9% compared to the sample in the as-built condition. For the specimen produced from powder mixture No. 2 (+0.2 wt.-% C-np), a mean value of 6.63 ± 0.8% for the RA content is obtained after carrying out S1-HT. Compared to the non-heat-treated sample, this corresponds to a decrease of about 15%.
Consequently, at this point, the following conclusion can be drawn. The RA content of the C-np-enriched tool steel samples significantly reduces with subsequent tempering using tempering parameters that are applicable for AISI H11. During this heat treatment process, the RA largely converts into ferrite and cementite, resulting in a homogenization and reconfiguration of the microstructure. By repeating the tempering step or by further adapting and successively improving the tempering parameters in consideration with the exact chemical compositions, it should be possible to achieve smaller RA contents in the C-np-reinforced specimens.
As follows from
Figure 8c and
Figure 14a–d, on the processed tool steel powder mixture, the domain size of the RA independently decreases after thermal post-processing using S1-HT (1× tempering). For the AISI H11 specimen, the domain size of the RA obtains a mean value of 159 ± 52 Å, which is on average 15 Å smaller than for the sample in the as-built condition. The domain size of the RA for the specimen manufactured with mixture No. 1 (+0.1 wt.-% C-np) corresponds to 179 ± 51 Å, which is a decrease of about 75 Å compared to the as-built sample. For the specimen manufactured out of mixture No. 2 (+0.2 wt.-%), the domain size of the RA lowers from 345 ± 17 Å (in as-built conditions) to 237 ± 95 Å (in a tempered state). The observed reduction of the domain size of the RA seems to be plausible, considering that the tempering causes a significant decomposition of the RA due to the transformation into ferrite and cementite. Oppositely, considering the standard deviation, the tempering process does not significantly affect the domain size of martensite in comparison to the specimens in the as-built conditions.
However, the dependency of the domain size of martensite and of the RA on the added C-np content or the total carbon content is still present in more or less the same manner as for the specimens in the as-built conditions.
After S2-HT, which consisted of hardening followed by one-time tempering, the microstructure of all samples is similar to an acicular martensitic microstructure, which is how these can also be found in conventionally machined and post-hardened AISI H11 components.
Figure 15a,b illustrate the characteristic martensitic structure for a specimen manufactured using a powder mixture with 0.2 wt.-% C-np. The process-specific dendritic microstructure and the characteristic Cr-mixing carbide network completely dissolves, and Cr tends to be homogenously distributed (see EDS mappings in
Figure 15). In contrast, as could be verified by the EDS area mappings shown in
Figure 15d,e, the VC- and Mo
2C-special carbides remain. First, during austenitization, the complete sample transforms into a high-temperature austenitic phase. It is only through quenching in an oil bath that the austenite transforms into martensite. Due to small differences in the local thermal boundary conditions (e.g., local cooling rates) during quenching, in the end, a homogeneous structure is formed and the layer-specific differences in the microstructure more or less disappear after the subsequent tempering.
Following
Table 6 and
Figure 8a, it is discernable that the RA content decreases significantly compared to the values of specimens in the as-built conditions. After carrying out S2-HT, the RA content for all samples takes on mean values that are in the order of or only slightly different from the values achieved after tempering (S1-HT). However, the effect of the C-np content or the total carbon content on the RA content persists after carrying out S2-HT, as in the case of the tempered or as-built samples.
Regarding the RA, a significant decrease in the domain size is detected for all specimens. The values of the domain size of the RA vary between 100 Å and 130 Å on average. The reduction of the austenite domain size results from the decomposition of the RA into ferrite and cementite.
Contrary to this, considering the standard deviation of the mean values, shown in
Figure 8b, no relevant changes in the domain size of martensite are induced in comparison to both the as-built and tempered specimens. However, the effect of the C-np on the domain size of martensite is still present after thermal post-processing is performed using 2S-HT. Consequently, the outcomes confirm that the addition of the nanoscale carbon additives favor the formation of a finer martensitic or ferritic structure in the builds, even after carrying out heat treatment strategies similar to their industrial use.
4.5. Influence of C-np on Vickers Hardness
As already reported in [
29], the inhomogeneity of the microstructure resulting from the in situ heat treatment of the lower weld track layers during the AM process also affects the hardness distribution.
Figure 16 shows the Vickers hardness HV 1 as a function of sample height h for the specimens in as-built conditions.
On the processed tool steel powder mixture, the hardness independently takes on the lowest values in the bottom region of samples. With the increasing number of deposited weld track layers, the hardness profile follows a steady increase (hardness gradient) and reaches a more or less constant level with maximum hardness in the middle and top section of the specimens (see
Figure 16). The characteristic inhomogeneous hardness profiles detected for each sample also correlate well with the local dependency of the RA content related to the in-process heat treatment, which was discussed in the previous section. Consequently, the local differences in the measured RA content deliver a plausible explanation for the inhomogeneous hardness profiles with lower hardness values in the bottom section and higher constant hardness values in the middle and upper sections of manufactured tool steel specimens. As already mentioned above, the high RA content in the lower section of the manufactured specimens results from the strong thermal cycling and multiple austenitization at the beginning of the manufacturing process.
By comparing the hardness profiles, it is obvious that the hardness noticeably increases with the increasing content of C-np added to the AISI H11 powder. While the pure tool steel sample has an average hardness of around 653 ± 10 HV1 in the upper sample area, it increases to around 767 ± 6 HV1 when just 0.1 wt.-% C-np (mixture No. 1) is added, and increases to a hardness mean value of 770 ± 14 HV1 to obtain a maximum C-np content of 0.2 wt.-% (
Figure 16a (right)).
There are several overriding reasons that are responsible for the hardness increase observed for the C-np-enriched tool steel samples. On the one hand, a higher total carbon content in the steel causes a stronger distortion of the formed martensite, resulting in an enhancement of the local hardness and strength. On the other hand, by adding the nanoscale carbon additives to the H11 powder, in the end, more carbon is present for the precipitation of hardness-increasing carbides. Additionally, a further reason for the high hardness of the C-np-enriched specimens is the fact that the nanoscale additives promote grain refinement and favor a finer martensitic/ferritic structure, as was verified with the EBSD and XRD investigations. Apart from the reasons mentioned, the formation of mixed martensite of high hardness and strength is favored from a total carbon content of about 0.5 wt.-%. Considering the measured chemical composition shown in
Table 3, this is the case for the specimen manufactured out of powder mixture No. 2 containing 0.2 wt.-% C-np. Nevertheless, considering the standard deviation, no substantial difference in hardness in the upper sample range can be identified between the specimens manufactured out mixture No. 1 (+0.1 wt.-% C-np) and mixture No. 2 (+0.2 wt.-% C-np) (see
Figure 16a). Rather, the hardness values of the samples enriched with C-np are at a comparably high level. The reason for this can be traced back to a higher content of the soft RA counteracting the higher hardness of the martensite in the sample produced from mixture No. 2, with a maximum C-np content of 0.2 wt.-% or a total carbon content, respectively.
Figure 16b shows the hardness profiles of the manufactured tool steel samples after the tempering process (S1-HT). With regard to the C-np concentration, a noticeable reduction of the hardness gradient between the bottom layers and the middle and top part was independently achieved for all specimens. Here, the hardness increases in the bottom section of the samples, resulting from the conversion of a soft RA into a (harder) ferrite and cementite. In contrast to this, the hardness in the top section is more or less uninfluenced, or only slightly decreases, compared to the as-built conditions. In result, the local decomposition of the RA induced by the tempering process leads to a homogenization of the microstructure and to an almost complete compensation of local hardness differences.
The hardness-increasing effect of the added C-np is retained, even after carrying out the one-time tempering process. In contrast to the samples in the as-built conditions, there is a clear correlation between maximum hardness and maximum C-np content or total carbon content, respectively. The maximum hardness was detected for the specimen manufactured with mixture No. 2 containing 0.2 wt.-% C-np. The value corresponds to 733 ± 8 HV1, which is a decrease of approx. 36 HV1 in comparison to the mean sample hardness in the as-built conditions. However, this value is about 118 HV1 larger than the mean hardness of the unmodified AISI H11 sample after carrying out S1-HT (see
Figure 16b (right)). For the specimen manufactured using mixture No. 1 (+0.1 wt.-% C-np), a mean value of hardness of 711 ± 11 HV1 is determined, which is a decrease of about 55 HV1 compared to the specimen in the as-built conditions.
Figure 16c illustrates the hardness profiles of the produced tool steel samples after carrying out S2-HT, consisting of a two-stage hardening process followed by one-time tempering (580 °C, 2 h). As already discussed in the scope of the microstructural analysis, the local differences in microstructure almost disappear, and the RA content is significantly lowered in specimens that are heat-treated using S2-HT. By analyzing the hardness dependently on the sample height, it becomes obvious that the homogenization of the microstructure provides a more or less constant Vickers hardness as a function of the sample height. By comparing the hardness profiles with each other, it is discernable that the hardness-increasing influence of the added C-np is still unchanged. Consequently, the maximum mean hardness is achieved with 789 ± 10 HV1 for the specimen with a maximum C-np content of 0.2 wt.-%. In comparison to the sample in the as-built conditions, this corresponds to a slight hardness increase of about 19 HV1. For the specimen built using mixture No. 1 (+0.1 wt.-% C-np), a mean hardness of 766 HV1 is measured. This value is in the same order of magnitude as the hardness achieved for the specimen in the as-built condition.
4.6. Influence of C-np on Compression Properties
As follows on from
Figure 17 (lower left), regardless of the performed heat treatment strategy, the addition of C-np leads to a significant enhancement in the compression yield strength Y
0.2% in comparison to the specimens manufactured out of pure AISI H11 powder. Under as-built conditions, Y
0.2% increases from 1839 ± 61 MPa for the unmodified steel samples to 2178 ± 58 MPa for the specimens from mixture No. 1 (+0.1 wt.-% C-np) and to 2143 ± 68 MPa for the samples manufactured with an additional 0.2 wt.-% of C-np (mixture No. 2) (see
Figure 17). By considering the standard deviations, no clear dependence of the maximum yield strength on the concentration of the nanoscale carbon additives can be derived. On the contrary, the maximum mean value of yield strength occurs when 0.1 wt% of carbon is added and not at the maximum C-np content of 0.2 wt%.
This can presumably be attributed to the increase in the RA-c with the increasing carbon concentration, as was discussed in the previous section for the hardness. Although the distortion in martensite and the associated martensite hardness enhances with a higher carbon content, a larger amount of carbon also stabilizes the softer austenitic high-temperature phase (γ-Fe). As a result, the observed phenomena are superimposed, which results in the characteristic yield strengths measured for the C-np reinforced specimens in the as-built conditions. In contrast to the increase in the mechanical strength and in hardness, however, the addition of C-np causes a deterioration in ductility. As presented in
Figure 17, this leads to a decrease in the maximum achievable degree of formation (maximal strain). While the maximum degree of deformation for the pure tool steel specimens is comparatively high (32 ± 1.3%), it drops down with the increase in the C-np concentration to an average of 12 ± 0.6% for the specimens manufactured with mixture No. 1 (+0.1 wt.-%) and to 10 ± 1.8% for the samples modified by adding 0.2 wt.-% of C-np to the H11 powder.
After the specimens are one-time tempered (S1-HT) and the microstructure was homogenized with a decomposition of RA-c and a reduction of process-inherent internal stresses, the yield strength of the tool steel specimens tends to increase slightly. In this case, Y
0.2% of pure H11 assumes a mean value of 1930 ± 25 MPa and increases by around 90 MPa in comparison to the samples without an additional heat treatment. The maximum Y
0.2% of the tempered specimens is determined for the samples enriched with 0.2 wt.-% of C-np and corresponds to a mean value of 2210 ± 24 MPa, which is equivalent to an increase of around 14% compared to the pure AISI H11 specimens. The determined Y
0.2% of the one-time tempered specimens is in a good agreement with the results presented in [
29]. Additionally, the one-time tempering process seems to improve the ductility of all samples, which manifests itself in an increase of the degree of deformation (maximum strain) in comparison to the specimens under the as-built conditions (see
Figure 17). The improvement in ductility can probably be correlated to the reduction of process-inherent stresses, which result from the reconfiguration of the microstructure through the tempering process. The maximum degree of deformation (maximum strain) occurs using the pure tool steel samples. It assumes a mean value of 40 ± 2.6% and is thus approx. 8% larger than in the non-heat-treated state. By adding the C-np, the degree of deformation also decreases in the tempered samples, as it has already been the case for the specimens under the as-built conditions. For the samples produced from mixture No. 1, a maximum degree of deformation of 32 ± 2.6% occurs on average and a minimum mean value of 24 ± 3.8% occurs when a maximum C-np content of 0.2 wt.-% is added.
At this stage, it can be highlighted that by performing tempering of the manufactured tool steel samples, both a high compression yield strength and an promising ductility is achieved.
After heat treating the specimens using S2-HT (two-stage hardening followed by one-time tempering at 550 °C for 2 h), the mean value of Y0.2% of the pure AISI H11 samples takes on 1859 ± 61 MPa, which is nearly equivalent to the value without an additional post-treatment. Oppositely, the value of Y0.2% of the NP-enriched tool steels decreases compared to specimens in the as-built conditions or the tempered samples. For the samples manufactured with 0.2 wt.-% C-np, the mean value of Y0.2% is determined to be 2110 ± 39 MPa, with a mean value of 2036 ± 16 MPa for the samples modified by adding 0.1 wt.-% C-np to the steel powder. A possible explanation for this could be that the relevant heat treatment parameters (e.g., the austenitizing temperature and holding time) selected for the H11 steel were not adapted to the enhanced total carbon content or changed chemical composition, respectively.
As already observed for the one-time tempered specimens, in comparison to the as-built specimens, the ductility of samples also improves after carrying out S2-HT (see
Figure 17). Here, the maximal degree of deformation for the pure tool steel specimens corresponds to 38 ± 1.1%, and takes on slightly smaller values of 34.3 ± 0.7% for the samples manufactured out of mixture No. 1 (+0.1 wt.-% C-np) and to 33 ± 0.9% for the samples manufactured using mixture No. 2 and enriched with 0.2 wt.-% C-np. Consequently, it can be stated that an improvement in ductility can also be realized with S2-HT.
To evaluate and verify the impact of the C-np on the forming behavior, the strain–stress curves, which are exemplarily presented in
Figure 17 (top), are extensively examined. Here, a steep increase in the compression stress, as a function of the degree of deformation (strain), is observed in the region of the plastic deformation for all specimens in the as-built conditions. This characteristic course is a direct consequence of the strain hardening effects which occurred during the deformation of the specimens. In particular, the transformation of the soft RA into a harder martensite, due to the plastic deformation-induced phase transformation, makes a decisive contribution to strain hardening. The strain hardening effect is therefore much more pronounced in the C-np-enriched samples due to their higher RA content.
After post-heat treating the specimens, the strain hardening effect is significantly weakened due to the reduction in the RA content and the homogenization of the microstructure caused by the induced metallurgical changes. This is also reflected through a significant reduction in the slopes of the strain–stress curves in the range of plastic deformation (see
Figure 17 (top)).