Introduction

Extensive efforts have been expended in doping MgB2 to enhance its superconductive properties, particularly its upper critical field, Bc2. The substitution of C for B has been shown to significantly increase MgB2’s Bc2 beyond that of the unalloyed sample1,2,3,4,5,6. Unfortunately, C doping is successful only at low temperatures (<20 K) since it reduces Tc and increases electron impurity scattering only in the σ band, leaving the high temperature (>20 K) Bc2 unchanged or even reduced. In order to enhance the Bc2 of MgB2 in the higher temperature regime, many attempts have been made to find effective dopants for Mg-site substitution to increase electron impurity scattering in both the σ band and the π band7,8, but without much success in terms of improved properties. Although the substitution of Al for Mg has been demonstrated, it was found to reduce Bc29,10.

Further studies focusing on the AlB2-like metal diborides ZrB2, TiB2 and NbB2, (e.g. refs 11, 12, 13, 14, 15, 16) yielded contradictory results. Feng et al.11,12 reported an enhancement in Bc2 in response to 10 mol% Zr doping; Bhatia et al.13 observed a significant increase in Bc2 (from 20.5 T to 28.6 T at 4.2 K) after adding 7.5 mol% ZrB2 to MgB2 bulks. On the other hand, Zhang et al.14 reported no Bc2 enhancement in ZrB2 doped MgB2 tapes. In any case, while Bc2 enhancements have been noted by various researchers working with MgB2 PIT or powder type processes, no one has reported enhanced transport current, suggesting that the effect may be in a surface layer. The one effort to date which has clearly injected Zr deeply into the grain, resulting in a pulsed laser deposition (PLD) synthesized ZrB2-doped MgB2 thin film15,16, showed a much stronger response to the presence of Zr and in this case a decrease of Tc and Bc2 with increasing Zr content. These various observations give rise to the question: what is the actual influence of Zr doping in MgB2? The possible roles of Zr in MgB2 can be summarized in terms of: 1) extrinsic effects, such as modified intergranular connectivity and reduced grain size11,12; 2) intrinsic effects, such as an influence on Bc2 of Zr substitution for Mg13,14,15,16, or increased flux pinning by a distribution of nano-sized ZrB2/Zr precipitates14. On the other hand, incomplete microscopic evidence of Zr substitution for Mg has been provided, at least for materials made by equilibrium processes (contrasting to the non-equilibrium processing of the films of15,16). Therefore further study on the limits of Zr doping was deemed necessary.

There are many roadblocks to clarifying the true role of chemical doping in MgB2. Chief among them is that homogeneous doping is very hard to achieve. Traditional powder synthesis is generally performed at 600–1000 °C –too low to form homogeneously doped samples. To overcome this problem a high temperature under pressure (HTP) route (see below) was developed to explore solubility limits of dopant species in MgB2 and maximize diffusion during reaction. MgB2 bulks synthesized by HTP should have a greater depth of dopant penetration into the MgB2 for any species introduced (if it is soluble) given the increase in diffusion rate at higher temperatures. HTP samples also have large grain size (over 5 μm) making chemical analysis easier. The purpose of the HTP method is to minimize any diffusion limitations so that we can explore the solubility limits of doping, rather than to fabricate MgB2 bulk samples with high Jc. Also, the use of existing metal diborides (MB2) with a structure isomorphous to MgB2 (P6/mmm) as a vector for effective metal element doping is a promising way to investigate possible changes of superconducting properties like Bc2 and Tc. Thus, three sets of MgB2 bulks doped with ZrB2, TiB2 and NbB2 powders were prepared.

Most additions which have been attempted for MgB2 tend to accumulate at the grain boundaries, with the exception of the above-mentioned C-bearing additions. On the other hand, several studies have shown that a very small amount of Dy2O317 can form nanosize precipitates within the MgB2 grains and thereby enhance flux pinning without changing Tc. Thus a Dy2O3 doped MgB2 bulk was also included in this study for comparison. The changes in microstructure and lattice parameter, as well as the superconducting properties Bc2, Tc and flux pinning were studied for all samples and are discussed below.

Results and Discussion

Influence of MB2 and Dy2O3 doping on XRD and lattice constants

The X-ray diffraction data for all HTP bulks are presented in Fig. 1 where the Bragg reflections are indexed for only the MgB2 phase for simplicity. MgO and Mg were present at some level in all samples (MgO was less than about 2 wt% for samples reacted below the peritectic). MgB4 peaks are also present for samples HT at 1700 °C. Only a very small peak shift (less than 0.2 degree) at both (110) and (002) was observed in MB2 doped samples while no peak shift was observed for the Dy2O3 doped sample. Peaks corresponding to MBx impurity phases were observed in all MB2 doped samples. The lattice parameters extracted from pseudo-Voigt fitting the MgB2 peak reflections and the calculated lattice parameters by Vegard’s law are given in Table 1. For the Dy2O3 added sample HTP-DY, similar to Chen’s report17, the lattice parameters a and c did not change with Dy2O3 addition. Similarly, for the MB2 added samples, even though both lattice parameters a and c were slightly changed, these changes were very small and none of the doped samples obeyed Vegard’s law, unlike C-doped MgB2 HTP bulks6. Therefore it seems that even under HTP processing the metal borides (ZrB2, TiB2 and NbB2) mainly acted as impurity phases and did not form homogeneous solid solutions with MgB2, at least not to an extent detectable by XRD. However, the small changes in lattice parameters suggest that a distortion of the MgB2 lattice was present. Such distortion, caused by strain generated around these dopant impurities (see below) rather than elemental substitution, appears to be the driver for the modified upper critical field Bc2 of the doped samples.

Table 1 Doping levels, heat-treatment parameters, lattice parameters and MBx impurity amounts for the MgB2 samples.
Figure 1
figure 1

(a) X-ray diffraction characterization of undoped MgB2 sample HTP-01 and all theMB2 doped samples; (b) Peaks (110) and (002), as these two peaks are directly related to lattice parameter a and c, respectively. Note for the MB2 doped samples, small peak shifting and MB2 peaks are observed, while no peak shifting is observed in the Dy2O3 doped sample.

Influence of MB2 and Dy2O3 doping on microstructure - SEM and TEM

The results of back-scatter (BSE) SEM characterization performed on the ZrB2 doped samples are presented in Fig. 2. The microstructures of HTP-Zr-01 (1500 °C, below the peritectic temperature) and HTP-Zr-02 (1700 °C, above the peritectic temperature) are shown in Figs 2(a) and 3(b) (for comparison, the microstructures of the undoped bulk can be found in6,18,19). Two phases are visible: MgB2 (majority phase, dark grey) and ZrB2 (minority phase, white) in HTP-Zr-01; while in HTP-Zr-02, Mg and MgB4 phases were present since it experienced (upon cooling) the reaction .

Figure 2
figure 2

(a) BSE image of HTP-Zr-01; (b) BSE image of HTP-Zr-02; (c) BF TEM image of HTP-Zr-02 (Microstructures of the undoped bulk can be found in6,18,19). Insets are SAD of two distinct MgB2 grains; (d) nano-size inclusions observed close to/at MgB2 grain boundaries; (e) EDS spectra of spot A, B and C; (f) HAADF image of a nano-size inclusion close to MgB2 grain boundaries. Inset is the intensity of Zr from STEM-EDS scanning across the inclusion (red dash line).

Figure 3
figure 3

(a) BSE image of HTP-Ti-01; (b) BSE image of HTP-Ti-02; (c) BF TEM image of HTP-Ti-01. Large amount of crystal defects in MgB2 grain close to MgB2/TiB2 interface; (d) DF TEM image of ingrain crystal defects in MgB2 grain close to TiB2; (e) HAADF image of MgB2 grains with high defect density and STEM-EDS scanning across one of the nano-inclusions (red dash line). Inset is the intensity of Ti; (f) EDS spectra of spot A-E from (c,e).

Figure 2(c) represents the bright-field (BF) TEM image obtained from a thin foil extracted from HTP-Zr-02 (1700 °C). This foil contains a cross section of several grains. The results from the energy dispersive spectroscopy analysis (EDS) and selected area diffraction (SAD) confirm that these large grains are MgB2. An intragranular crack and dislocation loops are present in one of the grains. Since HTP-Zr-02 was processed above the peritectic, the crack presumably resulted from volume expansion taking place during cooling as the MgB4 converted into MgB2. TEM examination revealed a number of impurity phases in the form of 30–80 nm inclusions around the MgB2 grains, Fig. 2(d). The results of EDS analysis performed on these inclusions are presented in Fig. 2(e). It is clear that these inclusions contain Zr, or possibly ZrB2, which is likely dispersed around the MgB2 grains in the bulk. Compared to other areas inside the MgB2 grains, the regions around these ZrB2 inclusions have much higher contrast under BF condition, which suggests that strain fields were generated around these inclusions and that the MgB2 lattice was distorted locally. This local lattice distortion may be the origin of the slight lattice parameter changes observed by XRD analysis in Section 3.1. Figure 2(f) shows HAADF imaging for one of these nano-size inclusions. Since HAADF imaging is sensitive to variations in the atomic number (Z-contrast), these white inclusions should have a higher average atomic number than MgB2. A STEM-EDS line scan applied using 21 distinct points over ~100 nm across this inclusion confirmed it was ZrB2. The Zr signal dropped to zero quickly outside of the inclusion, beyond the ZrB2/MgB2 interface. The spatial resolution of the STEM-EDS line scans is about 5 nm, thus these observation indicate that Zr did not notably penetrate into the MgB2 lattice.

Figure 3 shows the BSE images of the TiB2 doped samples, HTP-Ti-01 (1500 °C, below the peritectic) and HTP-Ti-02 (1700 °C, above the peritectic). Similar to the behavior of ZrB2, TiB2 mainly acts as an impurity phase (light grey in Fig. 4(a,b)) and is widely distributed. A TEM thin foil containing a cross section of both TiB2 and MgB2 grains was carefully extracted from HTP-Ti-01. A BF image including MgB2, TiB2 and their interface is represented in Fig. 3(c). A large number of defects can be observed inside the MgB2 grains close to the MgB2/TiB2 interface. Inclusions 100–200 nm in size are found at the interface as well as at MgB2 grain boundaries. EDS analysis confirms that these inclusions are MgO, Fig. 3(f). Dark-field (DF) imaging was also used to examine the dislocations and the interface since crystal defects have stronger contrast under DF conditions. In Fig. 3(d), the DF image clearly confirms that the MgB2 grain contains a high density of defects. Detailed in-grain analysis was performed using HAADF imaging. Figure 3(e) shows a HAADF image of MgB2 grains with high defect density. Nano-size inclusions (~10–30 nm, white) dispersed both in and around MgB2 grains were observed. EDS analysis performed on randomly selected white inclusions confirmed that they were TiB2 (EDS spectrum of spot E (inclusion) in Fig. 3(f)). No Ti was detected by EDS in the other regions of MgB2 grains (EDS spectrum of spot D (matrix) in Fig. 3(f)). STEM-EDS analysis was applied across inclusion E and beyond the TiB2/MgB2 interface the intensity of the Ti signal quickly dropped to zero, indicating that Ti did not dissolve into the MgB2 lattice. These nano-size TiB2 inclusions can also contribute to the high defect density observed in MgB2 grains in Fig. 3(c,d).

Figure 4
figure 4

(a) BSE image of HTP-Nb-01; (b) fractured SE images, insets are BSE image of the fractured area; (c) BF TEM image of HTP-Nb-01 contained NbB2 inclusions. Inset is the DF TEM image of NbB2 inclusions; (d) HAADF image of MgB2 grains with high density of defects; (e) STEM-EDS scanning across one of the nano-inclusions (red dash line). Inset is the intensity of Nb ; (f) EDS spectra of spot A-D.

In the NbB2-added sample HTP-Nb-01 (HT below the peritectic), three phases are visible in the BSE images of Figs 4(a) and 5(b): MgB2 (majority phase, dark grey), MgO (minority phase, light grey) and NbB2 (minority phase, white). In Fig. 4(b), based on fractured secondary electron (SE) imaging by ‘through the lens’ (TTL) detection, NbB2 particles are observed outside the MgB2 grains. These particles are small (~300–500 nm), well connected with the MgB2 grains and dispersed throughout the bulks.

Figure 5
figure 5

(a) BSE image of HTP-DY, four different phases (MgB2, MgB4, Mg and MgO) are labeled; (b) BF TEM image of nano-size inclusions (10–50 nm) are found inside MgB2 grains; (c) HAADF image of one MgB2 grain, white precipitates are Dy-contained; (d) EDS spectra of spot A-C in (c); (e) the intensity of Dy from STEM-EDS scanning (red dash line in (c)).

Further analysis was performed on a TEM thin foil sectioned from HTP-Nb-01. Both BF (Fig. 4(c)) and DF images (Inset of Fig. 4(c)) show nano-size inclusions (~300 nm) embedded in the MgB2 grain boundaries. Moreover, a large number of defects can be observed inside the MgB2 grains around these inclusions, while the other MgB2 grains have fewer intragranular defects. The EDS results in Fig. 4(e) confirm that these inclusions are NbB2. HAADF imaging performed on MgB2 grains with high density of defects is presented in Figs 4(d) and 5(e). Nano-size inclusions (~10–50 nm, white) were found inside these grains and high strain fields were observed around them. EDS analysis was applied on these distinct inclusions and several randomly selected spots in the matrix; those for spot C (matrix) and spot D (inclusion) are presented in Fig. 4(f). These EDS results confirm that these white inclusions were NbB2. A STEM-EDS line scan was applied across inclusion D. The intensity of Nb signals abruptly decreased from ~104 to 0 across the NbB2/MgB2 interface and no Nb was detected in other regions of the MgB2 grains.

The microstructure of HTP-DY (HT above the peritectic) was investigated by BSE in Fig. 5(a). Five phases are visible: MgB2 (majority phase, dark grey), Mg (main phase, grey), MgB4 (minority phase, black), MgO (minority phase, light grey) and Dy-containing inclusions (minority phase, white). These Dy- containing inclusions (DyB4 according to XRD results) with a size of ~100 nm were dispersed throughout the bulk. Bright-field TEM examination revealed a number of impurity phases in the form of ~10–50 nm inclusions inside the MgB2 grains in Fig. 5(b). A low density of large inclusions (over 100 nm) was also observed (Fig. 5(c)). HAADF imaging (Fig. 5(c)) showed that these nano-size inclusions had a higher average atomic weight. EDS indicates that these inclusions contained Dy and B suggesting they are the previously XRD-identified DyB4. STEM-EDS analysis was applied across inclusion B (the red dashed line in Fig. 5(c)) and the result is shown in Fig. 5(e). No Dy was detected outside the inclusion.

Influence of MB2 and Dy2O3 doping on superconducting properties - Magnetic Results

The superconducting transition temperature Tc and the distribution of Tc of all samples were extracted by magnetization measurements, Fig. 6. The onset Tcs and the full-width half maximum (FWHM) of all samples are listed in Table 2. The undoped sample HTP-01 shows a very sharp superconducting transition with Tc of 39.5 K and a FWHM of ~0.4 K. Below we describe the results for the doped samples.

Table 2 Comparison of the superconducting properties amount the MgB2 samples.
Figure 6
figure 6

(a) DC susceptibility χ vs T at 0.01 T; (b) the Tc distribution - dχ/dT vs T; and (c) the temperature dependent Bc2 (T) curves of the all MgB2 samples.

ZrB2 Doping

In the ZrB2 doped samples, the onset Tcs is ~39.2 K in HTP-Zr-01 and ~39.4 K in HTP-Zr-02, respectively. The unchanged Tcs suggest that a portion of these ZrB2 doped samples was unaffected with a Tc equal to that of the undoped sample. Figures 6(a,b), show very broad transitions and bi-modal peak in the Tc distribution. This effect became more severe in HTP-Zr-02 indicating the presence of regions with various Tcs.

TiB2 Doping

Similarly, in the TiB2 doped samples, the onset Tcs were unchanged, while their FWHMs were increased to ~0.5 K in HTP-Ti-01 and ~2.8 K in HTP-Ti-02, respectively.

NbB2 Doping

The NbB2 doped sample HTP-Nb-01 has an onset Tc of 39.8 K and a FWHM of ~1.0 K.

Dy2O3 Doping

The Dy2O3 doped sample HTP-DY has an onset Tc of 39.2 K and a FWHM of ~0.5 K.

The results for each of the MB2 samples was similar-after MB2 doping, the onset Tcs were relatively unaffected, however their transition widths were significantly enhanced. We interpret this effect in terms of the presence of nanoscale MB2 (where M = Zr, Nb, or Ti) second phases which produce locally distorted regions separated by large regions of unaffected MgB2, leading to broadened Tc distributions with a wide Tc variation ranging from 39 K to low values. Since these MB2 additives are isomorphous to MgB2 and their lattice parameters are close to those of MgB2, the localized distortion is probably due to the coherent strain generated around the MB2 inclusions. However, in HTP-DY both the onset Tc and FWHM did not change by adding Dy2O3, which is consistent with Chen’s observation17. The lattice parameters and crystal structures of Dy2O3 (cubic with space group Ia-3)20 and DyB4 (tetragonal with space group P4/mbm)21 are very different from MgB2 (hexagonal with space group P6/mmm)22, therefore it is unlikely that the Dy-contained inclusions in HTP-DY can generate coherent strain in the MgB2 grains.

As indicated above the MB2 dopants were mostly found as distinct impurity inclusions that only influenced the surrounding MgB2 grains through the MgB2/MB2 interfaces. Increasing the concentration of MB2 inclusions produced more “affected zones” leading to a wider Tc distribution, Fig. 6(b). The behaviors of MB2 doped samples are quite different from those of C-doped MgB2 bulks6. After doping with 6.2 at.% C Susner et al.6 observed a significant decrease in the onset Tc , from 39.5 K to ~24 K, while the FWHM changed from 0.65 K to 1.4 K6. Since C is known to be a substitutional defect, if homogeneous C doping is achieved, the onset Tc and the lattice parameter a will decrease simultaneously with increasing C doping levels6. Under MB2 doping, it seems that Zr, Ti and Nb did not substitute for Mg or form homogeneous solid solutions with MgB2, even under 1700 °C and 10 MPa. However, the properties of the host lattices in the vicinities of these dopants were indeed affected and their Tcs were clearly altered, possibly due to local compositional changes caused either by Mg diffusion into MB2 particles or by local strain. Both of these possibilities could cause Tc reduction, comparable to the effect of Al doping in MgB223,24,25. Based on the results in the previous section, the affected vicinities probably had thicknesses similar to or smaller than 5 nm-the resolution of STEM-EDS line scans used in this study.

The T dependencies of the upper critical fields Bc2 of all the samples are presented in Fig. 6(c). The Bc2s at 20 K linearly extrapolated from Fig. 6(c) are listed in Table 2. It is clear that Bc2 was increased by MB2 doping, but not by Dy2O3 doping. It is important to note, however, that since these MB2 doped samples were not homogeneous (as evidenced by the microstructure and the Tc distribution), the Bc2 values represent the properties of only a small fraction of the bulk samples. In other words, some “affected zones” inside these doped bulks have higher Bc2s than those in the unaffected MgB2, therefore their measured Bc2 was enhanced. The observed high defect densities in the MgB2 grains, which increase electron scattering and reduce the electron mean free path, are likely responsible for the Bc2 enhancement. These regions are at the edge of the grains and can therefore act as connected percolative paths. In the Dy2O3 doped sample, although nano-size inclusions were observed inside the MgB2 grains, Bc2 did not change. This observation together with the absence of changes in the lattice parameters and lack of change in the Tc and FWHM suggested that Dy2O3, unlike the metal diboride additions, did not cause a band of defect structure at the boundary of the MgB2 grain.

The magnetic critical current density Jcms and flux pinning behaviors of selected samples were calculated based on Bean’s critical state model:

where ΔM is the width of the hysteresis loop at a given field B, a and b are the edge lengths of the sample orthogonal to B (a > b). The results at 15 K are shown in Fig. 7. The Jcms for most of the field range were either not changed, or even reduced after MB2 doping; for HTP-DY, its Jcm was slightly increased at all measured fields. A “tail” in Jcm(B) can be observed in all MB2 doped samples, Fig. 7(a). Based on the microstructural evidence and results of Bc2 and Tc, this “tail” in Jcm of the MB2 doped samples is probably caused by regions in the samples with different Bc2s. However these regions were too small to have significant influence on the overall Jc (>100 A/cm2). The irreversibility field, Birr, defined as the point where flux pinning vanishes, is often taken as the field at which Jc (B) = 100 A/cm2. The results for these samples are given in Table 2; no increase in Birr is seen and in some cases there is a decrease. This definition of Birr does not capture the high field and super-low-Jc “tail” observed in Fig. 7(a).

Figure 7
figure 7

(a) Magnetic critical current density Jcm vs T; (b) Kramer plot Jc0.5B0.25 vs B; (c) The flux pinning density Fp vs B; and (d) Normalized flux pinning behavior fp vs b of the undoped sample HTP-01 and the doped samples HTP-Zr-01, HTP-Ti-01, HTP-Nb-01 and HTP-DY at 15 K.

MgB2 is primarily a grain boundary pinner and thus the starting place to describe its pinning is the Kramer function (although deviations are seen). In order to perform such analysis, a Kramer field is needed. A Kramer plot, Jc0.5B0.25 versus B, is shown in Fig. 7(b). The Kramer fields, Bk, taken at the cross-intercepts of linear fittings (Black dash lines in Fig. 7(b)) are listed in Table 2. The values of Bk, similarly to those of Birr, were stable or slightly reduced after MB2 doping, unlike the values of Bc2 which increased with doping. This effect is due to several factors: (1) the higher Bc2 region was apparently small, presumably restricted to the defected zones near the grain boundaries, these regions will not substantially influence the measured Bk; (2) the Bk is also affected by the sample connectivity, possibly reduced with second phases present. For samples with Dy2O3 additions, Bc2, Birr and Bk were not affected.

Figure 7 (c) is a plot of bulk pinning force density (Fp = Jcm × B) vs B; the maximum values, Fpmax, are listed in Table 2. Compared to the literature values, the Fp,max of these HTP processed bulks is quite small. Spark-plasma sintering26 gives a Jc at 2 T and 15 K of about 105 A/cm2, while our values here are closer to 2 × 103 A/cm2. The Fp,max of the undoped sample HTP-01 is only ~0.095 GN/m3 while according to Susner6 at 15 K the Fp,max of an undoped MgB2 wire was about 2 GN/m3. The main reason for the difference is that the MgB2 grains in these HTP bulks (>5 μm) are much larger than those in the traditional synthesized samples (typically 30–500 nm). Other factors could also contribute to this difference in Fp,max, including some reduction of connectivity by small amounts of MgO. However, as Jc ∝ 1/grain size and the grain size in our samples is roughly 50 times larger than the highest performing MgB2, the grain size effect is expected to be dominant. Among all samples, the highest value of Fp,max (~0.135 GN/m3) was observed in HTP-DY. The normalized bulk pinning force density is plotted against normalized magnetic field b = B/Bk in Fig. 7(d). The functions of grain boundary (GB) pinning from the Dew-Hughes model27, fp ∝ b1/2(1 − b)2, are also plotted for comparison. Although the undoped sample HTP-01 followed GB pinning function quite well, all doped samples show a deviation from the standard function. The peaks of fp in the doped samples were shifted from the value of b = 0.2 (the peak position of the GB pinning) to lower values. For example, the peaks in HTP-Zr-01, HTP-Ti-01, HTP-Nb-01 and HTP-DY were 0.12, 0.13, 0.11 and 0.10, respectively. This observation which was also reported by Matsushita et al.28 in C doped MgB2 bulk samples can be explained by two possibilities: (1) These doped samples might contain a set of local Bks instead of one distinct value (just like the Bc2s in the MB2 doped samples), which can lead to an artificial error in the estimation of the peak positions; while the variation in Bk of the undoped sample was small, thus the undoped sample followed the GB pinning function. (2) The deviation from bpeak = 0.2 might be caused by the operation of other pinning mechanisms (e.g., normal volume pinning in which fp maximizes at b → 0.0 27) in association with the GB pinning. As noted above, we clearly see a distribution of Dy-based second phases, consistent with17. Also present were modest levels of MgO, known both to act as a pinner and in some cases reduce connectivity29. However, any small MgO effects should be present in all samples. The parameter a of the flux line lattice (FLL) is given by a = 1.07(Φ0/B)1/2, where Φ0 is the quantum of magnetic flux (Φ0 = 2.07 × 10−15 Wb). Based on this expression, the values of a vary from ~50 nm at 1 T to ~20 nm at 6 T. By definition, the size of the volume pins needs to be larger than a. Considering the fact that these doped samples contained intragrain inclusions some of which were bigger than the FLL parameter a at every measured field, it appears that volume pinning contributed to these shifts in Fp,max.

In summary, after adding MB2, the Bc2 of MgB2 HTP bulks increased, the Tc distributions were broadened, but Tc, Bk and Jc remained unchanged (or slightly reduced). Considering the microstructural evidence, this observation can be explained as follows: only very small regions (possible ≤5 nm in thickness) around dopant particles of the MgB2 grains are influenced by doping, leaving the majority of MgB2 unaffected. To the contrary, the Dy2O3 doping did not change the Tc, Tc distribution and Bc2, instead it increased the Jc and flux pinning apparently associated with the nano-size precipitates in MgB2 grains.

Conclusion

In this work we have used our HTP method for synthesizing doped MgB2 bulks at high temperatures (up to 1700 °C) and at pressure (10 MPa) to explore solubility limits of dopant species in MgB2, maximize diffusion and (alternatively) attempt to form dense, nanoscale secondary phases during the sample synthesis. We explored both metal diborides (MB2, where M = Zr, Ti and Nb) for attempted Mg site substitution and Dy2O3 for nanoscale intragrain precipitate formation. Using the HTP process we conclusively show that the large increases in Bc2 with metal diboride additions are due to a highly defected band within the grain, rather than substitution or inclusion within the grain, or grain boundary effects. High defect densities observed in MgB2 grains around/with these MB2 inclusions, cause electron scattering and therefore contribute to the Bc2 enhancement and Tc distribution broadening. On the other hand, these regions (≤5 nm in thickness) were not large enough to significantly influence the high field Jc, Birr, or Bk. This observation explains the frequently observed increases seen for Bc2 in materials with no accompanying increase in transport current. We also confirm the previously observed but sparsely distributed intragrain precipitates formed with Dy2O3 additions. Dy2O3 additions not change the lattice parameters, Tc, Tc distribution and Bc2 of MgB2, but increased the Jc and flux pinning by forming an array of nano-size precipitates in MgB2 grains.

Methods

Sample Synthesis

Three sets of MgB2 bulks with various MB2 (M = Zr, Ti and Nb) dopants were fabricated at high temperatures and under pressure. This HTP process18,19 based on the reactive liquid Mg infiltration (Mg-RLI) method30. Three metal borides with a structure isomorphous to MgB2 (P6/mmm) were selected as vectors for Mg-site substitution: ZrB2 (99.5%, Alfa Aesar), TiB2 (99.5%, Alfa Aesar) and NbB2 (99.5%, Alfa Aesar). As the Dy2O3 additive, Dy2O3 (>99.9%, <100 nm particle size, ALDRICH) was used. Amorphous B powder (50–100 nm in size) manufactured by Specialty Metals Inc.31,32 was hand mixed with the dopant powder and high energy ball milled for 15 min in an Ar atmosphere. These dopants and B powder mixtures were then pressed into ~8 mm tall by ~13 mm diameter pellets and placed in an MgO crucible. Mg turnings (~4 mesh, 99.98%, Alfa Aesar) were packed on top. The Mg:B ratio in the crucible was about 1:1 to avoid possible Mg deficiencies during heat treatment. This crucible was capped and placed inside the HTP autoclave (see also18,19). All samples were heat treated at 10 MPa in an Ar atmosphere. Two heat treatment routes were used: (1) heating up to 1500 °C and soaking for 30 min; (2) heating up to 1700 °C and soaking for 20 min. A slow cooling rate of 5 °C/min was used in both HT routes to maintain thermal equilibrium. The first route was designed to limit the temperature to just below the peritectic decomposition point of the reaction, thus preventing decomposition while maximizing the diffusion of the dopant species. The second route was designed to allow the reaction to occur on the temperature upswing and hence to form MgB2 directly from MgB4 and Mg+ dopant species on cooling:

where Tp is the peritectic temperature (~1500 °C in our experiments).

Measurements

A Rigaku SmartLab X-ray diffractometer (using Cu Kα of 1.5406 Å) was used for structural characterization and the scanning angle 2θ ranged from 20° to 80°. A FEI/Philips Sirion scanning electron microscope (SEM) with a field-emission source and a through-the-lens (TTL) detector was used for microstructural imaging. An FEI Helios 600 dual beam focused ion beam instrument (FIB) with an Omniprobe micromanipulation tool was used to prepare TEM thin foils. The TEM imaging was performed on a FEI/Philips CM-200T transmission electron microscope (TEM) with a silicon drift detector (SDD) and energy-dispersive X-ray spectroscopy function (EDS). The high-angle annular dark-field imaging (HAADF) and EDS line scans with a resolution of ~5 nm were performed on a Tecnai F20 system field emission 200 kV scanning transmission electron microscope (STEM) with an X-TWIN lens and high brightness field emission electron gun (FEG).

The magnetic properties were measured by using a Quantum Design Model 6000 PPMS with 4.2 K < T < 300 K and −10 T < B < 10 T. The superconducting critical transition temperature Tc and Tc distribution were determined by DC magnetic susceptibility methods. The Tc was defined as the onset of superconductivity from the normal state at 10 mT and the Tc distribution was expressed in terms of dχ/dT, where χ is the DC susceptibility. M-T curves were taken at 1 T intervals from 0–14 T. The upper critical field, Bc2, was determined by the highest temperature point where the M-T curves deviated from M = 0 at each given field. The irreversibility field, Birr, was taken as the field at which Jc(B) = 100 A/cm2. The Kramer field, Bk, was taken as the point where Jc0.5B0.25 extrapolated to zero on a Kramer plot (Fig. 7(b)) and the bulk pinning force density, Fp, was calculated from Fp = JcB, where Jc was extracted from the magnetization results at various temperatures.

Additional Information

How to cite this article: Yang, Y. et al. Influence of Metal Diboride and Dy2O3 Additions on Microstructure and Properties of MgB2 Fabricated at High Temperatures and under Pressure. Sci. Rep. 6, 29306; doi: 10.1038/srep29306 (2016).