3.1. R–T results
Figure 4(a) shows the temperature (
T) dependence of electrical resistance (
R) under zero magnetic field, measured by the two-probe method. The upper panel shows that of a reference Nb sample which was mounted on the PPMS sample puck together with the PnM-Nb sample and was measured at the same time under the same condition. (The PPMS can measure three samples at the same time.) As shown, the reference sample undergoes the superconducting (SC) transition normally at the usual
for Nb (≈ 9 K) both for the first and second temperature cycles. Although there shows only the first and second cycles, the result is the same for all the rest of eight
R–T cycles. The standard deviation of the onset SC transition temperature for all the eight cycles is 9.0000±0.0006 K, indicating that during the whole cycles the thermometer of the PPMS was working properly and that the temperature profiles of
R–T results are accurate. Thus, whatever anomaly the PnM-Nb sample exhibits, any criticism arising from thermometry is invalid.
In the first cooling process (see bottom panel of
Figure 4(a)), the PnM-Nb sample undergoes the SC transition at 9 K, the usual
for Nb, and returns to the normal state at the same
in the subsequent warming process. During the first temperature cycle, the PnM-Nb sample thus exhibits usual SC properties. In the subsequent temperature cycle, by contrast, the sample begins to exhibit drastic changes. In the second cooling process (blue curve), the
R abruptly drops at 175 K. Because of the two-probe method adopted for this measurement, a tiny resistance residing in electrical pads (
) remains in the
R. The residual resistance aside, the resistance of the PnM-Nb sample part, i.e. Nb holey sheet, decreases by at least three orders of magnitude, giving evidence for the Nb holey sheet being superconducting below 175 K. In the subsequent warming process (red curve), the zero-resistance state keeps going through 175 K, and a finite resistance appears at 290 K.
One could, of course, speculate that the abrupt decrease in resistance at 175 K upon cooling is due to an electrical issue with measuring the sample. Indeed, weak points in the suspended holey structure, e.g. narrow bridge-like constrictions (see
Figure 2), might break after several thermal cycles because of thermal contractions. However, if some of the conducting paths between electrical pads abruptly open, the two-probe resistance would increase, not dropping to zero, as explained in Ref. [
13]. Additionally, such an electrical issue hardly explains why a finite resistance appears again in the subsequent warming process since that would imply that the supposedly opened conducting paths were connected again spontaneously by some means. Thus, such a scenario assuming electrical issues with measuring samples cannot explain either the zero resistance at high temperatures or the phase transition with thermal hysteresis; an alternative scenario is necessary.
Uninterruptedly,
R–T measurements were repeatedly performed. The result is shown in
Figure 4(b). The resistance drop, which was very sharp in the second temperature cycle, is obviously broadened by increasing the number of temperature cycles. The broadening of resistive transition is often observed when magnetic flux is involved in the SC transition, being typical for the so-called type-II superconductors. The more magnetic flux penetrating the sample, the wider the resistive transition becomes. Therefore, if the subject of this study is superconductivity, the experimental fact of resistive broadening indicates that the number of magnetic flux penetrating the sample is increased by repeating
R–T cycles.
The involvement of magnetic flux upon superconducting transition, despite the absence of an externally applied magnetic field, is a natural consequence of the theory of
hole superconductivity which describes the dynamics of superconducting transition. In Ref. [
14], the theory is applied to a wire conducting an electric current, and the dynamics of its self field is provided. Now, that situation is applied to this study, where an Nb holey sheet carries an excitation current for the
R–T measurement, as shown in
Figure 4(c). The theory of hole superconductivity predicts a significance of conduction through anions for the discovery of new high-
superconductors [
15], and the SEM-EDX image of the Nb holey sheet has proven that narrow bridge parts contain high amounts of oxygen (
Figure 3(b)). Hence, we focus on the conduction through narrow bridges and discuss its superconducting transition. In the normal state (upper panel), an excitation current flows parallel to a bridge, and it generates a self magnetic field in the interior of the bridge. Upon superconducting transition, the excitation current becomes a surface current that can only flow near the surface of a bridge (bottom panel). Because the motion of magnetic field lines is intimately tied to the motion of charge carriers (as proposed in the theory of hole superconductivity), the self field is expelled from the interior of a bridge. It is difficult to observe this effect clearly in conventional metals becoming superconducting, because they are aggregates of three-dimensional submicron grains, and the expelled flux would be averaged out among the grains with random size and random orientation. For a highly oriented system such as the one in this study, by contrast, the effect is significant, and the expelled flux would be trapped in the adjacent large void having the diameter of 20
m. By taking into account the heterogeneity of oxygen concentrations in bridges (see SEM-EDX image in
Figure 3), some of them would become superconducting and the others would not. In such a percolative superconducting network with low spatial symmetry, it is unlikely that the magnetic orientation of the trapped flux within the voids would be completely canceled out.
To conclude, each time the Nb holey sheet carrying current undergoes the superconducting transition of percolative nature, the self field is excluded from the body and almost concurrently trapped in the adjacent voids. In fact, even in the absence of an applied field, the width of its resistive transition broadens as the number of temperature cycles increases, as confirmed in
Figure 4(b). Further evidence for flux trapping in the holey sheet can be found in magnetic results shown in the next sections.
The superconducting transition, which evolves with magnetic flux, can be a first order phase transition showing hysteresis [
16,
17]. And the most important feature of such a resistive transition with thermal hysteresis is that a rise of resistance upon warming occurs
always at a
higher temperature than the drop of resistance upon cooling. The fact that the high-
transitions for
ALL the
R–T cycles shown in
Figure 4 share that key feature specific to a first-order phase transition validates this exotic kind of superconducting transition. The reason why warming is always beneficial to superconductivity of this kind can be found in the caption of
Figure 4(d), from Ref. [
18], using the sixth-order Ginzburg-Landau (GL) free energy density.
Figure 4.
R–Tresults and possible involvement of magnetic flux for the first order phase transition. (
a) Electrical resistance versus temperature of a reference Nb sample (upper panel) and that of a PnM-Nb sample (bottom panel) using the two-probe method, in the absence of an externally applied magnetic field. They were measured at the same time in the PPMS under the same condition. Only
R–T results for the first and second cycles are shown. (
b)
R–T results of the PnM-Nb sample for all the rest of eight cycles. (
c) Possible involvement of magnetic flux in the superconducting transition for an Nb holey sheet, based on the theory of
hole superconductivity. Center panel, a portion of an SEM image of the Nb holey sheet; a void with the diameter of approximately 20
m is surrounded by four narrow bridges. Upper and lower panel, a schematic normal bridge and a superconducting one with current streamlines and magnetic field lines (circles), respectively, from Ref. [
14]. Even in the absence of an externally applied magnetic field, a self field generated by an excitation current for measuring resistance is expelled from a bridge becoming superconducting. As the dynamics of magnetic flux is involved in the superconducting transition, it can be a first order phase transition. This effect is significant for a system wherein magnetic orientation of the expelled flux is not to be averaged out. For details, read text. (
d) Reason why warming is
always beneficial to the zero-resistance state presented in this study, based on the Ginzburg-Landau theory: the key feature of a first-order phase transition. Free energy density as a function of the order parameter, for various temperatures, is dictated by differing choices of
, as indicated in the legend, from Ref. [
18]. The highest temperature is for
and progresses downwards towards
. Note that hysteresis may arise because, on lowering the temperature from
to
the system may choose to remain in the
state. Only when the temperature is lowered to
(when
is no longer a local minimum), does the system have to transition to a state with non-zero
(we show the negative solutions only as square points for clarity), at which point the resistance would drop to zero. Upon increasing the temperature, a similar phenomenon may occur, whereby the non-zero
state remains, until finally
is reached, and then a transition has to occur to the
state with finite resistance.
Figure 4.
R–Tresults and possible involvement of magnetic flux for the first order phase transition. (
a) Electrical resistance versus temperature of a reference Nb sample (upper panel) and that of a PnM-Nb sample (bottom panel) using the two-probe method, in the absence of an externally applied magnetic field. They were measured at the same time in the PPMS under the same condition. Only
R–T results for the first and second cycles are shown. (
b)
R–T results of the PnM-Nb sample for all the rest of eight cycles. (
c) Possible involvement of magnetic flux in the superconducting transition for an Nb holey sheet, based on the theory of
hole superconductivity. Center panel, a portion of an SEM image of the Nb holey sheet; a void with the diameter of approximately 20
m is surrounded by four narrow bridges. Upper and lower panel, a schematic normal bridge and a superconducting one with current streamlines and magnetic field lines (circles), respectively, from Ref. [
14]. Even in the absence of an externally applied magnetic field, a self field generated by an excitation current for measuring resistance is expelled from a bridge becoming superconducting. As the dynamics of magnetic flux is involved in the superconducting transition, it can be a first order phase transition. This effect is significant for a system wherein magnetic orientation of the expelled flux is not to be averaged out. For details, read text. (
d) Reason why warming is
always beneficial to the zero-resistance state presented in this study, based on the Ginzburg-Landau theory: the key feature of a first-order phase transition. Free energy density as a function of the order parameter, for various temperatures, is dictated by differing choices of
, as indicated in the legend, from Ref. [
18]. The highest temperature is for
and progresses downwards towards
. Note that hysteresis may arise because, on lowering the temperature from
to
the system may choose to remain in the
state. Only when the temperature is lowered to
(when
is no longer a local minimum), does the system have to transition to a state with non-zero
(we show the negative solutions only as square points for clarity), at which point the resistance would drop to zero. Upon increasing the temperature, a similar phenomenon may occur, whereby the non-zero
state remains, until finally
is reached, and then a transition has to occur to the
state with finite resistance.
3.2. M–T results
After the eighth
R–T cycle, the sample chip was taken out of the PPMS, and its temperature (
T) dependence of magnetization (
M) was investigated using the MPMS under a perpendicular magnetic field of 1000 Oe, as shown in
Figure 5(a). First, a ZFC measurement was performed while warming the sample chip from 4.2 K to 300 K (orange). Subsequently, an FC measurement was performed while cooling it from 300 K to 5.1 K (purple). For comparison, another PnM-Nb sample chip, to which the
R–T procedure had not been applied yet and therefore that should not have zero resistance yet, was also measured (gray curves). Both sample chips have the same structural geometry. The only difference is whether the microprocessed part of the sample chip, i.e. Nb holey sheet, is in the anomalous zero-resistance state or not.
As the temperature was increased from 4.2 K, the
M of the as-fabricated sample (gray) increased from
emu to an approximate zero value of the order of
emu at 9 K, indicating an overall disappearance of superconducting diamagnetism of Nb constituting the sample. By contrast, the
M of the sample under study that contains the anomalous region (orange) does not reach zero at 9 K, remaining negative. As indicated by the
M–T curve above 9 K for the reference sample (gray), both diamagnetism of the bulk Si substrate and paramagnetism of the entire Nb film surrounding the microprocessed part (see
Figure 2) do not contribute to the magnitude of
M any more than the order of
emu. Hence, the anomalous behavior in
M above 9 K (orange) should be attributed to the microprocessed part with anomalous zero resistance. The magnitude of the remaining
M (
emu) is not trivial at all since the MPMS has the sensitivity of
emu. The remaining
M is approximately 50 times smaller than the diamagnetism below 9 K (
emu), because the area of the Nb holey sheet is only 0.3×0.5 mm
2, two-orders-of-magnitude smaller than that of the surrounding Nb film that yielded the perfect diamagnetism below 9 K.
As the temperature is increased from 9 K, the ZFC value (orange) decreases gradually from emu to more negative values. When considering the R–T result, the sample in this temperature range is in the superconducting (SC) state. Therefore, the M–T curve should show a flat temperature dependence if there were no magnetization other than SC diamagnetism. A possible reason for the discrepancy is flux trapping. As shown in a false-color SEM in the inset, there is a non-material part in the PnM sample—the void. Because of its large diameter of approximately 20 m, applied flux easily invades the void. Once it invades, it remains trapped and moves together with the sample. That is, the SQUID ring detects an extra magnetization in addition to the SC diamagnetism residing in the material part. Since the direction of the extra magnetization is parallel to the applied field, the value of the measured M increases, concealing the flatness of the temperature independent SC diamagnetism.
However, the extra magnetization owing to flux trapping is unfavorable to thermodynamic equilibration. As shown soon, the critical field for the PnM sample is very large. Because of its significantly large , the thermodynamic equilibrium state of this sample under the field of 1000 Oe during this M–T measurement is not the intermediate state but the perfect shielding state. For such a superconducting sample, the extra magnetization residing in the void is nothing but an unwanted source of thermodynamic nonequilibration. In other words, the extra magnetization owing to flux trapping decreases its magnitude as the temperature is increased, and the ZFC curve exhibits thus monotonically decreasing behavior.
As the temperature is increased further to 300 K, by contrast, the ZFC curve stops decreasing and alters its trend upward. The
M–T result in the temperature range of 170–300 K is enlarged on the upper panel of
Figure 5(b). The lower panel shows the eighth
R–T result in the same temperature range duplicated from
Figure 4(b). (The
R–T measurement was performed just before this
M–T measurement.) Precisely describing the complicated dynamics is difficult at this time. Yet it is remarkable that the
M measured in the ZFC “warming” process (orange) is flipping its trend upward in the temperature range of 250–270 K which is the same temperature range where the resistance
R started to rise in the
R–T “warming” process (orange). Based on the assumption that the appearance of
R is due to disappearance of superconductivity, it would be reasonable to conclude that the flip of
M across 250–270 K is the indication that the SC diamagnetism residing in the PnM material part is disappearing or at least weakening.
Uninterruptedly, the M–T measurement was continued while “cooling” the sample from 300 K, i.e., the FC measurement was performed. As the temperature is lowered, the FC value (purple) decreases. The lowering of M indicates that the applied magnetic flux is expelled from the interior of an examined specimen, and so far it cannot be accounted for by any physics other than superconductivity, the Meissner effect. The FC value reaches its minimum at the temperature below 230 K which is roughly consistent with the temperature range where the resistance R started to vanish in the R–T “cooling” process (purple).
The FC curve (purple) flips its trend upward below 210 K, which would be attributed to flux trapping in a non-material part, void. The flux trapping takes place after the PnM material part undergoes a superconducting transition. As the direction of the trapped flux is parallel to the applied field, the value of
M thus increases. Below 170 K, however, the FC curve exhibits a relatively flat temperature dependence in contrast to the ZFC curve (see orange and purple curves in
Figure 5(a)). For the ZFC measurement, the sample was already cold before the field was applied and, of course, did not know whether the field would be applied or not. Therefore, there was no choice for the sample other than to admit the extra magnetization in the voids. For the FC measurement, by contrast, the sample was cooled under the existence of the applied field. That is, there was a chance for the emergent superconducting screening currents to draw the most suitable geometric pattern that nullifies not only the field invading into the interior of the material part but also the other one trying to reside in voids. Hence, the thermodynamically unfavorable magnetization arising from voids was weakened during the FC measurement to the extent possible, and the FC curve exhibits thus relatively flat temperature dependence.
Finally, it is noteworthy that the value of
M still remains negative at 300 K, approximately
emu. Again, the magnitude is significant; such a large negative
M cannot be attributed to a bulk Si substrate or an Nb film covering the sample chip whose magnetization are only of the order of
emu. Magnetic impurity mixing as a source of the large negative
M is also impossible, because such a simple mixing cannot make the clear separation between the ZFC and FC curves observed below 170 K (see orange and purple curves in
Figure 5(a)). Hence, there is no alternative but to suppose that the observed large negative
M at 300 K is attributed to the SC diamagnetism of the Nb holey sheet. Because the minimum
M, approximately
emu, is reached at lower temperatures, it is reasonable to assume that not whole but just some part of the holey sheet is in the SC state at 300 K. This assumption is consistent with the percolative superconductivity discussed in the previous section, that the holey sheet as an anion network is composed of a mixture of superconducting and nonsuperconducting paths. It also explains why the resistance at 300 K was lowered from 28
to a several
after the manifestation of high-
superconductivity (
Figure 4(b)). Anyway, if really superconductivity survives at 300 K, then trapped flux should still be detectable at 300 K.
3.3. M–H results at 300 K
After the
M–T measurement, its magnetic field (
H) dependence of magnetization (
M) was measured at 300 K. As shown in the lower-left inset of
Figure 6, the
M–H measurement was started with the applied field increased from 0 Oe. It is remarkable that at
0 Oe the
M already shows a non-zero value, approximately
emu. After the FC measurement performed in advance of this
M–H measurement, the temperature was raised to 300 K with the applied field of 1000 Oe unchanged, then, after the temperature reached 300 K, the applied field was decreased to 0 Oe. Meanwhile, the decreasing magnetic field should reduce the magnitude of trapped flux in voids. Then, if really some part of the Nb holey sheet was still in the SC state, supercurrents flowing in the holey sheet should circulate around the voids in such an inductive way as to prevent the trapped flux from decreasing its magnitude. In other words, in a decreasing field a supercurrent circulating around a void must increase its magnitude. By taking into account the fact that such a circulating supercurrent consists of negative charge particles and by taking a look at the schematic illustration in the inset of
Figure 5(a), it can be found that the tangential direction of such circulating negative charge particles around the void is the same as that of negative charge particles circulating around the periphery of superconducting islands generating SC diamagnetism. That is, the SC diamagnetism is enhanced in a decreasing field, therefore, even when the field is decreased to zero, the value of
M still remains negative as shown at the beginning of this
M–H measurement.
As the virgin curve shows (lower-left inset of
Figure 6), the SC diamagnetism increases with
H increased. The low-field diamagnetism is a usual behavior of a superconductor, but the monotonic behavior is lasting even at
2000 Oe. As shown soon, the critical field
for the PnM sample is very large, and the 2000 Oe turns out to be indeed low enough when compared to the
.
When
H is subsequently decreased from 2000 Oe, the SC diamagnetism decreases, thus
M increases. It is remarkable that the hysteresis behavior begins at 1000 Oe that is the same magnitude of the magnetic filed applied to the PnM sample during the preceding
M–T measurement, implying that the flux indeed remain trapped in voids. The decreasing
H tries to reduce the magnitude of trapped flux in the array of voids. Due to the same physics explained above, in the decreasing field the SC diamagnetism is enhanced, thus, the measured
M becomes lower when
H is decreased versus when it is increased as can be seen in the lower-left inset. This is a universal feature of thermomagnetic hysteresis caused by flux trapping. For a good example, see Fig. 2 in Ref. [
19].
When
H is increased from
Oe and reaches 0 Oe (see main panel of
Figure 6), the PnM sample exhibits a positive
M, approximately
emu. This is a remnant magnetization. At 300 K and in the absence of an applied field, the time (
t) evolution of the remnant magnetization (
) was measured for 24 hours. The result is shown in the upper-right inset. At finite temperatures, thermal energy may allow flux lines to jump from one void to another in response to flux-density gradient, hence, there may be an observable decrease of the magnitude of trapped flux with time [
20]. Since any creep will relieve the gradient, the creep gets slower and slower. That is, the time dependence is logarithmic, and so is the observed
–t curve. Given that the driving force is proportional to the magnitude of trapped flux, the exponential dependence of the creep rate on the driving force is
which has the solution
, where the constants
and
can be obtained from the fitting curve shown in the figure. Then it is possible to estimate how long the trapped flux persists in the array of voids and hence how long the circulating supercurrents in the holey sheet will take to die out at 300 K. It will take
s
years.
When compared to the pioneering study for a tubular sample made of hard-superconductor NbZr measured at 4.2 K [
21] or to the recent study for a granular superconductor made of graphite powder measured at 300 K [
22], the lifetime is
or
times shorter. But still, when considering the thermodynamically unstable nature of trapped flux in the Nb holey sheet, which has a large
and hence prefers to remain the perfect shielding state, the lifetime of the unfavorable flux trapping is still long, long enough in any practical sense.
After the
–t measurement, namely 24 hours later, the
M–H measurement was performed again with
H increased from 0 Oe (red curve in
Figure 6). As shown, it retraces the initial hysteresis loop, indicating that the Nb holey sheet is indeed trapping magnetic flux. Whether or not it is due to superconductivity, the result evidences that the holey sheet sustains persistent current at 300 K. Of course, as of now, we do not know of any physics that makes this possible other than superconductivity.
3.4. Critical field and the Superconducting atom
As mentioned previously, the critical field
for the Nb holey sheet with zero resistance was investigated at 300 K. Under various perpendicular magnetic fields of
0, 5, 11 and 12 T, the output voltage was measured while applying current. The result is shown in
Figure 7. For
11 T, the voltage does not increase for applied currents lower than 1 mA. At
12 T, by contrast, a finite voltage is observed for a current exceeding a small value of 32.6 nA as shown in the inset. Hence, the intrinsic
at 300 K for the superconducting holey sheet without applied current will be a little bit larger than 12 T.
By taking into account the initial sharp resistance drop (gray curve in
Figure 4(b)), the Nb holey sheet is intrinsically the so-called type-I superconductor, therefore, its
corresponds to the thermodynamic critical field. As explained previously, the large thermodynamic critical field was responsible for the clear separation between the ZFC and FC curves in
Figure 5(a). Even if an unknown demagnetizing factor of the holey sample yields a 100 times larger magnitude than that of the actual field applied during the
M–T measurement (1000 Oe), the product is still lower than the
. That is, the perfect shielding state is indeed favorable for the superconducting holey sample during the
M–T measurement. In other words, flux trapping in the array of voids is indeed thermodynamically unfavorable. That’s why the ZFC curve exhibited the decreasing behavior as the temperature was increased. Also, whenever possible the flux trapping is suppressed as much as possible, that’s why the FC cooling curve exhibited the relatively flat temperature dependence. Nevertheless, once trapped under isothermal condition, it remains trapped for a sufficiently long time as proven in the previous section (
Figure 6).
By the way, the value of
12 T is much too large when considering the fact that the largest known critical fields for standard type-I superconductors are of order 0.05 T. According to the recent custom, such a potentially disruptive result is to be excluded from the science community [
23]. And the situation gets worse when such a result cannot be explained by an idol theory such as the BCS theory; unfortunately, that’s the case for this study. However, there is an alternative theory that explains the observed extraordinary value, although it was already forgotten a long time ago.
In 1937, following the logical argument raised by Fritz London [
24], John Slater immediately considered a magnetic susceptibility of an atom, and, in order for the atom to have perfect diamagnetism, he gave the minimum radius of the atom,
where
is the Bohr radius (0.529 Å) and the number 137 comes from the fine structure constant (
) [
25]. According to London and Slater, this is the element of any superconductor and cannot be separated into simpler units any further. It is approximately 137 times larger than the Bohr’s atom and has perfect diamagnetism with the magnitude corresponding to a single flux quantum,
(
Wb). The expanded atom is called the Slater’s atom.
Then, when applied flux invade the interior of a superconductor? It is when the applied flux are getting dense, so dense that the Slater’s atoms with the radius
having a diamagnetic
overlap each other. Hence the critical field can be given by
The experimentally confirmed extraordinary critical field, which will be a little bit larger than 12 T, is thus explained by the principle of the superconducting atom. This indicates that at least the superconductor presented in this study is composed of the superconducting atoms, as London and Slater expected. And what’s most remarkable is its simplicity; both Equations (
2) and (
3) do not include any single material parameter, implying that the pairing mechanism underlying this superconductivity as well as the forgotten theory are of pure physics. In other words, this study appears to be unrelated to BCS theory despite the initial motivation, because BCS superconductors do function only if material parameters such as an electron–phonon coupling constant
and a Coulomb pseudopotential
are empirically given [
1,
9].